astro-ph0412702/ms.tex
1: % revised version, 3/22/05
2: % copy-edited 1st version
3: 
4: %V11, 12/30/04
5: 
6: \documentclass[12pt,preprint]{aastex}
7: %\documentstyle[emulateapj]{article}
8: 
9: \usepackage{epsfig}
10: %\usepackage{psfig}
11: 
12: \received{2004 December 30}
13: \begin{document}
14: 
15: %% manuscript produces a one-column, double-spaced document:
16: %\documentclass[manuscript]{aastex}
17: %\documentstyle[11pt,apjpt4,psfig]{article}
18: %\documentstyle[11pt,aaspp4,psfig]{article}
19: 
20: \newcommand{\be}{\begin{equation}}
21: \newcommand{\ee}{\end{equation}}
22: \newcommand{\ba}{\begin{eqnarray}}
23: \newcommand{\ea}{\end{eqnarray}}
24: \newcommand{\siml}{\lower4pt \hbox{$\buildrel < \over \sim$}}
25: \newcommand{\simg}{\lower4pt \hbox{$\buildrel > \over \sim$}}
26: \newcommand{\Mesz}{M\'esz\'aros}
27: \newcommand{\epsd}{\epsilon_d}
28: \newcommand{\cm}{{\rm cm}}
29: \newcommand{\erg}{{\rm erg}}
30: \newcommand{\si}{{\rm s}^{-1}}
31: 
32: %\slugcomment{Submitted to The Astrophysical Journal}
33: 
34: %\shorttitle{}
35: %\shortauthors{}
36: 
37: 
38: 
39: \title{Dissipative Photosphere Models of \\ Gamma-ray Bursts and X-ray 
40: Flashes}
41: 
42: \author{M.J.~Rees\altaffilmark{1} and  P. \Mesz\altaffilmark{2,3}}
43: 
44: \altaffiltext{1}{Institute of Astronomy, Univ. of Cambridge, Madingley Rd, Cambridge CB30HA, U.K}
45: 
46: \altaffiltext{2}{Dpt.Astron.\&Astrophys., Dpt. Physics,
47: Pennsylvania State U., University Park, PA 16803}
48: 
49: \altaffiltext{3}{Institute for Advanced Study, Princeton, NJ 08540}
50: 
51: 
52: %\centerline {\date{Dec. 30, 2004} }
53: %\centerline {\date{\today} }
54: 
55: 
56: %\centerline {\bf Abstract}
57: 
58: \begin{abstract}
59: 
60: We consider dissipative effects occurring in the optically thick inner 
61: parts of the relativistic outflows producing gamma-ray bursts and X-ray 
62: flashes, emphasizing specially the Comptonization of the thermal radiation 
63: flux that is advected from the base of the outflow. Such dissipative 
64: effects --e.g. from magnetic reconnection, neutron decay or shocks -- 
65: would boost the energy density of the thermal radiation. The dissipation 
66: can lead to pair production, in which case the pairs create an effective 
67: photosphere further out than the usual baryonic one. In a slow dissipation
68: scenario, pair creation can be suppressed, and the effects are most
69: important when dissipation occurs below the baryonic photosphere.
70: In both cases an increased photospheric luminosity is obtained. 
71: We suggest that the spectral peak in gamma ray bursts is essentially 
72: due to the Comptonized thermal component from the photosphere, 
73: where the comoving optical depth in the outflow falls to unity. 
74: Typical peak photon energies range between those of classical bursts 
75: and X-ray flashes. The relationship between the observed photon peak 
76: energy and the luminosity depends on the details of the dissipation, 
77: but under plausible assumptions can resemble the observed correlations. 
78: 
79: \end{abstract}
80: 
81: \section{Introduction}
82: \label{sec:intro}
83: 
84: Most GRB models invoke a relativistic outflow, probably channeled into a 
85: jet, that is energized by a central compact object. The gamma-ray and hard 
86: X-ray emission is attributed to dissipative processes in the jet. The 
87: outflow is inferred to be unsteady, on timescales down to a millisecond -- 
88: indeed internal shocks are the most widely-discussed dissipative process, 
89: because of their ability to convert bulk kinetic energy into relativistic 
90: electrons which then radiate (e.g. via synchrotron emission) on very short 
91: timescales . The outflow would carry baryons, and also magnetic fields 
92: (which may carry as much power in Poynting flux as does the baryon kinetic 
93: energy). However, there is another inevitable ingredient of the outflow: 
94: thermal radiation. This radiation would originate near the base of the 
95: outflow, where densities are high enough to guarantee (at least 
96: approximate) thermal equilibration. This thermal radiation would be 
97: advected outward so long as the jet material remained opaque, and would 
98: emerge highly collimated, from a `photosphere' where the jet became 
99: optically thin. 
100: 
101: A laminar and steady jet, viewed head-on, would give rise 
102: to emission with a thermal spectrum peaking in the hard X-ray or gamma-ray 
103: band. Moreover, the comoving energy density of this black-body radiation 
104: could be at least comparable with that of the magnetic field. So, if 
105: dissipation generates relativistic electrons and supra-thermal pairs, their 
106: energy losses due to Compton scattering of the thermal radiation would be 
107: competitive with those from synchrotron emission -- perhaps even dominant. 
108: Consequently, when dissipation occurs (e.g. via internal shocks) the outcome 
109: may be a `hardened' (grey body) thermal component, along with a power-law 
110: component extending to higher photon energies. We suggest that the photon 
111: energy $E_{pk}$ at which GRB spectra reach a peak may be the (probably 
112: Comptonized) thermal peak. We discuss how, on this hypothesis, $E_{pk}$ 
113: would depend on the parameters characterizing the GRB. 
114: 
115: A key parameter in the outflow is plainly the photospheric radius -- 
116: the radius at which the comoving optical depth along the jet falls to 
117: unity. In calculating this radius, we must allow for the possibility 
118: that the electrons associated with the baryons are outnumbered by 
119: electron-positron pairs (e.g. Eichler \& Levinson, 2000). Moreover,
120: the number of pairs may be greatly increased by dissipative processes . 
121: The details depend on whether the photosphere  lies inside 
122: or outside the saturation radius at which the bulk 
123: Lorentz factor $\Gamma$ asymptotes to the dimensionless entropy $\eta 
124: =L_0/({\dot M} c^2)$, where $L_0$ and ${\dot M}$ are the total energy and 
125: mass outflow rates. For a spherical flow where the free energy emanates 
126: from a central region $r_0 \sim \alpha r_g =\alpha 2GM/c^2$ comparable to 
127: the Schwarzschild radius $r_g$ of a central object of mass $M$ (where 
128: $\alpha\geq 1$), the bulk Lorentz factor grows as $\Gamma(r)\sim r/r_0$ 
129: outside of $r_0$ up to a saturation radius $r_s\sim r_0\eta$, where it 
130: saturates at a value $\Gamma\sim\eta$. This simple behavior applies 
131: for a spherical outflow (or a conical one with jet opening half-angle 
132: $\theta_j <\Gamma^{-1}$) where there are no internal shocks. We focus 
133: on this as an illustrative case (bearing in mind that the effective value 
134: of $r_0$ may be increased by dissipation in the inner jet). Moreover, 
135: extensions to the cases of convergent or divergent jets are straightforward. 
136: Inside the saturation radius, the observer-frame photospheric luminosity 
137: $L_\gamma$ is approximately the total luminosity of the outflow $L_0$, 
138: since the increasing Doppler boost just cancels the adiabatic decay of the 
139: comoving characteristic photon energy. On the other hand, if the 
140: photosphere of an adiabatic flow occurs outside the saturation radius, 
141: $r>r_s$, the Lorentz factor no longer grows, and $L_\gamma 
142: (r)=L_0(r/r_s)^{-2/3} < L_0$, the greater part of the energy being in 
143: kinetic form, $L_k\sim L_0$ (e.g. \Mesz \& Rees, 2000). If this 
144: photospheric luminosity were the bulk of the observed radiation, the 
145: radiative efficiency would be low in the latter case.
146: 
147: However, the above scenario can change substantially due to dissipative 
148: effects such as magnetic reconnection (e.g. Thompson, 1994, Giannos \& 
149: Spruit, 2004), neutron decay (e.g. Beloborodov, 2003), or internal shocks. 
150: If the dissipation occurs below the photosphere, the adiabatic decrease of 
151: the radiative luminosity beyond the saturation radius can be compensated by 
152: reconversion of some fraction $\epsd \leq 1$ of the kinetic energy into 
153: radiation, which would re-energize the photospheric component. Moreover, 
154: dissipation outside the nominal photosphere may lead to sufficient pair 
155: formation to create a second photosphere, lying outside the original 
156: nominal photosphere which would have obtained in the absence of 
157: dissipation.
158: 
159: Thus, if there were sub-photospheric dissipation , the observable 
160: photospheric luminosity would be boosted by the energy recovered from the 
161: kinetic energy, which becomes available for converting into radiation or 
162: pairs. Moreover the dissipated energy would go mainly into Comptonized 
163: of the thermal radiation advected out from the central engine. Above $r_s$, 
164: the photospheric luminosity can be boosted to a value $L_\gamma=\epsd L_0 > 
165: L_0(r/r_s)^{-2/3}$, depending on the dissipation efficiency. We suggest 
166: that the peak energy of the photon spectrum of gamma ray bursts should be 
167: identified with the peak of this Comptonized spectrum.
168: 
169: \section{Photospheres, dissipation and pairs} \label{sec:phot}
170: 
171: In the dissipation regions of the flow, all suprathermal or relativistic 
172: electrons and pairs will lose energy by Compton scattering of the thermal 
173: radiation (which will be roughly isotropic in the comoving frame). 
174: Synchrotron losses might dominate for high $\gamma$ electrons, but for 
175: those with modest $\gamma$ , synchrotron emission will be inhibited by 
176: self-absorption; these will lose their energy primarily by Compton 
177: scattering even if the magnetic energy density exceeds that of the thermal 
178: radiation. They will cool down and thermalize in a time short compared to 
179: the dynamic time.
180: 
181: Relativistic electrons moving through black-body radiation Compton-boost
182: each scattered photon by $\gamma^2$, producing a power law rather than
183: just boosting each photon by a small amount. However, if the slope of
184: the injected power law is steeper than $-2$, most of the energy will be
185: at the low energy end, and all the energy of electrons with, say,
186: $\gamma \lesssim 3$ would go into what would look like a broad thermal
187: peak. They would emit no synchrotron radiation (because of
188: self-absorption) and they would not boost any of the thermal
189: photons by more than a factor $\epsilon_c \sim 10$.
190: 
191:  If the primary dissipation were mainly by strong shocks most of the energy 
192: might be channeled initially into very high-gamma electrons, which would 
193: produce photons with a power law spectrum extending to very high energies: 
194: production of photons above 1 Mev in the comoving frame would only require 
195: $\gamma\sim 10 -30$ for Compton scattering, and little more than $10^3$ for 
196: synchrotron emission. However pair production can change this situation, 
197: leading again to a situation where energy is ultimately dissipated via 
198: thermal Comptonization. If the compactness parameter is more than unity 
199: (which, as shown below, is often the case for the usual parameters 
200: considered) most of the energy in photons with $>$ MeV energies in the 
201: comoving frame will be converted into pairs with very modest $\gamma$. 
202: These pairs will then lose their energy (as described above) by Compton 
203: cooling, resulting in a quasi-thermal spectrum, whose characteristic peak 
204: is a factor $\epsilon_c \siml 10$ above the original thermal peak, i.e. in 
205: the tens to hundreds of keV. These pairs establish effectively a new 
206: photosphere, outside the one that would have been present in their 
207: absence, and the dissipation (or shocks) responsible for these pairs will 
208: effectively be a sub-photospheric dissipation, which has different 
209: characteristics from the more familiar shocks that occur well outside the 
210: photosphere (e.g. Ghisellini \& Celotti, 1999; Kobayashi, Ryde \& 
211: MacFadyen, 2002; Pe'er \& Waxman, 2004).
212: 
213: For a GRB outflow of radiative luminosity $L$ and bulk Lorentz factor 
214: $\Gamma$ in the observer frame, at a radius $r$ the comoving scattering 
215: opacity due to $e^\pm$ pairs in the high comoving compactness regime is 
216: \be 
217: \tau'_\pm \sim \ell'^{1/2} \sim (L \sigma_T/4\pi m_e c^3 \Gamma^3 r)^{1/2}, 
218: \label{eq:pairopacity} 
219: \ee 
220: where $\ell'$ is the comoving frame compactness 
221: parameter (e.g. Pe'er \& Waxman, 2004). Here we have approximated $L(>1{\rm 
222: MeV})\sim \epsd L_0$, where $L_0$ is the total luminosity in the observer 
223: frame, and we have taken other efficiency factors to be of order unity. The 
224: functional dependence of equation (\ref{eq:pairopacity}) can be obtained 
225: by considering in the comoving frame (primed quantities, as opposed to 
226: unprimed quantities in the observer frame) the balance between the rate at 
227: which pairs annihilate and the rate at which pairs are formed. The latter is 
228: the rate at which photons capable of pair-producing are introduced into the 
229: flow, i.e. the photon density above $m_e c^2$ divided by the comoving dynamic 
230: time, ${n'}_{\pm}^2 \sigma_T c \sim (L/4\pi r^2 m_e c^3 
231: \Gamma^2)(c\Gamma/r)$, from which follows the pair optical depth $\tau'_\pm 
232: \sim n'_\pm \sigma_T (r/\Gamma )$. The pair photosphere $r_{ph,\pm}$ is the 
233: radius where $\tau'_\pm \sim 1$, or 
234: \be 
235: r_{ph,\pm} \sim (\epsd /2 \alpha) (m_p/m_e)(L_0 /L_E)\Gamma^{-3} r_0
236:    \sim 2 \times 10^{14} L_{51}\epsilon_{d,-1}\alpha^{-1}\Gamma_2^{-3}~\cm
237: \label{eq:pairphot}
238: \ee
239: where $L_E=4\pi GM m_p c/\sigma_T\simeq 1.25\times 10^{39} m_1 \erg\si$
240: is the Eddington luminosity, $r_0=\alpha r_g$ where $\alpha \geq 1$ and
241: $r_g=2GM/c^2 \simeq 3\times 10^6 m_1~\cm$ is the Schwarzschild radius for
242: a central object (e.g. black hole) of mass $M\sim 10 m_1$ solar masses,
243: and $\eta=L/({\dot M}c^2$ is the dimensionless entropy of the relativistic
244: outflow.
245: 
246: On the other hand, the scattering opacity due to the ordinary electrons 
247: associated with baryons in the flow would give rise to a 'baryonic 
248: photosphere' at \be r_{ph,b}\sim (1/2\alpha)(L_0/L_E) \eta^{-1} \Gamma^{-2} 
249: r_0
250:         \sim 1.2 \times 
251: 10^{12}L_{51}\alpha^{-1}\eta_2^{-1}\Gamma_2^{-2}~\cm~, 
252: \label{eq:baryonphot} \ee with the same notation as above.
253: 
254: For an outflow which starts at $r_0 = \alpha r_g =
255: 3\times 10^6 \alpha m_1$ cm, where $\alpha \geq 1$ and the initial
256: Lorentz factor $\Gamma_0 \sim 1$, under adiabatic conditions
257: energy-momentum conservation leads  to a Lorentz factor which grows
258: linearly as $\Gamma(r) \propto r/r_0$ until it reaches a saturation
259: radius $r_s \simeq r_0\eta \simeq 3\times 10^8 \alpha m_1 \eta_2~\cm$,
260: beyond which the Lorentz factor saturates to $\Gamma\simeq \eta =$
261: constant. One can then define two critical limiting Lorentz factors
262: \ba
263: \eta_b= &\bigl( \frac{1}{2\alpha}\frac{L}{L_E}\bigr)^{1/4}=
264:        7.9\times 10^2(L_{51}/\alpha m_1)^{1/4} \cr
265: \eta_\pm=& \bigl(\epsd \frac{m_p}{m_e}\bigr)^{1/4} ~\eta_b=
266:        2.9\times 10^3 (L_{51}\epsilon_{d,-1}/\alpha m_1)^{1/4}
267: \label{eq:etacrit}
268: \ea
269: which characterize the behavior of the baryon and pair photospheres below
270: and above the saturation radius. The pair photosphere behaves as
271: \be
272: r_{ph,\pm}/r_0=\cases{
273:   \eta_\pm    & for $r<r_s$ \cr
274:   \eta_\pm (\eta/\eta_\pm)^{-3} & for $r>r_s$}
275: \label{eq:pairphot_eta}
276: \ee
277: and the pair photosphere occurs at $r<r_s$ for
278: $\eta > \eta_\pm$. The baryon photosphere behaves as
279: \be
280: r_{ph,b}/r_0 = \cases{
281:  \eta_b (\eta /\eta_b)^{-1/3} & for $r<r_s$ \cr
282:  \eta_b (\eta/\eta_b)^{-3}    & for $r>r_s$ }
283: \label{eq:barphot_eta}
284: \ee
285: and the baryon photosphere occurs at $r< r_s$ for
286: $\eta > \eta_b$. This is shown schematically in Figure 1,
287: %Fig. \ref{fig:photeta},
288: for values of $\alpha =1,~10^4$, i.e.  initial radii
289: $r_0=3\times 10^6 \alpha_0 m_1 \cm$ and
290: $r_0= 3\times 10^{10}\alpha_4 m_1 \cm$.
291: 
292: The pair photosphere (equation [\ref{eq:pairphot}]) will be above
293: the baryon photosphere provided that
294: \be
295: \epsd > (m_e/m_p)
296: \label{eq:etapairabovebar}
297: \ee
298: where $\epsd$ characterizes the dissipation efficiency producing
299: photons above energies $m_e c^2$ in the comoving frame.
300: 
301: \section{Characteristic Photon Luminosities and Temperatures}
302: \label{sec:lum}
303: 
304: When the conditions of equation (\ref{eq:etapairabovebar}) are satisfied, 
305: the real (outermost) photosphere is not the baryon photosphere but the pair 
306: photosphere. The pair photosphere will have a luminosity 
307: $L_{\gamma\pm}=\epsd L_0\leq L_0$. At $r<r_s$, the observed radiation is 
308: insensitive to the actual details of the photosphere: the decrease with $r$ 
309: in the comoving-frame luminosity is compensated by the observer-frame boost 
310: given by the increasing Lorentz factor $\Gamma$; moreover, there is less 
311: scope for dissipation (except in the case when Poynting flux far exceeds 
312: the radiative flux in the jet).
313: 
314: On the other hand, for a photosphere at $r>r_s$ the luminosity 
315: decays as $L_{\gamma } = L_0(r/r_s)^{-2/3}$ in the adiabatic regime. 
316: However, if dissipation  occurs  above $r_s$, this leads to an
317: effective luminosity
318: \be L_{\gamma} \sim \epsd L_0~. 
319: \label{eq:Lpm} \ee 
320: This luminosity is achieved at the baryon photosphere if dissipation 
321: occurs below this radius, even in the absence of significant pair formation, 
322: or above the baryon photosphere if dissipation above the baryon photosphere
323: leads to a pair photosphere radius $r_\pm$ such that $\epsd (r_\pm /r_s)^{2/3}
324: \geq 1$ (see Fig. 2). In such cases the effective photosphere luminosity 
325: exceeds what would have emerged from a non-dissipative outflow by 
326: $\epsd (r/r_s)^{2/3}$. 
327: 
328: The characteristic initial temperature of the fireball outflow is
329: \be
330: T_0=(L_0/4\pi r_0^2 c a)^{1/4}
331:    = 1.2 L_{51}^{1/4}(\alpha m_1)^{-1/2}~{\rm MeV}~,
332: \label{eq:T0}
333: \ee
334: which for a larger $\alpha=10^4$ (i.e. for a larger initial radius
335: $r_0=3\times 10^{10}\alpha_4 m_1$ cm) would be
336: $T_0=12.1~L_{51}^{1/4}(\alpha_4 m_1)^{-1/2}~{\rm keV}$.
337: 
338: For $r<r_s$ the observer-frame effective photospheric temperature (even in 
339: the presence of dissipation) remains as $T_\gamma =T_0$, being boosted by 
340: the growing Lorentz factor back to its initial value. For $r>r_s$, 
341: adiabatic effects (in the absence of dissipation) would cause the 
342: temperature to fall off as $T_\gamma=T_0(r_s/r)^{-2/3}$. 
343: Since, however, dissipation leads to a luminosity $\epsd L_0$ 
344: which can exceed the adiabatic value, this results in a temperature 
345: $T_{\gamma}$ which drops $\propto r^{-1/2}$.  If dissipation is 
346: maintained all the way to the (baryonic or pair-dominated) 
347: photosphere, the temperature is 
348: \be 
349: T_{\gamma,d}=\epsilon_c \epsilon_d^{1/4}(r/r_s)^{-1/2} T_0~, 
350: \label{eq:Tpm} 
351: \ee 
352: where a factor $\epsilon_c \simg 1$ accounts for possible departures 
353: from a black-body. This temperature is larger by a factor 
354: $\epsilon_c\epsilon_d^{1/4}(r/r_s)^{1/6}$ than the adiabatic 
355: photosphere temperature $T_\gamma$ (Fig. 2).
356: 
357: Thus, one consequence of dissipation is that, even for $\alpha=1$,
358: e.g. with $\epsd=10^{-1}$ and $r_d/r_s=10^2$, the  characteristic
359: temperatures can be  $kT_\gamma \sim 60$ keV, i.e. peak photon energies 
360: in the X-ray flash range, while for $\alpha=10^4$ this energy can 
361: easily be as low as a few keV.
362: 
363: If dissipation is important only for some range of radii,
364: starting at $r_s \epsd^{3/2}$ but ceasing, say, at some radius
365: $r_c$ below the photosphere $r_{ph}$, then the adiabatic decay
366: $L_\gamma \propto T_\gamma \propto r^{-2/3}$ resumes above $r_c$ until
367: $r_{ph}$ so the photospheric luminosity would be
368: lower than implied by eqs. (\ref{eq:Lpm},\ref{eq:Tpm}).
369: 
370: We should note that, in the present context, $r_0$ is essentially the 
371: radius beyond which the Lorentz factor starts to grow as $\Gamma \propto 
372: r/r_0$. Thus any dissipation in the inner ``cauldron" or along the inner 
373: jet in effect pushes out $r_0$. This could come about because of 
374: entrainment, or because of the oblique shocks that occur when the jet is 
375: initially poorly collimated (as exemplified in numerical calculations of 
376: collapsar models such as Zhang \& Woosley 2004, where in effect $r_0\simg 
377: 10^8$ cm). The initial reference temperature $T_0$ is correspondingly 
378: lower. (Another effect which could change the reference temperature is if 
379: the inner jet is Poynting-dominated, so that only a small fraction of the 
380: flux is in the radiation. In this case, pairs can be even more dominant 
381: inside $r_s$ . There could be modifications to the outflow dynamics if the 
382: field were tangled and did not obey the straightforward Bernouilli equation 
383: for a relativistic gas (cf Heinz and Begelman, 2000)).
384: 
385: Dissipation need not necessarily lead to pair formation.
386: For example, in a ``slow heating" scenario (such as that of Ghisellini 
387: and Celotti, 1999), the accelerated particles, and the photons associated 
388: with them, could all have energies substantially below $\sim 0.5$ MeV. 
389: Dissipation would then not enhance the photospheric radius, but, even so, 
390: as indicated above, the characteristic photon energies and photospheric 
391: luminosity could be substantially boosted over what their adiabatic value 
392: would have been.
393: 
394: An important feature of the model is that millisecond variations 
395: -- either at the photosphere or due to internal shocks further out -- may 
396: still be traced back to irregularities in the jet boundary at $r_0$, since 
397: the characteristic timescale for a nozzle of opening half-angle $\theta_j$ 
398: is $t_{var}\sim r_0 \theta_j /c$, rather than $r_0/c$ itself, which can be 
399: less than a millisecond even if $r_o$ is of order $10^8$ cm . 
400: If internal shocks are to develop, they must be induced by unsteady 
401: conditions near the base of the jet (resulting in changes in $\eta$ and 
402: the saturation Lorentz factor). 
403: While for the usual minimum variability timescale $t_{var}\sim r_0/c$
404: shocks would develop above the line marked $r_{sh}$ in Fig. 1, for 
405: $t_{var}\sim r_0 \theta/c$ the shocks can form at radii $r_{sh,j}$ a factor 
406: $\theta_j$ smaller then for the spherical case, see Fig. 1. Dissipation at 
407: such or smaller radii is also possible, e.g., in the case of oblique shocks  
408: induced by irregularities in the walls of the jet, or during the collimation 
409: of an initially poorly collimated jet, or in the case of dissipation due 
410: to magnetic reconnection.
411: 
412: Note also that any variability at $r_0$ would alter the conditions 
413: at the photosphere (and the value of the photospheric radius). Moreover, 
414: the photospheric changes can be rapid. Obviously this is true if the 
415: photosphere lies below the saturation radius; however, this condition is 
416: not necessary, and provided that the photospheric radius is within $\eta^2 
417: r_0$, there is no smearing of variability on any timescale down to $r_0/c$. 
418: We would therefore, generically, expect an internal shock to be slightly 
419: preceded by a change in the luminosity of the thermal component (and in 
420: $E_{pk}$). Indeed, one is led to conjecture that rapid variations in the 
421: photosphere could be at least as important as the associated internal 
422: shocks in causing rapid variability in GRBs. In contrast to shocks, 
423: variations in the photospheric emission could as readily account for a 
424: short dip as for a short peak.  Detailed evidence of spectral softening 
425: during both the rise and fall of individual sub-pulses (c.f. Ryde 2004), 
426: could clarify the relative contributions of these effects.
427: 
428: \section{Spectrum Formation}
429: \label{sec:spec}
430: 
431: When dissipation occurs, one expects the photospheric spectrum to be 
432: "grey" rather than an accurate blackbody, because there would not 
433: (except near the base of the jet) be processes capable of producing the 
434: new photons appropriate to a black body with the enhanced energy density. 
435: All photons emerging from the photosphere will, however, have undergone 
436: multiple scatterings. 
437: %
438: In the case of shock dissipation, a power law relativistic electron 
439: energy distribution can be formed, which would upscatter the thermal 
440: photons into a  power law photon distribution whose index is similar 
441: to that of synchrotron radiation, pair formation being possible at 
442: comoving energies $\simg m_e c^2$. At the pair photosphere the comoving 
443: inverse Compton cooling time is $t'_{IC}\sim 4\times 10^{-3} L_{51}
444: \epsilon_{d,-1}^2 \alpha^{-2}\gamma_{e,3}^{-1}$ s, while the dynamic time 
445: is $t'_{dyn}\sim 3\times 10^1 L_{51}\epsilon_{d,-1}\alpha^{-1} \eta_2^{-4}$ s.
446: The interplay between the electron and photon distributions requires a 
447: detailed analysis, and is discussed in Pe'er, et al, 2005b.  
448: %
449: For a slow heating scenario, such as that of Ghisellini and Celotti 
450: (1999) but with the added feature of dissipation (e.g. from magnetic 
451: reconnection, or from multiple shocks and/or MHD turbulence behind them), 
452: one expects the electrons to be heated to more modest values, say 
453: $\gamma_e \siml$ few, but the electrons could keep being reheated 
454: every Compton cooling time. In this case pair formation is at best 
455: modest (Pe'er, et al, 2005b), so the effects outside the baryonic 
456: photosphere are not significant. If slow dissipation occurs at or 
457: below the baryonic photosphere, where pair formation is suppressed, 
458: the IC cooling time is $t'_{IC}\sim 10^{-4} L_{51}\alpha^{-1}
459: \gamma_{e,0.5}^{-1}\eta_2^{-4}$ s while the dynamic time is 
460: $t'_{dyn}\sim 2\times 10^{-1} L_{51}\alpha^{-1} \eta_2^{-4}$ s. 
461: The dissipative baryonic photosphere thermal peak is at $3kT_{\gamma,b}
462: \sim 20 L_{51}^{-1/2}\epsilon_{d,-1}^{1/4}\alpha^{1/2}\eta_2^2$ keV, 
463: which (depending on $\gamma_e$) may get upscattered by factors $\sim 1-10$.
464: %
465: The schematic shape of the spectrum is shown in Fig. 3 (c.f. Pe'er et al, 
466: 2005a, 2005b), showing the original quasi-thermal Wien component, 
467: the up-scattered photospheric component resulting from sub-photospheric 
468: dissipation and Comptonization, and a possible additional synchrotron 
469: component from shocks outside the photosphere. The peak frequency scales 
470: with the amount of dissipation according to a power law which depends
471: on how many new photons are produced. (The photon production depends on the 
472: radial dependence of the dissipation and on the detailed dissipation mechanism.)
473: 
474: The dependence of the spectral peak energy on the burst parameters,
475: as observed in a given energy range by a given instrument,  depends on
476: the specific mechanism responsible for the spectrum in that energy range.
477: In the BATSE and Beppo-SAX energy range (roughly 20 keV to 0.5 MeV),
478: there is a quantitative relationship observed between the spectral
479: peak energy $E_{pk}$ and the isotropic-equivalent luminosity of the
480: burst in tat energy range, $L_{iso}$ (which requires a knowledge of
481: the redshift of the burst). This relationship (Amati et al, 2001) is
482: $E_{pk}\propto L_{iso}^{1/2}$, for a score of bursts with redshifts.
483: Our generic assumptions naturally yields a peak in the relevant range, 
484: but cannot predict any correlation with $L$ without a more specific model
485: which relates $L$ to the other significant parameters, in particular 
486: $r_0$ and $\eta$. Without going into details, we may consider several
487: possibilities.
488: 
489: If one seeks to explain this relationship by interpreting the peak
490: energy as the synchrotron peak in a simple standard internal shock
491: scenario, one expects the dependence (e.g.  Zhang \& \Mesz, 2001)
492: \be
493: E_{pk} \propto \Gamma^{-2} t_{var}^{-1}L^{1/2} ~,
494: \label{eq:Epkintsh}
495: \ee
496: where $L_{\gamma, iso} \sim L$. Here $\Gamma$ and $t_{var}$ are
497: the Lorentz factor of the outflow and its typical variability
498: timescale. If the  latter two quantities are approximately the same
499: for all bursts, this would reproduce the Amati et al (2001) relation.
500: However, it is not obvious why there should be a constancy of $\Gamma$
501: and $t_{var}$ across bursts, even if approximate.
502: 
503: If the spectral peak is of a quasi-thermal origin
504: determined by the photosphere (possibly shifted up by Comptonization,
505: e.g. from pair dissipative effects such as discussed above), and if
506: there are enough photons to guarantee an approximate black body
507: distribution, the peak photon energy in the observer frame is,
508: using equation (\ref{eq:pairphot}),
509: \be
510: E_{pk}\propto \Gamma kT'_{pk} \propto \Gamma (L/\Gamma r^2)^{1/4}
511:        \propto \Gamma^2 L^{-1/4} \propto L^{(8\beta -1)/4},
512: \label{eq:EpkBB}
513: \ee
514: which depends mainly on the Lorentz factor $\Gamma$. If the latter
515: in turn depends on $L$, e.g. as $\Gamma\propto L^{\beta }$, one obtains
516: the last part of equation (\ref{eq:EpkBB}).
517: For instance, taking the observed Frail et al (2001) relation
518: $L_{\gamma, iso}\propto \theta^{-2}$ between the jet opening half-angle
519: $\theta$ inferred from the light-curve break, and using the causality
520: relation $\theta \sim \Gamma^{-1}$, equation (\ref{eq:EpkBB}) becomes
521: $E_{pk}\propto L^{3/4}$.
522: %\label{eq:Epkbbfrail}
523: 
524: If dissipation occurs mainly very close to the central engine, this
525: could result in a larger radius $r_0$, where $r_0$ is defined as the
526: radius beyond which $\Gamma\propto r/r_0$. Assuming that the ``drag"
527: or dissipation at the base increases $r_0$ according to, e.g.,
528: $r_0\propto L^{-\beta'}$, for a photosphere occurring inside the
529: saturation radius, $r_0 < r_{ph,\pm} <r_s$, the growth of the
530: Lorentz factor $\Gamma \propto r/r_0$ cancels out the adiabatic
531: drop $T'\propto r^{-1}$ of the comoving temperature, and one has
532: \be
533: E_{pk}\propto r_0^{-1/2} L^{1/4} \propto L^{((2\beta' +1)/4}~.
534: \label{eq:Epksat}
535: \ee
536: Hence, for $\beta'=(0.5,~1)$ one has $E_{pk}\propto (L^{1/2},~L^{3/4})$.
537: 
538: In the extreme 'photon starved' case (likely to apply if the dissipation 
539: is concentrated not far inside the photosphere) where the photon number 
540: $N_\gamma$ is constant), one would have $E_{pk} \propto L/ N_\gamma \propto 
541: L$. 
542: 
543: Thus, a variety of $E_{pk}$ vs. $L$ dependences might in principle be 
544: expected, depending on the uncertain physical conditions just below the 
545: photosphere, some of which approximate the reported $L^{1/2}$ behavior. 
546: Ghirlanda et al (2004) have recently claimed an empirical correlation 
547: between $E_{pk}$ and a different quantity, the angle-corrected total energy 
548: $E_{tot}= E_{iso}(1-\cos\theta_j)\sim E_{iso}(\theta_j^2/2)$, where 
549: $E_{iso} \simeq L_{iso} t_{\gamma}$ and $t_\gamma$ is the burst duration. 
550: They find a tighter correlation between $E_{pk}$ and $E_{tot}$ than between 
551: $E_{pk}$ and $E_{iso}$, for bursts with observed redshifts and breaks. 
552: Furthermore, in contrast to the Amati et al (2001) $E_{pk}\propto 
553: E_{iso}^{1/2}$ dependence, they deduce from the data a steeper slope, 
554: $E_{pk}\propto E_{tot}^{0.7}$. Taking a standard burst duration and jet 
555: opening angle, this is of the form discussed in equation (\ref{eq:EpkBB}) 
556: or (\ref{eq:Epksat}). A critique of the methods for obtaining both types of 
557: correlations from the data is given by Friedman and Bloom (2004). We should 
558: note that such correlations are generally derived assuming that the 
559: efficiency of gamma-ray production is the same for all bursts, 
560: independently of the luminosity or the total energy. If, however, the 
561: efficiency were lower for the weaker (and therefore softer) bursts, then 
562: the correlation would have a flatter $E_{pk}$ vs. $E_{tot}$ slope than 
563: currently derived from the data. This is because, for a given gamma-ray 
564: isotropic luminosity, the momentum outflow per unit solid angle would be 
565: higher than they assume. This means that the standard jet-break argument 
566: would imply a narrower beam than inferred under the constant efficiency 
567: assumption, and therefore a lower $L_{tot}$ (for a given $E_{pk}$) than the 
568: values currently derived.
569: 
570: \section{Discussion}
571: \label{sec:disc}
572: 
573: We have considered dissipative effects below the photosphere of GRB
574: or XRF outflows, such as, e.g., due to magnetic reconnection or shocks.
575: Such dissipation can lead to copious pair formation, dominating the
576: photospheric opacity.  Alternatively, if dissipation occurs not too far
577: above an initial photosphere, it can result in a second effective
578: photosphere, situated outside the initial one.
579: 
580: Sub-photospheric dissipation can increase the radiative efficiency of the 
581: outflow, significantly boosting the quasi-thermal photospheric component, 
582: so that it may well dominate the much-discussed synchrotron component from 
583: nonthermal shocks outside the photosphere. The hypothesis that GRB emission 
584: is dominated by a Comptonized thermal component offers a natural 
585: explanation for the thermal GRB spectra discussed most recently, e.g., by 
586: Ryde (2004). It can also naturally explain the steeper than synchrotron 
587: lower energy spectral indices (Preece, et al, 2000; Lloyd, Petrosian \& 
588: Mallozi, 2000) noticed in some bursts.
589: 
590: The quasi-thermal peak of the photospheric spectrum is controlled by
591: the total luminosity $L_0$ and by the reference injection radius $r_0$
592: above which the Lorentz factor starts to grow linearly. Dissipation
593: near the central object or along the inner jet can result in an
594: increase $r_0$, thus lowering the reference temperature of the outflow
595: which characterizes the quasi-thermal photospheric component. The
596: characteristic variability timescales $r_0\sin\theta/c$ for jets
597: with the observationally inferred opening half-angles $\theta$ are
598: in the millisecond range. The spectral peak of the dissipation-enhanced
599: photospheric component, upscattered in energy by factors of $\sim 10$
600: due to electrons accelerated in the dissipation process, results in typical
601: photon energies ranging between those of classical bursts and X-ray flashes.
602: 
603: The relationship between the observed photon peak energy and the luminosity 
604: can have a variety of functional forms, which depends on a number of so-far 
605: poorly determined parameters. However, plausible assumptions can lead to 
606: relationships of the type $E_{pk}\propto L_{iso}^{1/2}$ (Amati, et al, 
607: 2000), or $E_{pk}\propto E_{tot}^{0.7}$ (Ghirlanda et al, 2004). Even 
608: though more physics and a more specific model will be needed before we can 
609: explain the correlations, the idea that $E_{pk}$ is essentially a thermal 
610: peak seems more readily able to account for a `standardized' value in a 
611: given class of objects, because there is not a steep $\Gamma$-dependence 
612: (and indeed to first order the $\Gamma$ factor cancels out, because 
613: adiabatic cooling in the comoving frame is compensated by the Doppler 
614: boosting).
615: 
616: In summary, our main result is that  a spectral  peak at photon  energies 
617: in the range of tens to hundreds of keV, typical of XRFs and GRBs, can 
618: naturally arise from an outflowing jet, in which dissipation below a 
619: baryonic or pair-dominated photosphere enhances the radiative efficiency 
620: and gives rise to a Comptonized thermal spectrum. On this hypothesis, 
621: the recently-discovered correlations between $L$ and $E_{pk}$ would be 
622: an important diagnostic of how the key jet parameters --  physics near 
623: the 'sonic point', baryon contamination, etc -- depend on $L$.
624: 
625: 
626: \acknowledgements
627: Research supported in part by NASA NAG5-13286, the Royal Society and the
628: Monell Foundation. We are grateful to S. Kobayashi, D. Lazzati, A. Pe'er
629: and the referee for useful comments.
630: 
631: \begin{references}
632: \def\refe{\reference{}}
633: 
634: \refe Amati, L. et al, 2002, A. \& A., 390, 81.
635: \refe Beloborodov, A, 2003, ApJ 585, L19
636: \refe Eichler, D. \& Levinson, A., 2000, ApJ 529, 146
637: \refe Friedman, A.S. and Bloom, J.S., ApJ subm, astro-ph/0408413
638: \refe Giannios, D \& Spruit, H, 2005, A\&A in press (astro-ph/0401109)
639: \refe Ghirlanda, G., Ghisellini, G. \& Lazzati, D, 2004, ApJ 616, 331
640: \refe Ghisellini, G. \& Celotti, A. 1999, ApJ 511, L93
641: \refe Heinz, S. \& Begelman, M., 2000, ApJ 535, 104
642: \refe Kobayashi, S, Ryde, F. \& MacFadyen, A, 2002, ApJ 577, 302
643: \refe Lloyd, N.M., Petrosian, V. \& Mallozzi, R.S., 2000, ApJ, 534, L227.
644: \refe \Mesz, P. \& Rees, M.J., 2000, ApJ 530, 292
645: \refe Pe'er, A \& Waxman, E., 2004, ApJ 613, 448
646: \refe Pe'er, A, et al, 2005a, 2005b, in preparation
647: \refe Preece, R.D.; Briggs, M.S.; Mallozzi, R. S.; Pendleton, G. N.;
648:   Paciesas, W. S.; Band, D. L., 2000, ApJS 126, 19.
649: \refe Ryde, F, 2004, ApJ 614, 827
650: \refe Thompson, C., 1994, MNRAS 270, 480
651: \refe Zhang B. \& \Mesz, P., 2002, ApJ, 581, 1236
652: \refe Zhang W. \& Woosley, S.E., 2004, ApJ 608, 365
653: \end{references}
654: 
655: \clearpage
656: %
657: \begin{figure}[htb]
658: \centering
659: %\epsfig{figure=f1.eps,width=5.in, height=4.in}
660: \epsfig{figure=par_phot_eta2.ps,width=5.in, height=4.in}
661: \figcaption{
662: Radii of the pair (full) and baryon (dashed) photospheres as a
663: function of $\eta$ for $\epsd=10^{-1}$, $L_0=10^{51}$ erg/s,
664: $\alpha=1$ ($r_0=3\times 10^6 m_1\cm$) and $\alpha=10^4$ ($r_0=
665: 3\times 10^{10}m_1\cm$).  Also shown are the saturation radius
666: $r_s=r_0\eta$, and the spherical minimum shock radius $r_{sh}=
667: r_0\eta^2$. Instabilities at the nozzle $\theta$ of a jet could
668: lead to shocks at a lower mininum radius $r_{sh,j}$, while magnetic 
669: dissipation could in principle occur both above and below $r_s$.
670: }
671: \label{fig:photeta}
672: \end{figure}
673: %
674: 
675: 
676: %
677: \begin{figure}[htb]
678: \centering
679: %\epsfig{figure=f2.eps,width=5.in,height=4.in}
680: \epsfig{figure=par_lum_rad.ps,width=5.in,height=4.in}
681: \figcaption{
682: Kinetic luminosity and photospheric radiation luminosity as a function
683: of radius. Beyond the saturation radius the luminosity decays as
684: $L_\gamma\propto r^{-2/3}$, but beyond $\epsilon_d^{3/2}r_s$ the
685: fraction $\epsilon_d$ of the kinetic energy reconverted into
686: radiative (pair) form becomes significant. Also shown is the value
687: of the observer temperature. Comptonization at the pair photosphere
688: (see text) could boost this by an additional factor $\siml 10$.
689: }
690: \label{fig:lumrad}
691: \end{figure}
692: %
693: 
694: %
695: \begin{figure}[htb]
696: \centering
697: %\epsfig{figure=f3.eps,width=5.in,height=4.in}
698: \epsfig{figure=parspect2.ps,width=5.in,height=4.in}
699: \figcaption{Schematic comoving frame spectrum, showing the photospheric
700: (thermal) spectrum and its Comptonized  component, as well as a shock
701: synchrotron component (assumed to arise further out). This is the
702: generic spectrum characterizing a slow dissipation model (see text).
703: Shocks with pair formation could lead to an additional component
704: at higher energies.
705: }
706: \label{fig:spectrum}
707: \end{figure}
708: %
709: 
710: 
711: 
712: 
713: \end{document}
714: 
715: