Articles

RELATIVISTIC SHOCK BREAKOUTS—A VARIETY OF GAMMA-RAY FLARES: FROM LOW-LUMINOSITY GAMMA-RAY BURSTS TO TYPE Ia SUPERNOVAE

and

Published 2012 February 17 © 2012. The American Astronomical Society. All rights reserved.
, , Citation Ehud Nakar and Re'em Sari 2012 ApJ 747 88 DOI 10.1088/0004-637X/747/2/88

0004-637X/747/2/88

ABSTRACT

The light from a shock breakout of stellar explosions, which carries a wealth of information, strongly depends on the shock velocity at the time of the breakout. The emission from Newtonian breakouts, typical in regular core-collapse supernovae (SNe), has been explored extensively. However, a large variety of explosions result in mildly or ultrarelativistic breakouts, where the observed signature is unknown. Here we calculate the luminosity and spectrum produced by relativistic breakouts. In order to do so, we improve the analytic description of relativistic radiation-mediated shocks and follow the system from the breakout itself, through the planar phase and into the spherical phase. We limit our calculation to cases where the post-breakout acceleration of the gas ends during the planar phase (i.e., the final gas Lorentz factor ≲ 30). We find that spherical relativistic breakouts produce a flash of gamma rays with energy, Ebo, temperature, Tbo, and duration, tobsbo, that provide the breakout radius (≈5 R(tobsbo/10 s)(Tbo/50 keV)2) and the Lorentz factor (≈Tbo/50 keV). They also always satisfy a relativistic breakout relation (tobsbo/20 s) ∼ (Ebo/1046 erg)1/2(Tbo/50 keV)−2.68. The breakout flare is typically followed, on longer timescales, by X-rays that carry a comparable energy. We apply our model to a variety of explosions, including Type Ia and .Ia SNe, accretion-induced collapse, energetic SNe, and gamma-ray bursts (GRBs). We find that all these events produce detectable gamma-ray signals, some of which may have already been seen. Some particular examples are: (1) relativistic shock breakouts provide a natural explanation to the energy, temperature, and timescales of low-luminosity GRBs. Indeed, all observed low-luminosity GRBs satisfy the relativistic breakout relation. (2) Nearby broad-line Type Ib/c (like SN 2002ap) may produce a detectable γ-ray signal. (3) Galactic Type Ia SNe may produce detectable γ-ray flares. We conclude that relativistic shock breakouts provide a generic process for the production of gamma-ray flares.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The first observable light from any stellar explosion is generated by the breakout of the shock that traverses the star. In most stellar objects the density drops sharply near the stellar edge. The shock that encounters this density drop accelerates, leaving most of its energy behind. Therefore, it breaks out at a velocity that is significantly higher than the typical shock velocity in the stellar interior, carrying only a small fraction of the total explosion energy. In typical core-collapse supernovae (SNe), where the total explosion energy is ∼1051 erg and the progenitor is not a compact object, the velocity of the shock at the breakout is Newtonian. For example, in a red supergiant progenitor the breakout velocity is in the range 5000–10,000 km s−1, while in a Wolf–Rayet progenitor it is ∼30, 000–100,000 km s−1. In cases where the explosion is more energetic or the progenitor is a compact object, the breakout can become mildly or even ultrarelativistic.

Before its breakout, the shock is dominated by radiation. During a Newtonian breakout, all the photons that are contained in the shock transition layer are escaping the system and are seen as short and bright flashes. The properties of Newtonian shock breakout flashes were explored by many authors (e.g., Colgate 1974; Falk 1978; Klein & Chevalier 1978; Imshennik et al. 1981; Ensman & Burrows 1992; Matzner & McKee 1999; Katz et al. 2010, 2011; Nakar & Sari 2010). These properties depend mostly on the breakout radius and on the shock velocity at the layer that dominates the breakout emission, v0. Among these properties the observed typical photon frequency, denoted here as the radiation temperature T0, can be especially sensitive to the shock velocity. In relatively slow shocks, v0 < 15, 000 km s−1, the radiation in the shock breakout layer is in thermal equilibrium and T0v1/20, producing typically a spectrum that peaks in the UV. However, as was shown by Weaver (1976), in faster radiation-mediated shocks the radiation is away from thermal equilibrium in the immediate shock downstream, and as a result the temperature in the shock transition layer is much higher than the one obtained assuming thermal equilibrium. As a result, the breakout in fast Newtonian shocks, v0 ∼ 30, 000–100, 000 km s−1, peaks in X-rays (Katz et al. 2010; Nakar & Sari 2010).

The emission following the breakout can be divided into two dynamically different regimes. At first, before the expanding gas doubles its radius, the dynamics is planar, whereas at a later time the evolution is spherical (Piro et al. 2010b; Nakar & Sari 2010; Katz et al. 2011). The evolution of the temperature and luminosity during the planar phase was studied by Nakar & Sari (2010), where we have shown that if the breakout emission is out of thermal equilibrium at the breakout, it remains out of equilibrium also during the planar phase and thermal equilibrium is achieved only during the spherical phase. The evolution during the spherical homologous phase was studied by many authors (e.g., Grassberg et al. 1971; Chevalier 1976, 1992; Chevalier & Fransson 2008; Piro et al. 2010b; Rabinak & Waxman 2011; Nakar & Sari 2010), where all use the assumption (which we prove to be adequate in Nakar & Sari 2010) of thermal equilibrium. All these studies considered only explosions where the shock breakout velocities are Newtonian and pairs are not significantly produced in the transition layer, i.e., β0 = v0/c < 0.5, where c is the speed of light.

A large range of cosmic explosions are expected to produce breakout velocities that are either mildly relativistic or ultrarelativistic. A few examples are a range of white dwarf explosions such as a Type Ia SN (Piro et al. 2010b), a Type .Ia SN (Shen et al. 2010), and possibly accretion-induced collapse (AIC; Dessart et al. 2006; Piro et al. 2010a). The energy involved in most of these explosions is ≲ 1051 erg, but the compact progenitor and relatively low ejecta mass result in relativistic breakouts. Another example is extremely energetic explosions of massive stars, with explosion energies in the range 1052–1053 erg. These were recently observed in growing numbers, e.g., SN2005ap and SN2007bi (e.g., Quimby et al. 2011; Gal-Yam et al. 2009). Yet another example is massive star explosions that are associated with long gamma-ray bursts (GRBs), where the breakout is ultrarelativistic. Relativistic breakouts may also take place in choked GRBs, where the relativistic jets fail to penetrate through the stellar envelope, and no regular long GRB is produced. As we show here, this mechanism is a promising source of all the observed low-luminosity GRBs.

The physical processes involved in producing the observed signature of a relativistic shock breakout are different from the Newtonian case in two fundamental ways. First, once the velocity of a radiation-mediated shock becomes large enough, βs > 0.5, and the temperature behind the shock exceeds 50 keV, pair production becomes the dominant source of leptons, which in turn are the main source of photons via free–free emission. The exponential sensitivity of the number of pairs to the temperature regulates the rest-frame temperature in the shock immediate downstream to an almost constant value of 100–200 keV, independent of the shock Lorentz factor (Katz et al. 2010; Budnik et al. 2010). The gas becomes loaded with pairs that dominate its optical depth, resulting in a temperature-dependent opacity per unit of mass. This is in sharp contrast to Newtonian shocks, especially when 0.05 < βs < 0.5, where the temperature of the shocked fluid depends strongly on the shock velocity but the opacity per unit of mass is constant. The second major difference of relativistic explosions is the hydrodynamics both before and after the breakout. A Newtonian shock that propagates in a power-law decreasing density assumes the self-similar solution of Sakurai (1960) before the breakout, and the gas velocity is roughly constant after the breakout (it doubles by about a factor of two). Once the shock becomes relativistic it follows the solution of Johnson & McKee (1971; see also Tan et al. 2001 and Pan & Sari 2006) before the breakout and each fluid element continues to accelerate significantly following the breakout. These two differences, together with the usual relativistic effects, such as relativistic beaming, govern the observed emission, which as we show here is very different than that in the Newtonian case.

Yet another difference between Newtonian and relativistic radiation-mediated shocks is the physical width of the shock. In radiation-mediated shocks τsc/vs, where τs is the total optical depth (including pairs and Klein–Nishina effects) seen by a photon going from the shock downstream toward the shock upstream. In Newtonian radiation-mediated shocks, where no pairs are produced, the optical depth of a layer remains similar before it is shocked and once it becomes the shock transition layer. Therefore, the shock breaks out at the layer in which τ ≈ τs, where τ is the Thomson optical depth to the stellar edge before the shock crossing (i.e., without pairs). However, in relativistic shocks, where pairs are produced and Klein–Nishina effects may become important, the optical depth of a layer is changed significantly when it is swept up by the shock and the relation between τs and τ is not trivial anymore. In order to find this relation one needs to know the structure of the shock transition layer. The structure of relativistic radiation-mediated shocks in different regimes was solved numerically by Budnik et al. (2010) and Levinson & Bromberg (2008). The solution of Budnik et al. (2010), where a significant number of photons are generated within the shock, is the one relevant to the type of shocks we consider here (Bromberg et al. 2011a). Budnik et al. (2010) also provide an approximate analytic description of the shock structure as a function of the optical depth. We derive a more accurate analytic description of the structure of the shock transition layer and use it to constrain the value of τ at the point that the shock breaks out. We find that while production of pairs results in a shock that breaks out of the star at τ ≪ 1, the observed emission is always dominated by the layer where τ ≈ 1. We therefore use the subscript "0" and the terminology "breakout shell" to denote the τ = 1 layer, which is inward of where the actual breakout takes place, but is the layer that dominates the energy released in the breakout flash.

In this paper, we calculate the evolution of the observed luminosity and temperature from explosions in which the shock becomes mildly or ultrarelativistic, i.e., γ0β0 > 0.5. We follow the light curve as long as the observed radiation is generated by gas that is moving at relativistic velocities. At later times the radiation is determined by Newtonian gas, whose light curve was discussed in Nakar & Sari (2010). Our solution is limited to cases where the breakout shell ends its post-shock acceleration before it doubles its radius. This limit is translated into a final (post-acceleration) Lorentz factor of the breakout shell ≲ 30. We also assume that the opacity of the progenitor wind is negligible and does not affect the breakout emission. Our calculations are applicable to a wide range of energetic and/or compact explosions including Type Ia and .Ia SNe, energetic core-collapse and pair instability SNe, and possibly part of the phase space of chocked GRB jets.

We find that, typically, relativistic breakouts produce flashes of gamma rays that without further information (e.g., the burst redshift) can be mistaken as cosmological GRBs. These results provide closure to Colgate (1968), who was the first to predict that shock breakouts produce flares of gamma rays, even before the detection of gamma-ray bursters was announced (Klebesadel et al. 1973). However, Colgate (1968) assumed that the gas behind the shock is in thermal equilibrium, and the observed γ-rays are due to blueshift of ultrarelativistic gas, brought to high Lorentz factors by shock that accelerates indefinitely as it approaches the stellar surface. It ignores deviation from thermal equilibrium and the fact that acceleration stops once the shock width is comparable to the distance to the stellar surface. As it turned out, the deviation from thermal equilibrium increases the photon's rest-frame temperature and compensates for the limited Lorentz factor of the gas, resulting in a burst of γ-rays. Since Colgate's prediction, breakouts were suggested to produce bright gamma-ray emission by several authors in the context of low-luminosity GRBs (Kulkarni et al. 1998; Tan et al. 2001; Campana et al. 2006; Waxman et al. 2007; Wang et al. 2007). All these calculations assume thermal equilibrium, underestimating the true radiation temperature. Katz et al. (2010) realized that the deviation from equilibrium, discussed by Weaver (1976), will lead to γ-ray breakouts. They did not follow, however, the post-shock dynamics and opacity evolution to find the properties of the observed emission. Here we provide, for the first time, a quantitative description of the predicted light curve from relativistic breakouts. We show that these light curves explain very well many of the observed features of low-luminosity GRBs and also that a variety of other stellar explosions are most likely producing a bright gamma-ray breakout emission.

In Section 2, we discuss the pre- and post-breakout dynamics. The dynamics is dominated by mass, energy, and momentum conservation, and is therefore largely independent of the photon number and spectrum, as long as the gas is optically thick. In Section 3, we calculate the temperature and opacity evolution of the expanding gas. Section 4 follows the observed light curve from the breakout to the spherical phase. Our results are applied in Section 5 to find the observed signature of known and hypothesized explosions. An analytic description of the structure of the transition layer in relativistic radiation-mediated shocks is presented in Appendix A.

2. DYNAMICS

2.1. Pre-breakout Dynamics

Consider a relativistic blast wave that propagates near a stellar edge pre-explosion mass density gradient of the form ρ∝zn, where z = (R*r), with R* being the stellar radius and r the distance from the star center. In compact stars, where shocks are more likely to become relativistic, the envelope is radiative and typically n ≈ 3, which is the value that we use throughout the paper. The shock acceleration stops when the width of the shock is comparable to z and the shock breaks out. As we show in Appendix A, this takes place in the shell where the pre-explosion optical depth to the stellar edge is τ ∼ 0.01γ3s or smaller, where γs is the shock Lorentz factor. Namely, breakout takes place at z ≲ 0.01γ3s/(ρκT), where κT ≈ 0.2 cm2 g−1 is the Thomson cross section per unit of mass.3 This sets the maximal Lorentz factor of the shock and we refer to that shell as the outermost shell. Unlike the Newtonian case, the energy released from a relativistic breakout is not dominated by the outermost shell. Instead it is dominated by the shell in which the pre-explosion optical depth is τ ∼ 1 (i.e., z ∼ 1/(ρκT)), which we therefore refer to as the breakout shell and denote its properties by the subscript "0." In the case of properties that evolve with time, this subscript refers to their value in the breakout shell right after it is crossed by the shock. Given the above assumptions the system is completely defined by the following parameters of the breakout shell.

  • 1.  
    γ0: the Lorentz factor of the shock when it crosses the breakout shell.
  • 2.  
    m0: the mass of the breakout shell.
  • 3.  
    d0: the post-shock width (i.e., after compression) of the breakout shell in the lab frame.
  • 4.  
    R*: the radius of the star.
  • 5.  
    n: the power law index describing the pre-explosion density profile, here we use n = 3.

Other characteristics of the breakout shell can be calculated from above. For example, the initial thermal energy of the breakout shell is E0 ≈ γ20m0c2.

The pre-breakout and post-breakout dynamics of a planar relativistic blast wave in a decreasing density profile was investigated by Johnson & McKee (1971), Tan et al. (2001), and Pan & Sari (2006). If the density follows a power law then the evolution is self-similar both before and after the breakout. Before the breakout a relativistic shock accelerates as γi∝ρ−μ, where $\mu =\left(\sqrt{3}-\frac{3}{2}\right) \approx 0.23$ and the subscript "i" indicates initial values (note that ρ is the pre-shock density). This propagation profile sets the entire pre-explosion hydrodynamic profile. Conceptually, it is useful to treat the whole expanding envelope as a series of successive shells. The properties of any shell deeper than the outermost shell can be characterized by its mass, m.

  • 1.  
    $\gamma _i=\gamma _0 (m/m_0)^{-\mu n \over n+1} \propto m^{-0.17}$ is the initial Lorentz factor.
  • 2.  
    $d_i \approx z/\gamma _i^2 \propto m^\frac{1+2n\mu }{(n+1)} \approx m^{0.6}$ is the width of a shocked shell at the shock breakout, in the lab frame.
  • 3.  
    $d_i^{\prime }=d_i \gamma _i \propto m^\frac{1+n\mu }{(n+1)} \approx m^{0.42}$ is the width of a shocked shell at the shock breakout, in the shell rest frame.
  • 4.  
    $E_i \approx \gamma _i^2 mc^2 \propto m^{\frac{1+n(1-2\mu)}{n+1}}\approx m^{0.65}$ is the internal energy of a shell at the shock breakout, in the lab frame.
  • 5.  
    $t_i \approx \frac{d_i}{c} \gamma _i^2 \propto m^\frac{1}{n+1} \propto m^{0.25}$ is the lab frame time for the shell expansion, which is also the time between the shock crossing and the breakout.

2.2. Post-breakout Dynamics

Following breakout all the shells expand and accelerate. The acceleration is faster than that of a freely expanding relativistic shell as outer shells gain energy from inner ones. In this subsection we follow the acceleration of shells with mass mm0. These shells remain opaque during the whole planar phase, also after their pairs annihilate. The final Lorentz factor of less massive shells depends on their evolving opacity and is discussed shortly in the next section.

The acceleration during the planar phase, i.e., before a shell doubles its radius, follows (Johnson & McKee 1971; Pan & Sari 2006):

Equation (1)

where t is the lab frame time (not to be confused with observer time) measured since the breakout. If the shell is optically thick at the end of acceleration and acceleration ends during the planar phase then the final Lorentz factor is (Johnson & McKee 1971; Tan et al. 2001; Pan & Sari 2006; Oren & Sari 2009)

Equation (2)

This relation is general as it does not depend on the exact density profile, as long as there is a large energy reservoir behind the accelerating shell. Equations (1) and (2) imply that the acceleration ends at

Equation (3)

Thus, more massive shells end their acceleration at earlier times and lower Lorentz factors.

In this paper, we restrict our treatment to cases where the breakout shell ends its acceleration during the planar phase, i.e., tf, 0 < ts where

Equation (4)

is the lab frame time of transition between the planar and spherical phases. Since pre-shocked optical depth of the breakout shell is of order unity (i.e., κTρ0z0 ≈ 1),

Equation (5)

where Mx and Rx are the progenitor star mass and radius in units of x solar masses and radii, respectively. Thus, there is a critical Lorentz factor, γ0, s, for which tf, 0 = ts:

Equation (6)

The corresponding critical final Lorentz factor of the breakout shell is

Equation (7)

If γ0 ⩽ γ0, s, then acceleration ends in the planar phase and the corresponding final Lorentz factor is given by Equation (2). Otherwise it ends during the spherical phase and Equation (2) is not valid anymore, a case that we treat elsewhere.

Acceleration continues beyond the breakout shell, up to the outermost shell, which lies at4 z ∼ 0.01γ3i/(ρκT). This shell achieves the maximal initial Lorentz factor

Equation (8)

Note that our calculation is limited to γ0 < γ0, s ≲ 4.5, where γ0 < γmax, i and the description above is consistent.

3. TEMPERATURE AND OPACITY

The rest-frame temperature behind a radiation-mediated shock with βs > 0.5 is roughly constant5 (Katz et al. 2010; Budnik et al. 2010):

Equation (9)

The reason is that at these velocities the radiation behind the shock is out of thermal equilibrium and is set by the ability to generate photons. The main photon source in the shock is free–free emission (Weaver 1976). Once the temperature behind the shock rises above ≈50 keV, pairs are generated and their number becomes dominant over the number of electrons that are advected from the upstream. These pairs significantly increase the rate of photon generation. The strong dependence of the number of pairs on the temperature in that range serves as a "thermostat" that sets the temperature in the immediate shock downstream to be nearly constant, regardless of the shock Lorentz factor. Although the radiation is out of thermal equilibrium, the photons and the pair loaded gas in the immediate downstream are in Compton-pair equilibrium and there is a single gas-radiation temperature (Budnik et al. 2010). This property of relativistic radiation-mediated shocks is dominating the emission at the shock breakout, during the planar phase and the relativistic part of the spherical phase.

The gas behind the shock is pair loaded and therefore its optical depth is very high, preventing any significant leakage of photons, even if the pre-shocked (unloaded) gas optical depth is below unity (the pairs set a diffusion time that is much larger than the dynamical time). On the other hand, the rest-frame temperature in the expanding gas drops, and with it the number of pairs and the photon generation rate. Therefore, the number of photons in the opaque expanding shells is roughly constant. The mildly relativistic temperature and the large optical depth ensure that photons and pairs interact many times (i.e., sharing energy, annihilation, and creation) over the dynamical time, thereby keeping the pair-photon gas in Compton-pair equilibrium.

The radiation of the breakout shell is confined to the gas during its expansion until the rest-frame temperature drops enough so the pair loading becomes negligible. During the expansion of a shell its rest-frame temperature falls as T'∝V'−ξ, where V' is the shell volume (in its rest frame) and 1 + ξ is an effective adiabatic index. The value of ξ depends on temperature and it can drop slightly (by up to 30%) below 1/3 because of pair production and annihilation. Here we neglect this small deviation from the typical relativistic equation of state and approximate ξ = 1/3. During the planar phase the volume of a fluid element grows as V'∝t/γ. Therefore during acceleration $V^{\prime } \propto t^{(3-\sqrt{3})/2}$, whereas after the acceleration ends V'∝t. During the spherical phase V'∝t3, implying

Equation (10)

The optical depth of the breakout shell drops to unity once the pair density, n±, is lower than the proton density, np (and their accompanied electrons), i.e., n±/np < 1. The initial value of this ratio is (e.g., Svensson 1984; Budnik et al. 2010)

Equation (11)

Since at 200 keV the number of photons and pairs is comparable, this is also the initial photon-to-proton ratio, nγ, 0/np, 0. Photons are not generated in a significant number following the shock breakout and therefore, the photon-to-proton ratio remains roughly constant in optically thick regions. Thus, the breakout shell becomes optically thin when its photon to pair ratio is nγ/n± ∼ 103γ0. As pairs are in annihilation–creation equilibrium their density (given in, e.g., Svensson 1984) drops exponentially with T' at T' < 100 keV and this value of the photon to pair ratio is satisfied at almost a constant temperature,

Equation (12)

where the dependence on γ0 is very weak. Around this temperature, the breakout shell, and any less massive shell, becomes transparent and its radiation escapes. We define tth, 0 as the time that the breakout shell becomes optically thin, i.e., its T' = T'th.

There are three critical times in the system, in which the dynamics and radiation change their behavior. These are the acceleration and transparency times of the breakout shell, tf, 0 and tth, 0, respectively, and the planar to spherical transition time, ts. The observed signal depends on the relative order of these three timescales. Here we restrict our analysis to the case that acceleration ends during the planar phase, i.e., tf, 0 < ts. Under this assumption, Equation (10) implies (note that T'0T'i)

Equation (13)

Now, T'0/T'th ∼ 4 and we consider only γ0 < γ0, s, which for any non-degenerate progenitor implies γ0 < γ0, s ≲ 4. Therefore, the breakout shell becomes transparent only after acceleration ends, i.e., tf, 0 < tth, 0. On the other hand, the opacity drops quickly during the spherical phase, so Equation (13) implies that the breakout shell becomes transparent during the planar phase, or at the beginning of the spherical phase, namely, tth, 0ts. Therefore for a large range of realistic explosion parameters tf, 0 < tth, 0 < ts, which is the case that we present in detail below. Finally, note that shells with γi > γ0 can be accelerated only as long as they are optically thick, namely, their T' > T'th. Therefore, only shells with γi ≲ min {4, γ0, s} satisfy the relation $\gamma _f=\gamma _i^{1+\sqrt{3}}$, shells with higher γi end their acceleration before this relation is satisfied.

4. LIGHT CURVE

4.1. Breakout Emission

The radiation in the expanding gas is trapped by pairs and γγ annihilation opacity at the time of the breakout. In the previous section, we have shown that for most progenitors in the regime that we consider, the breakout shell becomes transparent after it ended acceleration and before, or soon after, the transition to the spherical phase, i.e., tf, 0 < tth, 0ts. Therefore we derive below the observed light curve in that case. Note that the breakout shell is not the only shell that becomes transparent during the planar phase. All the lighter shells (m < m0) out to the outermost shell become transparent during the planar phase as well. Shells that are deeper (and more massive and slower) than the breakout shell remain opaque during the planar phase and become transparent only during the spherical phase. Figure 1 depicts a schematic sketch of the luminosity and temperature evolution during the breakout and planar phase.

Figure 1.

Figure 1. Schematic sketch of the luminosity and temperature evolution following a relativistic shock breakout. The earliest phase is dominated by breakout emission that is smeared by travel time effects over the planar phase. The energy radiated during this phase is dominated by the breakout shell, while the luminosity and temperature evolution is dictated by outer and faster shells that lay between the breakout shell and the outermost shell. The luminosity, temperature, and duration of the breakout shell emission depend only on the breakout radius and the Lorentz factor and are marked at the end of the planar phase. Following the planar phase, there is a short phase where high latitude emission from the breakout shell dominates the light curve, after which the emission from slower shells in the spherical phase takes over. The spherical phase is dominated at first by relativistic shells and later by Newtonian ones. The transition that takes place at tobsNW is prominent in the evolution of the luminosity and is hardly seen in that of the temperature. A sharp drop in the temperature is observed once the velocity of the luminosity shell drops below ≈0.5c, so that pair loading during the shock crossing becomes insignificant. See Section 4.1 for a discussion and details of the breakout, planar and high latitude emission phases, and Section 4.2 for a discussion of spherical phases.

Standard image High-resolution image

Since slower moving shells carry significantly more energy, Ei∝γ−3.75i, the breakout emission is dominated by the energy confined to the breakout shell at tth, 0, when its local frame temperature is T'th:

Equation (14)

where we use τ0 ≈ κTm0/(4πR2*) = 1 to find

Equation (15)

The observed temperature of this emission is

Equation (16)

and it is smeared by light travel time effects over a duration that is comparable to tobss:

Equation (17)

where tobss is the time of transition from the planar to the spherical phase as seen by the observer. Clearly, Equations (15)–(17) are general and are applicable to any gas density profile. In fact, Equation (14) is also applicable to any sharply decreasing density gradient, even if it is not a power law. The reason is that m0 does not depend on this profile and so does the relation between γ0 and γf, 0 (Equation (2)), as long as acceleration ends during the planar phase and there is a large energy reservoir behind the breakout shell. Together, Equations (14)–(17) provide three generic observables that depend on only two parameters γf, 0 and R*. Thus, the energy, temperature, and timescale over constrain R* and γf, 0, where any two observables out of these three are enough to determine R* and γf, 0. In case that all three are measured, they must satisfy

Equation (18)

if the source of the flare is a quasi-spherical6 relativistic shock breakout. This relativistic breakout relation can be used to test whether any observed gamma-ray flare is consistent with a relativistic breakout origin.

During t < tobsbo the observer receives also photons from shells that are outer of the breakout shell, which have higher observed temperature than Tbo. These photons generate a high-energy power law at frequencies Tbo < ν. Shells with γi < T'i/T'th ≈ 4 obtain their terminal Lorentz factor before they become optically thin, implying that $\gamma _f = \gamma _i^{\sqrt{3}+1}$ and EEiγfi∝γ−2.01i, while $T = T^{\prime }_{{\rm th}} \gamma _f \propto \gamma _i^{\sqrt{3}+1}$. Since the emission from each shell is spread over a duration tobs∝γ−2f, the luminosity from these outer shells dominates at early time: Lt−0.65obs and Tt−0.5obs. When integrating over the entire breakout flare (t < tobsbo), the spectrum exhibits a high-energy power law

Equation (19)

up to ν = 0.2γ1.750 MeV, which is at most a few times larger than Tbo.

During tobs < tobsbo(= tobss) the observer also receives photons from the breakout shell itself after it becomes transparent. During this phase adiabatic cooling dictates (e.g., Nakar & Sari 2010), Tt−1/3obs and Lt−4/3obs, implying

Equation (20)

The range of the low-energy power law is rather limited $T_{{\rm bo}}>\nu >T_{{\rm bo}}(c t_{{\rm th},0}/R_*)^{1/3}= 0.5 T_{{\rm bo}} \gamma _0^{1/\sqrt{3}} M_{5}^{-1/12} R_{5}^{1/6}$. Note that Equation (19) and the spectral range where the high- and low-energy power laws are observed do depend on the specific density profile and are given for n = 3.

At tobss the observed luminosity is still dominated by light emitted when the breakout shell becomes transparent, ∼tth, 0. This emission fades quickly until the emission from the spherical phase becomes dominant. The fading light curve can easily be derived in case γf, 0 ≫ 1, since then the observed luminosity is dominated by emission of the breakout shell at large angles (>1/γf, 0). The luminosity decays then as Lt−2obs and the temperature as Tt−1obs (Kumar & Panaitescu 2000; Nakar & Piran 2003), until the emission from the spherical phase becomes dominant.

4.2. Spherical Phase

During the spherical phase the radius of the expanding sphere is no longer constant. Instead, the radius of the shells is rvt and the luminosity is dominated by photons that are leaking from the shell where the optical depth to the observer is ∼c/v. We refer to this shell as the luminosity shell and denote its properties with the superscript " $ \, \widehat{ }$  " (see Nakar & Sari 2010 for a detailed discussion), such that, e.g., the luminosity shell satisfies, by definition, $\widehat{\tau }=c/\widehat{v}$. The evolution during the spherical phase depends on whether the luminosity shell is relativistic or not and on the initial temperature of the luminosity shell. As long as $\widehat{\gamma }_f\gg 1$, the initial temperature of the luminosity shell is ∼200 keV and its dynamics is well approximated by the relativistic self-similar solution. Shells with 0.5 < γiβi < 1 also have an initial temperature of ∼200 keV but their dynamics is better described by the Newtonian approximation, namely, shock propagation according to the self-similar solution of Sakurai (1960) and no significant acceleration after the shock breakout. The transition time between the relativistic and Newtonian phases is calculated below and it takes place around:

Equation (21)

During the Newtonian phase $t_{{\rm obs}} \propto \widehat{v}^{-4}$ (Nakar & Sari 2010). The end of the phase in which the initial temperature of the luminosity shell is constant takes place when $\widehat{v}\approx 0.5$c, which is at tobs ≈ 10tobsNW. Here we present luminosity and temperature evolution during the spherical phase up to that point ($\widehat{v}\approx 0.5$c). At later times the temperature drops quickly until thermal equilibrium is obtained. Therefore, there is a sharp break in the temperature evolution at 10tNW. The light curve evolution at t > 10tNW is covered by Nakar & Sari (2010). Figure 1 depicts a schematic sketch of the luminosity and temperature evolution during the spherical phase.

4.2.1. Relativistic Phase (tobss < tobs < tobsNW)

The relativistic velocity of the shells dictates $\widehat{\tau }\sim 1$ and since τ∝m/r2 the mass of the luminosity shell is $\widehat{m}\propto t^{2}$. The photons from the luminosity shell arrive to the observer at $t_{{\rm obs}} \approx t/\widehat{\gamma }_f^2$ and their arrival is distributed over a comparable duration. Therefore the luminosity shell satisfies $\widehat{m}\propto t_{{\rm obs}}^{0.69}$, $\widehat{\gamma }_i\propto t_{{\rm obs}}^{-0.12}$, $\widehat{\gamma }_f\propto t_{{\rm obs}}^{-0.33}$, and $\widehat{E}_i^{\prime } \propto \widehat{m}\widehat{\gamma }_i\propto t_{{\rm obs}}^{0.57}$, while $\widehat{r}\propto t \propto t_{{\rm obs}}\widehat{\gamma }_f^2 \propto t_{{\rm obs}}^{0.34}$ and Equation (10) dictates $\widehat{T}^{\prime } \propto t_{{\rm obs}}^{-0.36}$. The observed luminosity and temperature are

Equation (22)

Equation (23)

The relativistic phase ends once $\widehat{\gamma }_i\approx 1$ at a time given in Equation (30).

4.2.2. Newtonian Phase with Constant Initial Temperature ($t_{{\rm \rm NW}}^{{\rm obs}}<t_{{\rm obs}}<{\it 10}\, t_{{\rm \rm NW}}^{{\rm obs}}$)

The luminosity during this phase is independent of the temperature since pairs are long gone. It is therefore similar to the luminosity evolution of the spherical phase following a Newtonian breakout (e.g., Chevalier 1992), where (for n = 3)

Equation (24)

Therefore a significant flattening of the luminosity evolution is expected at tNW. From this point the luminosity decay is steady until recombination and/or radioactivity starts playing a role.

The initial volume of the luminosity shell is proportional to its initial width (all shells start at R*), $\widehat{d}_i \propto t^{0.44}$. The observed temperature evolves as

Equation (25)

This slope is not very different than that of the relativistic phase. Therefore there is no significant feature in the temperature evolution at tNW. Only at tobs ∼ 10tNW the temperature starts dropping rapidly until the luminosity shell gains thermal equilibrium.

5. PREDICTED SIGNAL OF VARIOUS EXPLOSIONS

We use the model developed in the previous section to calculate the signal expected from known and hypothesized explosions. There are a large number of explosions, such as various SN types, in which the bulk of the mass is moving in Newtonian velocities but the shock accelerates a minute amount of mass to a relativistic breakout. Such explosions are often observed mostly in the optical, where the kinetic energy, E, that is carried by the bulk of the ejected mass, Mej, can be measured. Therefore we first provide the characteristic properties of the breakout and following emission of such explosions as a function of E, Mej, and R*. Then we use these results to discuss the predicted signals from various explosions.

Given E, Mej, and R* the initial Lorentz factor of the breakout shell is determined by following the shock acceleration from the Newtonian phase (γsβs∝ρ−0.19) to the relativistic phase (γsβs∝ρ−0.23):

Equation (26)

where E53 = E/(1053 erg). The numerical coefficient is taken to fit the numerical results of Tan et al. (2001), which simulate the transition of the shock from Newtonian to relativistic. Equation (26) provides a good approximation to the breakout Lorentz factor for β0 > 0.5. The initial density of the breakout shell is ρ0 ≈ 5 × 10−9 g cm−3M0.255R−1.55 and its initial energy is

Equation (27)

and its final Lorentz factor is

Equation (28)

where we use the approximation7 γ0 ≈ γ0β0. Figure 2 shows a color map of the value of γf, 0 as a function of ejected mass and stellar radius for different values of explosion energies, where only the range where our results are valid, 0.5 < γ0β0 < γ0, s, is colored. The figure also includes the typical phase space location of various explosions in that range. It is evident that breakouts of white dwarf explosions, extremely energetic SNe, and possibly low-luminosity GRBs are all within the relevant velocity range.

Figure 2.

Figure 2. Color map of the value of γf, 0 as a function of the ejected mass and stellar radius for different values of explosion energies. Only cases with 0.5 < γ0β0 < γ0, s are colored. The approximate phase space locations of various explosions where the breakout velocity falls in this range are marked as well.

Standard image High-resolution image

Plugging Equation (28) into Equations (14)–(17) we find the observed breakout emission:

Equation (29)

Note that these equations are applicable only to a power-law density profile with n = 3 (unlike Equations (14)–(17), which are general) due to the dependence of γf, 0 on the density profile in case E, Mej, and R* are the physical parameters of interest.

The breakout emission dominates during the entire planar phase. At the end of the planar phase (tobss), the luminosity and temperature drop quickly (Lt−2obs, Tt−1obs if γf, 0 ≫ 1; see the discussion in the previous section) to join the spherical emission, which is dominated by relativistic ejecta until

Equation (30)

where we neglect a very weak dependence on R*. The luminosity and temperature at that time are

Equation (31)

where we take Mej = M*, ignore the very weak dependence8 on Mej/M*, and use the approximation γ0 ≈ γ0β0. The luminosity before and after tobsNW evolves as

Equation (32)

and the temperature as

Equation (33)

The energy emitted per logarithmic scale in time, Ltobs, is rising after tobsNW (∝t0.65obs), and most of the energy that is emitted in X-ray (for typical parameters) is emitted around tobs ∼ 10tobsNW. This energy may be comparable to, or even larger than, the breakout energy:

Equation (34)

At a later time, the luminosity evolution remains unchanged (until recombination or radioactive decay becomes dominant), while the temperature drops rapidly toward thermal equilibrium (Nakar & Sari 2010).

5.1. White Dwarf Explosions: Type Ia SN, Type .Ia SN, and AIC

There are several mechanisms that can result in a shock breakout from a white dwarf. The most famous is Type Ia SNe, where the whole star explodes. Other theoretically predicted scenarios are Type .Ia SNe (Shen et al. 2010), where helium accreted in an AM CVn is detonated, and AIC to a neutron star (e.g., Hillebrandt et al. 1984; Fryer et al. 1999; Dessart et al. 2006). All these explosions are expected to produce a rather constant ejecta velocity (2E/Mej)1/2 ∼ 10, 000 km s−1, with explosion energy that ranges between 1051 erg in Type Ia SNe, where the whole star is ejected (Mej ≈ 1.4 M), and 1049–1050 erg in the other explosions, where only 1%–10% of the stellar mass is ejected. There are also various observed SNe, which are probably generated by a range of white dwarf explosions. Some possible examples are SN 2005E (Mej ≈ 0.3 M, E ≈ 4 × 1050 erg; Perets et al. 2010), SN2010X (Mej ≈ 0.16 M, E ≈ 1.7 × 1050 erg; Kasliwal et al. 2010), and SN 2002bj (Mej ∼ 0.2 M, E ∼ 1050 erg; Poznanski et al. 2010).

Since E/Mej is similar for all these explosions and the dependence of γ0β0 on E (or Mej) when this ratio is constant is very weak (∝E0.17), shock breakouts from all these types of white dwarf explosions produce a similar signature. The breakout Lorentz factor is γ0β0 ≈ 1–3 for a white dwarf radius 3–10 × 108 cm. The total energy released in the breakout is 1040–1042 erg and the typical photon energy is ∼MeV. The breakout pulse is spread over ∼1–30 ms, having a peak luminosity of ∼1044 erg s−1. These results for the breakout emission are different, especially in the temperature prediction, than those of Piro et al. (2010b) who found a Type Ia breakout signal that peaks in the X-rays. The reason is that they assume thermal equilibrium behind the shock, which is not a good assumption for Type Ia SNe (Nakar & Sari 2010), and that they ignore relativistic effects.

Note that this signature is expected only if the line of sight to the observer is transparent. In case the explosion is triggered by white dwarf mergers, the optical depth of the debris that surrounds the exploding core can be high and the shock breakout may take place at a much larger radius than the white dwarf radius, carrying out much more energy at a much lower velocity. For example, Fryer et al. (2010) find that a Type Ia SN shock breakout in the case of a double degenerate merger takes place at a radius of ∼1013–1014 cm, where the shock is Newtonian. Thus, a breakout emission from a Type Ia SN can potentially differentiate between single and double degenerate progenitor systems.

A flare of 1041 erg in MeV γ-rays can easily be detected if it takes place in the Milky Way and possibly also in the Magellanic Clouds. Such detection will provide a constraint on the exact explosion timing as well as on the radius of the exploding white dwarf and explosion energy. It may also shed light on the explosion mechanism. This detection channel is extinction free and may be the main detection channel in case the burst takes place in an obscured Galactic environment. Given the estimated Milky Way rate of Type Ia SNe (∼0.01 yr−1), the chance to detect such a burst in the coming decade is low, but not negligible. The rate of events such as SN 2010X is very hard to constrain observationally, based on current optical surveys. In addition, there are no observational constraints on the rate of events with energy lower than 1050 erg, since their optical signatures, where these events are searched for, are too weak and evolve too fast for detection. Similarly there are no tight theoretical constraints on the rates of Type .Ia SNe and AICs. Since the gamma-ray shock breakout signatures of all these events are similar and detectable for any galactic event, it is worth looking for them in the current data of satellites such as BATSE, Konus-Wind, Swift, and Fermi. In fact, Cline et al. (2005) find that the sub-sample of several dozen BATSE and Konus-Wind GRBs, those with a duration shorter than 0.1 s, shows a significant anisotropy with concentration of events in the galactic plane (but not toward the Galactic center). This sample of events may contain some already observed shock breakouts from white dwarfs.

5.2. Broad-line Type Ibc SNe

Broad-line Type Ibc SNe are thought to be explosions of Wolf–Rayet progenitors, and in some cases are very energetic (E ≫ 1051 erg). Some Broad-line Type Ibc SNe are associated with GRBs (these are discussed later), but also those that are not associated with GRBs may produce a detectable γ-ray signal. For example SN 2002ap has a total explosion energy of ∼4 × 1051 erg and an ejected mass of ∼2.5 M (Maurer & Mazzali 2010), implying a mildly relativistic breakout with an energy output of ∼5 × 1044R0.75 erg within ∼10R5 s. At a distance of about 7 Mpc, the observed fluence on Earth is ∼10−7R0.75 erg cm−2. This fluence is detectable by sensitive gamma-ray detectors even for R* = R (although no large field-of-view detector with high sensitivity was operational in 2002). Note, however, that if a dense wind surrounds the progenitor, the flare properties may be very different. The total emitted energy is higher, while the typical photon energy may fall either within or below the observing band of soft γ-ray detectors.

5.3. Extremely Energetic SNe

Several types of extremely energetic SNe were detected in recent years. SN 2007bi has shown a typical velocity of 12, 000 km s−1 and >50 M of ejecta implying E ≳ 1053 erg (Gal-Yam et al. 2009). Most of the optical luminosity observed in this SN is generated by radioactive decay. The radius of the progenitor is unknown, but since there is no trace of either hydrogen or helium its radius is most likely in the range, R* ∼ 1011–1012 cm. These values imply γ0β0 ≈ 1. Therefore, the breakout is expected to produce a ∼100 keV flash that lasts seconds to tens of seconds and carries an energy of 1044–1046 erg.

Quimby et al. (2011) report on another type of extremely energetic SNe. They find a typical velocity of 14, 000 km s−1 and >5 M of ejecta implying E ≳ 1052 erg. The observed optical energy in these explosions is not generated by radioactive decay and the possible sources are interaction of the ejecta with matter that moves more slowly or a central engine. In the former case, all the energy is given to the ejecta during the explosion and a bright shock breakout is expected. These explosions also do not show any trace of hydrogen and helium implying that their progenitor is also rather compact. Assuming again R* ∼ 1011–1012 cm the signature of the shock breakout from these events is similar to that expected from SN 2007bi.

An energy output of 1044–1046 in soft gamma rays within a minute is detectable out to a distance of 3–30 Mpc. Therefore the detection of breakouts from such events is likely only if their rate is larger than 104 Gpc−3 yr−1. The rate of extremely energetic SNe is unknown, but it is most likely much lower than that since otherwise they should have been detected in optical searches in large numbers. We therefore do not expect current gamma-ray detectors to detect any of these SN types, although nature may surprise us.

5.4. Low-luminosity GRBs

The engine of long GRBs is thought to be produced during the collapse of a massive stellar core and thus to be launching relativistic jets into a stellar envelope. Bromberg et al. (2011b) show that in (at least some) low-luminosity GRBs, it is highly unlikely that the jets successfully punch through the star and emerge in order to produce the observed γ-ray emission. In this case, what can be the source of the observed gamma rays?

There are several common features observed in all low-luminosity GRBs. All show 1048–1050 erg of gamma rays or hard X-rays that are emitted in single pulsed light curves showing spectra that become softer with time (see Kaneko et al. 2007 and references therein). They also show radio afterglows that indicate on a similar energy in mildly relativistic ejecta (Kulkarni et al. 1998; Soderberg et al. 2004, 2006). This energy is only a small fraction of the total energy observed in the associated SNe (1052–1053 erg). The smooth light curves and the energy coupled to a mildly relativistic ejecta motivated several authors to suggest that the origin of this emission is shock breakouts (Kulkarni et al. 1998; Tan et al. 2001; Campana et al. 2006; Waxman et al. 2007; Wang et al. 2007). Later, Katz et al. (2010) showed that the deviation from thermal equilibrium can explain the observed γ-rays, compared with the expected X-rays when equilibrium is wrongly assumed. However, the question of whether shock breakouts are indeed the source of low-luminosity GRBs remained open, mostly since there was no quantitative model that could answer whether shock breakouts can indeed explain even the most basic observables, such as energy, temperature, and timescales, of these bursts. Below, we use our model to answer this question. In Section 4, we have shown that these three observables, Ebo, Tbo, and tobsbo, overconstrain the physical parameters since they depend in the case of a quasi-spherical relativistic breakout only on two parameters, γf, 0 and the breakout radius. Thus, in addition to finding the value of these two physical parameters, the three observables must satisfy the relativistic breakout relation (Equation (18)). This can be used as a test that any flare that originates from quasi-spherical relativistic breakout must pass. This result is generic as it is independent of the pre-breakout density profile (as long as it is steeply decreasing). Below we apply this test to the four observed low-luminosity GRBs.

The four well-studied low-luminosity GRBs are divided into two pairs with very different properties. GRBs 060218 and 100316D are soft and very long. Their rather soft spectrum is best fitted at early time with a cutoff around a peak energy of ∼40 keV (Kaneko et al. 2007; Starling et al. 2011) and both emitted a total energy of ∼5 × 1049 erg. Plugging these values into Equation (18) we find that if these bursts are flares from quasi-spherical breakouts then they should have durations of ∼1500 s, which is indeed comparable to the observed durations of these two bursts. The breakout radius in this model is ∼5 × 1013 cm and γf, 0 ≈ 1 implying γ0β0 ≈ 0.3–1. In the case of GRB 060218, radio emission suggests breakout velocities on the upper side of this range (Soderberg et al. 2006). The large radius inferred by the long duration implies that the breakout is most likely not from the edge of the progenitor star, but from an opaque material thrown by the progenitor to large radii (possibly a wind) prior to the explosion (Campana et al. 2006). The fact that a spherical model provides a good explanation implies that the suggested a-sphericity (Waxman et al. 2007), if exists, does not play a major role in determining the burst duration. Note though that it may play an important role in shaping the breakout spectrum above and below Tbo and the entire light curve following the breakout flare.

The other two low-luminosity GRBs, 980425 and 031203, are relatively hard, T ≳ 150 keV, and not particularly long, ≈30 s. Their hard spectrum made it difficult to explain these events as shock breakouts, before it was realized that the radiation will be away from thermal equilibrium.9 Plugging 1048 erg and 150 keV, the total energy and typical frequency of GRB 980425, respectively, into Equation (18) results in a breakout duration of ∼10 s, comparable to the observed one. The breakout radius in that case is ∼6 × 1012 cm and the final Lorentz factor is10 γf, 0 ≈ 3. The associated SN of GRB 980425, SN 1998bw, was very energetic, exhibiting Mej ≈ 15 M and a kinetic energy of ≈0.5 × 1053 erg (Iwamoto et al. 1998). Interestingly, setting R* ∼ 6 × 1012 and Mej = 15M in Equation (26) we find that an explosion energy of E ≈ 3 × 1053 produces a breakout with the observed properties of GRB 980425. This energy is slightly larger than the one seen in the SN 1998bw, but it is close enough to suggest that the energy source of the SN explosion is also the one of the shock breakout. The mild discrepancy between the two energies may be a result of the actual stellar density profile (compared to the power law with n = 3) or of a deviation of the explosion from sphericity. Using the energy and typical frequency of GRB 031203, 5 × 1049 erg, and >200 keV, Equation (18) predicts a duration ≲ 35 s. The observed duration, ∼30 s, is consistent with this value and indicates that the actual temperature is not much higher than 200 keV. The breakout radius in that case is ∼2 × 1013 cm and the final Lorentz factor is γf, 0 ≈ 5. The SN associated with GRB 031203, SN 2003lw, is similar in ejected mass and slightly more energetic than SN 1998bw (Mazzali et al. 2006). Setting R* ∼ 2 × 1013 and Mej = 15 M in Equation (26) we find that an explosion energy of E ≈ 1054 produces a breakout with the observed properties of GRB 031203. Again this value is larger by about an order of magnitude compared with the one observed in the associated SN.

The breakout radii that we find for GRBs 980425 and 031203 are larger than those of typical Wolf–Rayet stars, the probable progenitors of GRB–SNe. This implies that either the breakout is not from the edge of the progenitor star (e.g., from a wind) or that low-luminosity GRB progenitors are larger than commonly thought (which may help in choking the relativistic jet; Bromberg et al. 2011d). Finally, a prediction of relativistic shock breakouts is a significant X-ray emission that can be comparable in energy to that of the breakout γ-ray flare. The X-rays are emitted during the spherical phase, over a timescale that is significantly longer than that of the breakout emission. GRB 031203 has shown a bright signal of dust scattered X-rays, which indicates ∼2 keV emission with energy comparable to that of gamma rays, during the first 1000 s after the burst (Vaughan et al. 2004; Feng & Fox 2010). These values fit the predictions of a breakout model. For example, if the breakout is from a stellar surface at R* = 2 × 1013 and Mej = 15 M, then Equations (30)–(34) predict a release of ∼1050 erg at ∼10 keV within the first hour after the burst.

To conclude, we find that the energy, temperature, and timescales of all four low-luminosity GRBs are explained very well by shock breakout emission and that all of them satisfy the relativistic breakout relation between Ebo, Tbo, and tobsbo (Equation (18)). These observables are largely independent of the pre-shocked density profile and are therefore applicable also for a breakout from a shell of mass ejected by the progenitor prior to the explosion. Breakout emission can also explain many other observed properties, such as the smooth light curve, the afterglow radio emission, the small fraction of the explosion energy carried by the prompt high-energy emission, and the delayed bright X-ray emission seen in GRB 031203. We therefore find it to be very likely that indeed all low-luminosity GRBs are relativistic shock breakouts.

5.5. GRB 101225A

GRB 101225A is a peculiar explosion showing ∼an hour long smooth burst of gamma rays and X-rays followed by a peculiar IR-UV afterglow, which is consistent with a blackbody emission. Its origin is unclear. Thöne et al. (2011) argue for a cosmological origin and estimate its redshift to be ≈0.3, while Levan & Tanvir (2011) argue that its origin is local. Here we do not attempt to determine which, if any, of the two views is correct. Instead, given the similarity of the high-energy emission of GRB 101225A to some low-luminosity GRBs, we ask whether the shock breakout can be the source of this emission, assuming that the burst is cosmological and that the redshift estimate of Thöne et al. (2011) is correct.

We test whether its observed properties satisfy Equation (18), which with a total energy of ∼1051 erg and gamma-ray spectrum that peaks at ∼40 keV predicts a duration of ∼104 s. This value is consistent with the constraints on the actual duration of the flare (Thöne et al. 2011). The resulting breakout radius is ∼3 × 1014 cm. Interestingly Thöne et al. (2011) present a model that attempts to explain the IR-UV afterglow by an expanding opaque shell with velocities at least as high as β ≈ 1/3 and an initial radius of 2 × 1014 cm. A breakout of a shock that accelerates such a shell produces a high-energy signal that is similar to the observed one. We therefore conclude that if Thöne et al. (2011) redshift estimate is correct, then a relativistic shock breakout probably provides the simplest explanation for the high-energy emission.

5.6. Long Gamma-ray Bursts

In the collapsar model of long GRBs, the relativistic jets pierce through the envelope and emerge in order to produce the observed gamma-ray emission. During their propagation the jets drive mildly relativistic bow shocks into the envelope and inflate a cocoon that collimates the jets to an even narrower angle than their initial launching opening angle (e.g., MacFadyen & Woosley 1999; Matzner 2003; Morsony et al. 2007; Mizuta & Aloy 2009). Recently, Bromberg et al. (2011c) have derived an analytic description of this interaction while the jet is propagating deep in the envelope, before it encounters the sharp density gradient near the stellar edge.

The term shock breakouts from GRBs is used in the literature to describe several different phenomena. Here, similarly to the rest of the paper, we focus on the emission that is escaping from the breakout layer after the forward shock, which is driven by the jet head and the cocoon into the stellar envelope, breaks out. The structure of the shock that is driven into the stellar envelope is highly non-spherical, but at the time of the breakout the size of causally connected regions is small and a local spherical approximation is appropriate. Thus, the theory discussed above, together with the model presented in Bromberg et al. (2011c) and the results seen in numerical simulations (e.g., Mizuta & Aloy 2009), can be used to obtain a rough idea of the expected breakout signal.

We consider a typical GRB jet with a half-opening angle θ0 = 0.1 rad and a total luminosity of 1050 erg s−1. The jet is launched continuously into a 5 M progenitor with a radius of 5 R. At the point that the jet reaches the steep density drop (i.e., rR*/2) its velocity approaches the speed of light and it is narrowly collimated to an angle of ∼θ20 (Bromberg et al. 2011c). The opening angle of the cocoon at this point is ≈θ0. From that point the jet head and the cocoon shock start accelerating toward the stellar edge and to expand sideways. Numerical simulations show that the head is expanding sideways, becoming comparable in size to the cocoon, which does not spread significantly (e.g., Figure 17 in Mizuta & Aloy 2009). Once the fractional distance of the shock to the edge, z/R*, is comparable to the shock opening angle, in our case ∼θ0 = 0.1, the shock loses causal connection with the sides and it can be approximated by the spherical solution of an accelerating blast wave (Johnson & McKee 1971). Taking γsβs(z = 0.1 R*) ≈ 1 and using the relation γsβsz−0.69 with Equation (5), we obtain γ0 ≈ 10. This value is larger than γ0, s so the breakout shell accelerates also after the transition to the spherical phase. We do not provide here the solution to such a case, but the initial energy in the breakout shell is ∼1048(R*/5 R)2 erg and its energy increases during acceleration by at most an order of magnitude. Assuming that there is no significant wind surrounding the progenitor, this energy is released by a short, ∼ ms, pulse of >MeV photons.

The main distinguishable feature of the breakout pulse is its harder spectrum and low fluence compared with the total burst emission. The pulse luminosity depends strongly on the progenitor size. If the progenitor is rather large, R* ≳ 5 R, then the breakout can be detected by Swift and Fermi up to a redshift ∼0.1. If it is much smaller, then the breakout is too faint for detection. Even in case the breakout is detectable, it may be difficult to separate it from the GRB itself. Nevertheless, it is worth looking for an initial hard spike in nearby GRBs. Interestingly, there is a sub-sample of BATSE GRBs that show initial short hard spikes followed by a softer burst (Norris & Bonnell 2006).

6. SUMMARY

We calculate the emission from mildly and ultrarelativistic spherical shock breakouts. First, we determine the pre- and post-breakout hydrodynamical evolution. This requires knowing the shock width, which we obtain by deriving an improved analytic description to the structure of the transition layer in relativistic radiation-mediated shocks, based on the results of Budnik et al. (2010). After obtaining the hydrodynamical evolution we follow the temperature and opacity evolution to find the observed light curve. Our calculations are applicable to stellar explosions with shock velocity larger than 0.5c, in which the breakout shell ends its acceleration during the planar phase, i.e., its final Lorentz factor ≲ 30 for typical stars. We consider only cases where the progenitor wind density is low enough, so it has no effect on the observed emission.

A relativistic radiation-mediated shock brings the gas to a roughly constant rest-frame temperature, ∼200 keV, and loads it with pairs (Katz et al. 2010; Budnik et al. 2010). Photons remain confined to the post-shock expanding gas as long as pairs keep it optically thick. A significant number of photons are released toward the observer only from optically thin regions, which has a gas rest-frame temperature ≲ 50 keV (low pair load) and unloaded pair gas opacity, τ, smaller than unity. Our solution of the shock structure shows that the shock width is significantly smaller than τ = 1, implying that the shock also accelerates shells with τ ⩽ 1, which become transparent once their pairs annihilate. Among these shells the one that carries most of the energy is the most massive one, i.e., τ ≈ 1, which we refer to as the breakout shell. In the regime that we consider, the breakout shell becomes transparent during the planar phase (i.e., before its radius doubles), or soon after the transition to the spherical phase, and after it achieves its final Lorentz factor, light travel time and relativistic effects dictate that the emission of the breakout shell, once it becomes transparent, dominates the energy release during the entire planar phase. Slower shells release their photons only during the spherical phase. The processes described above produce the following main observables.

  • 1.  
    A γ-ray flare with a typical photon energy that ranges between ∼50 keV and ∼2 MeV. The flare typically contains a small fraction of the explosion energy and it may be significantly shorter than the progenitor light crossing time. This flare is generated by the breakout shell when it becomes transparent.The energy, temperature, and duration of the flare depend only on two parameters, γf, 0 and the breakout radius. Thus, in addition to providing the value of these two physical parameters, the three observables must satisfy the relativistic breakout relation (Equation (18)). This relation can be used to test whether an observed flare may have been produced by a quasi-spherical relativistic breakout. Note that our calculation of the breakout flare energy, temperature, and duration is independent of the exact profile of the pre-breakout density, as long as it is steeply decreasing.
  • 2.  
    Due to light travel time effects, the breakout flare contains photons from shells that are faster and lighter than the breakout shell. These shells dominate the luminosity at early times, but not the energy released in the breakout flare (which is dominated by the breakout shell). During the planar phase the luminosity evolves as t−0.65 and the temperature as t−0.5. The breakout flare-integrated spectrum shows a high-energy power-law spectrum, νFν∝ν−0.74. It also contains photons emitted by the breakout shell after it becomes transparent and cools adiabatically, producing a low-energy power-law spectrum, νFν∝ν. Both power laws are limited to about one order of magnitude in frequency.
  • 3.  
    The flare ends with a sharp decay at a time that coincides with the transition to the spherical phase. If the breakout is ultrarelativistic (i.e., γf, 0 ≫ 1) then Lt−2obs and Tt−1obs, until the spherical phase emission dominates.
  • 4.  
    After the flare ends, spherical evolution dictates a steady decay of the luminosity. At first, while relativistic shells dominate the emission, Lt−1.1obs. Later, at tobsNW, once Newtonian shells become transparent Lt−0.35obs. Note that during the latter phase the total emitted energy increases with time.
  • 5.  
    The post-flare temperature decays at first at a steady rate of about t−0.6obs, generated during the spherical phase. A sharp drop in the temperature is observed when shells that were not loaded by pairs during the shock crossing (vs ≲ 0.5) dominate the emission. The drop is typically from the X-ray range to the UV, and it is observed at about 10tobsNW.
  • 6.  
    The energy emitted before the steep temperature drop is often comparable to that emitted during the breakout flare. Thus, a signature of relativistic breakouts in many scenarios is a bright and short γ-ray flare and a delayed X-ray emission with comparable energy.

We apply our model to a range of observed and hypothesized explosions finding the following predictions.

  • 1.  
    Type Ia SNe and other white dwarf explosions. A range of white dwarf explosions, including Type Ia SNe and peculiar events such as SNe 2002bj, 2005E, and 2010X, as well as hypothesized Type .Ia SNe and AIC, are all predicted to produce a similar breakout emission. A ∼1–30 ms pulse of 1040–1042 erg is composed of ∼ MeV photons. Such a pulse from a Galactic explosion is detectable, but implies a low, yet not negligible, chance to detect such a flare from Type Ia SNe in the coming decade. The rate of the other types of explosions is not known. Moreover, there is almost no correlation between the properties of the γ-ray flare and the luminosity of the optical emission, implying that Galactic breakouts from white dwarf explosions may have already been observed, but not recognized as such. In this context, the suggested sub-class of very short GRBs (Cline et al. 2005) may contain such events.
  • 2.  
    Broad-line Type Ib/c SNe. Their progenitor is compact enough and the energy of some is high enough to form mildly relativistic breakouts. For example, SN2002ap-like events are predicted to produce a detectable breakout γ-ray signal in case there is no thick wind surrounding the progenitor.
  • 3.  
    Extremely energetic SNe. SNe such as 2007bi (Gal-Yam et al. 2009) and explosions of the type reported by Quimby et al. (2011) are energetic enough to generate mildly relativistic breakouts in very massive stars. These breakouts are likely to release 1044–1046 erg in soft γ-rays over seconds to minutes. Such flashes are detectable in the nearby universe, however these are rare events, and the chance to have the one in a detectable distance is low.
  • 4.  
    Low-luminosity GRBs. The origin of low-luminosity GRBs is unknown. Bromberg et al. (2011b) show that low-luminosity GRBs are very unlikely to be produced by relativistic jets that pierce through their host star, which is a necessary ingredient of the collapsar model for long GRBs. In that case what can be the source of the high-energy emission? The smooth light curve and the mildly relativistic ejecta that generate the radio emission of GRB 980425 lead Kulkarni et al. (1998) and Tan et al. (2001) to suggest that relativistic shock breakouts are related to low-luminosity GRBs. The detection of a thermal component in GRB 060218 (Campana et al. 2006) and of additional low-luminosity GRBs with similar properties to GRB 980425 supported this suggestion (Waxman et al. 2007; Wang et al. 2007), especially after it was realized that the shock emission strongly deviates from thermal equilibrium (Katz et al. 2010). However, there was no quantitative model of the breakout emission that includes deviation from thermal equilibrium and gas-radiation relativistic dynamics, which could test this suggestion. Moreover, it was unclear whether shock breakouts can generate the energies seen in these bursts, especially GRBs 980425 and 031203 due to their shorter duration (Katz et al. 2010).We find that relativistic shock breakouts can explain very well the energy, temperature, and timescales of the prompt emission of all observed low-luminosity GRBs. Moreover, all four known bursts satisfy the relativistic breakout relation between these three observables (Equation (18)). We also show that relativistic breakouts provide a natural explanation with the low gamma-ray luminosity compared for the total explosion energy and for the hard to soft spectral evolution (including the late energetic X-ray emission that inferred from dust echoes of GRB 031203). We therefore find shock breakouts to be the most promising sources of all low-luminosity GRBs observed so far. The connection between low-luminosity and long GRBs (both share a relation to a similar type of SNe) suggests that the energy source for the shock breakout is choked relativistic jets that are launched by a GRB engine but fail to punch through the star (Bromberg et al. 2011d).In this context we examine the high-energy emission of GRB 101225A, whose origin (local or cosmological) is undetermined yet, but its prompt emission resembles low-luminosity GRBs. We find that if the redshift determination of Thöne et al. (2011), z ≈ 0.3, is correct then this burst also successfully passes the test set by Equation (18), suggesting that the source of its high-energy emission is a mildly relativistic shock breakout at a relatively large radius of ∼3 × 1014 cm.
  • 5.  
    Long GRBs. According to the popular view of long GRBs, relativistic jets drill their way through the progenitor envelope, driving a forward shock into it. We estimate that a typical value of the initial breakout Lorentz factor of this shock is γ0 ∼ 10. Our theory cannot determine its final Lorentz factor, since it ends its acceleration after the evolution enters the spherical regime. Nevertheless, this value of γ0 indicates that the shock breakout from long GRBs releases ∼1048(R*/5 R)2 erg in a short pulse of >MeV photon. This pulse is too dim to be observed in the typical redshift of long GRBs, but it may be observable in nearby z ∼ 0.1 GRBs.

We conclude that relativistic shock breakouts are a generic process for the production of gamma-ray flares, which opens a new window for the study and detection of a variety of stellar explosions.

We thank Omer Bromberg, Amir Levinson, Hagai Netzer, and Tsvi Piran for helpful discussions. E.N. was partially supported by an ERC grant and by the Israel Science Foundation (Grant No. 174/08). R.S. was partially supported by ERC and IRG grants, and a Packard Fellowship.

APPENDIX A: THE TRANSITION LAYER STRUCTURE OF RELATIVISTIC RADIATION-MEDIATED SHOCKS

Here we derive an analytic description of the transition layer of relativistic radiation-mediated shocks, based on the results of Budnik et al. (2010). We use mostly similar notations to those of Budnik et al. (2010) and carry out the calculation in the shock frame, where the flow is assumed to be in a steady state. We approximate the gas radiation in the transition region as two counter flows. One is a flow of photons with energy ∼mec2 that are generated in the immediate downstream and flow toward the upstream at a drift velocity ∼c. The other is a gas flow that is moving toward the downstream. This is a flow of the fluid that is shocked, which is moving with a Lorentz factor Γu in the far upstream (this is the shock Lorentz factor as seen in the upstream frame) and is slowed down in the transition region to β = 1/3 (i.e., Γ ≈ 1). The downstream moving fluid is composed of protons with a proper density n and pairs11 with a proper density n±. We are interested in the shape of the transition, i.e., Γ and the pair fraction as a function of optical depth and physical length. Our final goal is to find the shock width in the upstream frame in units of the pre-shock (i.e., with no pairs) upstream Thomson optical depth.

The shock is generated by the interaction of the two counter-streaming flows. Every collision between an upstream moving photon and a downstream moving fluid lepton transfers energy and momentum to the photon and adds the upstream moving photon to the downstream moving fluid. The temperature of the pairs is ∼Γmec2, as seen in the fluid rest frame. In a steady state the flux of photons that are moving toward the upstream, nγ → us, is equal to the number flux of pairs that are moving toward the downstream (assuming that the latter dominate over the proton number flux). As long as Γ ≫ 1 this steady state equality implies nγ → us ≈ Γn±. Now, since an upstream moving photon that is scattered is joining the downstream moving fluid, dn±) = −nγ → usdτs, where τs is the optical depth for photons moving toward the upstream, defined to be increasing in the upstream direction.12 Defining x±n±/n we obtain

Equation (A1)

Number and energy flux conservation implies (assuming Γ ≫ 1)

Equation (A2)

Equation (A3)

In the energy flux equation, we assume that the energy flux of the photon field that flows toward the upstream is negligible, which is true as long as Γ2 ≫ 1. We also assume x± ≫ 1. These equations imply

Equation (A4)

and therefore

Equation (A5)

Define τ* as the Thomson optical depth of a photon moving toward the upstream then τ* ≈ τsTKN2)), where σT is the Thomson cross section and σKN2) is the Klein–Nishina cross section of electron at rest and photon with energy Γ2mec2. Thus,

Equation (A6)

Integrating this equation starting at the end of the transition layer, i.e., Γ(τs = 0) = 1 results in the structure of the transition layer. Comparison of this structure to the numerical results of Budnik et al. (2010) is presented in Figure 3.

Figure 3.

Figure 3. Comparison of the shock structure for four values of Γu (6, 10, 20, and 30), obtained by integrating over Equation (A6) (using τ* = τsTKN(1.4Γ2))) (thin solid line), to the numerical results of Budnik et al. (2010, their Figure 6; thick dashed line).

Standard image High-resolution image

In order to find the width of the shock in the upstream frame we use the relation

Equation (A7)

where z' is the length coordinate (positive toward the upstream) in the shock frame. Therefore

Equation (A8)

Integrating dΓ from the shock toward the upstream, the value of Γ/Γu (for values larger than 1/Γu) has a universal profile (independent of Γu) as a function of nuσTz'(1 + 2ln [Γ2u])/Γ2u. This profile is presented in Figure 4. It implies that the shock width in units of pair unloaded Thomson optical depth in the shock frame is roughly,13 ∝Γ2u. The value of the proportionality coefficient is somewhat arbitrary and it depends on the value of Γ/Γu that defines the "shock width." We chose Γ/Γu = 0.9, obtaining nuσTz' ≈ 0.01Γ2u (the value of the logarithmic factor is 5 − 10 for Γu = 2 − 10). Choosing a different value of Γ/Γu, to define the shock width, changes the coefficient, which in turn has a very weak effect on Equation (8). Now, since the length scale in the upstream frame, z, is shorter by a factor Γu than in the shock frame (i.e., z' = Γuz), and defining τu = nuσTz, we approximate the shock width in units of Thomson optical depth of the pre-shock upstream (i.e., with no pairs) as

Equation (A9)

Note that our analytic description is applicable to shocks where Γ2u ≫ 1 and is therefore not directly applicable to mildly relativistic shocks. We do expect, however, that pair creation in these shocks, which increases the optical depth per proton in the immediate downstream by a factor of the order of mp/me, will result in Δτu ≪ 1 and that Equation (A9) can be extrapolated to the mildly relativistic regime.

Figure 4.

Figure 4. Structure of the shock transition region of relativistic radiation-mediated shocks, presented in an optical depth scale (nuσTz'(1 + 2ln [Γ2u])/Γ2u), which makes it universal, i.e., the value of Γ/Γu is independent of Γu (for values larger than 1/Γu).

Standard image High-resolution image

When deriving Equation (A9) we assumed a steady state. A relativistic shock with Lorentz factor Γu and a width Δz (as seen in the upstream frame) contains gas collected over a length of ΔzΓ2u. Since near the edge of the star the density varies over a scale of the order of distance to the edge, z, the assumption of a steady state is valid as long as z > ΔzΓ2u or τ > ΔτuΓ2u ∼ 0.01Γ3u, where τ is the pre-shock Thomson optical depth to the stellar edge. As the shock propagates farther, the assumption of steady state breaks down and its structure must be modified. The shock may still be confined during this phase, which we do not attempt to address here. Our result shows that the shock must be confined to the star at least up to optical depth as low as τ ∼ 0.01Γ3u.

APPENDIX B: GLOSSARY OF MAIN SYMBOLS AND NOTATIONS

General variables:

  • t: time since breakout.
  • r: radius.
  • v: velocity.
  • β: velocity in units of the speed of light.
  • γ: Lorentz factor.
  • βs, γs: shock β and Lorentz factor.
  • m(r): mass at radius larger than r.
  • ρ: pre-shock stellar mass density.
  • d: post-shock shell width.
  • z: pre-shock distance to the stellar edge.
  • τ: pre-shock (pair un-loaded) Thomson optical depth to the stellar edge.
  • κT: Thomson cross section per unit of mass. Throughout the paper we use the value 0.2 cm2 g−1. The results are weakly sensitive to the exact value.
  • E: internal energy (not to be confused with E53).
  • L: observed luminosity.
  • T: temperature. Defined as the typical photon energy.

Notations and specific values:

  • breakout shell: the shell that dominates the energy released in the breakout flare (τ = 1).
  • outermost shell: the outermost shell that is accelerated.
  • luminosity shell: the shell that generates the observed luminosity.
  • For any quantity x, we use the following subscripts and superscripts:x: any quantity with no subscript or superscript is measured in the lab frame.xobs or xobs: an observer frame value (when different than the lab frame value).x': a rest frame value.xi: initial value, after shock crossing, of a shell.x0: a value at the breakout shell. For time-dependent quantities it is the value at the time of the breakout.$ \widehat{x}$: a value at the luminosity shell.xbo: an observed value of the breakout flash.
  • γ0: the Lorentz factor of the breakout shell right after the shock crossing.
  • γf, 0: the Lorentz factor of the breakout shell after acceleration ends.
  • γ0, sf, s]: the critical initial (final) Lorentz factor of a breakout shell that ends acceleration at the transition to the spherical phase.
  • γi, max: the initial Lorentz factor of the outermost shell, which is also the fastest initial Lorentz factor in the system.
  • ti: the initial time for shell expansion, which is also the time between the shock crossing and the shock breakout.
  • t0: the initial time for expansion of the breakout shell.
  • tf: the time acceleration ends.
  • tf, 0: the time that acceleration of the breakout shell ends.
  • ts: the transition time from planar to spherical geometry.
  • tth, 0: the time that the breakout shell becomes optically thin.
  • tobsNW: the observer time of transition of the luminosity shell from relativistic to Newtonian dynamics.
  • Tth: the temperature where the pair fraction becomes insignificant.
  • R*: the stellar radius.
  • M*: the stellar mass.
  • Mej: the ejecta mass.
  • R5: the stellar radius in units of 5 R.
  • M5 [Mej, 5]: the stellar (ejecta) mass in units of 5M.
  • E53: the explosion energy in units of 1053 erg.
  • n: the power-law index describing the pre-explosion stellar density profile near the edge. Throughout the paper we use n = 3.

Footnotes

  • Throughout the paper we use this value, which is appropriate for hydrogen free ionized gas. The dependence on the value of κT is weak and all our results are applicable also to hydrogen-rich ionized gas. Note that through most of the phases discussed here the temperatures in the system are larger than 10 keV, where the gas is fully ionized regardless of its composition, implying that the results are insensitive to metallicity.

  • We assume here that the shock breaks out as soon as its structure is affected by the proximity to the edge (i.e., the assumption of steady state used in Appendix A breaks down). If the shock remains confined to the star also beyond this point, i.e., z < 0.01γ3i/(ρκT), then γmax, i is larger than the value given in Equation (8).

  • This is true as long as the blackbody temperature that corresponds to the post-shock energy density is lower than T'i.

  • Ebo, Tbo, and tobsbo depend only on emission from a region within an angle of ∼1/γf, 0 with respect to the line of sight.

  • This approximation is good for γ0β0 > 2. For lower values of γ0β0 it is better to calculate $\gamma _{f,0}=\gamma _0^{1+\sqrt{3}}=(\gamma _0^2 \beta _0^2+1)^\frac{1+\sqrt{3}}{2}$ and plug it into Equation (29).

  • In some explosions Mej = M*, while in others Mej < M*. In any case the observations in all the regimes we discuss here depend on Mej and are almost independent of M*.

  • For example, Tan et al. (2001), who consider the shock breakout as the source of energy of GRB 980425, add an interaction of the expanding gas with external mass shell, at a radius where the gas is optically thin, in order to reprocess the gas energy into internal energy and radiate it non-thermally. Wang et al. (2007) suggested that the gamma rays are produced by bulk comptonization of low-energy thermal photons during the breakout. A process that does not work at the relevant temperatures due to Compton loses (Katz et al. 2010).

  • 10 

    Although this value of the Lorentz factor is below the lower limit derived by Lithwick & Sari (2001) for GRB 980425, there is no contradiction between the two studies. The reason is that Lithwick & Sari (2001) derived their limits assuming that there is a hard spectral power law that extends to very high energies, similarly to the one observed in long GRBs. However, such a power law does not exist in the case of relativistic breakout and therefore their constraints are not applicable to the model that we discuss here.

  • 11 

    Protons and pairs are assumed to be coupled by collective plasma processes on scales much shorter than any scale on interest in the system and thereby can be treated as a single fluid. The downstream moving fluid contains also a small number of photons that were part of the upstream moving photon flow, before they were scattered once by the fluid leptons. These downstream moving photons quickly produce pairs on upstream moving photons.

  • 12 

    Our definition of τs is similar to −τ in the notation of Budnik et al. (2010).

  • 13 

    Note that here we find the shock width in pre-shock, without pairs, Thomson optical depth, which can be directly related to physical width for a given upstream density. Budnik et al. (2010) find the shock width in units of total Thomson optical depth, including pairs, of a photon that crosses the transition layer toward the upstream, τ*. As it turns out, in both cases the width ∝Γ2 but the proportionality coefficient is different (Budnik et al. 2010 find Δτ* ≈ Γ2).

Please wait… references are loading.
10.1088/0004-637X/747/2/88