FAST RADIATION MEDIATED SHOCKS AND SUPERNOVA SHOCK BREAKOUTS

, , and

Published 2010 May 21 © 2010. The American Astronomical Society. All rights reserved.
, , Citation Boaz Katz et al 2010 ApJ 716 781 DOI 10.1088/0004-637X/716/1/781

0004-637X/716/1/781

ABSTRACT

We present a simple analytic model for the structure of non-relativistic and relativistic radiation mediated shocks. At shock velocities βsvs/c ≳ 0.1, the shock transition region is far from thermal equilibrium since the transition crossing time is too short for the production of a blackbody photon density (by bremsstrahlung emission). In this region, electrons and photons (and positrons) are in Compton (pair) equilibrium at temperatures Ts significantly exceeding the far downstream temperature, TsTd ≈ 2(εnu3c3)1/4. Ts ≳ 10keV is reached at shock velocities βs ≈ 0.2. At higher velocities, βs ≳ 0.6, the plasma is dominated in the transition region by e± pairs and 60keV ≲ Ts ≲ 200keV. We argue that the spectrum emitted during the breaking out of supernova (SN) shocks from the stellar envelopes (or the surrounding winds) of blue supergiants and Wolf–Rayet stars, which reach βs>0.1 for reasonable stellar parameters, may include a hard component with photon energies reaching tens or even hundreds of keV. Our breakout analysis is restricted to temperatures Ts ≲ 50keV corresponding to photon energies hν ≲ 150keV, where pair creation can be neglected. This may account for the X-ray outburst associated with SN2008D, and possibly for other SN-associated outbursts with spectra not extending beyond few 100keV (e.g., XRF060218/SN2006aj).

Export citation and abstract BibTeX RIS

1. INTRODUCTION

During core-collapse supernova (SN) explosions, strong shock waves traverse the exploding stars' mantle/envelope. These shocks move with high velocities, βsvs/c ≳ 0.01, through dense, n ∼ 1015–1020cm−3, optically thick plasma. Under such conditions, the shocked plasma's thermal energy density is dominated by radiation as long as

Equation (1)

where aBB = π2/15(ℏc)−3 is the Stefan–Boltzmann energy density coefficient. In such shocks, the photons that are generated in the downstream region diffuse upstream and decelerate the incoming plasma by colliding with the electrons, which in turn stop the nuclei through collective plasma interactions (e.g., an electrostatic field). The typical width of such radiation mediated shocks (RMSs) is the length that photons can diffuse into the upstream ld ∼ 1/nuβsσT, where σT is the Thompson cross section and the subscript u denotes the upstream values of parameters.

The structure of non-relativistic RMSs was studied by Colgate (1974a) and Weaver (1976), motivated by the idea that the post-shock plasma can reach high temperatures, T ≳ 10MeV, leading to the generation of deuterium, as an alternative to big bang nucleosynthesis. Non-relativistic numerical calculations were preformed by Weaver (1976) for shock velocities βs < 0.3. Self-consistent solutions of the hydrodynamic, radiation generation (dominated by bremsstrahlung) and diffusion equations were obtained, confirming that RMSs are consistent physical structures. At low velocities, βs ≲ 0.1c (compared to Equation (14)), thermal equilibrium is maintained throughout the shock. In such conditions, the shock temperature monotonically increases between the far upstream and the far downstream values (e.g., Zel'dovich & Raizer 1966). For higher velocities, it was found that the plasma in the shock transition region departs from thermal equilibrium, and the electron temperature reaches values of tens of keV, greatly exceeding the far downstream values (but much smaller than the suggested 10MeV temperatures). The results of the non-relativistic calculations are not applicable to shock velocities exceeding βs ≃ 0.2, where the high temperatures reached in the shock transition require inclusion of relativistic processes (e.g., pair production), and where the evolution of the photon distribution can no longer be simply described as diffusion in space and momentum.

Detailed studies of Compton scattering in the converging flow of non-relativistic RMSs (Blandford & Payne 1981; Lyubarskii & Sunyaev 1982; Becker 1988; and recently in the context of shock breakouts, Wang et al. 2007) have shown that photons traversing through RMS can significantly increase their energy while scattering back-and-forth in the converging flow in a manner similar to the Fermi acceleration of cosmic rays. The energy of photons that can be reached by such "bulk" acceleration is limited to

Equation (2)

above which the energy loss per collision due to the Compton recoil is larger than the energy gain due to the converging bulk flow. As we show here, for velocities βs ≳ 0.1 the electrons (and positrons) are "heated" in the immediate downstream (the region where the photons are generated) to energies exceeding those given in Equation (2) (see Equation (18)). Thus, photons are generated with typical energies larger than those given in Equation (2), implying that the "bulk" photon acceleration is only relevant for lower shock velocities, βs ≲ 0.1, and for sub-keV photons (and for densities low enough so that the downstream thermal temperatures are considerably smaller than hνmax ).

In this paper, we present a simple analytic model for the structure of RMSs. The model accurately reproduces the numerical results of Weaver (1976) for βs ≲ 0.2, and provides an approximate description of the shock structure at larger velocities, βs → 1. A detailed comparison of the results of the simple model presented here with the results of exact solutions of the shock structure (self-consistently solving the photon transport equation and the plasma flow equations for mildly and highly relativistic shocks) will be presented in a follow-up paper (Budnik et al. 2010). In our calculations, we assume that photons are generated mainly by bremsstrahlung emission. This assumption is valid in the absence of strong magnetic fields (note that direct bound–free transitions do not lead to net generation of photons, while the contribution of bound–bound transitions is negligible under the conditions relevant to shock breakout).

We first present, in Section 2, an analytical model for non-relativistic RMSs deriving simple analytic approximations for the non-equilibrium temperature and for the width of the temperature profile, and then extend our model in Section 3 to relativistic RMSs. In Section 4, we discuss the implications of our results to the expected properties of SN X-ray outbursts. We show that breakout velocities βs>0.1 may be reached in the explosions of blue supergiant (BSG) and Wolf–Rayet (W-R) stars (Section 4.2), and argue that the X-ray outbursts accompanying breakouts from such stars may include a hard component with photon energies reaching tens or even hundreds of keV (Section 4.3). The implications of our results to the interpretation of recent X-ray outbursts associated with SNe (XRO080109/SN2008D, XRF060218/SN2006aj) are discussed in Section 4.4. We summarize the results and discuss their implications in Section 5.

2. NON-RELATIVISTIC RADIATION MEDIATED SHOCKS

Consider an RMS in a plasma consisting of protons, electrons, and photons. The far upstream consists of cold protons and electrons moving with velocity βsc in the shock rest frame, while the far downstream consists of protons, electrons, and photons in thermal equilibrium with a temperature Td.

In this section, we construct an analytical, approximate model for the steady-state profile of such a shock. We start by writing down the downstream conditions in Section 2.1. Next, we discuss the velocity profile in Section 2.2 and argue that the deceleration scale is roughly (βsnuσT)−1. In Section 2.3, we show that for large shock velocities, βs ≳ 0.1, the temperature can have a non-trivial profile extending to distances much larger than the deceleration length. In Section 2.4, we estimate the temperature in the shock transition and show that it can greatly exceed Td. We show that our simple analytic results are in good agreement with detailed numerical calculations. Finally, we use the results derived in Sections 2.12.4 to describe the structure of fast non-relativistic RMSs in Section 2.5.

2.1. Downstream Conditions

The energy and momentum flux are dominated by the protons' momentum and kinetic energy in the upstream and by the photons in the far downstream. Proton number, momentum, and energy conservation require that

Equation (3)

where the subscript d denotes the downstream values of the parameters and we used the fact that the energy density in photons is eγ = 3pγ. Equations (3) imply βd = βs/7. Thermal equilibrium implies

Equation (4)

The far downstream temperature is thus

Equation (5)

where nu = 1015nu,15, ε = 0.5 β2smpc2 = 10 ε1MeV, and βs = 0.1βs,−1.

2.2. Deceleration

Next, we estimate the proton deceleration length scale. Once a proton reaches a point in the shock where the energy is dominated by photons, it experiences an effective force

Equation (6)

implying a deceleration length of

Equation (7)

We note that once the photons dominate the pressure, they cannot drift with the protons, as this will imply an energy flux greater than the total energy flux. Thus, the drag estimated in Equation (6) is unavoidable. The energy and momentum of the protons are thus transferred to the photons on a length scale Ldec irrespective of the details of the mechanisms that generate the energy in the radiation field. We note that the transition length, Ldec, is of the order of the distance a photon can diffuse upstream before being advected with the flow. The deceleration of the flow can be solved analytically (e.g., Blandford & Payne 1981, and references therein; see Equation (A6)). For the strong shocks considered here,

Equation (8)

where x is the distance measured in the shock frame and the point x = 0 corresponds to β/βd − 1 ≈ 0.25.

2.3. Temperature Profile Length Scale

The region of the shock profile over which the temperature changes before it reaches Td can be extended to distances that are much larger than Ldec.

To see this, consider the length scale that is required to generate the density of photons of energy ∼Td in the downstream, determined by thermal equilibrium, nγ,eqpγ,d/Td,

Equation (9)

where Qγ,eff is the effective generation rate of photons of energy 3Td. We use here the term "effective generation rate" due to the following important point. Photons that are produced by the emission mechanism at energies ≪Td may still be counted as contributing to the production of photons at Td, since they may be upscattered by inverse-Compton collisions with the hot electrons to energy ∼Td on a timescale shorter than that of the passage of the flow through the thermalization length, LTdc. Note that the relevant Compton y parameter,

Equation (10)

is much larger than unity for LTβsnuσT ≫ 1/4(T/100 eV)−1β2s,−1. Qγ,eff includes, therefore, all photons produced down to an energy that allows them to be upscattered to Td. For bremsstrahlung emission, which we assume to be the main source of photons, the number of photons generated diverges logarithmically at low energy, so that Qγ,eff may be significantly larger than the bremsstrahlung generation rate of photons at Td.

In order for the photon energy to significantly increase by scattering, it must be scattered ∼mec2/(4T) times before getting re-absorbed. This sets a lower limit to the frequency of photons that should be included in Qγ,eff,

Equation (11)

where n = 7nu = 7 × 1015n15cm−3. The generation rate should include all the photons that are generated with energy above hνa and are able to be upscattered during the available time LT/(βc). The bremsstrahlung effective photon generation rate is thus given by

Equation (12)

where geff is the Gaunt factor and Λeff ∼ log(T/(hνa)). As long as the Compton y parameter is large, y>log(T/hνa), all photons produced above νa are upscattered to ∼3T energies.

Using Equations (12) and (9), we obtain

Equation (13)

This implies that for large shock velocities,

Equation (14)

the length required to produce the downstream photon density is much larger than the deceleration scale. For lower shock velocities, thermal equilibrium is approximately maintained throughout the shock.

2.4. Immediate Downstream

The photons that stop the plasma at the deceleration region must be generated within a distance ∼(nuβsσT)−1 of the deceleration region in order to be able to reach it (before being swept downstream by the flow). Consider the following simple model for the first few diffusion lengths downstream of the deceleration region. We assume that the velocity is βd downstream from some point x = x0 and is much larger upstream of it. As the e-folding distance of the residual velocity δβ = β − βd is very short, =βsnuσT/21, this is a good approximation. We neglect any generation of photons upstream of the transition, $Q_{\gamma,\rm eff}(x<x_0)=0$. This is reasonable as the contribution of photons from a slab, one diffusion length wide, is roughly proportional to β−3 and β changes by a factor of 7 along the transition region. We further assume that the generation rate of photons downstream of the transition is roughly constant and equal to $Q_{\gamma,\rm eff}(x=x_0)$. The latter simplification is justified, to an accuracy of a few tens of percent, due to the fact that the temperature changes by order unity across the available (diffusion length) distance and QeffT0.5. For such a flow, the density of effective photons at x = x0 is

Equation (15)

As the flow at this point is close to the downstream velocity, we have nγ(x0)Ts = pγ,d, where TsT(x0) is the temperature in the immediate downstream. Using Equations (12) and (15), the momentum conservation equation can be written as

Equation (16)

where Λeff = 10Λeff,1 ∼ log(Ts/(hνa)) with hνa given by Equation (11). Here we assumed that the temperature does not change much within a distance of Ldec downstream of the point x = x0. The Compton y parameter, relevant for the upscattering of the photons that diffuse up to the deceleration region,

Equation (17)

is much larger than 1 for Ts ≫ 30β2s,−1 eV. We, therefore, assume that all photons above hνa are upscattered all the way to Ts.

Solving Equation (16) for βs, we find

Equation (18)

Equation (18) is in good agreement with the results of Weaver (1976) for Ts < 50keV. For βs>0.07 and nu = 1015–1022cm−3, the temperature (velocity) given by Equation (18) for a given velocity (temperature) agrees with the numerical calculations to within a factor of 1.5 (20%; with the gaunt factor calculated as in Weaver 1976). We note that the scaling with parameters in Equation (18) can be found by writing the RMS equations in dimensionless form (see Equations (A11)–(A16)).

Comparing Equation (18) with the equation for the far downstream temperature Td, Equation (5), we see that the shock temperature is much larger than the downstream temperature for shock temperatures Ts ≳ 1keV corresponding to shock velocities βs ≳ 0.1 (see also Weaver 1976).

Note that for temperatures T ≳ 1keV the typical photons in the immediate downstream have energies hν ∼ 3T higher than the maximal energies attainable by bulk Compton acceleration in non-relativistic flows, given by Equation (2). Indeed, the typical photon energies, as derived from Equation (18), rise much faster with velocity than hνmax  ∝ β2s, and at T ∼ 10keV, corresponding to βs ≈ 0.2, are already much higher (hν ∼ 3T ∼ 30keV compared to hνmax  ∼ 4keV). At temperatures exceeding T ∼ 50keV, where pair production becomes important, Equation (16) is not applicable anymore. However, the velocity required for bulk Comptonization to be able to accelerate to the typical photon energies at those temperatures, hν ∼ 150keV, approaches c. Thus, for non-relativistic and mildly relativistic shock velocities with βs ≳ 0.1, bulk photon acceleration cannot reach energies that are considerably higher than the typical "thermally" Comptonized photon energies in the immediate downstream.

2.5. Description of the Shock Structure

We can broadly divide the shock into four separate regions.

  • 1.  
    Near upstream: a few diffusion lengths, (βsσTnu)−1, upstream of the deceleration region. In this region, characterized by velocities that are close to the upstream velocity, β ≈ βs, and temperatures, TTu, the temperature changes from Tu to ∼Ts. It ends when the fractional velocity decrease becomes significant.
  • 2.  
    Deceleration region: a (βsσTnu)−1 wide region where the velocity changes from βu to βd and the temperature is roughly constant, TTs.
  • 3.  
    Immediate downstream: roughly a diffusion length, (βsσTnu)−1, downstream of the deceleration region. In this region, characterized by velocities close to the downstream velocity, β ≈ βd, and temperature, TTs, the photons that stop the incoming plasma are generated. Upstream of this region β>βd and the photon generation rate is negligible. Photons that are generated downstream of this region are not able to propagate up to the transition region.
  • 4.  
    Intermediate downstream: the region in the downstream where most of the far downstream photons are generated and T changes from Ts to Td. This region has a width LT given by Equation (13), much greater than (βsσTnu)−1. Thus, diffusion within this region can be neglected. The temperature profile is expected to follow Tx−2. To see this, note that the photon density at a distance x from the shock is proportional to the integral of the photon generation, nγT−1/2x. Since the photon pressure equals the downstream pressure, we have nγT−1 and Tx−2 (this is valid for a constant value of Λeffgeff and is somewhat shallower in reality). Using this dependence of the temperature on distance, we find that LTβsnuσT ∼ Λeffgeff|dT−1/2d((Λeffgeff)|sT−1/2s)−1, in agreement with Equations (13) and (18).

As an illustration, the profiles of the velocity and temperature of a non-relativistic RMS (βs = 0.25, nu = 1015cm−3) are shown in Figure 1. A description of a novel method used to find such solutions is described in the Appendix. As can be seen, the analytic estimates are in good agreement with the numerical calculation.

Figure 1.

Figure 1. RMS shock temperature and velocity profile for ε = 30MeV (βs = 0.25) and nu = 1015cm−3. The black full line is the velocity given by Equation (8). The blue full line is the temperature calculated numerically (see the Appendix). The dashed green line is the far downstream equilibrium temperature given by Equation (5). The dash-dotted red line is the analytic estimate for the immediate downstream temperature. The width of the temperature equilibrium length, given by Equation (13), for these parameters is LTβuσTnu ≈ 130.

Standard image High-resolution image

3. RELATIVISTIC RADIATION MEDIATED SHOCKS

3.1. General Discussion

The analysis presented above is valid for temperatures up to about Ts ∼ 30keV, where a number of effects that were not taken into account begin to affect the temperature profile. The main process that should be taken into consideration is the creation of electron–positron pairs at T ≳ 50keV. Additional corrections include relativistic corrections to the bremsstrahlung emission rate, inclusion of photon creation by double Compton scattering, Klein–Nishina (K-N) corrections to the Compton cross section, reduction of the Compton y parameter due to the smaller optical depth, τ ∼ β−2d, and the anisotropy due to the high velocities, βs ∼ 1.

The following important characteristic of the shock structure is likely unchanged: the plasma decelerates due to scattering with photon and pairs that are produced in the first β−1d optical depths of the downstream (a discussion of the deceleration velocity profile, neglecting K-N corrections, is given in Levinson & Bromberg (2008); note that K-N corrections and significant electron–positron production significantly affect the result; Budnik et al. 2010). As the downstream velocity is βd < 1/3 for all βs, the diffusion approximation in the downstream is reasonable and an equation similar to Equation (16), with the relativistic corrections included, can be used to constrain the temperature in the transition region.

Note that Weaver (1976) claimed that once a significant amount of pairs is created, the pairs will "short out" the electrostatic field which is responsible for the deceleration of the protons. He claimed that the electrons and ions will no longer be fully coupled, and suggested that once there are many pairs per proton, there will be a relative drift of order βs between the leptons and protons, and that the main mechanism for stopping the protons will be Coulomb collisions with the leptons. We find this picture to be unrealistic in real plasmas. A relatively weak magnetic field, which will be freezed into the pair plasma, will suffice to deflect the protons on scales which are much shorter than determined by Coulomb collisions. Several length scales may be important here. The gyroradius of the protons in a magnetic field B = 102B2G is

Equation (19)

the skin depth of the pair plasma is

Equation (20)

and the Compton mean free path is

Equation (21)

As long as the Compton mean free path is much larger than the other two scales, it is probably safe to assume that the protons and electrons are highly coupled on the dynamical scales. We thus assume in what follows that collective plasma effects will keep the pair plasma and the protons coupled without the need for an electrostatic field.

3.2. T–βd Relation

As long as the Compton y parameter is much larger than unity, the spectrum will be close to a Wein spectrum. For the entire parameter space considered here, the positron number is in equilibrium. To see this, note that the typical time a positron takes to annihilate is similar to the typical time a photon takes to Compton scatter, thus pairs can be annihilated (or generated) in βd Compton optical depths compared with the β−1d available.

A detailed discussion of the emission of plasmas in pair and Compton equilibrium is given by Svensson (1984), who shows that the main emission processes are bremsstrahlung and double Compton, with a domination of bremsstrahlung at temperatures above T ≈ 60keV.

A precise determination of the relation between Ts and βu is complicated in the range Ts ≳ 30keV due to the relativistic corrections to the emission rates and due to the order-unity y parameter. Such determination requires a careful analysis of these effects. We avoid such a detailed analysis by limiting our analysis to a demonstration of the validity of the following statements, which set stringent constraints on the relation between Ts and βd.

  • 1.  
    Temperatures Ts>30keV are reached for downstream velocities βd>0.03, with βd ≈ 0.03 for Ts ≈ 30keV (as shown in Section 2).
  • 2.  
    At βd = 0.1, the shock transition temperature satisfies 60keV < Ts.
  • 3.  
    Ts ≲ 200keV for all the possible downstream velocities, βd < 1/3.

The validity of statements 2 and 3 is demonstrated in Sections 3.3 and 3.4 below by comparing the photon-to-pair ratio that is determined by the balance of photon production and diffusion, Equation (15), with the ratio that is determined by pair production equilibrium. The complications due to the various corrections are avoided since, under the conditions we consider, the following hold: the number of pairs greatly exceeds the number of protons, nlnp, where nl = n+ + n ≈ 2n; double Compton emission is negligible; the Compton y parameter is large.

3.3. At βd ≈ 0.1, 60keV < Ts

Assuming pair production domination and neglecting double Compton emission, Equation (15) can be written as

Equation (22)

where we wrote the free–free emission in the form

Equation (23)

Here, $\hat{T}\equiv T/(m_ec^2)$, and $\bar{g}_{{\rm eff,rel}}$ is the total Gaunt factor (defined by Equation (23)) including all lepton–lepton bremsstrahlung emission. For 10 < Λeff < 20 and 60keV < T < mec2, the approximation

Equation (24)

agrees with the results of Svensson (1984) to an accuracy of better than 25%. Substituting the value of Equation (24) into Equation (22), we find

Equation (25)

We next estimate the absorption frequency. The absorption is dominated by e + e− bremsstrahlung and for T ≳ 0.1mec2 photons require a single scattering to change their energy considerably. The absorption frequency is thus roughly equal to the threshold frequency νC=B at which the Compton scattering rate equals the absorption scattering rate and is given by

Equation (26)

where nl = 1018nl,18. The lepton number density nl is roughly given by

Equation (27)

yielding

Equation (28)

where we used gff,ll ∼ 10, nl/nγ ∼ 10, and T ∼ 50keV (the result is not sensitive to the specific values adopted for these parameters).

The Compton y parameter is

Equation (29)

Equations (28) and (29) imply that for T ∼ 0.1mec2 the Compton y parameter is large enough, ya, to allow all the photons above the absorption frequency, hνa, to be Comptonized (photons are upscattered by a factor of ey in energy moving all photons above hνa to the Comptonized energies $3T=3e^{\Lambda _a}h\nu _a$). This implies, in turn, that photon production at all frequencies above hνa contributes to the effective generation rate and Λeff = Λa.

Let us now compare the photon–lepton ratio derived for the balance of photon production and diffusion, to the ratio derived from pair production equilibrium. For large chemical potential, μchT, pair production equilibrium gives

Equation (30)

where K2 is the modified Bessel function of the second kind of order 2. At the temperature range 60keV < T < 90keV, this ratio changes by an order of magnitude: nγ/nl|eq(T = 60keV) ≈ 130 while nγ/nl|eq(T = 90keV) ≈ 13. Comparing these values to the result given by Equation (25), we conclude that at βd = 0.1 the temperature must be above 60keV.

3.4. The Relativistic Limit βd = 1/3

In the regime 200keV < T < mec2, the pair production equilibrium is approximately given by

Equation (31)

The number of leptons is now given by (see Equation (27))

Equation (32)

and the absorption photon energy by

Equation (33)

yielding

Equation (34)

where we used gff,ll ∼ 10, nl/nγ ∼ 1, and Tmec2 (the result is not sensitive to the specific values adopted for these parameters).

We can rewrite Equation (25) for βd ≈ 1/3 and 60keV < T < mec2 as

Equation (35)

Comparing Equations (35) and (31), we see that if T ≳ 200keV there would be too many photons generated per lepton. At these high temperatures, the Compton y parameter is large and radiative Compton emission is negligible. At temperatures T ≳ 100keV, K-N corrections to the Compton cross section of photons with energy hν ∼ 3T become significant. The possible effect that K-N corrections have is to lower the temperature due to the larger photon generation depth. This in turn would suppress the K-N corrections. Hence, K-N corrections cannot change the conclusion that T ≲ 200keV. We conclude that for βd>0.1, 60keV < T ≲ 200keV.

4. SUPERNOVA SHOCK BREAKOUTS

Once an expanding RMS reaches a point where the residual optical depth is of order τ ∼ β−1, whether at the outer shell of a star's envelope or in the surrounding wind (in the case of an optically thick wind), the radiation escapes, and the shock no longer sustains itself. In this section, we summarize the relations between the stellar/wind parameters and the velocity, energy, and duration of the resulting shock-breakout radiation outburst. We first derive general relations between the energy, velocity, radius, and duration of the outburst in Section 4.1. Next, we discuss in Section 4.2 the dynamics of shock acceleration near the stellar surface, and its evolution within the wind, and derive characteristic breakout velocities. The X-ray characteristics of the breakout are discussed in Section 4.3, based on the results of Sections 23, and 4.2. Finally, the implications of our results to the interpretation of recent X-ray outbursts associated with SNe (XRO080109/SN2008D, XRF060218/SN2006aj), and possibly to other SN-associated outbursts (XRFs/GRBs), are discussed in Section 4.4. In this section, we restrict our discussion to non-relativistic shocks βs ≲ 0.6 and non-relativistic temperatures Ts ≲ 50keV, corresponding to photon energies hν ≲ 150keV, where pair creation can be neglected.

4.1. A General Relation between the Breakout Energy, Velocity, and Radius

Consider a shock with velocity βs ≳ 0.1 approaching the photosphere of a spherical mass distribution. The breakout will occur once the optical depth down to the photosphere is τ ∼ β−1 ≲ 10.

The energy carried by the photons in a shell of shocked material with a width of the order of the shock width is roughly given by (e.g., Waxman et al. 2007)

Equation (36)

where δm is the mass of the shell and fE is a coefficient that depends on the velocity and geometry. For a shocked shell in the downstream of a non-relativistic planar shock, fE = 36/49. fE is of order unity for non-relativistic and mildly relativistic shocks in spherical geometry, as long as the distance to the photosphere is not much larger than the radius. The mass of a shell that has an optical depth τ is roughly given by

Equation (37)

where κT = σT/mp ∼ 0.4cm2g−1, τ = 100.5τ0.5, and R = 1012R12cm. Assuming that the energy of shocked plasma with optical depth τ ∼ 3τ0.5 is emitted, we find that there is a simple relation between the released energy, breakout radius, and shock velocity at breakout:

Equation (38)

Once a significant fraction of the energy in the shock region is emitted, the steady-state shock solution is no longer applicable. Note that before the solution becomes invalid, a significant fraction of the energy in the shock region must be emitted. We thus expect that photons with energies of the order of 3Ts carrying an energy similar to the energy estimated in Equation (38) are inevitably emitted.

4.2. Breakout Dynamics

In this subsection, we discuss the dynamics of shock acceleration near the stellar surface, and its evolution within the wind, and derive characteristic breakout velocities. Many of the relations derived can be found in the literature (e.g., Matzner & McKee 1999).

At the outer layers of stars, where the enclosed mass, M, is approximately constant, M = M* = const, and the composition is approximately uniform, the density distribution is a power law in the distance from the edge of the star,

Equation (39)

where R* is the stellar radius and

Equation (40)

This relation is exact if we assume M = M* and a polytropic equation of state $p=k\rho ^{\gamma _n}$, in which case n = (γn − 1)−1 and

Equation (41)

For an efficiently convective envelope, the entropy is constant. Assuming an adiabatic index γn = 5/3, we have n = 3/2. For a radiative envelope, with radius-independent luminosity L = L*LEdd and mass M = M*, we have pγ = (L*/LEdd)p, where LEdd = 4πGMc/κ is the Eddington luminosity limit, resulting in a polytropic equation of state p ∝ ρ4/3, or n = 3. The optical depth from the edge of the star is given by

Equation (42)

The density at the outer edge is given by

Equation (43)

Denoting

Equation (44)

where Mej is the mass of the ejected envelope, we find

Equation (45)

for n = 3, and

Equation (46)

for n = 3/2.

As the shock wave approaches the stellar surface, it accelerates down the density gradient. In planar geometry, with density vanishing at x = x0 following ρ ∝ ((x0x)/x0)n = δn, the flow approaches a self-similar behavior as the shock velocity diverges (Gandel'Man & Frank-Kamenetskii 1956; Sakurai 1960). In this limit

Equation (47)

where λv = β1n/(n + 1) and β1 ≃ 0.19 for γ = 4/3 and n = 3, 3/2. The planar self-similar solution provides a good approximation for the shock behavior near the surface of the star, where the thickness of the layer lying ahead of the shock is small compared to the radius, i.e., for δ ≪ 1 (e.g., Matzner & McKee 1999). Denoting the mass M*M(r) of the shell lying ahead of radius r, at δ < (R*r)/R*, by m = M*M(r), and noting that near the surface the optical depth τ is τ ∝ m, Equation (47) may be written as

Equation (48)

with λv = β1n/(n + 1). The kinetic energy carried by the material having velocity greater than some vs scales with the velocity as

Equation (49)

where

Equation (50)

for n = 3 and n = 3/2, respectively.

The velocity of the shock at an optical depth τ is larger than the typical bulk envelope velocity, $v_{\rm ej,b}\approx \sqrt{E_{\rm ej}/M_{\rm ej}}$, where Eej is the energy deposited in the envelope, by a factor

Equation (51)

where MSes is the mass of the region in which the planar solution is applicable. For n = 3 we have λv ≈ 0.14, while for n = 3/2 we have λv ≈ 0.11. The velocity amplification factor for n = 3 is

Equation (52)

and for n = 3/2 it is

Equation (53)

The density and velocity profiles of the outer envelope are illustrated in Figure 2.

Figure 2.

Figure 2. Shock velocity amplification, and density typical of the edge of a BSG (n = 3, MSeS = M, and R = 1012cm), as a function of the optical depth from the edge of the star as given by Equations (52) and (45). The density is normalized to the value at τ = 1.

Standard image High-resolution image

If there is an optically thick wind surrounding the star, breakout occurs within the wind. Assuming a wind with a constant mass-loss rate $\dot{M}$ and velocity vw, $\rho _w=\dot{M}/4\pi v_{w}R^{2}$, the optical depth between R and infinity is related to the mass mw(<R) by τ(>R) = mw(<R)κ/4πR2 (for RR*), and breakout is expected to occur at a radius

Equation (54)

At early time, the velocity vs,w of the shock driven into the wind at radius R is approximately given by the velocity of the fastest shell of the ejecta. Since the final velocity reached by a fluid element shocked at vs, following its acceleration due to the post-shock adiabatic expansion, is vf ≃ 2vs (Matzner & McKee 1999), the velocity of the shock driven into the wind is initially $\approx 2 A_{v_s}v_{{\rm ej}}$. At later time, when the accumulated wind mass begins to decelerate the ejecta, the shock velocity is approximately obtained by equating the mass of the shocked wind, mw(<R), with the mass of the part of the ejecta that was accelerated to velocity v>vs,w. Using Equation (48) we have $m_w(<R)\approx M_{\rm ej}(v_{s,w}/2v_{\rm ej})^{-\lambda _v}$ or $v_{s,w}/v_{\rm ej}\approx 2(M_{\rm ej}/m_w(<R))^{\lambda _v}$. This gives

Equation (55)

for BSG (n = 3) and

Equation (56)

for RSG (n = 3/2).

The assumption that the shock velocity in the wind is equal to the velocity of the shocked ejecta is correct as long as the density encountered by the reverse shock in the ejecta is much larger than the density encountered by the forward shock in the wind.

Assuming typical bulk envelope velocities of $v_{{\rm ej,b}}\approx \sqrt{E_{\rm ej}/M_{\rm ej}}\sim (3\hbox{--}10)\times 10^8\mbox{ cm}\;{\rm s}^{-1}$, or βej,b ∼ 0.01–0.03, the typical shock velocities at the last few optical depths of BSGs and W-R stars may reach values of βs ≳ 0.2, for which temperatures exceeding T ∼ 10keV may be reached in the shock transition region. The shock velocities in the wind may be up to twice higher than the maximal shock velocity achieved within the star.

4.3. Breakout X-ray Characteristics

Let us consider the radiation that may be emitted during the breakout of a shock from the stellar surface. We consider the emission during a time interval R*/c following the time at which the shock reached the surface (as long as the star is not resolved, any emission will be spread in time over this timescale). On similar timescales, the outer part of the star expands significantly, over a distance βfR*R*, where βf ≈ 2βs is the final velocity reached by a fluid element shocked at βs (following its acceleration due to the post-shock adiabatic expansion, e.g., Matzner & McKee 1999). On the same timescale, R*/c, photons that originated at an optical depth τesc ∼ δ−1/2(τ = 1) are capable of escaping. Using Equation (42), we have

Equation (57)

This implies

Equation (58)

for BSG typical parameters, and

Equation (59)

for RSG parameters. For 1 ≳ βs ≳ 0.1, these values are not much larger than the shock widths, β−1s, and a non-negligible fraction, β−1sesc, of the photons in these regions is likely to escape over ∼R*/c. The amount of energy emitted during R*/c is thus not much larger than the estimate given in Equation (38).

Note that the radiation observed at a given time is the sum of radiation emitted from different positions of the star at different stages of the breakout (up to differences of R*/c). Photons that originated from shocked material in deeper layers, τ ≫ β−1s, will have characteristic energies which are lower than the photon energies at the transition region, hν ∼ 3Ts, due to photon production and possible adiabatic cooling. Therefore, we expect to see a spectrum which is a sum of Wein-like spectra with temperatures reaching the tens of keV shock temperatures Ts and starting from lower energies.

Let us consider next breakouts that take place within an optically thick wind. In this case, the optical depth of the shocked wind material, which is compressed to a thin shell, is comparable to the optical depth of the wind ahead of it. Thus, we expect all photons contained within this region to escape on a timescale similar to R/vs, the timescale on which the optical depth changes significantly. Again, the spectrum would be composed of Wein-like spectra with temperatures reaching Ts.

Finally, the following point should be stressed. As the shock approaches an optical depth of 1/β from the surface, the shock structure deviates from the steady-state solution due to the escape of energy from the system and due to the steep density gradient. Note that at this stage the shock propagates within a density profile which varies on a length scale comparable to the shock thickness. As for the effect of energy escape, a significant fraction of the energy in the shock transition region must be emitted before the structure is changed and thus the steady-state estimate for the photon energies will correctly describe the photons carrying most of the escaping energy. It is reasonable to expect that the effect of the density gradient on the temperature will be of order unity for fast shocks, as the pre-shock density across a shock width changes only by a factor of few—by ∼1/β in the case of stellar surface (see Equation (45)) and by ∼1 in the case of a wind. An exact calculation of the emerging spectrum requires a time-dependent calculation and is beyond the scope of this paper.

4.4. Implications to Recently Observed SN X-ray Outbursts

It was long been suggested that a strong outburst of ultraviolet or X-ray radiation may be observable when the RMS reaches the edge of the star (e.g., Colgate 1974b; Falk 1978; Klein & Chevalier 1978; Ensman & Burrows 1992; Matzner & McKee 1999; Blinnikov et al. 2000). These outbursts carry important information about the progenitor star (Arnett 1977) including a direct measure of the star's radius. During the past two years, the wide-field X-ray detectors on board the Swift satellite detected luminous X-ray outbursts preceding an SN explosion in two cases, SN2006aj and SN2008D (Campana et al. 2006; Soderberg et al. 2008). Analysis of the later optical SN emission revealed that both were of type Ib/c, probably produced by compact (W-R) progenitor stars (Pian et al. 2006; Mazzali et al. 2006, 2008; Modjaz et al. 2006; Maeda et al. 2007; Malesani et al. 2009; Tanaka et al. 2009). While some authors argue that the X-ray outbursts are produced by shock breakouts (Campana et al. 2006; Waxman et al. 2007; Soderberg et al. 2008), others argue that the spectral properties of the X-ray bursts rule out a breakout interpretation, and imply the existence of relativistic energetic jets penetrating through the stellar mantle/envelope (Soderberg et al. 2006; Fan et al. 2006; Ghisellini et al. 2007; Li 2007; Mazzali et al. 2008; Li 2008; some argue that the breakout interpretation holds for SN2008D, but not for SN2006aj; Chevalier & Fransson 2008).

The main challenge raised for the breakout interpretation (e.g., Mazzali et al. 2008) was that in both cases, SN2006aj and SN2008D, the observed X-ray flash had a non-thermal spectrum extending to ≳10keV, in contrast with the sub-keV thermal temperatures expected (Matzner & McKee 1999; note that while thermal components with low temperatures or fluxes cannot be ruled out, e.g., Mazzali et al. 2008; Modjaz et al. 2009, such thermal components carry only a small fraction of the observed flux). However, the analysis presented here implies that fast, βs>0.2, breakouts may be expected for compact, BSG or W-R progenitors, and that for such fast breakouts the X-ray outburst spectrum may naturally extend to tens or hundreds of keV.

Let us discuss in some more detail the possible interpretation of the X-ray outbursts associated with SN2006aj and SN2008D as breakouts. In the case of SN2008D, both the X-ray outburst and early UV/O emission that followed it are consistent with a breakout interpretation (see Soderberg et al. 2008, for detailed discussion). The X-ray burst energy and duration are consistent with the relation of Equation (38), and the UV/O emission is consistent with that expected from the post-shock expansion of the envelope. Moreover, the non-thermal X-ray and radio emission that follow the X-ray outburst are consistent with those expected from a shock driven into a wind by the fast shell that was accelerated by the RMS, for wind density and shell velocity, v/c ∼ 1/4, and energy, E ∼ 1047 erg, which are inferred from the X-ray outburst. The fast velocity inferred for the late radio and X-ray emission is consistent with that required to account for the non-thermal spectrum.

The case of SN2006aj is more complicated. The X-ray burst energy and temperature are consistent with a mildly relativistic, v/c ∼ 0.8, shock breakout from a wind surrounding the star, the UV/O emission is broadly consistent with that expected from the expanding envelope, and the non-thermal X-ray emission that follows the X-ray outburst is consistent with that expected from the shock driven into the wind by the fast shell that was accelerated by the RMS (Campana et al. 2006; Waxman et al. 2007). However, the duration of the X-ray outburst is larger than expected, the UV/O emission is not as well fit by the model as in the case of 2008D, and the non-thermal radio emission is higher than expected from the wind–shell interaction. Several authors (Campana et al. 2006; Waxman et al. 2007) have suggested that the deviations from the simple breakout model are due to a highly non-spherical explosion. Some support for the non-spherical nature of the explosion was later obtained by the polarization measurements of Gorosabel et al. (2006; see, however, Mazzali et al. 2007). Other authors (Soderberg et al. 2006; Ghisellini et al. 2007; Fan et al. 2006; Li et al. 2007) have argued that a relativistic jet is required to account for the observations. We believe that an analysis of the modifications to the simple spherical model, introduced by a highly non-spherical breakout, is required to make progress toward discriminating between the two scenarios.

There is an additional important point that should be clarified in the context of SN2006aj. Under the shock breakout hypothesis, the energy and velocity of the accelerated shell that is responsible for the X-ray emission are v/c ∼ 0.8 and E ∼ 1049.5 erg. This is similar to the parameters of the fast expanding shell inferred to be ejected by SN1998bw (associated with GRB080425), based on long-term radio and X-ray emission, which are interpreted as due to interaction with a low-density wind (Kulkarni et al. 1998; Waxman & Loeb 1999; Li & Chevalier 1999; Waxman 2004a, 2004b). Long-term radio observations strongly disfavor the existence of an energetic, 1051 erg, relativistic jet associated with SN1998bw (Soderberg et al. 2004). The similarity of SN1998bw & SN2006aj may therefore suggest that such a jet is not present also in the case of SN2006aj. However, the large energy deposited by the explosion in a mildly relativistic shell, v/c ∼ 0.8, is a challenge in itself: the acceleration of the SN shock near the edge of the star is typically expected to deposit only ∼1046 erg in such fast shells (Tan et al. 2001), as can be inferred, e.g., from Equation (49), which implies $E(>v_s)/E_{\rm ej}\sim\break (v_s/v_{\rm ej,b})^{-\delta _{v}}$.

5. DISCUSSION

We presented a simple analytic model for non-relativistic (Section 2) and relativistic (Section 3) RMSs. At shock velocities βs = vs/c ≳ 0.1(n/1015cm−3)1/30 (Equation (14)), where n is the upstream density, the plasma is far from thermal equilibrium within the transition region, where most of the deceleration takes place, since the plasma does not have enough time to generate the downstream blackbody photon density, ndaBBT3d/3. The electrons (and positrons) are heated in this region to temperatures Ts significantly exceeding the far downstream temperature, TsTd, and the photons are in Compton equilibrium with the electrons (and positrons). The transition temperature, Ts, is independent of the upstream density: it is given by βs ≈ 0.2(Ts/10keV)1/8 for βs ≲ 0.2 (Equation (18)); at velocities βs ≳ 0.6 (βd ≳ 0.1), where the plasma is dominated by electron–positron pairs in pair production equilibrium, the temperature is constrained to the range 60keV ≲ Ts ≲ 200keV. The thermalization length, the distance behind the transition region over which the plasma thermalizes and reaches the downstream temperature Td, is ∼(Ts/Td)1/2 times larger than the transition (deceleration) width, ∼1/βsσTn. Our simple model estimates are in agreement with the results of exact numerical relativistic calculations, which will be presented in a follow-up paper (Budnik et al. 2010).

We showed in Section 4 that RMSs breaking out of the stellar envelopes of BSGs and W-R stars are likely to reach velocities βs>0.2. We thus expect that for reasonable stellar parameters the spectrum emitted during SN shock breakouts from BSGs and W-R stars may include a hard component with photon energies reaching tens or even hundreds of keV. This implies that core-collapse SNe produced by BSGs/W-R stars may be searched for by using hard X-ray (wide field) detectors. The detection rate of such events is significantly different than is inferred assuming a thermal emission spectrum (e.g., Calzavara & Matzner 2004), due to both the modification of the intrinsic spectrum and to the reduced absorption of high-energy photons. A quantitative estimate of the increased detection rate is beyond the scope of this paper. As the escaping photon energies are highly sensitive to the shock velocity, the rate is sensitive to the unknown high-end part of the distribution of the shock velocities during breakouts.

We have argued in Section 4.4 that the X-ray outburst XRO080109 associated with SN2008D is most likely due to a fast breakout: the energy, ∼1047 erg, duration, ∼30 s, and fast ejecta velocity, βs ≃ 1/4, are all consistent with expected breakout parameters, and the hard X-ray spectrum is a natural consequence of the high velocity, which is independently inferred also from later X-ray and radio observations. Our analysis shows that the spectrum of the X-ray flash XRF060218, associated with SN2006aj, might also be explained as a fast breakout. However, the breakout interpretation of this event is challenged by the long duration of the X-ray flash, and by the high energy, E ∼ 1049.5 erg, deposited in this explosion in mildly relativistic, v/c ∼ 0.8, ejecta (see detailed discussion in Section 4.4).

Wang et al. (2007) have suggested, based on the breakout interpretation of the X-ray outbursts of SN2006aj and SN2008D, that all the low-luminosity gamma-ray bursts/X-ray flushes associated with SNe, which have smooth light curves and spectra not extending beyond few 100 keV (like those associated with SN1998bw, SN2003lw, and SN2006aj), are due to shock breakouts, and do not require the existence of energetic highly relativistic jets. The present analysis, which demonstrates that fast breakouts may indeed produce non-thermal spectra extending to hundreds of keV, supports the viability of the breakout interpretation of low-luminosity gamma-ray bursts/X-ray flushes associated with SNe. A major challenge for such a scenario is constituted by the requirement of large energy deposition, E ∼ 1049.5 erg, in the fastest, mildly relativistic (v/c ∼ 0.8), part of the expanding ejecta.

This research was partially supported by Minerva, ISF, and AEC grants.

APPENDIX: SOLVING THE NON-RELATIVISTIC RMS EQUATIONS

We describe below our numerical solution of the non-relativistic RMS equations. The derivation given below offers several advantages compared to the derivation given by Weaver (1976): the equations are written in a dimensionless form, making the dependence of shock structure on parameters (in particular, on plasma density) explicit, and the diffusion equation is solved analytically, making the numerical integration procedure much simpler.

We use the following approximations.

  • 1.  
    We neglect the pressure of the electrons and protons.
  • 2.  
    We assume that the radiation and the electrons are in Compton equilibrium with equal temperatures Tγ = Te.
  • 3.  
    We neglect the corrections to the Compton scattering cross section and assume it to be equal to the Thompson cross section.
  • 4.  
    We assume that the process generating the photons is bremsstrahlung emission, as given by Equation (12).
  • 5.  
    We assume that the photon distribution is described by the diffusion equation.

Under these assumptions, the equations determining the steady-state shock profile are the conservation of proton flux,

Equation (A1)

the conservation of momentum flux,

Equation (A2)

and the conservation of energy flux,

Equation (A3)

These equations, together with the relation

Equation (A4)

form a closed set of equations for the velocity β, particle number density, np = ne, and photon pressure, pγ, and energy, eγ.

eγ, pγ, and n may be eliminated from Equations (A1)–(A4) to obtain an equation for the velocity,

Equation (A5)

where we introduced the dimensionless variables $\tilde{x}=3\sigma _Tn_u\beta _s x$ and $\tilde{\beta }=\beta /\beta _s$. Equation (A5) can be solved analytically (e.g., Weaver 1976),

Equation (A6)

The photon pressure is related to the temperature through

Equation (A7)

where nγ is the photon number density. T, or nγ, is determined by the photon diffusion equation,

Equation (A8)

where

Equation (A9)

is the effective current of photons and the generation of photons is given by Equation (12):

Equation (A10)

Here, fabs = 1 − nγ/(aBBT3/3) is approximately the suppression factor due to absorption of the photons (Weaver 1976). Equations (A7) and (A8) determine T (and nγ) for given β and pγ.

Defining dimensionless variables $\tilde{T}$ and $\tilde{n}_\gamma$ that satisfy

Equation (A11)

and

Equation (A12)

Equation (A8) may be written as

Equation (A13)

where

Equation (A14)

and

Equation (A15)

Using Equation (A2), we have in addition

Equation (A16)

The solution for $\tilde{T}$ and $\tilde{n}_\gamma$ is nearly independent of β,  np,  T, since these parameters appear in the equations only through the product Λeffgefffabs, which is weakly dependent on parameters. In particular, Equation (18) satisfies the scaling given by Equations (A11) and (A12).

It is straightforward to show that the Green function, $G(\tilde{x},\tilde{x}_0)$, for the diffusion Equations (A13) and (A14), defined by the relation

Equation (A17)

is given by

Equation (A18)

where

Equation (A19)

The effective velocity $\tilde{\beta }_{{\rm eff}}$, calculated for the velocity profile given by Equation (A6), is shown in Figure 3.

Figure 3.

Figure 3. Effective diffusion velocity defined by Equation (A19) and calculated for the velocity profile given by Equation (A6). The full black line is the effective velocity and the red dotted line is the plasma velocity.

Standard image High-resolution image

A numerical solution to Equations (A13)–(A16) can be obtained by preforming iterations in which $\tilde{Q}_\gamma$ is calculated from $\tilde{T}$ using Equation (A15), then $\tilde{n}_\gamma$ is calculated using Equation (A17), and then a new value of $\tilde{T}$ is obtained using Equation (A16). The solution presented in Figure 1 was obtained for an interval $\tilde{x}_{\min }<\tilde{x}<\tilde{x}_{\max }$, ignoring photon contributions from outside of the interval, with $\tilde{x}_{\min }=\ln (m_e/m_p)-\ln (6)/7$ chosen as the point where 1 − β = me/mp and xmax  = 400. The solution is largely insensitive to the choice of the interval boundary.

Please wait… references are loading.
10.1088/0004-637X/716/1/781