A publishing partnership

SHOCK BREAKOUT IN DENSE MASS LOSS: LUMINOUS SUPERNOVAE

and

Published 2011 February 8 © 2011. The American Astronomical Society. All rights reserved.
, , Citation Roger A. Chevalier and Christopher M. Irwin 2011 ApJL 729 L6 DOI 10.1088/2041-8205/729/1/L6

2041-8205/729/1/L6

ABSTRACT

We examine the case where a circumstellar medium around a supernova is sufficiently opaque that a radiation-dominated shock propagates in the circumstellar region. The initial propagation of the shock front into the circumstellar region can be approximated by a self-similar solution that determines the radiative energy in a shocked shell; the eventual escape of this energy gives the maximum luminosity of the supernova. If the circumstellar density is described by ρ = Dr−2 out to a radius Rw, where D is a constant, the properties of the shock breakout radiation depend on Rw and Rd ≡ κDvsh/c, where κ is the opacity and vsh is the shock velocity. If Rw>Rd, the rise to maximum light begins at ∼Rd/vsh; the duration of the rise is also ∼Rd/vsh; the outer parts of the opaque medium are extended and at low velocity at the time of peak luminosity; and a dense shell forms whose continued interaction with the dense mass loss gives a characteristic flatter portion of the declining light curve. If Rw < Rd, the rise to maximum light begins at Rw/vsh; the duration of the rise is R2w/vshRd; the outer parts of the opaque medium are not extended and are accelerated to high velocity by radiation pressure at the time of maximum luminosity; and a dense shell forms but does not affect the light curve near maximum. We argue that SN 2006gy is an example of the first kind of event, while SN 2010gx and related supernovae are examples of the second.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Supernova shock breakouts from normal massive stars typically give rise to X-ray/ultraviolet bursts with a timescale ≲103 s (Klein & Chevalier 1978; Falk 1978; Ensman & Burrows 1992; Matzner & McKee 1999; Nakar & Sari 2010). The timescale is generally determined by R*/c, where R* is the stellar radius, because the radiative diffusion time at breakout is less than the light travel time across the star. The largest red supergiants have a radius of ∼1014 cm (Levesque et al. 2005) so that the longest time of the breakout event is about 1 hr. The short timescale of the bursts makes it difficult to detect them, other than with a wide-field X-ray telescope. The breakout detection of SN 2008D with Swift (Soderberg et al. 2008) was fortunate.

The situation changes if there is dense mass loss prior to the supernova that creates an optically thick region. Calculations of this case go back to early computer simulations of supernova light curves. Model 5 of Grassberg et al. (1971) and Model B of Falk & Arnett (1977) include a dense circumstellar shell with radius ∼1015 cm that determines the properties of the shock breakout. The peak luminosity occurs on a timescale of ∼15 days in the model of Grassberg et al. (1971). Grasberg & Nadyozhin (1987) considered a steady wind of limited duration just before the supernova. Chugai et al. (2004) calculated a model for SN 1994W which allows for dense mass loss around the supernova, leading to a peak luminosity at an age of ∼20 days. A systematic numerical study of such events has recently been carried out by Moriya et al. (2010). Because of the high surrounding density, the supernova can be especially bright and this type of model has been invoked for luminous supernovae. In addition to SN 1994W (Chugai et al. 2004), the application of the dense mass-loss model has been made to SN 2006gy (Smith & McCray 2007), PTF 09uj (Ofek et al. 2010), and SN 2009kf (Moriya et al. 2010) among others.

The aim here is to give an analytical description of the shock breakout process for the case where a radiation-dominated shock propagates into the mass-loss region. Progress on this front has been made by Ofek et al. (2010). The model is presented in Section 2 and compared to observations in Section 3.

2. INTERACTION WITH DENSE MASS LOSS

We assume that the dense mass loss can be described by a steady wind acting for some time tml before the supernova explosion; the extent is Rw = vwtml, where vw is the wind velocity. The actual case may be more complex, but is not well determined; the wind assumption has been made in other work (Grasberg & Nadyozhin 1987; Ofek et al. 2010; Moriya et al. 2010). If the mass loss is in a steady wind, the density $\rho _w=\dot{M}/4\pi r^2v_w\equiv Dr^{-2}$ can be specified by a density parameter, D*, scaled to a $\dot{M}=10^{-2} \,M_\odot \ \rm yr^{-1}$ and $v_w=10\,\rm km\, s^{-1}$ wind so that ρw = 5.0 × 1016D*r−2 in cgs units. For typical supernova parameters, the explosion drives a radiation-dominated shock through the star. A radiation-dominated shock in a uniform medium has a characteristic thickness described by an optical depth τshc/vsh, where vsh is the shock velocity (Weaver 1976). If the wind optical depth τw < c/vsh, then a radiation-dominated shock does not form in the wind and the radiation diffuses out through an optically thick wind (e.g., Nakar & Sari 2010). In the case that τw>c/vsh, the radiation-dominated shock front propagates into the mass-loss region. This is the case considered here.

After the shock wave passes through the star and the supernova ejecta tend toward free expansion, a shocked layer develops in the wind that is bounded by forward and reverse shock waves at Rfs and Rrs, respectively. The shock waves are initially energy conserving. Once the forward shock front reaches a place where the diffusion time equals the expansion time or, equivalently, τwc/vsh, radiation diffusion becomes important. We have τw = (R−1fsR−1wD, so the diffusion condition can be written as R−1dR−1fsR−1w, where Rd ≡ κDvsh/c = 5.7 × 1014vsh4kD* cm, where vsh4 is the shock velocity in units of 104 km s−1 and k is the opacity κ in units of 0.34 cm2 g−1. Two cases are of interest: Rw>Rd so that RfsRd when diffusion becomes important and Rw < Rd so that RfsRw at that time.

In the wind, for diffusion over a distance ΔR, the diffusion time is td = ΔR2c = κρwΔR2/c. Taking ΔRR and noting ρw = Dr−2, we have td = κD/c, independent of radius (see also Ofek et al. 2010). If the extent of the optically thick region (to Rph = min(Rw, κD)) is not much beyond Rd, then κD/c provides an estimate of the diffusion time for the radiation to escape. The time is longer if Rph>Rd. If the diffusion is assumed to proceed in steps with size ∝R, the timescale is increased by a term that is logarithmic in R. For the approximate results presented here, we neglect the logarithmic dependence and take

Equation (1)

We now examine a specific model for expansion into the dense wind. After the shock wave has expanded beyond the stellar radius, the ejecta tend toward free expansion (v = r/t) in which there is a steep power-law section on the outside and a flat profile in the inner region (e.g., Matzner & McKee 1999). Here we make the approximation that ρinr−1 at velocities v < vt and ρsnr−7 at v>vt. The outer profile is less steep than the profile inferred for the asymptotic profile expected at high velocities; the wind interaction causes a flatter part of the density distribution to be relevant. We then have vt = (2E/Me)1/2 and ρsn = Bt−3(r/t)−7, where B = 4E2/3πMe (Chevalier & Fransson 1994). The interaction between this power-law density distribution and a surrounding wind, with density ρw = Dr−2, can be described by a self-similar solution with adiabatic index γ = 4/3 (Chevalier 1983), which is appropriate for a radiation-dominated gas. The radius of the contact discontinuity is

Equation (2)

where E51 is the explosion energy in units of 1051 erg, Me1 is the ejecta mass in units of 10 M, and t1 is the age in units of 10 days. The highest velocities occur just inside the reverse shock wave, where the velocity is $v_{\rm rs}=0.978R_{\rm cd}/t=6.3\times 10^3E_{51}^{0.4}M_{e1}^{-0.2}D_*^{-0.2}t_1^{-0.2}\,\rm km\, s^{-1}$. The applicability of the solution requires that vt < vrs or E0.551M−1.5e1D*t1 < 32.

We initially assume that Rd < Rw. At time td,

Equation (3)

The amount of energy in the interaction shell above some velocity v0 is

Equation (4)

The internal energy, which is in the form of radiation, is Erad = 0.32Esh (Chevalier 1983). Using v0 = Rcd(td)/td at time td, we have

Equation (5)

for the radiative energy in the breakout shell. The timescale for the radiation to escape is td, so the luminosity at shock breakout is

Equation (6)

Taking the radiation energy Erad and the shocked volume of 4π(R3fsR3rs)/3 at time td, where Rfs and Rrs are the forward and reverse shock radii, respectively, the mean radiation energy density gives a radiation temperature

Equation (7)

where we have used Rfs/Rcd = 1.208 and Rrs/Rcd = 0.978 (Chevalier 1983), and td1 is td in units of 10 days. Another estimate of Trad can be obtained from the forward shock condition on the pressure: aT4/3 = (6/7)ρwv2fs for a γ = 4/3 gas, where a is the radiation constant (see also Ofek et al. 2010). In this case, the temperature coefficient in Equation (7) is 7% higher; the difference is due to the fact that there is a drop in the pressure behind the forward shock front in the shocked region. The results assume a smooth wind with an r−2 density profile, but the temperature of the breakout shell is not strongly dependent on this assumption. More generally, if ΔR is the size of the breakout region, the optical depth through the region is ρκΔRc/vsh and the shock crosses the shell in the diffusion time, or ΔRvshtd. The resulting density ρ = cv2shtd, when combined with the shock jump condition for the radiation pressure, gives a temperature comparable to that in Equation (7).

These results can be compared to the numerical results of Moriya et al. (2010). Taking their model s13w2r20m2e3 with E51 = 3, Me1 = 1.3, and D* = 1, the breakout radius in our model is Rd = 5.9 × 1014 cm, which is inside the outer radius of 2 × 1015 cm in the numerical model. Substituting the parameters into Equation (5), the radiated energy is Erad = 1.4 × 1050 erg, which is close to the 2.0 × 1050 erg found in the numerical model (Moriya et al. 2010). The numerical result may be larger, in part, because the shock wave generates power in the more extended circumstellar medium. We also examined the scaling of Erad with the parameters and found reasonable agreement with the numerical results. The scaling depends on the supernova density gradient, which is only approximately treated here.

Estimates of the observable color temperature of the radiation require considerations of whether radiation equilibrium is attained in the emitting region. Following Nakar & Sari (2010), we define a thermal coupling coefficient $\eta =n_{{\rm BB}}/(t_d\dot{n}_{\rm ph,ff}(T_{{\rm BB}}))$, where nBBaT4BB/3kBTBB is the photon number density in thermal equilibrium, kB is Boltzmann's constant, and $\dot{n}_{\rm ph,ff}(T_{{\rm BB}})=3.5\times 10^{36}\rho ^2T^{-1/2}$ s−1 cm−3 is the production rate of photons by the free–free process. If sufficient photons are produced to maintain the blackbody number density, or η ≲ 1, thermal equilibrium is achieved. Using Equations (1) and (7) for td and TBB, and ρ = 7ρw(Rfs) (taking into account the factor 7 compression in the shock wave), we estimate η for the breakout shell:

Equation (8)

For the standard parameters, the breakout shell is marginally in thermal equilibrium. As the radiation propagates into the unshocked mass-loss region, the lower density results in a deviation from thermal equilibrium. The frequency dependence of the opacity can play a role (Moriya et al. 2010) and we do not treat the details of spectrum production here.

The loss of radiative energy from the shocked region results in the formation of a dense shell at radius R, as seen in numerical simulations (Grassberg et al. 1971; Falk & Arnett 1977). The expansion of the shell into additional mass loss produces continuing power for the supernova, L = 2πR2ρw(vfsvw)3, where the wind velocity vw may be affected by preshock radiative acceleration. The simulations of Moriya et al. (2010) show some evidence for acceleration, but it is only significant near the breakout radius because of the r−2 dependence of the radiative flux, and we neglect it here. The expansion of R can be described by the thin shell approximation (Chevalier 1982), yielding R = 0.94Rcd. The resulting power is

Equation (9)

The magnitude of the luminosity is similar to that produced by the initial breakout radiation, as is seen in numerical simulations (Grasberg & Nadyozhin 1987; Moriya et al. 2010). The luminosity lasts until the shock wave at R reaches the edge of the dense wind, Rw, at tw = 1.1E−0.551M0.25e1D0.25*R1.25w16 yr, where Rw16 is in units of 1016 cm. Figure 1(a) illustrates the luminosity evolution with the late flattening from the shell interaction.

Figure 1.

Figure 1. Luminosity curves of two types of supernova interaction with dense mass loss: (a) wind extent Rw greater than the characteristic diffusion radius Rd and (b) Rw < Rd. In each case, there is a time from the explosion at t = 0 to the shock wave reaching a place where the radiation can escape and the luminosity rises. In the case Rw>Rd, there is a later, slower luminosity decline due to continued interaction of the shock wave (velocity vsh) with slow wind material.

Standard image High-resolution image

In the case of shock breakout from a red supergiant, the shock front takes ∼1 day to traverse the star and the time for shock breakout is ∼103 s (Klein & Chevalier 1978). The shock breakout timescale is much less than the time since explosion. In the dense mass-loss case considered here, the time for the shock front to move to the breakout region is ∼Rd/vsh, which is also the timescale for the breakout event. This property of the luminosity evolution can be seen in simulations of such events (Grassberg et al. 1971; Falk & Arnett 1977; Chugai et al. 2004; Moriya et al. 2010). The rise to maximum light can have complications due to variations in the gas opacity. At the high circumstellar densities considered here, it is likely that the gas is initially neutral and that most of the opacity is due to dust in the presupernova environment. The radiation-dominated shock wave from the supernova has a precursor in the mass-loss region that is expected to heat the circumstellar dust. As the temperature rises through 1000–2000 K, the dust evaporates, giving a decrease in the opacity and the photospheric radius drops to where there is a sharp gradient in the opacity as the gas becomes ionized. In Type IIP supernovae, this property of the opacity causes a recombination wave to back into the expanding envelope with a constant photospheric temperature T ∼ 5000–6000 K. Here, the process is inverted and the photosphere is expected to follow the ionization wave moving out through the dense circumstellar gas. This phase of approximately constant temperature can be seen in simulations (Grassberg et al. 1971). Once the circumstellar mass is ionized, the photospheric expansion slows and the temperature rises. The light curve rises fairly sharply due to the temperature rise until the maximum temperature, and luminosity, is reached.

We now consider the case that Rw < Rd. The shock breakout process begins at tbRw/vsh. The rise time for the light curve is tr ≈ δR/vsh, where δR is the distance in from Rw across which the diffusion time equals the shock crossing time. We thus have (δR)2κDR−2w/c = δR/vsh, leading to tr ≈ (Rw/vsh)(Rw/Rd). In this case, the rise time can be considerably shorter than the time for initial heating of the envelope (Figure 1(b)). Since δR/Rw < 1, the region involved in shock breakout is not radially extended, so the pressure of the escaping radiation can accelerate the gas out to Rw (Ensman 1994). High gas velocities are expected around the time of maximum luminosity. A dense shell forms when the radiation can escape, but does not produce continuing high luminosity because the dense gas does not extend far beyond the breakout point and it has been radiatively accelerated.

For our specific supernova model, the shock breakout begins when the forward shock (Rfs) reaches Rw, which occurs at tw = 16R1.25w15E−0.551M0.25e1D0.25* days, where Rw15 is Rw in units of 1015 cm. The free expansion velocity at the reverse shock is $v_0=5.7\times 10^3 R_{w15}^{-0.25}E_{51}^{0.5}M_{e1}^{-0.25}D_*^{-0.25}\,\rm km\, s^{-1}$, which leads to a radiated energy of $E_{{\rm rad}}=0.65\times 10^{50} R_{w15}^{0.5}\linebreak [2] E_{51}M_{e1}^{-0.5}D_*^{0.5}$ erg. The rise time for the light curve is $t_r=t_wR_w/R_d=41 k^{-0.8} R_{w15}^{2.25}E_{51}^{-0.9}\linebreak [2] M_{e1}^{0.45}D_*^{-0.35}$ days, so that $L\approx E_{{\rm rad}}/t_r=1.8\times 10^{43}k^{0.8}R_{w15}^{-1.75}E_{51}^{1.9}M_{e1}^{-0.95}D_*^{0.85}\rm erg\, s^{-1}$. The temperature in the shocked shell is like that given in the first part of Equation (7) except that td is replaced by tw. The temperature is lowered in the escape out to the photosphere if there is sufficient photon production in this region (Nakar & Sari 2010).

3. COMPARISON WITH OBSERVATIONS

The radiated energy from SN 2006gy from the initial optical rise to around the peak is ∼1 × 1051 erg (Ofek et al. 2007; Smith et al. 2007), which makes it a good candidate for the physical situation described here (with wind optical depth τw>c/vsh). In this case, the observed rise time gives an estimate of td, the diffusion time. The observations indicate a rise time of 60 days. Using Equation (1), the indicated wind density is D* ≈ 10. The observed peak luminosity has sensitivity to the supernova energy. At maximum light, the observed luminosity was 4 × 1044 erg s−1 (Smith et al. 2010), which implies E51 ≈ 3 for the other standard parameters. With these parameters, Equation (3) gives Rd = 2.5 × 1015M−0.2e1 cm while the dense medium may extend to ∼1 × 1016 cm (Smith et al. 2010), so that Rw>Rd. We attribute the flattening of the observed light curve (Smith et al. 2010) to the continuing interaction in the extended region. With Rw = 1016 cm, the implied circumstellar mass is ∼30 M. These parameters are close to those found by Smith & McCray (2007) and Smith et al. (2010), which is expected because the basic physical picture is similar in the two cases. Depending on the ejecta mass, the condition that vt < vrs may be violated, but we do not expect the results to be strongly affected.

Smith et al. (2007, 2010) estimated an explosion date of 2006 August 20 for SN 2006gy; this time is just before the beginning of a rise in optical luminosity from 0.01Lmax to Lmax, where Lmax is the maximum optical luminosity. In the shock breakout view, the sharp rise of optical radiation occurs after a time Rd/vsh during which the shock is traveling in the optically thick region (Figure 1). The explosion date is thus ∼60 days earlier than the estimate of Smith et al. (2007, 2010), and mean velocities of uniformly expanding ejecta are lower than the estimates of Smith et al.

Another aspect of the rise to maximum in shock breakout is that the increasing temperature as the shock breaks out is an important component of the rising luminosity. In the case of SN 2006gy, Smith et al. (2010) find a temperature T ≈ 11, 000 K on days 65 and 71 (from 2006 August 20; their Figure 4), which is close to the time of optical maximum light. On day 36, Smith et al. (2010) estimate T ≈ 9500 K, using the same extinction value as that used for other epochs (their Figure 4). However, they find that with an assumed larger extinction, a photosphere with T ≈ 15, 000 K provides a better fit to the observed spectrum. The higher temperature would bring the temperature evolution more in line with that observed in other supernovae, i.e., a decreasing temperature with time, and is advocated by Smith et al. (2010). However, in the shock breakout view, the increasing temperature is expected and there is no need for a time-dependent extinction.

In our model, the dense shell and shock wave are deep within the circumstellar envelope, outside of which is the last equilibrium shell where the spectrum is formed. The radius of this shell is not the blackbody radius Rbb because of the nonequilibrium conditions in the medium. Outside of the shell is an electron-scattering region of moderate optical depth (between c/vsh and 1) where the peaked Hα profile with a broad base can be formed (Chugai 2001; Smith et al. 2010); the broadening is due to scattering in the thermal gas as opposed to the Doppler effect.

An example of a different type of luminous supernova is SN 2010gx and related objects (Pastorello et al. 2010; Quimby et al. 2009). In this case, there is a sharp drop from peak luminosity without a flattening, broad lines of O ii are present at maximum light, and the narrow H or He lines often observed in Type IIn supernovae are not present. The evidence points to an explosion in a more compact circumstellar medium for this class of objects. The rise time, ∼50 days, and peak luminosity, $(3\hbox{--}4)\times 10^{44}\,\rm erg\, s^{-1}$, are comparable to the case of SN 2006gy, while the temperature is higher, 15,000 K. The rise time suggests that Rw is not much less than Rd, so RwRd. The density is comparable to SN 2006gy. The higher temperature in this case can be attributed to the lower opacity due to the lack of H and He, and the smaller extent of the opaque circumstellar medium.

Although the case for dense mass-loss years before the supernova appears good, the cause of the mass loss is not known. Woosley et al. (2007) suggested that SN 2006gy was the result of pair instability eruption before the supernova. Since the radiated energy produced in this case is ∼1050 erg and the radiated energy in observed luminous supernovae is ≳1051 erg, we have not specifically treated that case here, although if the energy were larger similar physical arguments would presumably apply. The dense mass loss is typically attributed to luminous blue variable eruptions (Smith & McCray 2007), although such eruptions are not understood in the context of stellar evolution. An eruption was observed two years before the explosion of SN 2006jc, which was an H-poor, Type Ib supernova (Pastorello et al. 2007; Foley et al. 2007). The mass loss in this case was too weak to produce the high density phenomena discussed here, but it does show the possibility of a presupernova outburst, even in the H-poor case.

We thank Claes Fransson for discussions. This research was supported in part by NSF grant AST-0807727.

Note added in proof. A model similar to that developed here was presented by Balberg & Loeb (2011) and applied to the Type Ib SN 2008D.

Please wait… references are loading.
10.1088/2041-8205/729/1/L6