Abstract
Transients powered by interaction with the circumstellar medium (CSM) are often observed in wavelengths other than the optical, and multiwavelength modeling can be important when inferring the properties of the explosion and CSM, or for distinguishing from other powering mechanisms. We develop a model calculating the time-dependent emission spectrum of interaction-powered transients. We solve the energy equations of electron–proton plasma in the shocked supernova ejecta and CSM and the radiation transfer equation out to the outer edge of the CSM, incorporating collisional relaxation and the Comptonization of bremsstrahlung radiation. We compare our model to observations of Type IIn supernovae covering a frequency range from the optical to X-rays. For SN 2010jl the observed optical and X-ray light curves can be consistently explained if a clumpy or asymmetric structure in the CSM is assumed, in agreement with previous studies. For SN 2014C our model successfully reproduces the X-ray bremsstrahlung component and the emergence of Hα emission at 400 days after explosion. Finally we find a parameter space where the supernova is extremely X-ray bright, reaching 1043–1044 erg s−1, for up to 100 days. Such X-ray transients are likely to be detectable with all-sky surveys by, e.g., eROSITA.
Export citation and abstract BibTeX RIS
1. Introduction
Massive stars with initial masses of >8M⊙ are considered to end their lives as supernovae (SNe). Among the diverse photometric and spectroscopic detections, there have been a growing number of transients that show signs of interaction between the SN ejecta and a preexisting dense circumstellar medium (CSM).
Type IIn SNe (Schlegel 1990; Filippenko 1997), which display narrow (10–1000 km s−1) hydrogen emission lines in optical spectra, are representative of this kind of transient. Such emission lines are attributed to the existence of a slowly moving dense CSM (e.g., Chugai 1991). The dense CSM also efficiently dissipates the kinetic energy of the SN ejecta (e.g., Grasberg & Nadezhin 1986), which are considered to power the light curve at least in the early phases. In cases where the CSM is dense, shock breakout from the progenitor itself is delayed and prolonged (e.g., Chevalier & Irwin 2011; Förster et al. 2018).
The collision of SN ejecta with the CSM leads to the formation of shock waves. The shock waves, propagating at a velocity of 103–104 km s−1, heat the upstream gas to temperatures of 1–100 keV. By radiative cooling this plasma emits copious amounts of X-ray photons, which are reprocessed in the shocked region and the CSM. X-ray emission is detected in some Type IIn SNe, e.g., in SN 2005ip (Katsuda et al. 2014), 2005kd (Dwarkadas et al. 2016; Katsuda et al. 2016), 2006jd (Chandra et al. 2012a; Katsuda et al. 2016), and 2010jl (Ofek et al. 2014; Chandra et al. 2015). The X-ray luminosity from these SNe ranges from 1040 to 1042 erg s−1, much brighter than those in other classes of SNe (Chandra 2018).
Observations in soft and hard X-rays enable us to extract the column depth of neutral hydrogen ahead of the shock. This has been done, e.g., in SN 2010jl (Chandra et al. 2015; Katsuda et al. 2016), which showed an evolving column density by orders of magnitude over a timescale of a few years. With the natural assumption that the column density is dominated by the CSM, this demonstrates that simultaneous observations in X-rays as well as in the optical band can be important for tracing the activity of massive stars at the end of their lives.
There are previous works on modeling X-ray emission from interaction-powered transients upon and after shock breakout (Katz et al. 2010; Balberg & Loeb 2011; Chevalier & Irwin 2012; Svirski et al. 2012; Pan et al. 2013). While these works enable us to obtain a rough estimate of the X-ray emission, no previous work has been done to obtain a long-term broadband spectrum that can realize detailed comparison with observations.
In this work we attempt to model the multiwavelength emission of interaction-powered transients from the optical band to X-rays. In our model we take into account the basic radiative processes governing the inside of the shocked region as well as the CSM. We find that the density (mass-loss rate) of the CSM is crucial in shaping the optical–X-ray spectra, and that the X-ray emission can be quite bright and detectable when the CSM ahead of the shock front becomes transparent.
This paper is constructed as follows. We describe our emission model in Section 2. We present the resulting spectra and optical/X-ray light curves in Section 3, and compare our model with observations of Type IIn SNe 2010jl and 2014C in Section 4. We list some caveats for our model in Section 5, and conclude in Section 6.
2. Model
We first outline the basic idea of our emission model, next summarize the basic equations, and then describe how to solve them. A schematic view of our modeling is shown in Figure 1.
2.1. Overview
We consider a core-collapse SN occurring in a dense CSM. When the SN ejecta collide with the CSM, two shocks are formed: a forward shock (FS) separating the shocked and unshocked CSM and a reverse shock (RS) separating the shocked and unshocked SN ejecta (Chevalier 1982). The interface of the shocked CSM and SN ejecta is called a contact discontinuity (see Figure 1). We consider the case where homologously expanding cold SN ejecta with a power-law density profile (ρej ∝ v−n with n > 5) run through a CSM of a wind density profile
where constants and vw are, respectively, the mass-loss rate and wind velocity. We adopt a normalization parameter for the density, . The number density and velocity at the immediate downstream can be obtained from the self-similar solution of the shock dynamics (Chevalier 1982). The details of the shock propagation are described in Appendix A. We here overview how the dissipated kinetic energy at these shocks is converted to radiation and how the radiation emanates from the CSM.
The immediate upstream, either the unshocked ejecta or the unshocked CSM, is a cold but ionized gas due to strong UV/X-ray emission from the shocked region. When this crosses the shock, thermal relaxations between particles of the same species first occur; for a gas of density ρ the electron–electron and proton–proton collision timescales (Padmanabhan 2000),
are generally much shorter than the dynamical timescales,
Here kB is the Boltzmann constant, me (mp) is the electron (proton) mass, e is the electron charge, and is the Coulomb logarithm. Then the temperature Te (Tp) of electrons (protons) at the immediate downstream of the shock fronts is obtained as a function of the shock velocities vfs(rs) from the jump condition with an adiabatic index γ = 5/3 as
Most of the energy dissipated at the shock goes to protons at the immediate downstream.
The FS (RS) downstream is slightly slower (faster) than the contact discontinuity, so for both shocks the incoming gas is gradually advected to the contact discontinuity. The advection timescale is roughly equivalent to the dynamical timescale tad ≈ tdyn. 4 During the advection, the electrons in the shocked region exchange their energies with the protons due to Coulomb interaction with an equipartition timescale of
The electrons also either gain or lose energy by emitting or absorbing photons via bremsstrahlung processes and by inelastic Compton scattering (Weaver 1976; Katz et al. 2010). Since the timescales in Equations (2) and (3) are shorter than that in Equation (7) and other relevant cooling and heating timescales, the shocked regions can be treated as two temperature plasmas. On the other hand, since the advection of the shocked plasma (Equation (4)) is the slowest process, we can approximate that the radiative shock structure is quasi-steady. For each epoch we obtain the hydrodynamic parameters at the shock interface (shock radius, shock velocity, and upstream density), and use these to solve the electron/proton temperatures inside each region swept by the FS and RS.
Photons are predominantly generated by bremsstrahlung at the immediate shock downstream of the FS and RS. The spectral energy distributions are modified by Compton scattering and free–free and bound–free absorption while propagating through the shocked regions. Radiation escapes from both regions after a diffusion time passes, and half of the radiation that escapes the FS region reaches the observer (see Figure 1). We solve the radiation transfer to calculate the escaping photon spectrum consistently with the two temperature structures in the shocked regions.
2.2. Basic Equations
In this section, we review the basic equations that govern the structure of the shocked region.
As the scale heights of the downstream parameters are much smaller than r, the geometric factor r2 can be treated as a constant in a narrow shocked region. Assuming a steady state in the shock rest frame the hydrodynamic equations are described as (e.g., Takei & Shigeyama 2020)
where ρ is the mass density of the shock downstream, and the pressure p and specific internal energy U are defined as
where x is the ionization fraction. We set x = 1 when protons and electrons have not reached equipartition yet. Once Te = Tp, assuming local thermodynamic equilibrium the value of x can be obtained from the Saha equation. The gas is assumed to consist purely of hydrogen for simplicity, with the temperatures of neutral and ionized hydrogen assumed to be equal, TH i = Tp, in the entire shocked region. Fν and erad,ν are the energy flux and density of radiation, respectively. Flux-limited diffusion (e.g., Levermore & Pomraning 1981) gives the flux as a function of erad,ν by the following equation:
where κν is the opacity, c is the speed of light, and λν is a parameter between 0 and 1/3, the limits corresponding to the optically thin and thick cases, respectively.
When electrons and protons have different temperatures, Coulomb interaction works between them to reduce the difference. The energy exchange rate of protons and electrons Pe−p (erg cm−3 s−1) is (Katz et al. 2011)
where ni = xρ/mp is the number density of protons and electrons, and σT is the Thomson cross section. Noting that electrons are dominantly affected by the photon field, the energy equation can be rewritten into the following equations for Te and Tp:
The equation of radiation transfer in spherical symmetry along the advection flow of the downstream plasma is
where Iν is the specific intensity, θ is the angle between the light ray's propagation direction and the radial direction, and
are, respectively, the emissivity and absorption coefficient, which are related to the free–free and Comptonization processes. Taking the zeroth moment of Equation (17) we obtain
Here we neglect the geometric factor 2Fν /r because the relevant scale lengths are much shorter than the radius r. Transferring to the Euler description and in the shock's rest frame, ∂/∂t → (∂/∂t + v∂/∂r). From the steady-state assumption we obtain
For simplicity we assume the free–free process is the dominant emission/absorption process and neglect free–bound/bound–free and bound–bound contributions. The emissivity (erg s−1 cm−3 Hz−1) and absorption coefficient (cm−1) are given in cgs units by Rybicki & Lightman (1979):
where h is Planck's constant and gff is the Gaunt factor
For the photon frequency we sample a wide range of 0.1 eV < hν < 100 MeV, with 80 bins evenly spaced on a logarithmic scale.
For inverse Compton (IC) scattering, we obtain the rate of change in photon energy density jν,IC (erg s−1 cm−3 Hz−1) as
where σKN(ν) is the Klein–Nishina cross section given by Rybicki & Lightman (1979):
with β = h ν/me c2, and is the typical frequency of photons whose frequency becomes ν after scattering.
For a nonrelativistic electron (kB Te ≪ me c2) and a photon energy , the typical frequency ν after scattering of a photon of energy is (Rybicki & Lightman 1979)
Since we consider a photon spectrum at high energies exceeding me c2 as well, we refine the second term as follows. For a photon scattering with an electron at rest, from the kinematics of the scattering
where θ is the angle between the momenta of scattered and incident photons. Averaging over θ
Repeating the same argument that led to Equation (26), we obtain
The factor 4 is valid only when electrons are nonrelativistic, which is generally satisfied for our model parameters. For each ν we solve Equation (27) implicitly and obtain .
2.3. Two-zone Approximation
By solving Equations (8), (9), (13), (15), (16), and (21) for both the FS and RS regions, one can in principle obtain the shock structure, i.e., Te(r), Tp(r), ρ(r), v(r), Fν (r), and erad,ν (r). These equations take into account the radiation feedback onto the dynamics of the advected gas, but this will be self-consistent only if the propagation of the shock is simultaneously modified (Takei & Shigeyama 2020). Our formulation relies on the self-similar solution for the shock propagation and parameters at the immediate downstream, and taking into account the deviation from this self-similar evolution drastically complicates the problem.
In order to simplify the problem, in this work we adopt an assumption that the pressure gradient is zero in each shocked region, i.e., the dynamics is assumed to be one-zone. The plasma that crosses the shock is thus advected by a constant advection velocity, given by the downstream velocity in the shock's rest frame. Takei & Shigeyama (2020), who solved the above equations with the assumption of local thermodynamic equilibrium in the shocked region and by adopting the diffusion approximation, showed that a nearly constant pressure is indeed seen (see their Figure 1).
On the other hand we follow the evolution of Te, Tp, and erad,ν in the region, which are important for the emission. With our assumption the energy equation (10) is simplified to
Under the one-zone assumption, we average the flux that appears in the energy equation, over the shocked region's volume. The volume average then becomes a surface integral, and the energy equation is modified as
where eesc,ν (erg cm–3 Hz–1) is the energy density of radiation taken away from the shocked region, and jν,in is the source term, which consists of two components explained later in this section. From our constant-pressure assumption v is independent of r, and we can set vd/dr = d/dt, where t is the time that has passed since the downstream plasma crossed the shock. Then with Tp > Te, the equations for Te and Tp are
We make Te and Tp equal by hand (with the total energy conserved) if these two approach each other within 10%. This is merely for numerical purposes, as a finite difference tends to make the Coulomb term Pe−p erroneously large after electrons and protons have cooled down. Once Te = Tp, thermal relaxation is fast enough that this equality is kept. Then the equations for Te and Tp are further simplified to
We first obtain [1 + x(Te)]Te with this equation, then use the Saha equation and implicitly obtain Te (= Tp).
We next present our method of obtaining erad,ν and eesc,ν . The equation for radiation transfer becomes
where we redefine jν ≡ jν,ff + jν,in. This equation reduces to the standard equation of radiation transfer if is assumed. In this case the formal solution can be obtained by expressing erad,ν as , and solving for F(t) by substituting this into Equation (33).
To determine and separately we need another constraint, which should be related to the physics of diffusion. As an approximation we treat the diffusion as a hard cutoff over a typical diffusion timescale tdiff. In other words, photons created at time t = s are in the shocked region and contribute to erad,ν at time s < t < s + tdiff, then completely escape and contribute to eesc,ν at t > s + tdiff. Then by analogy of the standard case described above, we find the solution of the form
for t > tdiff, and
for 0 < t ≤ tdiff. To calculate the integrals, we keep records of jν and throughout the dynamical time with time step tdyn/24,000. This resolution is limited by our computational resource. In the RS region, we observe cases where the cooling timescale (Equation (46)) becomes shorter than this time resolution, thereby preventing us from obtaining these integrals accurately. In these cases we switch to an analytical treatment for calculating erad,ν and eesc,ν , explained in Appendix B.
We calculate Te, Tp, erad,ν , and eesc,ν from t = 0 to tend, the timescale for the gas advection from the shock to the contact discontinuity, in our case equivalent to the dynamical time (see Section 2.1). Thus tend is roughly equivalent to the epoch , or the time from explosion. 5 We do this integration for a series of epochs , with . For each epoch the conditions at t = 0 (i.e., at the shock) are obtained by the equations in Appendix A. The series should be set so that the dynamics and radiation flow do not significantly change between neighboring epochs. We set the epochs by the following equations:
The parameter is the diffusion timescale in the CSM, which controls the light-curve evolution:
where κscat ≈ 0.34 cm2 g−1 is the Thomson scattering opacity. The series continues up to 250 days or when the RS sweeps the entire outer ejecta, whichever is earlier. The variables are, respectively, the diffusion timescales in the RS and FS regions. The diffusion time in the FS region is given by its optical depth multiplied by the light-crossing timescale:
where the width of the FS region ΔrFS relates to the radius of the contact discontinuity rsh derived in Equation (A4) as ΔrFS ≈ [(γ − 1)/(γ + 1)]rsh. The diffusion time in the RS region depends on the parameters for the self-similar solution, but is generally much shorter than due to the much smaller width. For example if we assume n = 10 and γ = 4/3, we find .
In our model jν,in has two components. One is radiation entering from the other shocked region, shown as black arrows in Figure 1. For radiation escaping from the FS region, assuming the escaping radiation is isotropic escapes outward (which is observed), while the other half goes inward to the RS region, and is possibly reprocessed there. Under the assumption of cold SN ejecta in which only the vicinity of the shocked region is ionized, we assume that the bulk of the ejecta is transparent and all the escaping radiation from the RS region goes to the FS region as a source term. We calculate the source term from the eesc,ν in the previous epoch, and inject this at a constant rate for tdyn.
The other component, which is rather artificial, is erad,ν , which should be present in the shocked region at t = 0. Our formulation of erad,ν (Equation (36)) implicitly assumes erad,ν = 0 at t = 0, which neglects this contribution. We instead include this as a source term estimated from the previous epoch, i.e., leftover erad,ν in the end of the previous epoch, erad,ν (t = tend,N−1), is carried over in the present epoch as a source term. We inject this at a constant rate for the first diffusion time at epoch .
When setting the injection term from the previous epoch, we make sure that the total injected energy in the shocked region, eesc,ν (or erad,ν ) times the volume of the shocked region, is conserved. Mathematically, jν,in is given for the FS and RS regions as
where Θ(t) is the Heaviside step function, and is the volume of the FS (RS) region at epoch . Variables with superscript FS (RS) denote values in the FS (RS) region. In the self-similar solution the ratio between the volumes of the FS and RS regions is a constant of time. We neglect additional energy sources that can appear as a source term, such as radioactive decay of nickel inside the ejecta.
For the first epoch , the calculation is iterated until erad,ν becomes stable. To obtain jν,in for the next iteration, we use the values of erad,ν and eesc,ν at t = tend,0 because there is no previous epoch. The energy density at t = tend,0 becomes stable within 1% after about 10 iterations, from which we move on to epoch .
2.4. Calculating the Observed Spectrum
To obtain the observed spectrum, we need to solve the radiation transfer through the unshocked CSM. At each epoch, the luminosity of radiation that escapes from the FS region toward the observer is
where the factor 1/2 is due to the aforementioned geometrical effect. Once we obtain this luminosity Lν , we use version 17.02 of CLOUDY (Ferland et al. 2017) to obtain the spectrum after it is processed through the unshocked CSM Lν,CSM. We assume the CSM extends with a spherically symmetric wind profile from r = rsh out to 1017 cm. The metallicity of the CSM is assumed to be solar.
This treatment also assumes a steady state, i.e., the photoionization at a given epoch is determined by the luminosity at that time. For Type IIn SNe this is justified, as the recombination timescale is generally much shorter than the dynamical timescale (Chevalier & Irwin 2012). For example, for a gas of Te ∼ 104 K the recombination coefficient is αrec ∼ O (10−13) cm3 s−1, and the recombination timescale is of the order of 103 (105) s at day 10 (100) for our fiducial model parameters.
We also include the effect of diffusion inside the CSM delaying the light curve, by the following formula (Chatzopoulos et al. 2012; Equation (4)):
For the contribution of Lν,CSM at , which is absent in our modeling, we adopt the value at for simplicity. The actual contribution is uncertain, as this also depends on when the CSM interaction starts. However, the exponential dependence makes this uncertainty unimportant after about a few times .
3. Results
We consider the model parameters listed in Table 1, which are within the range of typical inferred parameters for Type IIn SNe (Smith 2014). Model EMμ represents a model with ejecta energy × 1051 erg, and CSM density parameter . We assume n = 10 and adiabatic index γ = 4/3 for all of our parameter sets when solving the shock dynamics from the self-similar solution. We assume the ejecta mass to be 10 M⊙ for all cases.
Table 1. Summary of Model Parameters Adopted in This Work
Model | Eej (erg) | MCSM(r < 1016 cm) (M⊙) | |
---|---|---|---|
E1M1 | 1051 | 1 × 10−2/100 | 0.3 |
E3M1 | 3 × 1051 | 1 × 10−2/100 | 0.3 |
E10M1 | 1052 | 1 × 10−2/100 | 0.3 |
E1M10 | 1051 | 1 × 10−1/100 | 3 |
E3M10 | 3 × 1051 | 1 × 10−1/100 | 3 |
E10M10 | 1052 | 1 × 10−1/100 | 3 |
Notes. The parameters are explosion energy Eej and the wind density parameter . The last column shows the enclosed mass of the CSM within 1016 cm.
Download table as: ASCIITypeset image
We first discuss the shock downstream, such as the evolution of the electron temperature and the contribution of various heating/cooling processes. These processes are also important for shaping the radiation spectrum. We then discuss the spectra and the optical and X-ray light curves obtained from our calculations.
3.1. Shock Downstream
We show the change of electron temperature Te as a function of time in Figure 2. The value of Te initially rises due to Coulomb interaction with the protons (hotter by mp/me), but then the temperature drops first due to free–free emission, then by IC scattering. The detailed evolution differs for the RS and FS, which we discuss in detail below.
Download figure:
Standard image High-resolution image3.1.1. RS Region
In the RS region bremsstrahlung tends to be the dominant process for cooling electrons, owing to the high density and low shock velocity (Chevalier & Irwin 2012). The available energy density at the shock downstream is
where α is a parameter that appears in the self-similar solution of shock dynamics, as can be seen in Equation (A4). This thermal energy density corresponds to 5 × 106 K for the E1M1 model at 10 days. The cooling timescale is dominant at the immediate downstream when Te is highest, and scales with time and model parameters as
where is the frequency-averaged value of gff (Rybicki & Lightman 1979). This timescale is consistent with the curves in the left panel of Figure 2. This tcool is generally much shorter than the dynamical time t, and the plasma efficiently cools down and is eventually balanced with heating by free–free absorption. In this case the gas energy density is converted to blackbody radiation, whose temperature is given as
where a is the radiation density constant. There is also a contribution from the FS region, which can further increase Te due to absorption and interaction of FS photons with the electrons. Compton heating is more important for models with higher ejecta energy, as it makes the seed photons harder. For some cases this balances free–free cooling and halts the decrease of Te.
The escape of photons after a diffusion time contributes to the lowering of Te, as the supply of hard photons emitted in the RS region or streaming from the FS region would be limited. This generally makes Te < TBB, and for some cases even reaching around 6000 K, where hydrogen recombines.
3.1.2. FS Region
The situation can be different for the FS region, due to the higher downstream energy density and, more importantly, the much lower density at the shock front. The free–free cooling time is orders of magnitude longer than in the RS region, with a scaling
This roughly explains the drop at later epochs, such as at day 103. However at earlier epochs, due to inefficient free–free cooling, IC can take over the cooling when there is enough supply of seed photons (Chevalier & Irwin 2012).
To see this more quantitatively, in the bottom panel of Figure 3 we plot the time evolution of the inverse of various electron heating and cooling timescales in the E1M1 model. The Coulomb heating timescale is te−p ∼ (3ni kB Te/2)/Pe−p, the IC cooling timescale is tIC ∼ (3ni kB Te/2)/[∫d ν jν,IC], and the free–free absorption heating timescale is tabs ∼ (3ni kB Te/2)/[∫d ν αν,ff cerad,ν ]. The aforementioned transition of IC cooling being the dominant cooling process can be seen in the bottom panels, at ∼104 s in day 10.5. For early phases, this is likely to be responsible for significantly reducing the electron temperature (see top panel of Figure 3) before free–free cooling takes back the role.
Download figure:
Standard image High-resolution imageIn the early epochs we see a convergence of Te, which is due to (i) a balance of free–free absorption and cooling given that the density is high enough, (ii) a balance of free–free cooling and Compton heating by photons from the RS, or (iii) a recombination of hydrogen for those that converge around 6000 K.
3.2. Spectral Properties
Figure 4 shows the resulting spectra of photons escaping the FS region for our choice of model parameters. We find that for the E1M1 and all the M10 models soft X-rays emitted from the shock are significantly absorbed by the unshocked CSM, and the observed spectrum is much different from the spectrum emergent from the shock front. A fraction of the absorbed X-ray photons are converted to optical and UV emission, with a large number of emission lines, including Hα at ≈1.9 eV. This matches the naive expectation that for low Eej and/or high , the shock expands more slowly and the optical depth of the CSM is kept large. On the other hand, the observed spectra for models E3M1 and E10M1 are largely unmodified from the spectrum from the shock. In addition to the CSM quickly becoming optically thin, higher explosion energies (and hence higher shock velocities) extend the FS spectra to hard X-rays, which avoid significant absorption by the CSM.
Download figure:
Standard image High-resolution imageAn important aspect that determines the spectrum from the shock is the optical depth to Thomson scattering inside the FS region. For a wind-profile CSM this is equivalent to that in the unshocked CSM, and is given as
where we assume κscat = 0.34 cm2 g−1.
For an optically thick (τfs ≫ 1) case, Compton downscattering can drastically reduce the photon energy. The number of Thomson scatterings per unit time is κscat × 7ρw c, and photons stay in the FS region for a diffusion time , where the factor 7 comes from the compression ratio for adiabatic index 4/3. Thus there are scatterings in the FS region before escaping, and as a result the energy of photons reaching the observer is reduced to . This is consistent with the cutoff in the spectra at the FS for the E1M10 model on day 10.6, when τfs ≈ 38 (blue dotted line in the top-right panel of Figure 4).
The degree of absorption of soft X-rays depends on this cutoff frequency and the luminosity of ionizing photons Lion, as discussed in Chevalier & Irwin (2012). Qualitatively, in order for soft X-rays to reach the observer the following two conditions must be satisfied: (i) τfs should be at most a few tens, so that the cutoff is in or beyond the soft X-ray band, and (ii) the flux of ionizing photons is high enough that the main elements that contribute to absorption of soft X-rays (oxygen and iron) are completely ionized. The latter is quantified as the ionization parameter in cgs units. Chevalier & Irwin (2012) find that ξ greater than 5000–104 will ionize all the elements, a precise value depending on the characteristic temperature of the spectrum. For our parameters
which is consistent with our results within an order of magnitude, and explains the tendency seen in our spectra that models with higher Eej (i.e., higher Lion) and smaller are visible in soft X-rays.
Another notable feature is that the spectrum νLν below the cutoff is nearly flat for the E1M1 and E3M1 cases, while it becomes steeper up to νLν ∝ ν for the E10M1 case. The latter is simply due to the fact that the free–free emissivity jν,ff is flat at h ν ≪ kB Te, and the Compton and IC scatterings are ineffective. Repeated scattering between electrons and photons can convert X-ray photons to lower energies, which flattens the spectrum from the free–free one.
3.3. Optical and X-Ray Light Curves
We calculate the optical and X-ray light curves to compare them with observations. For the optical light curves, we use the νLν from our model and adopt the Johnson–Cousins filter response function to obtain the (absolute) B- and V-band magnitudes with
where e(λ) is the filter response function for each band, 6 fν,ref is the reference flux that defines the zero-point, and we use the conversion fλ = (c/λ2)fν . For fν,ref we adopt the values 4260 Jy and 3640 Jy for the B and V bands, respectively (see Table 4 of Bessell 1979). The resulting light curves in the B and V bands are shown in Figure 5. The peak magnitude of average Type IIn SNe is estimated to be MV = −18.4 ± 1.0 (Kiewe et al. 2012), with more recent estimates (Richardson et al. 2014; Nyholm et al. 2020) giving consistent results. Our results that roughly reproduce this magnitude favor mass-loss rates of the order of 0.1 M⊙ yr−1. This is at odds with the much lower derived for some SNe IIn in previous works (Kiewe et al. 2012; Taddia et al. 2013). This should be partly due to the shortcomings of our model with incomplete opacity in the shocked region and an assumption of a one-zone density profile, which are elaborated in Section 5. The discrepancy can also be from the previous works' assumption of a high conversion efficiency from the kinetic energy dissipation rate to the Hα luminosity, Hα = 0.1. Inspecting Figure 4, some spectra show Hα ≪ 0.1, especially for the M1 models. A low Hα is also implied from a more recent SN IIn sample (Kokubo et al. 2019). As the value of derived from this method scales as the inverse of Hα , this implies that the actual can be much higher than previously estimated. A more sophisticated modeling of Hα may give a better understanding of , although it is beyond the scope of this work.
Download figure:
Standard image High-resolution imageFor the X-ray luminosity, most observations of SNe IIn, using, e.g., Chandra and XMM-Newton, have been done in the soft X-ray band covering energies up to 10 keV. Here we define the X-ray luminosity as
This is to be compared with the bolometric luminosity integrated over the frequency range in our model:
We show the resulting light curves in Figure 6. For the E1M1, E1M10, and E3M10 models, CSM absorption is significant enough to make X-rays observable only at later times. The E1M10 model is extreme, as soft X-ray photons are completely absorbed until the end of our simulation of ∼200 days. For most SNe IIn X-rays have been observed only around a year after explosion (Chandra 2018). This observational fact may favor mass-loss rates of the order of 0.1 M⊙ yr−1 for SNe IIn, independent of the previous argument on the optical luminosity. For models E3M1 and E10M1, where the CSM quickly becomes transparent to X-rays, the X-ray luminosity can become comparable to the bolometric luminosity. In fact, they show very bright X-ray emission of luminosity 1042–1044 erg s−1 for ∼100 days. While the rate of such energetic explosions inside a moderate CSM is uncertain, they are good targets for all-sky X-ray surveys carried out with, e.g., eROSITA (see Figure 5.7.2 of Merloni et al. 2012).
Download figure:
Standard image High-resolution image4. Comparison with Type IIn SNe
As a demonstration of our emission model, we test our model with observations of two well-studied Type IIn SNe, SN 2010jl and SN 2014C.
4.1. SN 2010jl
SN 2010jl was detected in a nearby star-forming galaxy, UGC 5189A (Newton & Puckett 2010), whose distance is measured to be 49 Mpc. Observations from the UV to the near-IR have revealed it to be a luminous Type IIn SN with bolometric luminosity of at least 3 × 1043 erg s−1 (Zhang et al. 2012; Fransson et al. 2014; Ofek et al. 2014). Various studies converge on the CSM being massive, with estimates of the mass-loss rate ranging from 4 × 10−2 to 1 M⊙ yr−1 (Zhang et al. 2012; Moriya et al. 2013; Fransson et al. 2014; Ofek et al. 2014; Chandra et al. 2015).
X-ray emission was detected from around 40 days after the explosion (Chandra et al. 2012b). Long-term X-ray observations by Chandra et al. (2015) found an evolving column density from NH ∼ 1024 cm−2 to 1021 cm−2 over a few years, and radio emission was first detected on day 570. This implies that the shock has been running through a dense CSM that efficiently absorbs X-ray and radio emission.
We compare our model to the bolometric and X-ray light curves of SN 2010jl. We adopt parameters close to those estimated by previous works: ejecta with energy 1.6 × 1051 erg for mass of 10 M⊙, an outer ejecta slope of n = 7, and a CSM mass-loss rate of 0.13 M⊙ yr−1. As an input of CLOUDY we adopt a CSM metallicity of 0.35 solar, which is around the upper limit inferred from observations of the host galaxy (Stoll et al. 2011). We compare the light curves only up to around day 200. The bolometric light curve shows a break between days 200 and 400 (Fransson et al. 2014), which can be because either the FS has reached the outer edge of the dense CSM or the RS has reached the inner ejecta. In either case, our assumption of self-similarity in deriving the shock dynamics is inapplicable beyond this time.
The result is shown in Figure 7. We find the bolometric light curve matches well the best fit obtained from observations in Fransson et al. (2014). Both the FS and RS are radiative throughout the time range we consider, which is consistent with the analytical estimate by Fransson et al. (2014).
Download figure:
Standard image High-resolution imageWhile the bolometric light curve is consistent, the X-ray luminosity in the right panel is not reproduced. Our predicted X-ray luminosity (orange dashed line) falls far below the observed one (black points). Furthermore, the X-ray luminosity at the FS front obtained from our model (solid line) overshoots the unabsorbed luminosity, calculated from spectral analysis of observational data (gray points) by Chandra et al. (2015).
These inconsistencies can be reconciled by abandoning the simplest assumption of the CSM being spherical, and by introducing, e.g., the existence of clumps or asphericity as mentioned in previous works (Fransson et al. 2014; Chandra et al. 2015; Katsuda et al. 2016). To show this, we consider a case where 1% of the emission from the shock front can escape, possibly due to incomplete coverage of the CSM. The bolometric light curve in the left panel of Figure 7 is hardly affected with this small solid angle.
The X-ray light curve in this case is shown as green dotted lines, which roughly matches the observed X-ray luminosity up to day 200. With this 1% escape fraction, the X-ray spectrum at around 60 days (Figure 8) extends to about 10 keV, consistent with Chandra observations at a similar epoch (Chandra et al. 2015; Katsuda et al. 2016). Intriguingly, an order of 1% escape fraction is also inferred independently from the modeling of the radio emission (Murase et al. 2019), though the radio data were taken long after the time range we consider (at around day 600). Nonetheless, SN 2010jl demonstrates that observations in both the optical and X-ray bands would be an effective way of diagnosing the presence of asymmetry in the CSM.
Download figure:
Standard image High-resolution image4.2. SN 2014C
SN 2014C was first detected by the Lick Observatory Supernova Search (Kim et al. 2014). It was originally classified as an SN Type Ib in a host galaxy, NGC 7331, at a distance of 15 Mpc. However, from ∼300 days after explosion the source showed strong hydrogen emission lines analogous to those of Type IIn (Milisavljevic et al. 2015), and became bright in X-rays (Margutti et al. 2017). This transition from a free expansion to a strong interaction indicates that the progenitor exploded inside a shell-like massive CSM embedding a low-density cavity.
We compare our model with X-ray observations of SN 2014C. However our dynamical modeling, which assumes an extended CSM with a power-law density profile, cannot be applied to the shell-like CSM argued for SN 2014C. Instead, we focus on one epoch, at day 400, and construct an X-ray spectrum at this epoch.
The X-ray spectrum at this epoch obtained by Margutti et al. (2017) shows a clean bremsstrahlung emission with a characteristic temperature of 18 keV. From this spectrum Margutti et al. (2017) deduced that the X-ray emission mainly originates from the optically thin FS region, which is close to adiabatic and is moving at a velocity of Vsh ∼ 4000 km s−1. The shock radius at this epoch is estimated to be 6.4 × 1016 cm from very long baseline interferometry imaging (Bietenholz et al. 2018). We adopt these values and use our model with only the FS component to obtain the spectrum, as the RS component is argued to be negligible (Margutti et al. 2017). We vary the downstream number density ndown as a parameter. We put a CSM of solar metallicity (Milisavljevic et al. 2015) with a flat density profile with number density ndown/4. The 4 in the denominator is the compression ratio for γ = 5/3, as this shock is close to adiabatic. The width of the CSM is determined so that the hydrogen column density is 3 × 1022 cm−2, as obtained from X-ray spectral fitting. For this case the advection time rsh/Vsh ∼ 5 yr is much longer than 400 days. Thus we only calculate with our model up to 400 days, and obtain Lν as
This time we do not include the factor 2 accounting for direction, as we do not consider the reprocessing by the (presumably optically thin) RS region. We also neglect the effect of diffusion since the diffusion timescale is negligible at this epoch.
We compare our spectrum (solid lines) and the observed spectrum by Margutti et al. (2017) (blue points) in Figure 9. Our spectrum, with a downstream density of ndown = 1.4 × 106 cm−3, reasonably matches the bremsstrahlung component in Margutti et al. (2017). The prominent Hα line observed in Milisavljevic et al. (2015) is also seen in our spectrum, although a detailed comparison is subject to uncertainties in the profile and geometry of the CSM, and is beyond the scope of this work. One caveat is that the strong excess at 6.7–6.9 keV that Margutti et al. (2017) attribute to Kα transitions of H- and He-like ions of iron is not seen in our model. This excess is seen in SN 2006jd as well, and may be explained by an enhancement of iron as compared to the solar metallicity, or by the coexistence of a cooler component, e.g., from the RS region or from clumps (see Chandra et al. 2012a; Margutti et al. 2017 for a discussion).
Download figure:
Standard image High-resolution image5. Caveats
Our emission model ignores some physics that may have impact on the emission. As this work is the first step to derive detailed optical to X-ray spectra, incorporation of these is to be dealt with in future work. In this section we note the caveats of our work, and discuss qualitatively in what cases these may become important.
5.1. Opacity in the Shocked Region
As mentioned in Section 2.2 we have included only free–free processes as the emission/absorption mechanism of photons in the shocked region. Although this may be a good approximation at low-metallicity environments, there are additional contributions from bound–free (and bound–bound) processes that can be important at metallicities around and beyond the solar metallicity. As seen in Figure 2 the plasma temperature in the shocked region can drop down to ∼104–105 K. At this temperature range metals are not fully ionized and thus can efficiently convert soft X-rays to UV and optical radiation. 7
We defer the precise modeling of this to future work, but the qualitative effect of this may be seen by enhancing the opacity by a factor η = [1 + (κbb + κbf)/κff], where κbb, κbf, and κff are the bound–bound, bound–free, and free–free opacities (Svirski et al. 2012). We crudely assume that η is independent of frequency and temperature, i.e., the free–free emissivity and absorption coefficient (Equations (22) and (23)) are multiplied by η, with everything else unchanged. We compare our results (assuming η = 1) for the E1M1 model to the case where η = 10.
The comparison can be seen in Figure 10. The left panel shows the spectrum at the shock radius, and the right panel shows the spectrum that is reprocessed through the CSM. We generally see that for a larger η there is enhancement of the optical component at the cost of reducing the X-ray component. The difference is most clearly seen at late epochs. Thus we conclude that inclusion of a more realistic bound–free opacity would be important for characterizing emission in the optical and soft X-rays, although this is included in the radiation transfer through the unshocked CSM. However, we should note that the hard X-ray spectrum should be close to the η = 1 one, as plasma reaching temperatures hotter than ∼10 keV is fully ionized and free–free emission should dominate there.
Download figure:
Standard image High-resolution image5.2. Feedback from Radiative Cooling
Another caveat of our work is that we neglect the effect of radiative cooling on the dynamics. The effect of this can be summarized in two aspects.
One is the shock propagation. In this work we have adopted the self-similar model of shock evolution by Chevalier (1982), which assumes the shock downstream is adiabatic. We observe in some cases the shock becoming radiative, where solving the momentum equation assuming a thin shell (Chevalier & Fransson 2003) is more appropriate. In this case we obtain the time evolution of the shock radius as , which differs from Equation (A4) by a factor of [α(n − 4)(n − 3)/2]1/(n−2). In the case of our model (α = 0.0595, n = 10) this is merely a ∼3% difference, which affects the bolometric light curve ( for a radiative case) by ≲10%. We conclude that refinement of the shock dynamics will only have a small impact on our results.
The other point, perhaps more important, is that we neglect the density gradient in the shocked region for simplicity. The radiative processes adopted in our model depend on the plasma density, and thus cooling and heating will accelerate if this compression is taken into account. This can be partially taken into account by adopting a lower adiabatic index close to 1, which enhances the downstream density.
To see the effects of lowering γ we compare our E1M1 model (assuming γ = 4/3) with a result assuming γ = 1.2. With this low γ the downstream density is enhanced due to the higher compression ratio (factor 11/7), and slightly more due to the lower value of α by order 10%. Figure 11 shows the comparison of these two values of γ at the same epochs. In the early phase, the density enhancement helps convert more UV/X-ray photons to optical, as shown in the left panel. This can have a large impact on the absorption of soft X-rays around 1 keV, as shown in the right panel for day 24.7. In the late phase, the density has sufficiently dropped and the resulting spectra become almost the same.
Download figure:
Standard image High-resolution image5.3. Collisionless Relaxation
We have considered Coulomb collision as a relaxation process for protons and electrons. However it is known that for interactions of the ejecta and CSM, the FS turns collisionless around breakout (Katz et al. 2011; Murase et al. 2011). As a result collisionless relaxation can occur, which can modify the relaxation timescale of the plasma. Furthermore, a non-negligible fraction of the post-shock energy may be used to accelerate electrons and protons to relativistic energies through the diffusive shock acceleration mechanism (e.g., Blandford & Eichler 1987).
The latter is proposed to give rise to various emissions in both photons and neutrinos (Katz et al. 2011; Murase et al. 2011, 2019, 2014; Petropoulou et al. 2016; Zirakashvili & Ptuskin 2016; Murase 2018). The particles and their emissions also interact with optical/X-ray photons, e.g., through IC processes and two-photon annihilation processes. Detailed modeling of the spectrum such as done in this work may thus be important for deriving nonthermal emission from these relativistic particles. In future work we plan to apply our current emission model to calculate the nonthermal emission of interaction-powered transients. We note that these photons are expected to be prominent in radio and gamma rays, and we expect inclusion of this will not greatly modify our present results, which cover the optical to X-ray bands.
6. Conclusion
In this work we constructed for the first time a theoretical model for obtaining multiwavelength spectra of interaction-powered SNe. Our model takes into account the important physical processes in the shocked region, such as the collisional relaxation of electrons and protons, and the Comptonization/absorption of bremsstrahlung radiation.
We focused on the parameter space relevant to Type IIn SNe, and found that the mass-loss rate has great impact on shaping the spectra and the relative brightness of X-rays compared with the optical/UV. In agreement with previous studies, we found that for high mass-loss rates X-ray emission is largely suppressed due to reprocessing in the shocked region and the CSM, and the transient becomes bright in the optical/UV. On the other hand our model extends previous works by considering a lower mass-loss rate, and predicts that these events can become luminous X-ray sources, with luminosity 1042–1044 erg s−1, for up to 100 days. We summarize our results in Figure 12, with quantitative values roughly estimated from the results of our work. This paradigm would be testable by current/future all-sky surveys in the X-ray wavelength.
Download figure:
Standard image High-resolution imageWe furthermore tested our model with observations of SN 2010jl and SN 2014C. The comparison with SN 2010jl shows that an asymmetric structure in the CSM is required to explain simultaneously the optical and X-ray light curves, in agreement with previous studies (Fransson et al. 2014; Chandra et al. 2015). For SN 2014C the X-ray and Hα emission seen in the late phase are roughly reproduced.
Our model only requires the shock radius, shock velocity, and upstream density as input to calculate the emitted spectrum. These parameters can be obtained not only from the self-similar solutions we adopted, but also by utilizing hydrodynamical simulations. Coupling our model with hydrodynamical simulations will greatly enhance the capability of our model, because we can relax various restrictions of the ejecta and CSM. We plan to explore such a possibility in future work.
The authors thank Kohta Murase for many valuable comments throughout the course of this work, and Satoru Katsuda for important comments on SN 2010jl. D.T. is supported by the Advanced Leading Graduate Course for Photon Science at the University of Tokyo. This work is also supported by JSPS KAKENHI grant Nos. JP19J21578, JP16H06341, JP20K04010, JP20H01904, and JP20H05639, MEXT, Japan.
Appendix A: Shock Dynamics
The dynamics of SN ejecta colliding with a CSM of a wind profile () was first obtained as a self-similar solution in Chevalier (1982). This is applicable when the swept-up mass is much smaller than the ejecta mass. The SN ejecta is assumed to be cold (i.e., negligible thermal pressure) and to have a density profile with functions of radius and time of (Matzner & McKee 1999)
where is the elapsed time since explosion introduced in Section 2.3, and g and vt are given by the ejecta mass Mej and energy Eej as
The outer ejecta have a steep density profile with n = 7–12. In this case the radius at the contact discontinuity follows a power-law time evolution (Chevalier 1982):
where α is a parameter obtained numerically and depends on n and the adiabatic index γ. The factor α1/(n−2) is of order unity in most cases. The velocities of the FS and RS upstreams in the lab frame are from the Rankine–Hugoniot relations (Tsuna et al. 2019):
where we assume a strong shock and adopt a thin-shell approximation ().
In this work we adopt n = 10, δ = 1, γ = 4/3, and Mej = 10 M⊙ unless otherwise noted. By numerically obtaining the self-similar solutions we get α ≈ 0.0595, and
where we adopt a fiducial mass-loss rate for normalization , corresponding to 10−2 M⊙ yr−1 for a wind velocity of vw = 100 km s−1. For a wind CSM (ρ ∝ r−2), the downstreams of the two shocks typically have completely different densities (Chevalier 1982). The FS downstream density ρfs is (γ + 1)/(γ − 1) = 7 times the upstream density,
From the Rankine–Hugoniot relations and adopting the thin-shell assumption, the density ratio at the RS and FS downstreams is ρrs/ρfs ≈ α−1 (Tsuna et al. 2019), which is ≈17 for our parameters.
At a sufficiently late time, the outer ejecta are completely swept up by the RS, after which the self-similarity breaks down. For the above parameters, this occurs at
We use these formulations to obtain the shock radius rsh, shock velocities (vfs and vrs), and downstream densities (ρfs and ρrs) at a given epoch {}. These values will be given as inputs to calculate the emitted radiation spectrum.
Appendix B: Ultrafast Cooling Case
When solving electron cooling in the early epochs, the free–free cooling timescale tcool in the RS region (Equation (46)) became very small, in extreme cases less than 10 s at 10 days after explosion. This resulted in difficulties in accurately obtaining the radiation energy densities from Equations (34) and (35), because to compute these equations accurately one has to keep records of jν,ff with resolution of tcool for the past tdiff (which increases with ). The need to allocate this amount of memory made it impossible with our computational resource to numerically obtain the radiation energy densities.
Fortunately though, this heavy computation is not required. In this regime, the free–free processes dominate over any other processes, and the timescale for the plasma to approach thermal equilibrium with the radiation is much shorter than the diffusion or dynamical timescale. Thus we can safely assume that the plasma is always in thermal equilibrium with radiation, with the blackbody temperature given by the total energy density ∫d ν erad,ν available at that time in the RS region. Motivated by this, we analytically obtain the approximate erad,ν and eesc,ν at the end t = tdyn in the following way.
The thermal energy initially carried by the downstream eth,rs ≡ (3/2)ni kB(Tp,rs + Te,rs) is instantaneously converted to radiation within t = tcool, and is radiated away within tdiff + tcool ≈ tdiff. There are also photons injected from the FS region at rate (erg s−1 cm−3 Hz−1); what is injected from 0 < t < tdyn − tdiff will be radiated away while the contribution over the last tdiff will remain at t = tdyn.
The bolometric values ∫d ν erad,ν and ∫d ν eesc,ν are thus
The spectrum erad,ν is a blackbody spectrum with the total energy density given in Equation (B1). On the other hand, the spectrum eesc,ν will slightly differ for (i) 0.5tdyn < tdiff < tdyn and (ii) tdiff < 0.5tdyn. For case (i), eesc,ν is a single-component blackbody spectrum with the total energy density given in Equation (B2). For case (ii), eesc,ν is a sum of two components, the first being a blackbody spectrum with total energy density , and the second being a blackbody spectrum with total energy density , multiplied by (tdyn − 2tdiff)/tdiff. This is because for case (ii), there will be a time range tdiff < t < tdyn − tdiff where the total radiation energy density is limited to .
Footnotes
- 4
The width of the downstream is compressed by a factor of (γ + 1)/(γ − 1), but from the jump condition the downstream velocity in the shock rest frame is reduced by the same factor with respect to the upstream velocity.
- 5
This is not to be confused with the variable t, which was previously defined for the calculation of the shock structure at a given epoch .
- 6
- 7
We note that these processes are taken into account in the radiation transfer calculation through the CSM done with CLOUDY, as observed in Figure 4.