A publishing partnership

The following article is Open access

A Dynamical Model for IRAS 00500+6713: The Remnant of a Type Iax Supernova SN 1181 Hosting a Double Degenerate Merger Product WD J005311

, , , , , , , , and

Published 2024 July 5 © 2024. The Author(s). Published by the American Astronomical Society.
, , Citation Takatoshi Ko et al 2024 ApJ 969 116 DOI 10.3847/1538-4357/ad4d99

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/969/2/116

Abstract

IRAS 00500+6713 is a hypothesized remnant of a Type Iax supernova SN 1181. Multiwavelength observations have revealed its complicated morphology; a dusty infrared ring is sandwiched by the inner and outer X-ray nebulae. We analyze the archival X-ray data taken by XMM-Newton and Chandra X-ray Observatory to constrain the angular radius, mass, and metal abundance of the X-ray nebulae, and construct a theoretical model describing the dynamical evolution of IRAS 00500+6713, including the effects of the interaction between the SN ejecta and the intense wind enriched with carbon-burning ashes from the central white dwarf (WD) J005311. We show that the inner X-ray nebula corresponds to the wind termination shock while the outer X-ray nebula to the shocked interface between the SN ejecta and the interstellar matter. The observed X-ray properties can be explained by our model with an ejecta kinetic energy of Eej = (0.77–1.1) × 1048 erg, an ejecta mass of Mej = 0.18–0.53 M, if the currently observed wind from WD J005311 started to blow tw ≳ 810 yr after the explosion, i.e., approximately after 1990 CE. The inferred SN properties are compatible with those of Type Iax SNe and the timing of the wind launch may correspond to the Kelvin–Helmholtz contraction of the oxygen–neon core of WD J005311 that triggered a surface carbon burning. Our analysis supports that IRAS 00500+6713 is the remnant of SN Iax 1181 produced by a double degenerate merger of oxygen–neon and carbon–oxygen WDs, and WD J005311 is the surviving merger product.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

A double white dwarf (WD) merger is an important event because it might lead to a supernova (SN) explosion if the total mass is close to or exceeds the Chandrasekhar limit (e.g., Iben & Tutukov 1984). Another possibility was proposed (Nomoto & Iben 1985; Saio & Nomoto 1985) that it results in a gravitational collapse to form a neutron star without a violent explosion such as an SN even when the total mass exceeds the Chandrasekhar limit. We do not know what will happen as a consequence of the merging especially when the total mass is close to or exceeds the Chandrasekhar limit.

IRAS 00500+6713, a Galactic infrared (IR) nebula hosting an extremely hot WD J005311 at the center, which was reported by Gvaramadze et al. (2019), may provide definitive testimony as to the fate of this kind of event. Optical spectroscopic observations identified an intense wind blowing from the WD (Gvaramadze et al. 2019; Lykou et al. 2023): the wind velocity reaches vw = 15,000 km s−1, which is significantly faster than the escape velocity from an ordinary mass WD, and the mass-loss rate is as high as those of massive stars, ${\dot{M}}_{{\rm{w}}}\sim {10}^{-6}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$. The chemical composition of the wind is dominated by oxygen, and the neon mass fraction is significantly larger than the solar abundance, indicating an enrichment of carbon-burning ashes. Given the peculiarities of the wind, WD J005311 has been proposed to be a super- or near-Chandrasekar limit WD with a strong magnetic field and a fast spin, formed via a double degenerate WD merger (Gvaramadze et al. 2019; Kashiyama et al. 2019). In fact, the photometric properties, i.e., an effective temperature of Teff ∼ 200,000 K and a bolometric luminosity of Lbol ∼ 40,000 L (Gvaramadze et al. 2019; Lykou et al. 2023), are broadly consistent with a theoretically calculated remnant WD ∼1000–10,000 yr after the merger (Schwab et al. 2016; Yao et al. 2023; Wu et al. 2023).

If indeed WD J005311 is a remnant of a double degenerate WD merger, the surrounding IRAS 00500+6713 nebula should hold the information on the merger dynamics and the postmerger evolution of the remnant. In fact, multiwavelength observations of IRAS 00500+6713 have revealed its complicated morphology (e.g., Gvaramadze et al. 2019; Ritter et al. 2021; Fesen et al. 2023). From the IR image taken by the Wide-field Infrared Survey Explorer (WISE) reported by Gvaramadze et al. (2019), it can be inferred that a dust-rich ring with an angular radius of ${\theta }_{{\rm{d}}}\sim 60^{\prime\prime} $ is surrounded by a diffuse IR halo extending to ${\theta }_{\mathrm{out}}\sim 140^{\prime\prime} $. Ritter et al. (2021) conducted a long-slit spectroscopic observation with OSIRIS on board the Gran Telescopio Canarias, and measured the expansion velocity of a layer surrounding the dust-rich ring at ${\theta }_{\mathrm{SII}}\sim 90^{\prime\prime} \,$ as vSii ∼ 1100 km s−1 from the Doppler shift of the [S ii] doublet. Given θSii , vSii , and the distance to the source d = 2.3 kpc (Bailer-Jones et al. 2021), the age of IRAS 00500+6713 can be estimated to be ∼1000 yr. On the other hand, the observation with XMM-Newton identified the X-ray counterpart of IRAS 00500+6713, consisting of the inner nebula enveloped within the dust-rich ring and the outer nebula whose angular radius is comparable to the IR halo (Oskinova et al. 2020). The X-ray spectra indicate that both the inner and outer nebulae are also enriched with carbon-burning ashes.

The outer X-ray nebula likely corresponds to the shocked interface between the interstellar matter (ISM) and an expanding matter ejected at the merger of the progenitor binary. From the angular radius θout and the emission measure (EM) of the outer X-ray nebula EMout, Oskinova et al. (2020) estimated the ejecta mass as Mej ∼ 0.1 M. Assuming that the bulk of the nebula is expanding with a velocity comparable to the dust-rich ring, i.e., ∼1000 km s−1 (Ritter et al. 2021), the kinetic energy is estimated to be Eej ∼ 1048 erg. Such properties are compatible with so-called Type Iax supernovae (SNe Iax; Foley et al. 2013), which can be accompanied by the merger of carbon–oxygen and/or oxygen–neon WDs (see, e.g., Kashyap et al. 2018). Interestingly, Ritter et al. (2021) and Lykou et al. (2023) pointed out that the sky location and the timing of the hypothesized merger are in accord with the ancient records on a historical Galactic SN, SN 1181. The documented apparent magnitude of SN 1181 is also consistent with SNe Iax, although the uncertainties are large (Schaefer 2023). In addition, recently, Fesen et al. (2023) analyzed deep optical [S ii] λ λ6716, 6713 images and revealed radially aligned filaments with angular scales of 5''–20''. The authors proposed that these filaments can be formed via the photoionization of the ejecta if the ionizing photons are produced by the dissipation of the wind luminosity and injected into a clumpy region.

The origin of the inner X-ray nebula is uncertain; the XMM image was reported to be consistent with a point source (Oskinova et al. 2020). Given that the intense wind with vw = 15,000 km s−1 is blowing (Lykou et al. 2023) toward the IR halo expanding at a slower velocity vSii ∼ 1100 km s−1 (Ritter et al. 2021), the wind termination shock would be the most likely possibility. As such, the intense wind can have significant impacts on the structure of IRAS 00500+6713, and accordingly, on the values of the physical parameters estimated from the observed morphology. In addition, the history of the wind mass loss is directly connected to the time evolution and the fate of the merger product WD. Thus, it is crucial to construct a dynamical model that cohesively contains the wind, the SN ejecta, the ISM, and intervening multiple shocks.

In this paper, we first analyze the X-ray data on IRAS 00500+6713 obtained with XMM-Newton and Chandra X-ray Observatory (Chandra) to determine the angular radius, EM, and metal abundance of the X-ray nebulae, taking into account the possible non–collisional ionization equilibrium (non-CIE) effects (Section 2). In particular, we newly obtain a constraint on the angular radius of the inner X-ray nebula from the Chandra data. Next, we construct a dynamical model based on the hypothesis that IRAS 00500+6713 is the remnant of SN 1181 (Section 3). We consistently solve the dynamics of the inner and outer X-ray nebulae, allowing that the currently observed wind started to blow a finite time after SN 1181. We use the dynamical model to constrain the relevant physical parameters of the system, i.e., the kinetic energy and mass of the SN ejecta, the timing of the launch of the wind, from the currently observed multiwavelength morphology. We discuss the implications of the results of our dynamical model in Section 4.

2. X-Ray Data Analysis

2.1. Observations and Data Reduction

IRAS 00500+6713 was observed with XMM-Newton in 2019 and 2021 and with Chandra in 2021. We use all the available archival data obtained with the CCDs on board both observatories, XMM EPIC (MOS1, MOS2, and pn; Turner et al. 2001; Strüder et al. 2001; Jansen et al. 2001) and Chandra ACIS (Garmire 1997) for the imaging analysis. For the spectroscopy, we simultaneously use imaging spectroscopic data obtained with XMM EPIC and grating data obtained with XMM Reflection Grating Spectrometer (RGS; den Herder et al. 2001). Table 1 lists the XMM and Chandra observation logs. We use the XMM Science Analysis Software (SAS; v19.1.0; Gabriel et al. 2004) package, which also includes the Extended Source Analysis Software (ESAS; Snowden et al. 2004). We process the EPIC data with the SAS tasks pn-filter, epchain, emchain, and mos-filter. The RGS data are processed using the standard pipeline tool rgsproc in the SAS package. We exclude one observation (ObsID 0872590301) because it shows high count rates over the whole field of view and is affected by background flares. The total effective exposures for the EPIC and RGS are ≈76 and ≈102 ks, respectively. For the Chandra data, we process the data with the CIAO (v4.15; Fruscione et al. 2006) tool chandra_repro. The total effective exposure is ≈144 ks. Considering the relatively low statistics of the Chandra data, we use them only for the imaging analysis (Section 2.2.1). All the five observations are merged for use in the analysis.

Table 1. XMM and Chandra Observation Logs of IRAS 00500+6713

 ObsIDDateExposure a PI Name
   (ks) 
   MOS2RGS1 
XMM08416401012019 Jul 814.219.5L. Oskinova
 08416402012019 Jul 2413.416.7L. Oskinova
 08725901012021 Jan 86.811.8L. Oskinova
 08725902012021 Jan 1010.912.9L. Oskinova
 0872590301 b 2021 Jan 144.07.9L. Oskinova
 08725904012021 Jan 167.717.2L. Oskinova
 08725905012021 Jan 2010.412.7L. Oskinova
 08725906012021 Jan 1812.214.6L. Oskinova
Chandra234192021 May 1227.7L. Oskinova
 243422021 May 1714.8L. Oskinova
 243432021 May 1427.7L. Oskinova
 243442021 Oct 2138.0L. Oskinova
 243452021 Dec 1520.8L. Oskinova
 250452021 May 1714.8L. Oskinova

Notes.

a Effective exposures of MOS2 and RGS1 are presented representing the instruments. b Excluded from the analysis due to a high background rate.

Download table as:  ASCIITypeset image

In the data analysis, we use HEASoft (v6.20; HEASARC 2014), XSPEC (v12.9.1; Arnaud 1996), and AtomDB (v3.0.9; Smith et al. 2001; Foster et al. 2012). The cstat (Cash 1979) and wstat 12 implemented in XSPEC are used in our spectral studies. In this section, uncertainties in the text, figures, and tables indicate 1σ confidence intervals. The upper and lower limits presented in this section are at a 95% confidence level.

2.2. Analysis and Results

2.2.1. Spatial Distributions

Figure 1 shows 0.5–5.0 keV X-ray images obtained with XMM EPIC-pn and Chandra ACIS. One can see a central source (≲20'') and a faint diffuse emission around it (≲100''). We exclude data in the lower energies, considering less accurate detector calibrations, and in the higher energies because of higher background rates. The central source is found to be slightly extended with the spatial resolution of Chandra (≈0farcs5). The diffuse emission extending to ∼100'' is not visible with Chandra because of the high detector background rate, which is ≈10 times higher than the expected diffuse flux. In order to evaluate the spatial distribution of the source, we extract radial profiles around the central source. The central position for extraction is set to (R.A., decl.)2000 = (13fdg2967, 67fdg5007) by referring to the SIMBAD catalog. 13 We take into account spatially dependent exposures (depending on off-axis angles), bad pixels, and gaps between the CCDs. For XMM EPIC, we obtain radial flux profiles for each instrument for each observation. For Chandra ACIS, we merge all the observations and extract radial profiles considering almost the same pointing directions and observation dates close to each other. In Figure 2, we show flux profiles extracted from one XMM observation in 2019 (ObsID 0841640101) and the merged Chandra data.

Figure 1.

Figure 1. 0.5–5.0 keV images of IRAS 00500+6713 obtained with XMM EPIC-pn (left) and Chandra ACIS (right). For the XMM image, all the observations but one with ObsID 0872590301 are merged. All the observations are used to make the Chandra image. In the left panel, circles with radii of 20'', 100'', and 200'' are overlaid in green. The former two correspond to the approximate spatial extents of the central and diffuse sources. The central black square in the XMM image shows the range of the Chandra image. The images are shown on a logarithmic scale.

Standard image High-resolution image
Figure 2.

Figure 2. Radial flux profiles of IRAS 00500+6713 obtained with XMM MOS1, MOS2, pn, and Chandra ACIS. The XMM data are extracted from one observation in 2019 (ObsID 0841640101). All the observations are merged for the Chandra data. The energy range for extraction is 0.5–5.0 keV. The solid lines represent the best-fit model components.

Standard image High-resolution image

We compare the observed distributions with a phenomenological model. The instrumental point spread functions (PSFs), i.e., the radial profiles of an ideal point source on the CCDs, are calculated for the source position with the SAS tool psfgen and CIAO tool simulate_psf. We assume a simple emission model for the central and diffuse sources: a sphere with a uniform and isotropic emissivity over the whole volume. The model for the radial flux profile (F(r), where r is the radius) is described as

Equation (1)

with P(r, θin), D(r, θout), and B is the fluxes of the central source, diffuse source, and background (assumed to be uniform), respectively. The parameters θin and θout indicate the spatial extents of the central and diffuse sources, respectively. P(r, θin) and D(r, θout) are spherical emission convolved with the instrumental PSFs. We mainly focus on the evaluation of θin and θout.

We use the Markov Chain Monte Carlo algorithm 14 to determine the best-fit model parameters and their confidence ranges. To perform a precise evaluation, the profiles extracted from all the instruments in all the observations (21 in total) are modeled simultaneously via linked θin and θout for the XMM data. The normalizations of P(r, θin) and D(r, θout), and B for individual instruments are treated as free parameters. Radial ranges of 0–300 and 0''–10'' are used for the modeling of the XMM and Chandra data, respectively. As a result, we obtain θin < 1farcs5 and ${\theta }_{\mathrm{out}}\,=\,{131}_{-2-2}^{+2}$'' 15 with the XMM data. We obtain the corresponding $-2\mathrm{log}L/\mathrm{dof}\,=\,2597/2026$, where L and dof are the likelihood value and degree of freedom, respectively. This likelihood value is marginally acceptable, considering that we are using a simple phenomenological model. The central source is consistent with a pointlike source for XMM. The spatial extent of the diffuse source is similar to that in IR (Gvaramadze et al. 2019), as pointed out by Oskinova et al. (2020). On the other hand, we are able to constrain the radius of the central source as ${\theta }_{\mathrm{in}}\,=\,{1.52}_{-0.08-0.67}^{+0.08}$'' 16 with Chandra. The corresponding $-2\mathrm{log}L/\mathrm{dof}\,=\,29.0/9$, is also marginally acceptable. The best-fit models are overlaid in Figure 2. We also search for a possible time variation of the spatial distributions by separately modeling the XMM flux profiles in 2019 and 2021. We find results consistent with each other, θin < 1farcs5 and θout = 132'' ± 2'' in 2019 and θin < 1farcs5 and θout = 129'' ± 3'' in 2021, which means that there is almost no time variation in the sizes of the X-ray emitting regions. Throughout this paper, we adopt the value ${\theta }_{\mathrm{out}}={131}_{-2-2}^{+2}$'' obtained using all the observations simultaneously.

The background rates B determined from the XMM data, e.g., ≈0.0023 counts s−1 arcmin−2 (0.5–5.0 keV) for MOS1 in 2021, are consistent with or slightly higher than the pure particle background rate (Kuntz & Snowden 2008). This is reasonable considering the contribution of the sky background. The diffuse plus background flux D(r, θout) + B for the Chandra data are also consistent with the sum of the diffuse flux determined from the XMM data and the particle background rate (Bartalucci et al. 2014; Suzuki et al. 2021). Using the observed constraints on ${\theta }_{\mathrm{in}}={1.52}_{-0.08-0.67}^{+0.08}$'' with 1σ uncertainties, the minimum and maximum values of θin are 0farcs77 and 1farcs60, respectively. Adopting the Gaia distance of 2.3 kpc, the minimum and maximum physical radii are 0.0087 and 0.018 pc. These values are used to compare with our model. The spread of the inner source rin = 0.0087–0.018 pc is much larger than the photospheric radius ∼1 × 1010 cm (Gvaramadze et al. 2019). This indicates that the inner shocked region has certainly spread, likely due to the wind from the central WD. However, since the wind-crossing time rin/vw ≲ 1 yr is much shorter than the SN age, this region may have been formed very recently.

2.2.2. Spectral Properties of the Thermal Plasmas

We here extract spectra and estimate the temperatures, metal abundances, ionization timescale, and EMs of the central and diffuse sources. We use the XMM data, which have more than twice greater statistics than the Chandra data. Spectral extraction regions for the EPIC data are selected based on the imaging analysis results (Figure 2) and are shown in Figure 1. A circle with a 20'' radius is used for the central source, and an annulus with inner and outer radii of 30'' and 100'', respectively, is assumed for the diffuse source. As the background, we use the diffuse source and an annulus region with inner and outer radii of 140'' and 200'', respectively. The contamination of the diffuse emission to the background region is <1% of the total flux, according to the results of our imaging analysis. The central position for extraction is the same as that used for the extraction of the radial profiles. Since the RGS spectrum has much poorer statistics (<1% of the EPIC data), we only use it as a consistency check for the spectral analysis of the bright central source. We merge all the spectra in each year and for each instrument: the first- and second-order spectra are also combined for the RGS data. The spatial variation of the detector background is negligible, considering the on-axis position and sizes of our analysis regions (Kuntz & Snowden 2008).

As the spectral model, we consider optically thin thermal plasmas in two different assumptions of metalicity by referring to Oskinova et al. (2020): near solar and pure metal. Under each assumption, we try several models with different free parameters and finally adopt the simplest model that explains the data sufficiently well. Note that we do not pre-assume the abundance patterns reported in other wavelengths. We find an absorbed ionizing plasma model (tbabs × nei in XSPEC) explains both central and diffuse sources. The treatment of the model parameters for the central source is as follows: in the near-solar case, the hydrogen column density (NH), electron temperature (kB Te), ionization timescale (τ), EM, and abundances of O, Mg, and Fe (=Ni) are treated as free parameters. The other metal abundances are fixed to solar. Here, kB denotes the Boltzmann constant. In the pure-metal case, the same model but without H, He, N, Ar, and Ca is used. 17 For the diffuse source, the free parameters are the same as above but the abundance of O is fixed to solar and Ne is set free. The spectra obtained with all the instruments in 2019 and 2021 (six in total) are modeled simultaneously. We use the energy ranges of 0.5–5.0 keV and 0.5–2.0 keV for the spectral fit of the central and diffuse sources, respectively, considering their statistics.

The spectra and best-fit models are presented in Figure 3. Table 2 shows the best-fit parameters. Both models explain well both central and diffuse sources with acceptable fit statistics. The EMs (hereafter EMin and EMout for the central and diffuse sources, respectively) derived for the pure-metal and near-solar plasmas differ by 2 orders of magnitude because of the difference in the total ion emissivity. The abundance ratios are C/O = 0.014 ± 0.003, Ne/O = 0.005 ± 0.001, Mg/O = 0.0010 ± 0.0004, Fe/O = 0.005 ± 0.002 for the central source, and C/O = 0.43 (fixed), Ne/O = 0.41 ± 0.07, Mg/O = 0.04 ± 0.01, Fe/O < 0.003 for the diffuse source (all in number). The mass fractions in the pure-metal case are presented in Table 3. Oskinova et al. (2020) applied a two-temperature, CIE thermal plasma model to these sources. Their assumption of the abundance pattern is partly based on the IR observations (Gvaramadze et al. 2019) and thus different from ours. Although the mass fractions in this work differ from previous studies (Gvaramadze et al. 2019; Oskinova et al. 2020; Lykou et al. 2023), the tendency is similar: C/O < 1 and Ne/O < 1 for the central source and C/O∼1 and Ne/O∼1 for the diffuse source. Compared to Oskinova et al. (2020), the electron temperatures we obtain here lie between their higher and lower temperatures. The absorption column densities are similar.

Figure 3.

Figure 3. Energy spectra extracted from the central pointlike source (panels (a) and (b)) and surrounding diffuse source (panels (c) and (d)). Multiple observations in each year are merged. The solid lines show the best-fit models in the two cases (near solar and pure metal). The background spectra are subtracted from all the data. The XMM RGS spectrum, which is used for a consistency check, is presented in panel (a).

Standard image High-resolution image

Table 2. Best-fit Spectral Parameters for the Pointlike and Diffuse Sources

   NH kB Te τa O b Ne b Mg b Fe (=Ni) b EM c C-stat/dof d
  (1022 cm−2)(keV)(1010 s cm−3)    (1052 cm−3) 
Point sourceNear solar0.78 ± 0.051.30 ± 0.087.9 ± 1.331 ± 81 (fixed)0.68 ± 0.252.81 ± 0.63(1.8 ± 0.4) × 102 4406/5409
 Pure metal0.75 ± 0.051.21 ± 0.069.5 ± 1.437 ± 91 (fixed)0.70 ± 0.252.62 ± 0.626.3 ± 1.14405/5409
Diffuse sourceNear solar0.51 ± 0.050.56 ± 0.125.9 ± 4.01 (fixed)2.67 ± 0.451.04 ± 0.270.04 ± 0.04(3.7 ± 1.0) × 102 2113/1809
 Pure metal0.53 ± 0.050.29 ± 0.03>531 (fixed)1.57 ± 0.270.74 ± 0.240.04 ± 0.043.6 ± 0.82105/1809

Notes.

a Ionization timescale. b Metal abundance normalized by that of C, i.e., (O/O)/(C/C). c EM, ne nion V. The parameters ne, nion, and V represent electron and ion number densities, and plasma volume, respectively. A distance of 2.3 kpc is assumed. d C-stat and dof represent the C-statistic value and degree of freedom, respectively.

Download table as:  ASCIITypeset image

Table 3. Mass Fractions of the Central and Diffuse Sources Derived in the Pure-metal Case

ElementCentralDiffuse
C ${0.008}_{-0.008}^{+0.072}$ 0.18
O ${0.97}_{-0.24}^{+0.03}$ 0.56
Ne ${0.005}_{-0.005}^{+0.045}$ 0.16 ± 0.03
Mg ${0.004}_{-0.004}^{+0.037}$ 0.09 ± 0.02
Fe ${0.010}_{-0.010}^{+0.092}$ 0.004 ± 0.004

Download table as:  ASCIITypeset image

Note that we use the data with approximately twice better statistics than those used in Oskinova et al. (2020). Oskinova et al. (2020) suggested that an additional nonthermal component might contribute to the central region. If we add an additional nonthermal component described with a power law with the fixed index of 2.4 following Oskinova et al. (2020), we indeed obtain an improved fit with C-statistic/dof = 4294/5408. However, since the fit statistics before adding a nonthermal component are already acceptable, we cannot claim that the central source requires a nonthermal component. We also search for possible spectral changes as a function of time but find no such evidence.

3. Dynamical Model

In the previous section, we have determined the amounts and spatial extents of the X-ray emitting plasma in the inner and outer nebulae of IRAS 00500+6713, i.e., the emission measures (EMin and EMout) and the angular radii (θin and θout). Here, we construct a dynamical model of IRAS 00500+6713 for interpreting the currently observed X-ray properties, assuming that this nebula is the remnant of SN 1181. The schematic picture of the overall structure is shown in the right part of Figure 4. The inner shocked region is produced by the interaction between the ejecta of SN 1181, set to be launched at t = 0, and the wind from the central WD that started blowing at t = tw. On the other hand, the outer shocked region is produced by the interaction between the SN ejecta and the surrounding ISM. A comparison between this schematic picture and the observed image is shown on the left of Figure 4. We follow the dynamical evolution of the inner and outer X-ray nebulae from the onset of the SN explosion (t = 0) through t = 845 yr (t = 840 yr corresponds to 2021 CE) to determine the characteristic quantities of the system, i.e., the ejecta kinetic energy Eej and the ejecta mass Mej of SN 1181 and the timing tw of the launch of the wind that reproduce the X-ray properties.

Figure 4.

Figure 4. A comparison between an observed image (left panel; X-ray image (XMM) and IR contours (WISE)) and our schematic picture of IRAS 00500+6713 (right panel). The schematic picture includes the morphology of IRAS 00500+6713 inferred from multiwavelength observations (the left semicircle) and our dynamical model (the right semicircle). A dusty IR ring is sandwiched by the inner and outer X-ray nebulae. We assume that the inner X-ray nebula corresponds to the wind termination shock while the outer X-ray nebula to the shocked interface between the ejecta of SN 1181 and the ISM, and the dust ring corresponds to the unshocked SN ejecta.

Standard image High-resolution image

We note that other than EMs and θs, there are observed properties of IRAS 00500+6713 that might be useful to constrain the model. For example, the electron temperatures kB Te and the ionization timescales τ have been estimated. In addition, the previous mid-IR observations determined the angular radius θd of the dust-rich ring (Gvaramadze et al. 2019), and the expansion velocity of the shock-excited ejecta vSii (Ritter et al. 2021). We mainly use these pieces of information to verify the model that fits the EMs and θs of the X-ray nebulae. The ranges of physical parameters that best reproduce the X-ray observations are listed in Table 4.

Table 4. Parameters in Our Dynamical Model for IRAS 00500+6713

 Parameters Estimated from Observations 
Distance d [kpc] a ${2.3}_{-0.1}^{+0.1}$
Wind mass-loss rate ${\dot{M}}_{{\rm{w}}}$ [M yr−1] b 7.5 × 10−7–4 × 10−6
Wind mechanical luminosity Lw [erg s−1] b (0.53–2.8) × 1038
Angular radius θin [arcsec] c 0.77 − 1.6
  θout [arcsec] c ${131}_{-2-2}^{+2}$
  θd [arcsec] d ${60}_{-12}^{+12}$
Expansion velocity vS ii [km s−1] e ${1100}_{-100}^{+100}$
EMEMin [1052 cm−3] c 180 ± 40 (near solar)
  6.3 ± 1.1 (pure metal)
 EMout [1052 cm−3] c 370 ± 100 (near solar)
  3.6 ± 0.8 (pure metal)
 Parameters Estimated from Our Model 
Power-law index of the SN ejecta density profile n 6 (fixed)
  δ 1.5 (fixed)
Number density of ISM nISM [cm−3]0.097–0.13
Ejecta kinetic energy Eej [1048 erg]0.77–1.1
Ejecta mass Mej [M]0.18–0.53
Wind launching time tw [yr]810–828

Notes. The references to the parameters in the upper table are as follows.

a Gaia Collaboration et al. (2021). b Lykou et al. (2023). c This work. d WISE image reported by Gvaramadze et al. (2019). e Ritter et al. (2021).

Download table as:  ASCIITypeset image

3.1. Density Profiles

In order to calculate the evolution of this system, the density profile of each region should be given. The density profile of the SN ejecta is often described by a double power law (Chevalier & Soker 1989; Matzner & McKee 1999; Moriya et al. 2013);

Equation (2)

where rbr is the radius where the exponent changes, and ρej,br is the density at r = rbr. These parameters are obtained from the total mass (Mej) and kinetic energy (Eej) of the SN ejecta as follows (Chevalier & Soker 1989; Matzner & McKee 1999; Moriya et al. 2013):

Equation (3)

Equation (4)

Equation (5)

Here, g is a constant obtained for a given δ, n, Eej, and Mej. In order to have a finite total mass and energy, n > 5 and δ < 3 are required (Chevalier 1982). We assume that the SN ejecta are surrounded by ISM with a constant mass density of

Equation (6)

where

Equation (7)

is the mean molecular weight assuming the solar abundance, mu is the atomic mass unit, and nISM is the number density of the ISM. On the other hand, WD J005311 resides inside the SN ejecta and blows an intense wind with a mass-loss rate of ${\dot{M}}_{{\rm{w}}}=(7.5\mbox{--}40)\times {10}^{-7}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ and velocity of vw = 15,000 km s−1 (Lykou et al. 2023). Accordingly, the mechanical luminosity of the wind can be estimated as ${L}_{{\rm{w}}}={\dot{M}}_{{\rm{w}}}{v}_{{\rm{w}}}^{2}/2=(0.53\mbox{--}2.8)\times {10}^{38}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$. We assume that the currently observed wind from WD J005311 started blowing at t = tw with a constant mass-loss rate. The density profile in the wind region is described as

Equation (8)

The radial density profile is shown schematically in Figure 5. While the SN ejecta expand in the ISM, the stellar wind catches up and collides with the SN ejecta. Two shocked regions form: the outer one between the SN ejecta and the ISM and the inner one between the wind and the SN ejecta. The outer and the inner X-ray nebulae are powered by these two shocked regions, which are analyzed in Sections 3.2 and 3.3, respectively. Even though the near-surface temperature of WD J005311 is expected to be as high as ≳107 K (Kashiyama et al. 2019), the observed effective temperature is ${T}_{{\rm{eff}}}={\text{211,000}}_{-23.000}^{+\text{40,000}}\,{\rm{K}}$ (Gvaramadze et al. 2019). This is due to significant adiabatic cooling in the optically thick wind, and thus the emission from the photosphere cannot explain the observed X-rays from the inner nebula. In between the two shocked regions, there can be an unshocked layer. In our view, both the dusty IR ring detected in the WISE image and the [S ii] line emission region detected by OSIRIS correspond to this unshocked layer; when the shocks sweep through this layer, dust grains are expected to be destroyed.

Figure 5.

Figure 5. A schematic picture of the radial density profile of IRAS 00500+6713. The curved regions represent shocked regions and the dotted lines surrounding them represent the forward or reverse shocks.

Standard image High-resolution image

3.2. Outer X-Ray Nebula

Let us first consider the outer X-ray nebula. For homologously expanding SN ejecta sweeping a uniform ISM described in Section 3.1, we can employ the self-similar solution of Chevalier (1982) to describe the evolution of the shocked region. The actual values of the exponents of the density profile should depend on the progenitor system and the mechanism of the SN explosion. While the exponent in the range 7 ≲ n ≲ 10 has been inferred from simulations for Type Ia SNe (e.g., Colgate & McKee 1969; Kasen 2010), the exponent may become smaller for weaker explosions such as Type Iax SNe. Actually, simulations of the mass eruption from a massive star before core-collapse have found a profile of n ∼ 6 (e.g., Kuriyama & Shigeyama 2020; Tsuna et al. 2020). Given that, we consider a range of 6 ≤ n ≤ 10 and set n = 6 as the fiducial value. In this case, the radii of the forward shock rout,f, the contact discontinuity rout,c, and the reverse shock rout,r (see Figure 5) all evolve with the same power law in time,

Equation (9)

The proportionality factors Bi for these radii, as well as the density, pressure, and velocity profiles inside both the shocked ejecta and the shocked ISM, are obtained for a given set of parameters (Eej, Mej, nISM). Here, we use the code developed in Tsuna et al. (2019) that calculates the hydrodynamical profiles of the shocked region, for an arbitrary value of the adiabatic index and power-law density profiles of the ejecta and ISM for which a self-similar solution exists. We assume an adiabatic index of γ = 5/3, because this region consists of approximately adiabatic, nonrelativistic ionized gas.

3.2.1. Angular Radius

As we have shown in Section 2, the angular radius of the outer X-ray nebula is ${\theta }_{\mathrm{out}}={131}_{-2-2}^{+2}{^{\prime\prime}} $ based on the XMM-Newton observation. In our model, this corresponds to the forward shock radius rout,f at t = 840 yr:

Equation (10)

Note that the estimated error of rout,f is smaller than the errors of the other parameters, thus we neglect the error hereafter. Applying this condition to the self-similar solution described above, we obtain the condition

Equation (11)

3.2.2. EM

The EM of the outer X-ray nebula can be described as

Equation (12)

The first term in the right-hand side of Equation (12) represents the contribution from the reverse shock or the shocked ejecta; μr is the mean molecular weight and ρ(r) is the density. The second term represents the contribution from the forward shock, or the shocked ISM, where μf is the mean molecular weight. Using the shock structure obtained from the self-similar solution, the integrals of Equation (12) can be calculated as

Equation (13)

Equation (14)

Note that they do not depend on either Eej or Mej for a given rout,f. The mean molecular weights, μr and μf, are evaluated as follows. When the composition is pure metal, which is suggested by our analysis of the observed X-ray spectrum in Section 2, the mean molecular weights can be expressed as follows: in the outer regions

Equation (15)

where

Equation (16)

Equation (17)

Here, the subscript "pm" represents the physical quantities for the case of pure-metal abundance, Xi denotes the mass fraction of element i, Ai its mass number, and 〈ni is the average number of free electrons originating from element i. We adopt these values shown in Table 3. The number of free electrons is obtained from AtomDB 3.0.9 (Foster et al. 2012) using the abundance ratios and the ionization degree for each element. We have used kB Te and τ obtained from our X-ray analysis to estimate the degree of ionization. On the other hand, the mean molecular weight for the solar abundance is as follows:

Equation (18)

where

Equation (19)

Equation (20)

Here, X and Y are the mass fractions of hydrogen and helium, respectively.

The EM obtained from the X-ray analysis assumes a one-zone uniform shocked gas, while our EM model is a two-zone with different abundances. Though this makes a straightforward comparison difficult, the abundance of the mixture of the two regions is expected to be close to the ISM composition in the above self-similar solution because the mass ratio of the shocked ejecta to the shocked ISM is determined to be 0.28 (Chevalier 1982). Thus, we compare the calculated EM (Equation (12), adopting μr = μout,pm, μf = μout,solar) with the value obtained by the X-ray analysis assuming the near-solar abundance (EMout = 3.7 ± 1.0 × 1054 cm−3, Table 4). By substituting Equations (13)–(20) into Equation (12), we can determine the ISM number density as

Equation (21)

Here, we have used μISM in Equation (7). Substituting this into Equation (11), we obtain the condition

Equation (22)

Figure 6 shows the obtained relation between Eej and Mej. Given that IRAS 00500+6713 hosts a WD remnant, the ejecta mass would not be larger than Mej ≲ 1 M. In this case, the ejecta kinetic energy is constrained to be Eej ≲ 1048 erg only from the outer X-ray nebula. These results are broadly consistent with the previous work (Oskinova et al. 2020).

Figure 6.

Figure 6. The range of the ejecta kinetic energy Eej and the ejecta mass Mej of SN 1181 obtained from our modeling of the outer X-ray nebula of IRAS 00500+6713 (Equation (22)). The blue line and the green shaded region correspond to the best-fit value and 2σ uncertainties of the EMout obtained in Section 2.

Standard image High-resolution image

3.3. Inner X-Ray Nebula

Next, we consider the inner X-ray nebula. We calculate the position of the contact surface between the shocked wind and shocked ejecta using the thin-shell approximation (e.g., Chevalier & Fransson 1992). The mass, momentum, and energy conservations of the shell can be described as

Equation (23)

Equation (24)

Equation (25)

respectively. Here, Msh,in, rsh, and vsh are the mass, radius, and velocity of the shocked shell, respectively. The mass of the central star is denoted by M*, psh is the pressure in the wind bubble, ρej is the unshocked ejecta density, and vej is the velocity. In Equations (23) and (25), we assume that the tail part of the SN ejecta are falling back toward the central WD with a velocity of $-\sqrt{2{{GM}}_{* }/{r}_{\mathrm{sh}}}$. This can be justified given that the inferred ejecta kinetic energy from the outer X-ray nebula, Eej ∼ 1048 erg (see Figure 6), is much smaller than the gravitational binding energy of the central WD, ${{GM}}_{* }^{2}/{R}_{* }\sim {10}^{50}\,\mathrm{erg}$ (Kashiyama et al. 2019). Here, R* refers to the radius of the central WD, whose value is around 109 cm (Kashiyama et al. 2019). Note that R* is different from the photospheric radius (∼1 × 1010 cm) since the wind is optically thick, and most of the WD mass is contained within R*. The density at the forward shock front is estimated from Equation (2) with fixing δ = 1.5, which is appropriate for a fallback tail (e.g., Tsuna et al. 2021b). 18

In order to follow the evolution of the inner nebula, we integrate Equations (23)–(25) from t = tw for a given set of (tw, Eej, Mej). Substituting the value of EMout measured from the X-ray data for the outer nebula into Equation (22), we have two independent parameters (tw, Mej). The initial conditions are given as $[{v}_{\mathrm{sh},0},{p}_{\mathrm{sh},0},{M}_{\mathrm{sh},0}]=[-\sqrt{2{{GM}}_{* }/{r}_{\mathrm{sh},0}}$, $3(\gamma -1){L}_{{\rm{w}}}/4\pi {r}_{\mathrm{sh},0}^{2}{v}_{{\rm{w}}}$, and ${\int }_{{R}_{* }}^{{r}_{\mathrm{sh},0}}4\pi {r}^{2}{\rho }_{\mathrm{ej}}(r){dr}]$. We set the initial shell radius sufficiently small, rsh,0 = 1010 cm so that the initial conditions do not affect the evolution on a timescale of 100 yr. We investigate a range of wind mass-loss rate ${\dot{M}}_{{\rm{w}}}\,=(7.5\mbox{--}40)\times {10}^{-7}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ (Lykou et al. 2023), inferred from the optical spectroscopic observations.

3.3.1. Angular Radius

As we show in Section 2, the angular radius of the inner X-ray nebula is constrained as $0\buildrel{\prime\prime}\over{.} 77\lt {\theta }_{\mathrm{in}}\lt 1\buildrel{\prime\prime}\over{.} 6$ based on the Chandra observation in 2021. In our model, this corresponds to the thin-shell radius, i.e., the radius of the wind termination shock rsh at t = 840 yr:

Equation (26)

Figure 7 shows the time evolution of the inner nebula radius calculated for several parameter sets. One can see that rsh at t = 840 yr becomes larger for a smaller Mej, a larger ${\dot{M}}_{{\rm{w}}}$, and/or smaller tw. This is simply because the wind keeps pushing the fallback ejecta since its launch. Therefore, too small Mej or too small tw leads to a too-expanded shell, which is inconsistent with the Chandra observation. This can give a constraint on the parameter set $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})$.

Figure 7.

Figure 7. The time evolution of the radius of the inner X-ray nebula rsh calculated for several parameter sets $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})$ using our dynamical model. The red line indicates the value obtained by the Chandra observations.

Standard image High-resolution image

3.3.2. EM

By modeling the expansion of the inner nebula, we can also estimate the total mass in the shocked region. Figure 8 shows the time evolution of the shocked mass for a case with $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})=(1\times {10}^{-6}\,{M}_{\odot }\,{\mathrm{yr}}^{-1},820\,\mathrm{yr},0.3\,{M}_{\odot })$. The solid and dashed lines indicate the contribution from the shocked wind swept by the reverse shock and the shocked-SN ejecta swept by the forward shock, respectively. One can see that the latter is 2–3 orders of magnitude larger than the former. Importantly, however, the shocked-SN ejecta component will not contribute to the EM of the inner X-ray nebula; as we show in the Appendix, the electron temperature is kept below 3–4 eV in the shocked-SN ejecta. Thus, we only consider the shocked wind as the source of the inner X-ray nebula. In this case, the EM of the inner nebula can be estimated as

Equation (27)

where ${M}_{\mathrm{sh},{\rm{w}}}={\dot{M}}_{{\rm{w}}}(t-{t}_{{\rm{w}}})$ is the total mass of the shocked wind, and

Equation (28)

is the volume of the shocked-wind region. Here, we estimate the positions of the reverse shock fronts as

Equation (29)

With an approximation that the shock region is plane-parallel and adiabatic, the compression ratio is ${ \mathcal C }=(\gamma -1)/(\gamma +1)=4$ for γ = 5/3 in the strong shock limit. Since we consider the inner X-ray region to be formed by the interaction between the C-O-rich wind from the central WD and the unshocked ejecta, we consider the abundance in this region to be pure metal, and adopt the value of EMin obtained from the analysis under the pure-metal assumption. The mean molecular weights in the shocked-wind region are calculated for pure-metal abundance with ionization inferred from the X-ray spectroscopy in Section 2:

Equation (30)

Equation (31)

Figure 8.

Figure 8. The time evolution of the mass in the shocked-wind (red solid line) and the shocked-SN ejecta (blue dashed line) of the inner X-ray nebula, calculated for a parameter set $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})=(1\times {10}^{-6}\,{M}_{\odot }\,{\mathrm{yr}}^{-1},820\,\mathrm{yr},0.3\,{M}_{\odot })$ using our dynamical model. The green dotted line indicates the year 2021 CE when the X-ray observations were done.

Standard image High-resolution image

By comparing the calculated EMin of the shocked wind (Equation (27)) with the observed value for the pure-metal model (Tables 2 and 4), we can obtain an additional constraint on our model parameters. Here, as shown in Table 4, the value of EMin is constrained from the X-ray analysis as follows:

Equation (32)

Figure 9 shows the time evolution of EMin calculated for several parameter sets $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})$ =(2 × 10−6 M yr−1, 820 yr, 0.6 M), (2 × 10−6 M yr−1, 820 yr, 0.3 M), and (4 × 10−6 M yr−1, 820 yr, 0.6 M). As expected, a larger ${\dot{M}}_{{\rm{w}}}$ gives a larger shocked mass and thus a larger EMin. In addition, a larger Mej also gives a larger EMin; this is because the inner nebula expansion becomes slower for a larger Mej and thus the volume of the inner nebula Vsh,w becomes smaller (see Equation (27)).

Figure 9.

Figure 9. The time evolution of the EM of the inner X-ray nebula EMin calculated for several parameter sets $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})$ using our dynamical model. The red bar indicates the value obtained by our spectral analysis of the XMM-Newton observations in Section 2.2.2.

Standard image High-resolution image

3.4. Model Parameter Determination

Combining all the conditions shown in the previous subsections, we here determine the parameters of our dynamical model. First, we constrain the ISM density nISM (Equation (21)) and obtain a relation between Eej and Mej (Equation (22)) from the observed size and EM of the outer X-ray nebula. Then, we search sets of the residual parameters (tw, Mej) that can consistently explain the observed size and EM of the inner X-ray nebula. In doing so, we allow the wind mass-loss rate ${\dot{M}}_{{\rm{w}}}$ to vary in the range inferred from the optical spectroscopy. In addition, we perform parameter search (tw, Mej) in the range of 12 yr < Δtw < 840 yr and Mej < 3 M. Here, Δtw = 840 yr − tw [yr] refers to the timing of the wind launch. The lower limit of Δtw > 12 yr comes from the constraints of Swift XRT detection in 2009 CE, which identified an X-ray counterpart comparably bright as the currently observed inner X-ray nebula (Evans et al. 2013; Lykou et al. 2023). We conservatively adopt a loose upper limit of the ejecta mass Mej < 3 M, as it cannot be heavier than twice the Chandrasekhar limit for a WD merger.

In this work, we search for the possible range of parameters that satisfy the observational constraint (Equations (26) and (32)). Note that we do not conduct any optimization methods, and therefore we aim to obtain the allowed range of parameters rather than a statistical constraint on them.

Figure 10 shows the allowed range of the parameters with respect to the SN ejecta mass Mej (lower horizontal axis), the SN ejecta kinetic energy Eej (upper horizontal axis), and the timing of the wind launch Δtw = 840 yr − tw [yr] (vertical axis) consistent with the X-ray observations by XMM-Newton and Chandra. We find that the size constraint of the inner nebula is tight; to confine the wind termination shock as observed, it only allows a very recent launch of the wind and a sufficiently large ejecta mass. On the other hand, the combination of the EMs of the inner and outer nebulae set an upper limit on the ejecta mass. The earlier Swift X-ray Telescope (XRT) detection is considered as a lower limit for Δtw in our models. In summary, we obtain the following constraints on the model parameters: 19

Equation (33)

Equation (34)

Equation (35)

and

Equation (36)

or

Equation (37)

Figure 10.

Figure 10. Constraints on the parameters of our dynamical model. The shaded region indicates the allowed range of the parameters with respect to the ejecta mass (lower horizontal axis) and the ejecta kinetic energy (upper horizontal axis) of SN 1181, and the timing of the launch of the currently observed wind (vertical axis). The relation between the upper and lower horizontal axis is based on Equation (22). The red dashed line indicates the timing of the XRT observation of IRAS 00500+6713.

Standard image High-resolution image

3.5. Consistency Check with Other Observations

The model parameters (Equations (33)–(37) are obtained by mainly fitting the angular radius and the EMs of the X-ray nebulae. We here test the validity of the fitted model by comparing it with other relevant observations of IRAS 00500+6713.

3.5.1. The Plasma Properties in the X-Ray Nebulae

In addition to measuring angular radius and EMs, our X-ray spectral analysis has determined relevant quantities of the X-ray-emitting plasma, such as ionization timescale and electron temperature (see Table 2). The ionization timescale is estimated as τne × Δt, where ne is the post-shock electron number density and Δt is the dynamical time of the plasma, and it provides an indicator of the degree of ionization.

In the allowed parameter space obtained by our analysis, the ionization timescale in the inner shocked region can be estimated as τinρsh,w/(μe mu) × Δtw ∼ 1010–11 s cm−3. Here, ρsh,w is the density of the inner shocked region obtained by ρsh,w = Msh,w/Vsh,w. The ionization timescale in the outer shocked region has a value of τout ≈ 4nISM × 840 yr ∼ 1010 s cm−3. Both timescales are broadly consistent with the values obtained from the X-ray spectroscopy. In order to achieve CIE in the nebulae with the estimated temperatures, τ ≳ 1011 s cm−3 is required for oxygen and ≳1012 s cm−3 for heavier elements (e.g., Smith & Hughes 2010). Hence, we infer that both the inner and outer nebulae are likely not in CIE. To estimate the electron temperature accurately, a detailed plasma calculation coupled with shock dynamics is required (e.g., Hamilton et al. 1983; Masai 1994; Tsuna et al. 2021a), which is beyond the scope of this paper. Still, if the shock downstream is adiabatic and the electrons and ions are far from CIE, which are good approximations, particularly for the shocked wind or the inner nebula, the electron temperature can be estimated as ${k}_{{\rm{B}}}{T}_{{\rm{e}},\mathrm{in}}\approx 511\,\mathrm{keV}\times {({v}_{{\rm{w}}}/c)}^{2}\sim 1.3\,\mathrm{keV}$, which is also consistent with X-ray spectroscopy. Here, c denotes the speed of light.

3.5.2. The IR Observations

IR observations of IRAS 00500+6713 revealed a dust-rich ring with an angular radius of θd ∼ 1' or a physical size of rd ∼ 0.66 pc. The ring is surrounded by a diffuse IR halo (Gvaramadze et al. 2019; Lykou et al. 2023), whose expansion velocity was assumed to be similar to that of the shocked ejecta vSii ∼ 1, 100 km s−1 based on the [S ii] line signatures observed for ${\theta }_{\mathrm{SII}}\sim 90^{\prime\prime} $ or physical size of rSii ∼ 0.99 pc (Ritter et al. 2021; Lykou et al. 2023; Fesen et al. 2023).

In our model, we expect that there are relatively low-temperature layers in the unshocked region, rsh < r < rout,r, where the reverse shock radius is calculated as rout,r = 1.04 pc. Note that the ratio between the forward shock radius and the reverse shock radius is rout,f/rout,r = 1.39, according to the self-similar solution with n = 6 and s = 0 (Chevalier 1982). Therefore, we have rsh < rd, rSII < rout,r, which is consistent with the observed sizes of the dust-rich ring and the diffuse IR halo. Since the inferred expansion velocity of the outer edge of the unshocked region is ∼rout,r/840 yr ∼ 1200 km s−1, which is comparable to vSii = 1100 km s−1, the region with the [S ii] line signature should be close to the reverse shock front. On the other hand, the outer edge of the dust-rich ring, which is at a slightly smaller radius, may correspond to the sublimation front of the dust that is irradiated by the radiation from the reverse shock. Although a dynamical calculation including the formation and destruction of dust is beyond the scope of this paper, it is important, particularly in light of the recently identified radially aligned filamentary structure around this region (Fesen et al. 2023).

4. Summary and Discussion

In this paper, we construct a dynamical model for IRAS 00500+6713 to determine the physical parameters of the system and consistently explain the multiwavelength data. We first analyze the archival X-ray data obtained by XMM-Newton and Chandra to extract information about the central (pointlike) component and diffuse components, such as their angular radius, EMs, ionization timescales, and electron temperatures. In particular, we confirm from the Chandra data that the central component has a finite angular radius of θin = 0.77–1farcs6 or a physical size of rsh = 0.0087–0.018 pc. Assuming that IRAS 00500+6713 is a remnant of SN 1181, we interpret that the central component originates from the shocked region between the carbon-burning ash-enriched wind from the central WD J005311 and the SN ejecta, while the diffuse component corresponds to the shocked region where the SN ejecta collide with the ISM. Based on this picture, we construct a dynamical evolution model for the inner and outer shocked regions and find that the X-ray properties of IRAS 00500+6713 can be reproduced by a case with an SN ejecta kinetic energy of Eej = (0.77–1.1) × 1048 erg, an SN ejecta mass of Mej = 0.18–0.53 M, if the currently observed intense wind started to blow a few decades ago, Δtw = 12–30 yr. In other words, the wind began blowing approximately after 1990 CE. We have confirmed that our model is also consistent with the IR geometry, including the dust-rich ring. In the following sections, we discuss the implications of these results for the progenitor system and the evolution of the remnant including the mechanism of the wind.

4.1. The Progenitor System

The SN parameters obtained by our dynamical model are broadly consistent with previous independent estimates: Oskinova et al. (2020) estimated the mass of the X-ray-emitting gas as ∼0.1 M from the XMM-Newton data, while Lykou et al. (2023) assessed the mass of ionized gas based on the tentative Hα detection, concluding it to be 0.15 ± 0.05 M, both of which are considered lower limits of the SN ejecta mass. In our model, this tentative Hα detection reported by Lykou et al. (2023) corresponds to our outer shocked ISM region. By assuming an expansion velocity of vSII ∼ 1100 km s−1, the ejecta kinetic energy was estimated to be a few ×1048 erg (Lykou et al. 2023). These SN parameter values, coupled with the ejecta's metal abundance inferred from the observed X-ray spectra, suggest that the explosion corresponds to a weak type of thermonuclear event from a degenerate system, consistent with type Iax SNe (e.g., Foley et al. 2013; Jha 2017). This suggestion is consistent with the comment by Fesen et al. (2023) that the nebula is metal rich and H poor. Our study, which newly incorporates the effects of the strong wind on the ejecta and nebula dynamics, further supports the assertion that IRAS 00500+6713 is the remnant of Type Iax SN 1181, which emerged approximately 840 yr ago.

Type Iax SNe are considered to be produced either from single (e.g., Livne et al. 2005; Kromer et al. 2013; Fink et al. 2014) or double (e.g., Kashyap et al. 2018) degenerate systems, although other possibilities, such as core-collapse systems (e.g., Moriya et al. 2010) and core-degenerate systems (e.g., Canals et al. 2018; Soker 2019), have also been proposed. Since a stellar companion of WD J005311 has not been identified, the double degenerate merger model is likely for SN 1181. From the observed wind properties, the mass of the remnant WD J005311 has been estimated to be M* ≳ 1.1 M (Gvaramadze et al. 2019; Kashiyama et al. 2019). In this case, the total mass of the progenitor binary, ∼M* + Mej, is comparable to or exceeds the Chandrasekhar limit. To produce a Type Iax SN from such a system, a binary consisting of a primary WD with M1 ≳ 1.1, M would be preferred (Shen 2015). This primary WD resides in the heavier end of the WD mass distribution (Kepler et al. 2007; Kilic et al. 2018) and is likely to originate from an intermediate-mass star with M* ≳ 5 M (e.g., Cummings et al. 2018; Temmink et al. 2020). Systematic studies of double degenerate merger simulations aiming to reproduce IRAS 00500+6713 are desirable, for which our model predictions, i.e., Eej, Mej, and the ejecta density profile, will be useful.

4.2. The Remnant Evolution

If IRAS 00500+6713 is the remnant of SN 1181, the age of WD J005311 is approximately 840 yr. Stellar evolution calculations have shown that the observed properties of WD J005311 and the surrounding IR halo can be consistent with a remnant of either CO + CO or ONe + CO binary merger occurring ∼1000–10,000 yr ago (Schwab et al. 2016; Yao et al. 2023; Wu et al. 2023). Note that, even in the case of CO + CO binaries, the remnant can become an ONe WD as a result of an off-center carbon burning.

An unsettled point is the origin and mechanism of the currently observed wind. We have shown that, in order to be consistent with the size constraint on the inner X-ray nebula obtained by Chandra, the wind started blowing only a few decades ago. Before that, the mass-loss rate and/or the wind velocity should have been significantly smaller than those today. Given the remnant age of ${ \mathcal O }({10}^{3})$ yr, the timing of the wind launch appears to be finely tuned. Carbon burning is currently ongoing around the surface of WD J005311 since the wind is enriched with the ashes. Considering the progenitor system shown in the previous section, the fuel of the wind can be unburned carbon accreted from the secondary CO WD at the merger, and the launch of the wind corresponds to the onset of shell burning of the fuel triggered by the Kelvin–Helmholtz contraction of the ONe core of WD J005311, as predicted by the evolution calculations of double degenerate merger remnants (Schwab et al. 2016). More dedicated studies on the timing of the wind launch will provide us with hints for understanding the mechanism of the wind and the fate of WD J005311, whether it will collapse into a neutron star or not.

4.3. Possible Caveats

We conclude by briefly discussing several possible caveats of our work. We have calculated the evolution of both the outer and inner shocks using a 1D model under the assumption of spherical symmetry. Though observations indicate that the overall shape of both regions is indeed spherical, filamentary structures are reported by Fesen et al. (2023). Therefore, it is necessary to include multidimensional effects in the calculation in future works. In addition, to reveal the origin of asphericity, radio observations with good spatial resolutions are required to see the shape of the termination shock of the wind in the future (Ko et al. 2024).

The mechanism of the fast wind is still unknown. We also suggest that there may be a relation between the fact that the wind started blowing recently and the fact that the optical brightness has been decreasing over the last 100 yr or so (Schaefer 2023). The wind mechanism and its relation to the luminosity decrease should be investigated in the future.

Acknowledgments

This research has made use of data and software provided by the High Energy Astrophysics Science Archive Research Center (HEASARC), which is a service of the Astrophysics Science Division at NASA/GSFC. This work was financially supported by the Japan Society for the Promotion of Science Grants-in-Aid for Scientific Research (KAKENHI) grant No. JP21J00031 (H.S.), JP20K04010 (K.K.), JP20H01904 (K.K.), JP22H00130 (KK), JP22H01265 (H.U.), JP19H01936 (T.T.), JP21H04493 (T.T.), JP20K14512 (K.F.), JP23H01211 (A.B.), JP24KJ0672 (K.T.), JP22K03688 (T.S.), JP22K03671 (T.S.), and JP20H05639 (T.S.). D.T. is supported by the Sherman Fairchild Postdoctoral Fellowship at the California Institute of Technology. T.K. is supported by the RIKEN Junior Research Associate Program.

Facilities: CXO (ACIS) - Chandra X-ray Observatory satellite, XMM (MOS, pn, RGS) - Newton X-Ray Multimirror Mission satellite

Software: HEASoft (v6.20; HEASARC 2014); CIAO (v4.15; Fruscione et al. 2006); SAS (v19.1.0; Gabriel et al. 2004).

Appendix: Emission from the Inner Shocked Ejecta

In this appendix, we show that the electron temperature in the shocked ejecta is kept below ≲3–4 eV, and thus, although the mass is significantly larger than the shocked wind, the shocked ejecta do not contribute to the EM of the inner X-ray nebula.

At the immediate shock downstream, electrons and ions are first independently heated by collision to the following temperatures, respectively,

Equation (A1)

Equation (A2)

For t ∼ 840 yr, these temperatures are determined from the observed size of the inner nebula, and kB Te,down ∼ 1 eV and kB Tion,down ∼ 20 keV, where we adopt vsh = 1100 km s−1 and rsh = 0.018 pc for simplicity. Then, the electrons experience collisional heating by the ions and radiative cooling through recombinations of ions and bremsstrahlung radiation. The corresponding heating and cooling timescales can be estimated as

Equation (A3)

Equation (A4)

respectively. Here, Equation (A3) gives a timescale for the temperature equilibrium between electrons and ions (e.g., Spitzer 1956), with ne being the electron number density in cgs units, and Tion,down and Te,down being the ion and electron temperature in kelvin. We assume that the shocked-SN ejecta are dominated by oxygen (Aion = 16 and Ae = 1/1840). A singly ionized state is inferred from the electron temperature at the immediate downstream; thus, we set ne = 4ρej(rsh)/16mu, Zion = 1 and Ze = 1, while for the cooling timescale (Equation (A4)), we employ the cooling function Λ(Te) for the oxygen-dominated gas given in Kirchschlager et al. (2019) (Figure 11).

Figure 11.

Figure 11. Cooling function of an oxygen-dominated gas from Kirchschlager et al. (2019).

Standard image High-resolution image

The solid lines in Figure 12 show the evolution of the heating and cooling timescales of electrons at the immediate shock downstream calculated for a parameter set $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})=(1\times {10}^{-6}\,{M}_{\odot }\,{\mathrm{yr}}^{-1},820\,\mathrm{yr},0.3\,{M}_{\odot })$. It can be observed that teq(Te,down) < tcool(Te,down) at t ∼ 840 yr. This implies that electrons at the immediate shock downstream with a temperature Te,down will gradually approach equilibrium with ions, meaning that the electron temperature begins to increase due to collisional heating with ions.

Figure 12.

Figure 12. The time evolution of the heating (red) and cooling (blue) timescales of the inner shocked ejecta, calculated for a parameter set $({\dot{M}}_{{\rm{w}}},{t}_{{\rm{w}}},{M}_{\mathrm{ej}})=(1\times {10}^{-6}\,{M}_{\odot }\,{\mathrm{yr}}^{-1},820\,\mathrm{yr},0.3\,{M}_{\odot })$ using our dynamical model. The solid lines are for the immediate shock downstream with an electron temperature of Te = Te,down and the dotted lines are estimated for a peak temperature of the cooling function kB Te = kB Te,pk = 3.7 eV.

Standard image High-resolution image

On the other hand, the dotted lines in Figure 12 are the cooling and heating timescales of electrons at Te = Te,pk. We calculated these values in the same way as the Equations (A3) and (A4). We find that the electron temperature is not likely to get into the X-ray regime, i.e., kB Te,down ≪ 0.1 keV; the radiative cooling becomes relevant as the cooling function Λ(Te) steeply increases with electron temperature for kB Te,down ∼ 1 eV < kB Te < 3.7 eV. It is shown that teq(Te,pk) > tcool(Te,pk) is always satisfied, which means that the collisional heating of the electron saturates before reaching kB Te,pk = 3.7 eV. After the electron temperature becomes saturated at Te,down < Te < Te,pk, the ion temperature starts to decrease from Tion,down via the radiative cooling. Since these heating and cooling timescales of the plasma are much shorter than the dynamical timescale of the shock ≈rsh/vsh ≈ Δtw = a few 10 yr, the shock downstream is radiative, i.e., the internal energy produced by the shock is predominantly lost via thermal radiation with a temperature of Te,down < T < Te,pk.

Although a radiative transfer calculation including the photo- and collisional ionization of the downstream plasma is required to accurately calculate the electron temperature, it can be roughly estimated in the radiative shock limit from the following condition:

Equation (A5)

where

Equation (A6)

is the EM of the shocked ejecta with

Equation (A7)

being its volume. The left and right-hand sides of Equation (A5) represent the emission luminosity of and the energy injection rate into the shocked ejecta.

Figure 13 shows the electron temperature obtained from Equation (A5) for several sets of model parameters: To estimate the electron temperature Te from Equations (A5)–(A7), we need to begin by solving the shock dynamics using the model outlined in Section 3, which is characterized by model parameters Mej, ${\dot{M}}_{{\rm{w}}}$, and tw. As expected from the timescale arguments above, the emission temperature is in the range of ≲3–4 eV, suggesting that the inner shocked-SN ejecta emit mainly UV photons, not X-rays. These UV photons are expected to be absorbed by the surrounding dust-rich ring and reemitted in the IR bands.

Figure 13.

Figure 13. The time evolution of the electron temperature of the inner shocked-SN ejecta obtained by the radiative shock condition (Equation (A5)).

Standard image High-resolution image

Footnotes

Please wait… references are loading.
10.3847/1538-4357/ad4d99
  翻译: