A Sparkler in the Fireworks Galaxy: Discovery of an Ultraluminous X-Ray Transient with a Strong Oxygen Line in NGC 6946

, , and

Published 2019 September 19 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Chen Wang et al 2019 ApJ 883 44 DOI 10.3847/1538-4357/ab3c4d

Download Article PDF
DownloadArticle ePub

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

0004-637X/883/1/44

Abstract

We discovered and studied an ultraluminous X-ray source (CXOU J203451.1+601043) that appeared in the spiral galaxy NGC 6946 at some point between 2008 February and 2012 May and has remained at luminosities ≈2–4 × 1039 erg s−1 in all observations since then. Our spectral modeling shows that the source is generally soft but with spectral variability from epoch to epoch. Using standard empirical categories of the ultraluminous regimes, we found that CXOU J203451.1+601043 was consistent with a broadened disk state in 2012 but was in a transitional state approaching the supersoft regime in 2016, with substantial down-scattering of the hard photons (similar, for example, to the ultraluminous X-ray source in NGC 55). It has since hardened again in 2018–2019 without any significant luminosity change. The most outstanding property of CXOU J203451.1+601043 is a strong emission line at an energy of of (0.66 ± 0.01) keV, with an equivalent width of ≈100 eV and de-absorbed line luminosity of ≈2 × 1038 erg s−1, seen when the continuum spectrum was softest. We identify the line as O viii Lyα (rest-frame energy of 0.654 keV); we interpret it as a strong indicator of a massive outflow. Our finding supports the connection between two independent observational signatures of the wind in super-Eddington sources: a lower temperature of the Comptonized component and the presence of emission lines in the soft X-ray band. We speculate that the donor star is oxygen-rich: a CO or O–Ne–Mg white dwarf in an ultracompact binary. If that is the case, the transient behavior of CXOU J203451.1+601043 raises intriguing theoretical questions.

Export citation and abstract BibTeX RIS

1. Introduction

Off-nuclear, pointlike sources with X-ray luminosities >1039 erg s−1 are generally interpreted as the high-luminosity tail of stellar-mass X-ray binaries (Feng & Soria 2011; Kaaret et al. 2017). As such, they must be emitting at or above their Eddington limit, which is ≈2 × 1038 erg s−1 for an accreting neutron star (NS) or ≈1.3 × 1039 erg s−1 for a 10 M black hole (BH); this justifies the name "ultraluminous X-ray sources" (ULXs) given to that subpopulation. The identification of ULXs as stellar-mass X-ray binaries (most likely with a massive donor star) is based both on studies of individual sources, in rare cases where the compact object can be identified as an NS (Bachetti et al. 2014; Motch et al. 2014; Fürst et al. 2016; Israel et al. 2017a, 2017b; Carpano et al. 2018; Rodríguez Castillo et al. 2019; Sathyaprakash et al. 2019), and on their statistical population properties, which are consistent with an extension of the high-mass X-ray binary luminosity function (Swartz et al. 2011; Mineo et al. 2012; Lehmer et al. 2019). This is also the reason why the majority of ULXs are found in star-forming galaxies rather than old ellipticals (Swartz et al. 2004; Earnshaw et al. 2019b). It is still possible that a very small fraction of ULXs (Wiersema et al. 2010; Sutton et al. 2012; Earnshaw et al. 2019b), especially those located in dwarf galaxies (Moran et al. 2014; Mezcua et al. 2018), are sub-Eddington intermediate-mass BHs, but we are not concerned with that population here.

Several physical properties of ULXs either remain unexplained or are consistent with more than one scenario. In this paper, we focus on two unanswered questions. The first question is about the properties of transient ULXs. A census of transient ULXs is largely incomplete because of the relatively short history of X-ray observations and sparse monitoring cadence of individual galaxies with ULXs. For sub-Eddington X-ray binaries (Remillard & McClintock 2006), the rule of thumb is that systems with low-mass donors accreting via Roche lobe overflow are generally transient, while wind-accreting systems with a high-mass donor are generally persistent X-ray sources. For ULXs, we still cannot confidently predict which systems may exhibit a thermal-viscous instability (invoked to explain the outburst cycle in sub-Eddington sources) or what other mechanisms (e.g., accretor/propeller transitions or precession of the polar funnel in and out of our line of sight) may cause an apparently transient behavior. Transient ULXs have been observed in early-type galaxies (M86: van Haaften et al. 2019; NGC 5128: Burke et al. 2013) but more often in spiral galaxies (M31: Middleton et al. 2012; M83: Soria et al. 2012; M101: Kuntz et al. 2005; NGC 5907: Walton et al. 2015; Pintore et al. 2018; the pulsar ULX in NGC 300: Carpano et al. 2018) and even starburst galaxies (NGC 3628: Strickland et al. 2001; the pulsar ULX in M82: Feng & Kaaret 2007). We do not know yet whether there are particular types of donor stars or age ranges that are more likely to be associated with transient ULXs. This is, in turn, part of the more general identification problem of the donor star in any ULX. Even in systems where a pointlike UV/optical/IR counterpart is clearly identified, it is often difficult to distinguish the contributions from the donor star, irradiated accretion disk, disk outflows, and perhaps circumbinary material (Tao et al. 2011; Gladstone et al. 2013; Heida et al. 2014; Fabrika et al. 2015; López et al. 2017; Lau et al. 2019).

The second issue we discuss in this paper is the spectral evolution of individual ULXs. Individual and population studies have shown clear evidence of X-ray spectral and time variability differences between softer and harder ULXs (Sutton et al. 2013). There is also solid theoretical (Poutanen et al. 2007; Ohsuga & Mineshige 2011; Kawashima et al. 2012; Narayan et al. 2017) and observational (Pinto et al. 2016, 2017; Walton et al. 2016; Kosec et al. 2018) evidence that super-Eddington sources launch massive, radiation-driven winds. These winds are expected to have a lower-density funnel along the polar direction and a higher optical depth at higher viewing angles (closer to the disk plane). The wind down-scatters the harder photons emitted from the innermost part of the flow, introducing a characteristic spectral curvature and high-energy downturn and enhancing short-term X-ray variability. Based on those findings, it is now commonly accepted that softer and harder ULX spectral shapes correspond to sources seen through a thicker or thinner wind, respectively (Middleton et al. 2015a, 2015b). However, it is not clear whether those differences can be explained mostly as an effect of our viewing angle (a similar scenario to the so-called active galactic nucleus unification model) or instead correspond to intrinsically different physical regimes with different wind properties. To reduce this ambiguity, more discoveries and studies of spectral state changes in individual sources would be very useful (in parallel to population studies of spectral differences between different sources). To date, spectral state transitions have been well studied (and related to wind properties) only in a handful of ULXs, for example, NGC 247 X-1 (Feng et al. 2016), NGC 55 X-1 (Pinto et al. 2017), IC 342 X-1 (Shidatsu et al. 2017), and a few other examples discussed in Pintore et al. (2017) and Weng & Feng (2018).

Given the variety of alternative scenarios for spectral evolution in ULXs, and the limited number of X-ray state transitions identified in individual systems so far, any new identifications of such behavior can help constrain the models. In this paper, we report on our discovery of transient behavior, spectral evolution, and wind signatures in a ULX (CXOU J203451.1+601043) in NGC 6946 ("the fireworks galaxy"); four other ULXs have already been identified in this galaxy (Earnshaw et al. 2019a).

Between 2001 and 2008, CXOU J203451.1+601043 was undetected in multiple Chandra, XMM-Newton, and Swift observations; it was first detected by Chandra in 2012 and has subsequently remained in the ultraluminous state. We monitored its behavior using the long series of X-ray observations that have covered the highly star-forming host galaxy NGC 6946. We searched for an optical counterpart using archival Hubble Space Telescope (HST) images. For the host galaxy, we assumed the recently determined distance of 7.7 ± 0.3 Mpc (Anand et al. 2018; Eldridge & Xiao 2019), higher than the distances of ≈5.5 Mpc (Tully 1988) or ≈5.9 Mpc (Karachentsev et al. 2000) commonly used in the literature before 2018; for example, the new distance implies an increase in the intrinsic source luminosities by a factor of ≈1.7 when we compare our results with the Chandra study of Fridriksson et al. (2008).

The paper is organized as follows. Observations and data analysis are described in Section 2. In Section 3, we present the X-ray spectral and time variability properties and constrain a possible optical counterpart. In Section 4, we discuss how this system fits into our current knowledge of ULXs, in particular those with evidence of a strong wind, and we propose a white dwarf scenario for the donor star. Finally, prospects for follow-up studies are summarized in Section 5.

2. Observations and Data Analysis

2.1. Chandra

Chandra observed NGC 6946 nine times between 2001 and 2017 (in most cases, to follow supernova explosions, which occur quite frequently in this galaxy). The observation log is shown in Table 1. All observations were made with the S3 chip of the Advanced CCD Imaging Spectrometer (ACIS) array, except for the ACIS observation of 2012 (ObsID 13435), in which the target was placed on the S2 chip. After downloading the data from the public archive, we reprocessed and analyzed them with the Chandra Interactive Analysis of Observations (ciao) software version 4.10 (Fruscione et al. 2006) with calibration database version 4.7.9. Specifically, we rebuilt level 2 event files with the task chandra_repro and filtered out background flares with the task deflare. We created images in multiple energy bands for each epoch with dmcopy. We searched for point sources in each epoch with with wavdetect. Our target was undetected at every epoch until its first appearance in 2012 May. It was also later detected in the subsequent Chandra observations of 2016 September and 2017 June.

Table 1.  Chandra, XMM-Newton, and Swift Observations of NGC 6946 and Luminosity of CXOU J203451.1+601043

ObsID Good Time Interval Observation Date Observed X-Ray Fluxa X-Ray Luminositya
  (ks)   (erg cm−2 s−1) (erg s−1)
Chandra/ACIS
1043 58.3 2001 Sep 7 <1.2 × 10−15 <2.1 × 1037
4404 28.7 2002 Nov 25 <1.4 × 10−15 <2.5 × 1037
4631 28.4 2004 Oct 22 <1.1 × 10−15 <2.0 × 1037
4632 25.2 2004 Nov 6    
4633 26.6 2004 Dec 3    
13435 20.4 2012 May 21 ${4.0}_{-0.6}^{+0.6}\times {10}^{-14}$ ${1.1}_{-0.2}^{+0.2}\times {10}^{39}$
17878 40.0 2016 Sep 28 ${2.7}_{-0.2}^{+0.2}\times {10}^{-13}$ ${4.8}_{-0.4}^{+0.4}\times {10}^{39}$
19887 18.5 2016 Sep 28    
19040 9.8 2017 Jun 11 ${0.6}_{-0.1}^{+0.1}\times {10}^{-14}$ ${1.7}_{-0.3}^{+0.3}\times {10}^{39}$
XMM-Newton/EPIC
0093641501 0.6 2003 Apr 18 $\lt 1\times {10}^{-14}$ $\lt 2\times {10}^{38}$
0093641601 2.2 2003 May 17    
0093641701 1.2 2003 Jun 18    
0200670101 3.9 2004 Jun 9 <3.1 × 10−15 <3 × 1037
0200670201 12.7 2004 Jun 11    
0200670301 11.3 2004 Jun 13    
0200670401 8.8 2004 Jun 25    
0401360101 18.7 2006 May 23 <2 × 10−15 <1.5 × 1037
0401360201 4.7 2006 Jun 2    
0401360301 4.9 2006 Jun 18    
0500730101 26.0 2007 Nov 8 <1.2 × 10−15 <2 × 1037
0500730201 31.7 2007 Nov 2    
0691570101 109.3 2012 Oct 21 ${1.6}_{-0.1}^{+0.1}\times {10}^{-13}$ ${1.6}_{-0.1}^{+0.1}\times {10}^{39}$
0794581201 43.1 2017 Jun 1 ${0.6}_{-0.1}^{+0.1}\times {10}^{-13}$ ${2.8}_{-0.1}^{+2.0}\times {10}^{39}$
Swift/XRT
31113001–31113004 10 2008 Feb 4–2008 Feb 14 $\lt 2\times {10}^{-14}$ < 4 × 1038
49820001–49820003 7 2013 May 31–2013 Jun 4 ${1.4}_{-0.4}^{+0.4}\times {10}^{-13}$ ${2.4}_{-0.5}^{+0.5}\times {10}^{39}$
10130001–10130029 44 2017 May 13–2017 Sep 17 ${1.1}_{-0.2}^{+0.2}\times {10}^{-13}$ ${3.6}_{-0.4}^{+0.4}\times {10}^{39}$
94059001–94059044 43 2018 Apr 1–2018 Dec 27 ${2.2}_{-0.3}^{+0.3}\times {10}^{-13}$ ${2.8}_{-0.4}^{+0.4}\times {10}^{39}$
94059045–94059072 17 2019 Jan 6–2019 Apr 6 ${1.7}_{-0.3}^{+0.3}\times {10}^{-13}$ ${3.2}_{-0.4}^{+0.4}\times {10}^{39}$

Note.

aFor the Chandra observations, we estimated observed fluxes and intrinsic luminosities (or their respective upper limits) in the 0.3–10 keV band with the ciao task srcflux, assuming a power-law model with photon index ${\rm{\Gamma }}=2.5$ and total column density ${N}_{{\rm{H}}}=4\times {10}^{21}$ cm−2 (twice the Galactic line-of-sight value). This model was chosen because it approximates the best-fitting power-law model for the 2016 Chandra spectrum (Table 2). For XMM-Newton observations, we also used a simple power-law model to convert from count rates to fluxes and luminosities in this table. When we had enough counts for a detailed fit (2012 and 2017 XMM-Newton observations), we used the best-fitting values (Table 2); for the nondetections, we assumed Γ = 2.5 and NH = 4 × 1021 cm−2. For Swift, we fixed NH = 4 × 1021 cm−2, then used pimms to estimate the power-law photon index that best approximates the observed (1.5–10)/(0.3–1.5) hardness ratio of each observation. We used those indices to infer fluxes and luminosities of the respective observations; when not enough counts were available, we again assumed Γ = 2.5. See Table 2 for a comparison of flux and luminosity conversions using power laws versus more complex spectral models.

Download table as:  ASCIITypeset image

In order to check whether the source is spatially extended, we computed the point-spread functions (PSFs) at the off-axis locations of the source in 2012 and 2016 using the Chandra Ray Tracer (ChaRT5 ) and simulated PSF files with the marx software.6 We compared the profiles of the source with those of the respective simulated PSFs with the ciao tool srcextent. We confirm that the transient source is consistent with being pointlike. This is further confirmed by the 2017 observation (ObsID 19040), the only one in which the target is almost on-axis: its observed FWHM of ≈2'' is consistent with the expected width of an on-axis PSF.

For each observation (both those with a detection and those with no detection), we extracted the source events from a circular aperture of either 5'' or 2'' radius (for ObsID 19040) to match the size of the PSF at the respective locations. Background events were extracted from nearby source-free regions at least three times the size of the source region. We used the ciao task scrflux to estimate an upper limit to the source flux in the epochs when it was not detected and an approximate flux in the two epochs (2012 and 2017) when it was detected but with a small number of counts. For the 2016 data set, we combined the spectra from the two exposures taken on 2016 September 28 and obtained enough counts for meaningful spectral fitting; we extracted source and background spectra with specextract, which also generates the appropriate auxiliary response files (ARFs) and response matrix files (RMFs) for the subsequent spectral analysis. Spectra were grouped to a minimum of 15 counts bin–1 for χ2 fitting. We also repeated our spectral analysis on the same spectra grouped to 1 count bin–1 using the Cash statistics (Cash 1979). Moreover, to check for short-term variability in each observation in which the source was detected, we extracted background-subtracted light curves using the ciao tool dmextract.

We did subsequent data analysis with NASA's High Energy Astrophysics Software (HEASOFT): ds9 (Joye & Mandel 2003) version 8.0 for imaging and photometry, ftools/Xronos (Blackburn 1995) version 6.25 for timing analysis, and xspec (Arnaud 1996) version 12.9.1 for spectral modeling. The reported errors are 90% confidence intervals for the fitting parameters. For the 2016 Chandra spectra, all of the best-fitting parameters obtained from χ2 and Cash statistics fitting (cstat in xspec) agree well within their error ranges; thus, in Section 3.3 and Table 2, we report only the values from χ2 fitting for simplicity.

Table 2.  Best-fitting Parameters of the EPIC Spectra from 2012 and 2017 and the ACIS Spectrum from 2016

Model Parameters Values
  2012 2016 2017
tbabs × tbabs × (po + Gaussian)
NH,Gal (1022 cm−2) [0.20] [0.20] [0.20]
NH,int (1022 cm−2) ${0.14}_{-0.03}^{+0.03}$ ${0.20}_{-0.12}^{+0.13}$ ${0.23}_{-0.06}^{+0.07}$
Γ (keV) ${2.17}_{-0.06}^{+0.07}$ ${2.63}_{-0.20}^{+0.22}$ ${3.50}_{-0.26}^{+0.31}$
${N}_{\mathrm{po}}$ (10−5 ph keV−1 cm−2 s−1 at 1 keV) ${5.8}_{-0.4}^{+0.4}$ ${14.2}_{-2.9}^{+3.9}$ ${6.3}_{-1.1}^{+1.4}$
${E}_{\mathrm{line}}$ (keV) ${0.61}_{-0.02}^{+0.02}$ ${0.66}_{-0.01}^{+0.01}$
${\sigma }_{\mathrm{line}}$ (keV) [0] $\lt 0.030$
${N}_{\mathrm{line}}$ (10−5 ph cm−2 s−1) ${1.2}_{-0.7}^{+1.2}$ ${2.7}_{-1.4}^{+2.3}$
${\chi }^{2}/$dof 346.8/230 (1.51) 68.0/61 (1.12) 89.3/64 (1.40)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)a ${1.68}_{-0.06}^{+0.06}$ ${2.49}_{-0.20}^{+0.22}$ ${0.62}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)b ${2.23}_{-0.14}^{+0.16}$ ${4.86}_{-0.97}^{+1.50}$ ${3.10}_{-0.84}^{+1.45}$
tbabs × tbabs × (bbodyrad + diskbb + Gaussian)
${N}_{{\rm{H}},\mathrm{Gal}}$ (1022 cm−2) [0.20] [0.20] [0.20]
${N}_{{\rm{H}},\mathrm{int}}$ (1022 cm−2) ${0.08}_{-0.08}^{+0.12}$ $\lt 0.15$ ${0.26}_{-0.18}^{+0.23}$
${{kT}}_{\mathrm{bb}}$ (keV) ${0.15}_{-0.02}^{+0.03}$ ${0.27}_{-0.09}^{+0.08}$ ${0.15}_{-0.02}^{+0.03}$
${N}_{\mathrm{bb}}$ (km2)c ${11.3}_{-9.2}^{+41.5}$ ${1.7}_{-1.0}^{+8.7}$ ${89}_{-79}^{+908}$
${{kT}}_{\mathrm{in}}$ (keV) ${1.25}_{-0.07}^{+0.07}$ ${1.14}_{-0.20}^{+0.40}$ ${0.72}_{-0.11}^{+0.14}$
${N}_{\mathrm{dbb}}$ (10−3 km2)d ${3.7}_{-0.9}^{+1.1}$ ${6.3}_{-4.8}^{+10.0}$ ${13.3}_{-8.1}^{+19.7}$
${E}_{\mathrm{line}}$ (keV) ${0.61}_{-0.04}^{+0.04}$ ${0.66}_{-0.02}^{+0.02}$
${\sigma }_{\mathrm{line}}$ (keV) [0] $\lt 0.033$
${N}_{\mathrm{line}}$ (10−5 ph cm−2 s−1) ${0.48}_{-0.26}^{+1.44}$ ${3.2}_{-2.1}^{+13.5}$
${\chi }^{2}/$dof $235.5/228$ (1.03) $68.3/59$ (1.16) $72.5/62$ (1.17)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)a ${1.57}_{-0.05}^{+0.05}$ ${2.28}_{-0.10}^{+0.11}$ ${0.61}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)b ${1.75}_{-0.30}^{+0.76}$ ${2.22}_{-0.17}^{+0.76}$ ${2.52}_{-1.44}^{+6.62}$
tbabs × tbabs × (bbodyrad + comptt + Gaussian)
${N}_{{\rm{H}},\mathrm{Gal}}$ (1022 cm−2) [0.20] [0.20] [0.20]
${N}_{{\rm{H}},\mathrm{int}}$ (1022 cm−2) ${0.05}_{-0.01}^{+0.01}$ ${0.23}_{-0.10}^{+0.33}$ ${0.28}_{-0.02}^{+0.03}$
${{kT}}_{\mathrm{bb}}$ (keV) ${0.16}_{-0.01}^{+0.01}$ ${0.07}_{-0.07}^{+0.06}$ ${0.13}_{-0.01}^{+0.06}$
${N}_{\mathrm{bb}}$ (km2)c ${7.3}_{-1.0}^{+13.7}$ (unconstrained) ${117}_{-12}^{+258}$
kT0 (keV)=${{kT}}_{\mathrm{bb}}$ $[{0.16}_{-0.01}^{+0.01}]$ $[{0.07}_{-0.07}^{+0.06}]$ $[{0.13}_{-0.01}^{+0.06}]$
kTe (keV) ${1.01}_{-0.10}^{+0.15}$ $\gt 0.98$ ${0.65}_{* }^{+0.89}$
τ ${13.7}_{-0.3}^{+0.4}$ ${5.0}_{-5.0}^{+3.4}$ ${13.6}_{-0.8}^{+7.3}$
${N}_{{\rm{c}}}$ (10−5) ${9.7}_{-0.2}^{+0.2}$ ${63}_{* }^{+606}$ ${11.9}_{-1.1}^{+2.4}$
${E}_{\mathrm{line}}$ (keV) ${0.61}_{-0.04}^{+0.03}$ −– ${0.66}_{-0.02}^{+0.01}$
${\sigma }_{\mathrm{line}}$ (keV) [0] $\lt 0.025$
${N}_{\mathrm{line}}$ (10−5 ph cm−2 s−1) ${0.36}_{-0.27}^{+0.27}$ ${3.7}_{-1.4}^{+1.4}$
${\chi }^{2}/$dof $234.6/227$ (1.03) $64.1/58$ (1.11) $72.6/61$ (1.19)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)a ${1.56}_{-0.05}^{+0.05}$ ${2.59}_{-0.32}^{+1.85}$ ${0.61}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)b ${1.63}_{-0.04}^{+0.05}$ ($\gtrsim 3.8$)e ${2.83}_{-0.08}^{+2.03}$
tbabs × tbabs × (diskpbb + Gaussian)
NH,Gal (1022 cm−2) [0.20] [0.20] [0.20]
NH,int (1022 cm−2) < 0.02 ${0.03}_{-0.03}^{+0.06}$ ${0.03}_{-0.02}^{+0.18}$
kTin (keV) ${1.49}_{-0.11}^{+0.13}$ ${1.40}_{-0.31}^{+0.29}$ ${0.68}_{-0.09}^{+0.11}$
p ${0.60}_{-0.02}^{+0.01}$ ${0.50}_{* }^{+0.04}$ ${0.50}_{* }^{+0.04}$
Ndpbb (10−3 km2)d ${0.92}_{-0.31}^{+0.44}$ ${0.89}_{-0.53}^{+2.17}$ ${6.7}_{-3.4}^{+8.1}$
${E}_{\mathrm{line}}$ (keV) ${0.61}_{-0.04}^{+0.04}$ ${0.67}_{-0.02}^{+0.01}$
σline (keV) [0] $\lt 0.029$
Nline (10−5 ph cm−2 s−1) ${0.28}_{-0.19}^{+0.23}$ ${0.80}_{-0.36}^{+0.45}$
χ2/dof 252.7 /229 (1.10) 65.9/60 (1.10) 102.5/63 (1.63)
f0.3−10 (10−13erg cm−2 s−1)a ${1.58}_{-0.05}^{+0.06}$ ${2.41}_{-0.20}^{+0.19}$ ${0.59}_{-0.04}^{+0.04}$
L0.3−10 (1039 erg cm−2 s−1)b ${1.50}_{-0.05}^{+0.06}$ ${1.71}_{-0.14}^{+0.13}$ ${0.93}_{-0.06}^{+0.06}$

Notes.

aObserved fluxes in the 0.3–10 keV band. bFor all spectral models, the de-absorbed luminosities L0.3−10 (0.3–10 keV band) were defined as 4πd2 times the de-absorbed fluxes. cNbb = (Rbb/D10)2, where Rbb is the source radius in km and D10 is the distance to the source in units of 10 kpc (here D10 = 770). dNdbb = (Rin/D10)2 cos θ, where Rin is the apparent inner disk radius in km, D10 is the distance to the source in units of 10 kpc, and θ is our viewing angle (θ = 0 is face-on). Ndpbb is defined exactly as Ndbb. eUpper limit unconstrained because of the degeneracy between absorption column density and seed blackbody luminosity.

Download table as:  ASCIITypeset images: 1 2

2.2. XMM-Newton

There are 14 XMM-Newton observations of NGC 6946 with the European Photon Imaging Camera (EPIC) in the full-frame mode. We downloaded the data from the XMM-Newton Science Archive and reduced them with the Science Analysis Software (sas) version 17.0. The EPIC-pn and EPIC-MOS events were processed using the XMM-Newton pipeline and associated calibration files. We removed time intervals of high background from the event files with evselect using a "RATE ≤ 0.4" threshold for the pn and "RATE ≤ 0.35" for MOS1 and MOS2.

Between 2003 and 2007, CXOU J203451.1+601043 was not detected in any of the observations. We grouped the nondetection observations by year to increase the signal-to-noise ratio. For each year, we built a stacked pn and MOS image and measured the total counts inside a circle with 20'' radius at the position of the source and the background counts from surrounding regions. We then applied the Bayesian method of Kraft et al. (1991) for Poisson-distributed counts to obtain the 90% upper limit to the net counts and count rates.7 For the conversion from count rates to fluxes, we used the EPIC exposure maps to account for vignetting, and we assumed that for an equal effective exposure time between the three instruments, the pn contributes to about 61% of the count rate and MOS1+MOS2 to about 39%. We obtained this relative ratio between the EPIC instruments with the Portable, Interactive Multi-Mission Simulator (pimms) version 4.9 for a range of plausible spectral models; the relative contribution of pn and MOS changes at most by 2% or 3% from model to model, which is negligible for the purpose of our analysis. We converted the upper limits on the pn+MOS count rates to flux upper limits with pimms using a power-law model with photon index Γ = 2.5 and total absorbing column density NH = 4 × 1021 cm−2 (Table 1).

In contrast, CXOU J203451.1+601043 was detected in 2012 and 2017 with luminosities exceeding 1039 ergs s−1. For the 2012 and 2017 detections, source events were extracted from a circular aperture of 20'' radius. We generated background-subtracted light curves binned to 0.1 s with the sas task evselect followed by epiclccorr. Background events were extracted from point source–free, circular regions on the same chip approximately three times larger than the source region. We created spectral files and associated instrumental responses with the script multiespecget, which extracts the spectra of all three detectors and combines them into a single EPIC spectrum. We chose to combine the three EPIC spectra to increase the signal-to-noise ratio of possible line features in the soft X-ray band. For spectral extraction, we used the standard filtering conditions "(FLAG==0) && (PATTERN ≤ 4)" for the pn and "(#XMMEA_EM && (PATTERN ≤ 12)" for MOSs. Spectra were grouped to a minimum of 25 counts bin–1 for χ2 fitting.

As for the Chandra data, subsequent imaging, timing, and spectral analysis were carried out with HEASOFT packages (ds9, ftools, and xspec) and astropy (Astropy Collaboration et al. 2018) specifically for period searches (Section 3.2).

2.3. Swift and NuSTAR

The X-ray Telescope (XRT) on board Swift has monitored NGC 6946 frequently over the years, but typical exposure times are very short (∼1 ks). For this work, we used all of the XRT observations from 2008 onward, stacked into data sets for individual years (Table 3). The 2008 observations are the last ones from any X-ray observatory in which the transient ULX is not detected; it is detected in all subsequent Swift observations. We used standard HEASOFT packages (version 6.25) for spectral extraction; we created a combined spectrum and exposure map with xselect and an ancillary response function with xrtmkarf, and the ready-made response file comes from the XRT Calibration Database.8

Table 3.  Alternative Set of Models for the 2017 EPIC Spectra

Model Parameters Values
tbabs × tbabs × (bbodyrad + diskbb + vapec)
NH,Gal (1022 cm−2) [0.20]
NH,int (1022 cm−2) ${0.13}_{-0.13}^{+0.19}$
${{kT}}_{\mathrm{bb}}$ (keV) ${0.14}_{-0.04}^{+0.06}$
${N}_{\mathrm{bb}}$ (km2)a ${22}_{-21}^{+358}$
${{kT}}_{\mathrm{in}}$ (keV) ${0.72}_{-0.12}^{+0.14}$
${N}_{\mathrm{dbb}}$ (10−3 km2)b ${7.5}_{-4.3}^{+9.5}$
kTvapec (keV) ${0.41}_{-0.10}^{+0.09}$
O abundance (solar units) >10
Ne = Mg = Si abundances (solar units) >10
Other abundances (solar units) [1]
Nvapec (10−6)c ${0.07}_{-0.03}^{+4.3}$
χ2/dof $68.6/61$ (1.13)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)d ${0.62}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)e ${1.32}_{-0.57}^{+2.51}$
tbabs × tbabs × (bbodyrad + comptt + vapec)
${N}_{{\rm{H}},\mathrm{Gal}}$ (1022 cm−2) [0.20]
${N}_{{\rm{H}},\mathrm{int}}$ (1022 cm−2) ${0.12}_{-0.02}^{+0.16}$
${{kT}}_{\mathrm{bb}}$ (keV) ${0.15}_{-0.02}^{+0.01}$
${N}_{\mathrm{bb}}$ (km2)c ${17.5}_{-2.8}^{+39.9}$
kT0 (keV)=${{kT}}_{\mathrm{bb}}$ $[{0.15}_{-0.02}^{+0.01}]$
kTe (keV) ${0.59}_{* }^{+0.46}$
τ ${26.4}_{-3.1}^{+4.1}$
${N}_{{\rm{c}}}$ (10−5) ${4.3}_{-0.5}^{+0.7}$
${{kT}}_{\mathrm{vapec}}$ (keV) ${0.43}_{-0.04}^{+0.05}$
O abundance (solar units) ${385}_{-140}^{+155}$
Ne = Mg = Si abundances (solar units) ${230}_{-95}^{+110}$
Other abundances (solar units) [1]
${N}_{\mathrm{vapec}}$ (10−6)c ${0.18}_{-0.05}^{+0.65}$
${\chi }^{2}/$dof $68.4/60$ (1.14)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)d ${0.61}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)e ${1.23}_{-0.07}^{+2.71}$
tbabs × tbabs × (diskpbb + vapec)
${N}_{{\rm{H}},\mathrm{Gal}}$ (1022 cm−2) [0.20]
${N}_{{\rm{H}},\mathrm{int}}$ (1022 cm−2) ${0.05}_{-0.05}^{+0.06}$
${{kT}}_{\mathrm{in}}$ (keV) ${0.93}_{-0.16}^{+0.23}$
p ${0.50}_{* }^{+0.12}$
${N}_{\mathrm{dpbb}}$ (10−3 km2)b ${1.2}_{-0.6}^{+6.7}$
${{kT}}_{\mathrm{vapec}}$ (keV) ${0.40}_{-0.04}^{+0.07}$
O abundance (solar units) >5.5
Ne = Mg = Si abundances (solar units) >6.3
Other abundances (solar units) [1]
${N}_{\mathrm{vapec}}$ (10−6)c ${2.8}_{-2.7}^{+4.5}$
${\chi }^{2}/$ dof $71.7/62$ (1.16)
${f}_{0.3-10}$ (10−13 erg cm−2 s−1)d ${0.62}_{-0.04}^{+0.04}$
${L}_{0.3-10}$ (1039 erg cm−2 s−1)e ${0.95}_{-0.20}^{+0.23}$

Notes.

a ${N}_{\mathrm{bb}}={({R}_{\mathrm{bb}}/{D}_{10})}^{2}$, where ${R}_{\mathrm{bb}}$ is the source radius in km and D10 is the distance to the source in units of 10 kpc (here ${D}_{10}=770$). b ${N}_{\mathrm{dbb}}={({R}_{\mathrm{in}}/{D}_{10})}^{2}\cos \theta $, where Rin is the apparent inner disk radius in km, D10 is the distance to the source in units of 10 kpc, and θ is our viewing angle ($\theta =0$ is face-on). Ndpbb is defined exactly as Ndbb. c ${N}_{\mathrm{vapec}}=\displaystyle \frac{{10}^{-14}}{4\pi \,{d}^{2}}\int {n}_{e}{n}_{{\rm{H}}}{dV}$, where d is the angular diameter distance to the source (cm), and ne and nH are the electron and hydrogen densities (cm−3). dObserved fluxes in the 0.3–10 keV band. eFor all spectral models, the de-absorbed luminosities ${L}_{0.3-10}$ (0.3–10 keV band) were defined as 4πd2 times the de-absorbed fluxes

Download table as:  ASCIITypeset image

We also searched for NuSTAR observations that covered the position of the transient and found two: one from 2017 May (ObsID 90302004002; 66.8 ks) and the other from 2017 June (ObsID 90302004004; 47.8 ks). We examined the focal plane module A and focal plane module B data, processed by the pipeline task nupipeline; however, we did not detect any source at the ULX position, either in the individual images or in a stack of the two images.

2.4. HST

The field containing our target source was observed (Table 4) with the Advanced Camera for Surveys (ACS) Wide Field Channel (WFC) on 2004 July 29 with the F814W filter (exposure time of 120 s). It was then reobserved with ACS-WFC on 2016 October 26 in the F606W and F814W filters (exposure times of 2430 and 2570 s, respectively). It was later imaged with the Wide Field Camera 3 (WFC3) Ultraviolet and VISible light camera (UVIS) on 2018 January 5 in the F555W band for 710 s and in F814W for 780 s.

Table 4.  HST Observations of a Candidate Counterpart of CXOU J203451.1+601043

Observation Date Detector Filter Exposure Time Apparent Brightness Absolute Magnitude
      (s) (Vegamag) (Vegamag)
2004 Jul 29 ACS-WFC F814W 120
2016 Oct 26 ACS-WFC F606W 2430 26.35 ± 0.15 −3.95 ± 0.15
2016 Oct 26 ACS-WFC F814W 2570 26.05 ± 0.15 −3.90 ± 0.15
2018 Jan 5 WFC3-UVIS F555W 710
2018 Jan 5 WFC3-UVIS F814W 780

Download table as:  ASCIITypeset image

We retrieved calibrated, geometrically corrected images (.drc files) from the Mikulski Archive for Space Telescopes. We used ds9 to inspect the images and perform aperture photometry of the candidate counterparts, with a source extraction radius of 0farcs15 and a local annular background region. We then converted these small-aperture measurements to infinite-aperture values with the help of the online tables of encircled energy fractions for ACS-WFC and WFC3-UVIS. Finally, we applied the corresponding zero-points to convert our photometric measurements into magnitudes in the Vega system.

3. Results

3.1. Detection and Location of the X-Ray Transient

The first detection of this transient as a bright new X-ray source was in the Chandra observations of 2012 May; its luminosity was 2 orders of magnitude higher than typical previous nondetection limits (Figure 1 and Table 3). It has remained very luminous in all subsequent Chandra, XMM-Newton, and Swift observations, including the most recent ones (Figure 2). The transient resides in one of the spiral arms (Figure 3); the projected galactocentric radius is ≈90'', which is ≈3.4 kpc at the assumed distance of 7.7 Mpc.

Figure 1.

Figure 1. Left panel: adaptively smoothed Chandra/ACIS image of NGC 6946, based on the stacked data from 2001 to 2004. Red represents the 0.3–1 keV band, green is for 1–2 keV, and blue is for 2–8 keV. Right panel: adaptively smoothed Chandra/ACIS image, based on the stacked data from 2012 to 2017, showing the appearance of the transient ULX investigated in this paper.

Standard image High-resolution image
Figure 2.

Figure 2. Gaussian-smoothed Swift/XRT image of the field around the transient ULX CXOU J203451.1+601043 (labeled as t-ULX) in NGC 6946, based on the stacked data from 2018 April to 2019 April, showing that the source is currently still ultraluminous. Red represents the 0.3–1 keV band, green is for 1–2 keV, and blue is for 2–10 keV. The other bright off-nuclear sources labeled "1," "2," and "3" correspond to ULX-1, ULX-2, and ULX-3 in Earnshaw et al. (2019a; see their Figure 1); in particular, ULX-3 is the well-studied ULX inside the MF16 nebula.

Standard image High-resolution image
Figure 3.

Figure 3. Left panel: archival Gemini-North i-band image; the box marks the location of the transient ULX and is zoomed in on the right. Right panel: stellar field around CXOU J203451.1+601043, from an HST/ACS image in the F814W band. The blue circle shows the location of the transient ULX and has a 90% error radius of 0farcs2.

Standard image High-resolution image

At first, we improved the astrometry of the stacked Chandra/ACIS images with the help of a small sample of sources that have a counterpart in the Gaia and 2MASS catalogs (available in ds9). We estimate that our transient source is located at R.A. (J2000) $=\,{20}^{{\rm{h}}}{34}^{{\rm{m}}}51\buildrel{\rm{s}}\over{.} 1$, decl. (J2000) = 60°10'43farcs6, but the uncertainty remained large, ≈0farcs5. The reason for this large positional error is caused by the off-axis location of the X-ray source (hence, a distorted PSF) in three of the four observations in which it is detected; the only Chandra observation in which the ULX is almost on-axis is also short (ObsID 19040; 10 ks), and there are no direct X-ray/Gaia or X-ray/2MASS associations for that short exposure alone.

Therefore, we used a different method in our search for an optical counterpart. In the on-axis observation 19040, there are two bright X-ray sources within 2' of the transient ULX, with a well-identified, pointlike optical counterpart in the HST images. One is SN 2017eaw (Wiggins 2017), located at R.A. (J2000) = 20h34m44fs24, decl. (J2000) = 60°11'35farcs9; the other is the ULX inside the MF16 nebula (Roberts & Colbert 2003).9 We determined the centroids of the X-ray emission from the three sources with wavdetect applied to the ACIS image from ObsID 19040. The uncertainty in the centroid position is ≲0farcs1 for each of the three sources. The transient ULX is located 51farcs4 east and 52farcs6 south of SN 2017eaw. It is also located 71farcs8 west and 47farcs3 south of the MF16 ULX.

Using those relative offsets in the HST images, we constrained the error circle of the transient ULX. The SN 2017eaw (but not MF16) is in the field of view of the WFC3 image from 2018; its relative offset gives the transient ULX location in those images. In the 2004 and 2016 ACS images, both MF16 and the position of SN 2017eaw are in the field of view10 ; this gives us two reference offsets for the relative position of the transient ULX. Both offsets point to the same location with a difference of <0farcs1 between them. In summary, we constrained the relative position of the transient ULX on the HST images with an error radius of ≲0farcs2.

Finally, we also refined the absolute astrometry of the HST images based on Gaia and Sloan Digital Sky Survey associations. This was a much simpler and straightforward task and reduced the uncertainty on the absolute astrometry of the HST images to ≲0farcs1. The most accurate position for the transient ULX is then R.A. (J2000) = 20h34m51fs12, decl. (J2000) = 60°10'43farcs3 (±0farcs2).

3.2. X-Ray Light Curve

First, we studied the long-term X-ray variability of CXOU J203451.1+601043. We determined the net count rate or 90% upper limit in each Chandra/ACIS observation in the 0.3–7 keV band. In some cases, we stacked observations taken a few weeks apart to reach a deeper detection limit. We converted 0.3–7 keV count rates to 0.3–10 keV fluxes assuming the same model for all Chandra observations: a power law with photon index Γ = 2.5 and intrinsic neutral absorption column density NH = 2 × 1021 cm−2, in addition to the line-of-sight Galactic NH = 2 × 1021 cm−2. We chose these model parameters because they are a good approximation to those found from a detailed fit to the 2016 Chandra data (Table 2 and Section 3.3). If we use a "standard" photon index Γ = 1.7 and only line-of-sight absorption, the inferred luminosities or upper limits will be ≈75% of those reported in Table 3.

We did a similar analysis for the XMM-Newton/EPIC observations, stacking the exposures from individual years to increase the signal-to-noise ratio. Count rates were extracted from the 0.3–10 keV band. For the 2012 and 2017 observations, we had enough counts to fit multicomponent models to the data (as we shall discuss in Section 3.3). We adopted the best-fitting Comptonization model to convert from net count rates to the fluxes and luminosities listed in Table 3. For the XMM-Newton observations in which the source was not detected, we determined upper count-rate limits from the combined EPIC images, and we used our fiducial power-law model (Γ = 2.5 and total column density NH = 4 × 1021 cm−2) to convert to flux and luminosity limits.

For the stacked Swift/XRT observations from 2008 (total of ≈10 ks between February 4 and 14) and 2013 (total of ≈7 ks between May 31 and June 4), we used our fiducial power-law model (Γ = 2.5, NH = 4 × 1021 cm−2) to determine fluxes and luminosities or their upper limits. For the other three sets of stacked Swift observations (44 ks in 2017, 43 ks in 2018, and 17 ks so far in 2019), we had enough counts to determine the hardness ratios between the 1.5 and 10 keV band and the 0.3–1.5 keV band. We fixed the total column density to NH = 4 × 1021 cm−2 and used pimms to estimate the power-law photon indices that most closely reproduce the observed hardness ratios; the values are Γ = 3.2 ± 0.5 in 2017, Γ = 2.1 ± 0.4 in 2018, and Γ = 2.6 ± 0.5 in 2019. We then used those photon indices to convert from net count rates to fluxes and luminosities in the respective observations.

The resulting long-term light curve is shown in Figure 4. In all observations between 2001 and 2008, CXOU J203451.1+601043 was always undetected, and it was always detected from 2012 to 2019. The upper limit to the nondetections is typically ≈few × 1037 erg s−1 for individual years and <1037 erg s−1 if all of the observations with nondetections are stacked up. We also know that the source was undetected in the ROSAT High Resolution Imager in 1994 May (Schlegel et al. 2000), with an upper limit to the de-absorbed 0.3–10 keV luminosity of ≈1038 erg s−1 (after converting from the model and distance used in that paper to those used for this work). Since 2012, the source has been hovering at luminosities ≈1–4 × 1039 erg s−1 (depending on the choice of spectral model).

Figure 4.

Figure 4. Long-term X-ray luminosity evolution of CXOU J203451.1+601043 in the 0.3–10 keV band (data from Table 3). When sufficient counts were available, we estimated the luminosities from detailed spectral modeling of individual observations (Table 2); when only a few counts were available or for nondetection limits, we used a fiducial power-law model with photon index Γ = 2.5 and intrinsic column density NH = 2 × 1021 cm−2, which approximates the average spectrum of the source. Each of the Swift data points in this plot is a stack of several short observations over intervals of a few months.

Standard image High-resolution image

We also determined and examined the background-subtracted light curves from individual Chandra and XMM-Newton observations with detections; for this, we used the ftools tasks lcurve and lcstats. We found that the 2016 Chandra observations, binned to 500 s, show statistically significant intra-observation variability by a factor of 2 (Figure 5, top panels), with a χ2 probability of constant rate <1%. The intra-observational variability during the 2012 and 2017 XMM-Newton observations is more marginal (Figure 5, bottom panels). We searched for periods or quasi-periodic oscillations in both the Chandra and XMM-Newton light curves using powspec, efsearch, and efold, but none of the signal peaks is significant above the noise level. In particular, we searched for periods of ∼1 s in the 2012 and 2017 EPIC-pn light curves (binned to 0.1 s), by analogy with typical periods found in ULX pulsars (e.g., Bachetti et al. 2014; Rodríguez Castillo et al. 2019; Sathyaprakash et al. 2019). For this, we used the LombScargle routine in astropy version 3.2.1 (Astropy Collaboration et al. 2018). For the 2012 light curve, we found a probability >40% that any of the peaks in the Lomb–Scargle periodogram are due to random fluctuations of photon counts; for the 2017 data set, that probability is >50%. Thus, we cannot detect any significant period in this source. The relatively low number of counts and signal-to-noise ratio of these observations are not sufficient for any more detailed variability analysis on such short timescales.

Figure 5.

Figure 5. Top left panel: background-subtracted Chandra/ACIS-S light curve from the first of the two exposures on 2016 September 28, binned to 500 s. It shows moderate intra-observational variability. Top right panel: same as the top left panel but for the second ACIS-S exposure on 2016 September 28. Bottom left panel: background-subtracted XMM-Newton/EPIC-pn light curve from 2012 October 21, binned to 1000 s. Bottom right panel: same as the bottom left panel but for the EPIC-pn observation of 2017 June 1.

Standard image High-resolution image

3.3. Spectral Properties

We have enough counts for detailed modeling of the spectra from XMM-Newton in 2012 and 2017 and Chandra in 2016. As a first step, we tried a simple power-law model (tbabs × tbabs × pow) with a fixed Galactic absorption component NH,Gal = 2 × 1021 cm−2 (Kalberla et al. 2005) and a free intrinsic component. The model provides a good fit for the 2016 spectrum (partly because of the more limited band coverage of ACIS) but leaves significant systematic residuals in the 2012 and 2017 spectra. One finding that is already obvious from a power-law fit is that the 2017 spectrum is significantly softer (Γ ≈ 3.5) than the other two. Another finding is a strong residual feature consistent with an unresolved emission line at E ≈ 0.66 keV in the 2017 and (at a weaker level) 2012 data. The presence of this feature is significant for every choice of continuum model (power-law and every other continuum component tested in the rest of this section); we will discuss this line in more detail later.

Next, we tried another simple one-component model: tbabs × tbabs × diskbb. The disk–blackbody model does not improve on the power-law model for any of the three spectra; in simple terms, it is too curved for the observed data points. We improved on the disk–blackbody model in two alternative ways: by adding a second (softer) thermal component (tbabs × tbabs × (diskbb + bbodyrad)) and by using a p-free disk model (tbabs × tbabs × diskpbb) with p < 0.75.

The double thermal model provides a good fit (χ2ν < 1.2) at all three epochs (Table 2). The lower-temperature blackbody has a characteristic temperature kTbb ≈ 0.15–0.3 keV, typical of the soft excess observed in many ULXs (Gladstone et al. 2009; Kajava & Poutanen 2009; Kaaret et al. 2017; Zhou et al. 2019), and a characteristic radius of ∼103 km for all three epochs (again, a common feature for this class of systems). The best-fitting inner disk color temperatures in 2012 (kTin ≈ 1.25 keV) and 2016 (kTin ≈ 1.14 keV) are consistent with the temperature expected from the inner disk of a stellar-mass BH at the Eddington limit. The physical inner disk radius Rin is assumed to be the innermost stable circular orbit and is defined as Rin ≈ 1.19rin, where rin is the best-fitting radius in the diskbb model in xspec (Kubota et al. 1998); the standard scaling factor 1.19 accounts for the hardening factor (≈1.7) and the fact that the peak effective temperature occurs slightly outward from the inner disk radius, at R ≈ (49/36) Rin (Kubota et al. 1998). In our case, we find ${R}_{\mathrm{in}}\sqrt{\cos \theta }\approx 55$ km in 2012 and ${R}_{\mathrm{in}}\sqrt{\cos \theta }\approx 70$ km in 2016; again, these values are typical of stellar-mass BHs at high luminosities. Instead, the best-fitting temperature kTin ≈ 0.7 keV, determined for the 2017 spectrum, is too low for the observed luminosity. For comparison, in other BH X-ray binaries in which the accretion disk emission is still the dominant component of the X-ray spectrum at LX ≈ LEdd ≈ 1039 erg s−1, the peak color temperature kTin ≈ 1.2–1.5 keV (Kubota & Makishima 2004; Soria 2007; Sutton et al. 2013). This suggests that although the curvature of a disk model provides a formally good fit to the data, it does not provide the correct physical interpretation of the emission at that epoch. The de-absorbed 0.3–10 keV luminosity, defined as 4πd2 times the de-absorbed flux f11 , is L0.3−10 ≈ 2 × 1039 erg s−1 in both 2012 and 2016 (Table 2).

The p-free disk model is a simple approximation to slim disk models (Abramowicz et al. 1988; Watarai et al. 2001), suitable for stellar-mass BHs near or slightly above their Eddington limit. The definition of the parameter p is that the effective temperature on the disk scales as T(R) ∝ R−p; the standard disk case is p = 0.75, while near-Eddington BHs tend to have p ≲ 0.6 (Sutton et al. 2017). In our case, the p-free disk provides statistically equivalent fits to the double thermal model for the 2012 and 2016 spectra (${\chi }^{{2}_{\nu }}\approx 1.1$ at both epochs). The best-fitting temperature kTin ≈ 1.5 keV in 2012 and kTin ≈ 1.4 keV in 2016, consistent with typical values expected for slim disks near the Eddington limit. For a p-free disk, the physical inner radius is usually defined as Rin ≈ 3.19rin (Vierdayanti et al. 2008); in our case, we obtain ${R}_{\mathrm{in}}\sqrt{\cos \theta }\approx 75$ km in both 2012 and 2016. De-absorbed luminosities are ≈1.5–2 × 1039 erg s−1 in both epochs. Instead, a p-free disk is not a good fit for the 2017 spectrum (χ2ν ≈ 1.6); specifically, it does not model well the characteristic downturn seen in the data above 2 keV.

For our fourth attempt to model the data, we used a Comptonization model: tbabs × tbabs × (bbodyrad + comptt). Here the bbodyrad component plays the role of seed thermal emission. Replacing bbodyrad with a diskbb seed component does not change our results, because we are only looking at the Wien section of the curve. This model provides a good fit for all three epochs. For 2012 and 2016, it is statistically equivalent to the double thermal and p-free models; for 2017, it is as good as the double thermal model. The seed photon temperature is kT0 ≈ 0.15 keV in 2012 and 2017, while it is unconstrained (≲0.1 keV) in 2016. The electron temperature in the Comptonizing cloud, which determines the location of the high-energy downturn, is ∼1 keV in 2012 and 2017, while it is unconstrained in 2016, where the presence of a high-energy downturn is not statistically significant (mostly because of the more limited band coverage in Chandra). Low temperatures and high optical depths are among the defining properties of ULXs (Gladstone et al. 2009) when their spectra are fitted with Comptonization models. The most common interpretation for this kind of spectra is that the harder photons from the inner disk region are down-scattered in a thick disk outflow, seen at high inclination angles. The de-absorbed luminosity for the Comptonization model in 2012 and 2017 is ≈1.5–3 × 1039 erg s−1, similar to the luminosity inferred from the other models. Instead, the luminosity is unconstrained in the 2016 Chandra spectrum because the temperature of the seed thermal component is too low and its normalization is unconstrained; if we neglect the contribution of the seed photons, the luminosity is ≈4 × 1039 erg s−1 in 2016. Alternatively, we fixed the temperature of the blackbody and seed thermal components at kT0 = kTbb ≡ 0.10 keV to reduce the degeneracy between temperature, normalization, and absorption column. The best-fitting parameters are essentially unchanged; the de-absorbed luminosity has a best-fitting value of ≈5.4 × 1039 erg s−1 (of which ≈3.5 × 1039 erg s−1 is from the Comptonized component), with a 90% lower limit of ≈2.6 × 1039 erg s−1 and a badly constrained upper limit of ≈3.3 × 1040 erg s−1.

In summary, all three spectra are consistent with a mildly super-Eddington stellar-mass BH. The EPIC spectrum from 2012 is more consistent with a broadened disk regime (Gladstone et al. 2009; Sutton et al. 2013) but can also be the result of Comptonization. The 2016 ACIS spectrum is softer than that in 2012 but cannot be reliably classified because of the limited band coverage. The 2017 EPIC spectrum is significantly softer than the other two, is best modeled with a Comptonization model, and belongs to the soft ultraluminous regime. The de-absorbed luminosity was ≈1.5–2 × 1039 erg s−1 in 2012, ≈2–4 × 1039 erg s−1 in 2016, and ≈2–3 × 1039 erg s−1 in 2017. The softness of the 2017 spectrum compared with the other two is obvious when we plot unfolded spectra (Figure 6), based on Comptonization models for consistency.

Figure 6.

Figure 6. Unfolded X-ray spectra of CXOU J203451.1+601043 at three different epochs, based on the best-fitting Comptonization models listed in Table 2 (tbabs × tbabs × (bbodyrad + comptt + Gaussian)). Blue data points are for the 2012 XMM-Newton spectrum, green data points are for the 2016 Chandra spectrum, and red data points are for the 2017 XMM-Newton spectrum (including a strong oxygen emission line).

Standard image High-resolution image

For all continuum spectral models described above, we found a significant emission line residual at E = 0.66 ± 0.01 keV in the 2017 spectrum (Table 2); a similar residual but with much lower significance is also seen in the 2012 spectrum. We modeled the residual with a Gaussian emission line. We found an equivalent width of ${95}_{-60}^{+55}$ eV in 2017 (almost independent of the choice of continuum model) and an intrinsic line luminosity of $\approx 2.4$ $\times {10}^{38}$ erg s−1, which is ≈10% of the total intrinsic luminosity in the EPIC band. The FWHM of the line (σline in Table 2) is consistent with zero (that is, the intrinsic velocity broadening of the line is much less than the instrumental broadening) and constrained to be ≲30 eV (≈14,000 km s−1) to the 90% confidence level in the 2017 spectrum, with only small differences between the different continuum models.

To assess the significance of the line component, first of all, we note that the 90% lower limit of the line normalization parameter (photon flux) is >0 in both the 2012 and 2017 spectral fits (Table 2). The improvement in the fit statistic when we include the Gaussian emission line component is ${\rm{\Delta }}{\chi }^{2}\gt 11.4$ for all of the continuum spectral models. We obtained a more rigorous statistical constraint with the likelihood ratio test lrt in xspec: we ran 10,000 simulations for various pairs of models with and without the line component. The significance of the line in 2012 is only ≈70%, while in 2017, it is >99.8%, regardless of the continuum model. In fact, Figure 7 shows that the 0.66 keV emission line in the 2017 spectrum is evident even simply by eye. We also built and inspected EPIC-pn and MOS images in the 0.60–0.70 keV band (Figure 8) for the 2017 data set to make sure that the emission is dominated by the pointlike ULX and not contaminated by, for example, diffuse hot gas in a spiral arm. We compared the narrow-band images from 2017 with those from earlier epochs, and from Chandra when the ULX was not detected, to exclude the possibility of a preexisting supernova remnant at that location. As a result of these tests, we are confident that the ≳1038 erg s−1 line emission seen does come from the ULX.

Figure 7.

Figure 7. Top left panel: best-fitting spectrum and χ2 residuals for the 2012 XMM-Newton/EPIC data set (pn and MOS combined), fitted with a Comptonization model (see Table 2 for the fit parameters). Top right panel: same as the top left panel but for the 2016 Chandra/ACIS-S spectrum. Bottom left panel: same as the top left panel but for the 2017 XMM-Newton/EPIC spectrum. Notice the strong line at 0.66 keV fitted with a Gaussian. Bottom right panel: zoomed-in view of the soft X-ray band for the 2017 XMM-Newton/EPIC spectrum, fitted this time only with a continuum model, without the addition of any lines. Systematic residuals are clearly significant.

Standard image High-resolution image
Figure 8.

Figure 8. Top panel: XMM-Newton/EPIC-MOS image from the 2017 data set, filtered to the 0.60–0.70 keV band. It shows that the O viii line emission is associated with the pointlike ULX (labeled as t-ULX) and is not due to the contamination from diffuse hot gas. The other bright off-nuclear sources labeled "1," "2," "3," and "4" correspond to ULX-1, ULX-2, ULX-3, and ULX-4 in Earnshaw et al. 2019a. Bottom panel: same as the top panel for the EPIC-pn image, consistent with the MOS image. It shows that the line is not an instrumental artifact in one of the EPIC detectors. (The bright streak on the left is a bad column, flagged out for spectral analysis.)

Standard image High-resolution image

The most likely identification is the O viii Lyα line (rest-frame energy of 0.654 keV). Such a line was strongly detected in other soft ULXs and is interpreted as a signature of fast outflows (Pinto et al. 2016, 2017; Kosec et al. 2018). Although it is not a surprise to detect this line in a ULX, it is unusual to see it so strong and significant even in a moderate-quality spectrum at CCD resolution (see also Section 4.2 for a comparison with other ULXs).

Another line that has been found associated with ULX outflows is the O vii triplet at 0.56 keV (Pinto et al. 2016, 2017; Kosec et al. 2018). In our spectra, we do not have enough counts to detect or place meaningful constraints on the presence of this line. The main reason for the lower signal-to-noise ratio is that photoelectric absorption is stronger at 0.56 than at 0.65 keV, and, conversely, the pn and MOS effective areas are lower. Nonetheless, we tried adding a narrow line with energy fixed at 0.56 keV and estimated the 90% upper limit to its normalization for each of the four spectral models described before. The result of this test is that a 0.56 keV with the same photon flux as the 0.66 keV line (i.e., ∼10−5 photons cm−2 s−1, or EW ≲ 75 eV) is still consistent with our data, even though this additional line would not stand out "by eye" in the spectra. The lack of strong constraints on the strength of the O viii triplet means that we cannot exclude a (minor) contribution to the 0.66 keV line from the O viii Heβ line at 0.67 keV.

The 2017 EPIC spectrum also shows hints of other emission line features in the soft band (Figure 7, bottom right panel), for example, the Mg xi triplet at 1.35 keV. The presence of other possible line features motivated us to try and replace the single Gaussian component in our 2017 spectral fits with a thermal plasma component, vapec in xspec, suitable for collisionally ionized gas. Keeping the metal abundances of all elements at the solar value does not provide an improvement compared with the simpler model fits without a Gaussian line. Similarly, leaving all abundances free but locked together does not improve the fit. Instead, we found that we can obtain good fits (Table 3) by leaving the abundance of some elements (O, Ne, Mg, and Si) free and keeping the other abundances at the solar level. The very high relative abundances of the α elements required to fit our spectrum may be a clue for the type of donor star, as we shall discuss later (Section 4.3). The de-absorbed luminosity in the vapec component is ≈3 × 1038 erg s−1.

For the 2017 spectrum, we then replaced the solar-metallicity intrinsic absorption component (tbabs) with variable abundance absorption models (we tried tbvarabs and vphabs). We did not find any significant improvement or change in the emission line properties when the oxygen abundance in the intrinsic absorber is a free parameter. In fact, the oxygen abundance in the intrinsic absorber is constrained to be ≲2 times solar at the 90% confidence level. This suggests that the gas responsible for the intrinsic absorption is not the same oxygen-rich gas responsible for the strong 0.66 keV emission line; the intrinsic absorption component may be located in the halo and disk of NGC 6946, rather than within the binary system.

Finally, we modeled the stacked spectra from the Swift/XRT observations in 2017, 2018, and 2019. Because of the low number of counts, the detailed models described above are degenerate and cannot be adequately constrained; however, we can still get useful information on the hardness evolution. To do so, we fixed the intrinsic column density NH = 2 × 1021 cm−2, as derived from most of the Chandra and XMM-Newton spectral fits, and fitted the power-law slope and normalization. We found that the ULX spectrum has hardened in recent years, from Γ = 3.2 ± 0.5 in 2017 (consistent with the contemporaneous XMM-Newton observations), to a more moderate Γ = 2.1 ± 0.4 in 2018, to Γ = 2.6 ± 0.5 in 2019. At the same time, the X-ray luminosity has remained approximately constant at ≈3 × 1039 erg s−1.

3.4. Constraints on the Optical Counterpart

We have already described (Section 3.1) how we used the relative offsets to MF16 and SN 20017eaw to pinpoint the location of the X-ray transient in the HST images. We detect only one faint optical source inside the error circle (Figure 9), in the 2016 ACS images, in both the F606W and F814W bands. The same source is too faint to be detectable in the 2004 and 2017 observations (shorter exposure times). Thus, we cannot constrain the source variability. Aside from the positional coincidence, there is no direct evidence that this optical source is the actual donor star for the transient X-ray source.

Figure 9.

Figure 9. Top panel: 2016 HST/ACS image in the F606W band. The yellow circle represents the 90% confidence limit of 0farcs2 for the ULX position. The only source marginally detected inside the circle has an apparent brightness of ≈26.4 mag and an absolute magnitude of ≈−4.0 mag. Bottom panel: 2016 HST/WFC3 image in the F814W band.

Standard image High-resolution image

We measured an apparent brightness mF606W = (26.35 ± 0.15) and mF814W = (26.05 ± 0.15) mag (corrected to infinite aperture). After correcting for line-of-sight extinction (AF606W = 0.85, AF814W = 0.52 mag; Schlafly & Finkbeiner 2011), we obtain absolute magnitudes MF606W = (−3.95 ± 0.15) and MF814W = (−3.90 ± 0.15) mag (Table 4). This is consistent with an early B star (main-sequence or subgiant), while it rules out O stars and supergiants. It is a type of optical counterpart very common in ULXs (Gladstone et al. 2013; Motch et al. 2014). However, in Section 4.3 we will propose a white dwarf scenario for the donor star, in which case this faint optical source is unrelated to the ULX.

4. Discussion

4.1. Transient Behavior

There are at least two types of transient X-ray sources that reach the ultraluminous regime. One type, well exemplified by CXOU J203451.1+601043, turns on as a ULX after years of nondetection (typically at least 2 orders of magnitude fainter) and then remains in the super-Eddington regime for many years. Another example of this behavior is the transient ULX in M83 (Soria et al. 2012, 2015). In our galaxy, GRS 1915+105 turned on in 1992 (Castro-Tirado et al. 1992) and has been near the Eddington limit ever since, at least until mid-2018, when it started to decline toward the low/hard state (Negoro et al. 2018; Miller. et al. 2019; Rodriguez et al. 2019). The second type of ultraluminous transient is well represented by another source recently discovered in NGC 6946, labeled ULX-4 in Earnshaw et al. (2019a); its outburst lasted only ∼10 days. Another example of a short-duration ULX transient is the first ULX discovered in M31, which went back to quiescence after a few months (Middleton et al. 2013) . In the Milky Way, V404 Cyg provides the most notable example of a short-duration outburst reaching the Eddington luminosity and then declining to quiescence, between 2015 June and August (Sivakoff et al. 2015; Kimura et al. 2016; Muñoz-Darias et al. 2016; Motta et al. 2017).

The most commonly accepted theoretical interpretation of X-ray outbursts in Galactic X-ray binaries is based on the thermal-viscous disk instability (King & Ritter 1998; Dubus et al. 2001; Lasota 2001), which depends on the existence of an outer disk region where hydrogen is mostly neutral. The natural extension of this model for long-duration super-Eddington transients (such as CXOU J203451.1+601043) requires the simultaneous presence of two ingredients: a sufficiently high irradiation of the outer disk to keep it in the ionized state and a sufficiently high accretion rate to keep the source at or above the Eddington limit for years. If the donor is a low-mass star, which cannot keep a persistently high mass transfer rate through the L1 Lagrangian point, a long-duration phase of super-Eddington accretion may still be achieved if the disk is very large12 and able to store enough mass prior to an outburst (the mass stored in the disk is proportional to Rout3, and the peak luminosity scales as Rout2; King & Ritter 1998). For example, in the case of GRS 1915+105, the estimated disk size is ∼1012 cm, and the mass in the disk is ∼1028 g, which suffices to keep the X-ray source at or above 1039 erg s−1 for decades (Done et al. 2004). As for the other requirement (i.e., a strong irradiation of the outer disk), broadband spectral modeling of observational data suggests (Sutton et al. 2014) reprocessing fractions of ∼10−3 for sub-Eddington stellar-mass X-ray binaries and some ULXs and ∼10−2 for other ULXs, specifically those with soft X-ray spectra. The reason for the enhanced irradiation factor in some ULXs may be that their strong disk wind intercepts and scatters a fraction of photons emitted in the polar funnel and redirects them onto the outer disk (Sutton et al. 2014).

This is not the only viable explanation for transient behavior in ULXs. If a ULX is powered by an NS rather than a BH, transitions between the propeller and the accretor state will cause transient behavior (Dall'Osso et al. 2015; Tsygankov et al. 2016; Earnshaw et al. 2018). The drop in luminosity happens when the magnetospheric radius of the NS (the radius at which the magnetic pressure "stops" the inflowing matter) becomes larger than the corotation radius of the disk, thus creating a centrifugal barrier (Illarionov & Sunyaev 1975; Stella et al. 1986). The location of the magnetospheric radius depends on the NS spin, mass accretion rate, and strength of the NS magnetic field. The existence of a luminosity gap in the long-term light curve of a transient ULX, between ∼a few 1037 and ∼1039 erg s−1, would be a clue in favor of an accretor/propeller switch model, rather than a thermal-viscous disk instability (where we expect the system to evolve smoothly between higher and lower luminosities without discrete jumps). For CXOU J203451.1+601043, all we can say at the moment is that the system was never detected at any other luminosity <1039 erg s−1; thus, the accretor/propeller model is still viable. Future observations of CXOU J203451.1+601043 will be needed to determine whether, when, and how it will decline.

4.2. The Soft Ultraluminous Regime and the Oxygen Line

We showed (Section 3.3) that the EPIC spectrum from 2012 is consistent with either a broadened disk regime or Comptonization, while the 2017 spectrum is significantly softer, is not consistent with disk models, and suggests down-scattering of the direct X-ray emission in a cooler medium, such as the disk outflow. A high-energy downturn already at a photon energy of ≈3 keV (or, equivalently, a steep photon index Γ ≈ 3.5 when fitted in the 0.3–8 keV band) makes the 2017 state of this ULX even softer than NGC 5408 X-1, i.e., the source usually taken as a standard for the soft ultraluminous regime (Sutton et al. 2013; Middleton et al. 2014). The 2017 spectrum puts it in the same class as NGC 55 X-1 (Stobbart et al. 2004; Pinto et al. 2017) and NGC 247 X-1 (Feng et al. 2016), i.e., the two sources that are in a transitional state between the classical ULX and supersoft regimes. The very soft spectral appearance found in 2017 is also reminiscent of the eclipsing ULX CXOM51 J132940.0+471237 in M51 (Urquhart & Soria 2016). The high scattering optical depth (τ ≈ 13) fitted to the comptt model is consistent with the relation between optical depth and coronal temperature found by Pintore et al. (2014), again at the very soft end of the sequence.

The spectral and timing properties of the soft ultraluminous regime are generally attributed to our viewing angle passing through the thick disk outflow (Kawashima et al. 2012; Sutton et al. 2013; Pintore et al. 2014; Middleton et al. 2015a; Narayan et al. 2017; Pinto et al. 2017). This reduces or suppresses our detection of hard X-ray photons (directly emitted from the inner disk region) and increases the contribution of down-scattered soft X-ray photons. Spectral softening may be caused by an increase of the radiatively driven wind mass-loss rate and, therefore, its optical depth, which is a function of, for example, the accretion rate (Poutanen et al. 2007). In the case of CXOU J203451.1+601043, the de-absorbed luminosity fitted to the observed spectral data points varies only by a factor of 2 between the Chandra, XMM-Newton, and Swift observations; however, those luminosity estimates correct only for the effect of cold absorption, not for the down-scattering of hard X-ray photons into the soft X-ray or far-UV band. It is plausible that the intrinsic luminosity in 2017 would be much higher if we could see the direct emission unaffected by the down-scattering outflow (for example, if we had a pole-on view). Precession of the viewing angle may also change the scattering optical depth seen by distant observers for constant wind properties.

If our interpretation of CXOU J203451.1+601043 as an extreme example of the soft ultraluminous regime is correct, we expect two other properties associated with this regime. One is the large amount of short-term variability (Middleton et al. 2015a). Moderate intra-observational variability is observed in our X-ray timing analysis; however, the observed count rate is too low to constrain the rms fractional variability at high frequencies. Thus, we cannot make any firm conclusions on whether this source is more variable than other ULXs.

The second property that should be associated with a thick down-scattering outflow is the presence of spectral residuals in the soft X-ray band, caused by blends of emission and absorption lines (Middleton et al. 2015b; Pinto et al. 2016, 2017). Indeed, we have shown the presence of at least one strong O viii emission line at E = (0.66 ± 0.01) in the 2017 EPIC spectrum, with a luminosity of ≈2 × 1038 erg s−1 and an equivalent width of ≈100 eV. The line was significantly stronger than that in the 2012 spectrum. We have also shown hints of other likely residuals around 1.35 (Mg xi) and 1.7 (Mg xii) keV. The luminosity of the O viii emission line in 2017 was an order of magnitude higher than that of analogous lines detected in other ULXs, such as NGC 1313 X-1 and NGC 5408 X-1 (Middleton et al. 2015b; Pinto et al. 2016), and also an order of magnitude higher than that of the O viii line in NGC 55 X-1 (Pinto et al. 2017), despite the similarity in the X-ray continuum.

4.3. An Ultracompact ULX?

Why is the oxygen line so strong? We speculate that the donor star is oxygen-rich, well above solar abundance. One scenario is that the donor star is an oxygen-rich Wolf-Rayet (WO subclass; Crowther 2007; McClelland & Eldridge 2016). The optical luminosity of a WO star is low enough (Sander et al. 2019) to be consistent with our upper limit of ≈−4 mag for the optical counterpart. Our spectral modeling suggests that the source is likely viewed at a high inclination angle. Consequently, if the donor star was a Wolf-Rayet, we would expect to see strong sinusoidal variability or even eclipsing behavior in the X-ray flux, by analogy with other Wolf-Rayet X-ray binaries (Qiu & Soria 2019; Qiu et al. 2019). Instead, no such variability is detected in the light curves of this ULX (Figure 5). A more plausible scenario is that the donor star is a CO or (preferably) an O–Ne–Mg white dwarf; the latter are the most massive subclass of white dwarfs (Truran & Livio 1986; Shara & Prialnik 1994), formed from B-type stars with initial masses just below the limit (≈8 M) for supernova explosions. In order to form a luminous X-ray binary, the white dwarf must be filling its Roche lobe in an ultracompact system (e.g., van Haaften et al. 2012). Strong, relativistically broadened O viii Lyα lines attributed to Compton reflection have been seen from some (sub-Eddington) ultracompact X-ray binaries (UCXBs) in the Milky Way (Madej et al. 2010, 2014; Madej & Jonker 2011). The origin of the emission line in CXOU J203451.1+601043 may be different (an outflow rather than Compton reflection on the inner disk surface), but we mention this analogy simply as an indication of an oxygen-rich accretion flow. It is also theoretically possible that UCXBs reach super-Eddington luminosities: ULXs in globular clusters, such as the one in the RZ 2109 cluster of NGC 4472 (Maccarone et al. 2007), have been interpreted as UCXBs. In particular, a strong [O iii] λ5007 emission line was detected in the optical spectra of the RZ 2109 source (Steele et al. 2014) and interpreted as evidence of a hydrogen-poor, oxygen-rich donor and an outflow powered by the accreting compact object. In the soft X-ray band, several XMM-Newton and Chandra spectra of the UCXB in RX 2109, presented by Dage et al. (2018), also show emission residuals at 0.6–0.7 keV consistent with O viii Lyα emission.

A possible issue with the UCXB scenario is that the candidate ULX UCXBs suggested in the literature so far are all in globular clusters, where dynamical formation is strongly enhanced; instead, CXOU J203451.1+601043 is in the field, in or near a spiral arm, almost certainly not in a globular cluster, because its optical counterpart is fainter than MI ≈ −4 mag. The typical optical luminosity distribution of old globular clusters spans $-11\lesssim {M}_{I}(\mathrm{mag})\lesssim -5$ (e.g., Secker 1992; Barmby et al. 2000; Jordán et al. 2007). On the other hand, several Galactic UCXBs are also found in the field outside globular clusters or the bulge (Cartwright et al. 2013). The transient nature of CXOU J203451.1+601043 is also a puzzle: the disk in an ultraluminous UCXB is too hot (and therefore fully ionized) to undergo thermal-viscous instabilities. Instead, the transient behavior could be caused by mass transfer instabilities from the donor star or an accretor/propeller switch if the compact object is a magnetized NS.

5. Conclusions and Future Prospects

Using archival Chandra and XMM-Newton observations, we have identified a previously unrecognized transient ULX in NGC 6946. The source was undetected at luminosities ≲a few 1037 erg s−1 in all observations until 2008 and always detected at luminosities of ≈1.5–3 × 1039 erg s−1 in all observations between 2012 and 2019. We pointed out a few interesting properties that help our understanding of the ultraluminous regime. First, we showed that the source is extremely soft: if modeled with a Comptonization spectrum, the electron temperature is ≈0.7 keV in the 2017 EPIC spectrum. Spectral evolution between different epochs suggests a change in the down-scattering wind properties. In its softest state (2017 XMM-Newton observations), CXOU J203451.1+601043 is in the transitional regime between standard ULXs and ultraluminous supersoft sources. This finding supports the argument that soft ULXs and ultraluminous supersoft sources are fundamentally similar systems, distinguished by the optical depth of the scattering wind along our line of sight. Second, we showed that CXOU J203451.1+601043 has another property associated with super-Eddington outflows: strong line residuals in the soft X-ray band. In particular, the strong 0.66 keV emission line (likely to be the O viii Lyα line) is the most outstanding feature of this source in its softest state. Very few ULXs display wind emission lines so strong that can be easily identified and modeled even at CCD resolution. We speculate that the strong oxygen line is evidence of an oxygen-rich donor star, such as an O–Ne–Mg white dwarf. If so, it would be the first example of an ultracompact ULX outside a globular cluster, adding more variety of formation channels to the already heterogeneous ULX population.

Future follow-up studies of CXOU J203451.1+601043 may provide important constraints on at least three unsolved problems. First, it will be important to monitor the duration of the super-Eddington regime, which has already lasted at least 7 yr and shows no sign of decline. When (if) the source does decline, the crucial test will be whether the outburst decline follows the characteristic hardness–luminosity tracks of stellar-mass BHs (Fender et al. 2004) or shows the sudden disappearance expected for accretor/propeller transitions in NSs. Second, future observations of this ULX offer a chance to determine a quantitative relation between the strength of the 0.66 keV emission line (and other line residuals) and the energy of the downturn in the continuum, linking the imprint of the wind on line and continuum emission. Third, we speculate that if CXOU J203451.1+601043 is a UCXB, X-ray light curves may reveal a characteristic period of ∼10 minutes, but we may have to wait until deeper observations with Athena to find out.

We thank the referee for a careful reading and useful suggestions. We are grateful to Song Wang (NAOC) and Jianfeng Wu (XMU) for assistance with the timing analysis and Hua Feng (Tsinghua University) for helpful discussion. J.W. was supported by the National Key R&D Program of China (2016YFA0400702) and the National Science Foundation of China (U1831205, 11473021, 11522323). For this research, we used data obtained from the Chandra Data Archive and ciao software provided by the Chandra X-ray Center. We also used data obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. Our optical results were based on images from the NASA/ESA Hubble Space Telescope, obtained from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555.

Footnotes

  • The confidence interval tables and plots provided by Kraft et al. (1991) only cover the range of ≈zero to 10 counts. The (raw) source and background counts in our EPIC images are typically higher than that. To obtain the Bayesian confidence intervals for our data, we used the Bayesian Analysis Toolkit package (Caldwell et al. 2009) version 1.0.0, downloaded from https://bat.mpp.mpg.de.

  • The MF16 nebula is obviously extended, but the optical counterpart for the peak of the X-ray emission is a pointlike blue star in the center of the nebula.

  • 10 

    Obviously, SN 2017eaw was not visible at those epochs, but its precise location on the ACS chip is easily determined from the relative position of the surrounding stars, compared with the 2018 images.

  • 11 

    The luminosity L of a disk emission component is more properly defined as L = 4πd2f/cos θ; however, it is also customary to adopt L = 2πd2f in the absence of information about the viewing angle θ.

  • 12 

    A large disk is probably a necessary but not sufficient condition for a long outburst. For example, the transient Galactic stellar-mass BH V404 Cygni, mentioned before, also has a large disk despite its short-duration outburst. This has been interpreted (Muñoz-Darias et al. 2016) as the effect of a strong outflow that disrupted the supply of accreting matter from the outer disk to the inner disk.

Please wait… references are loading.
10.3847/1538-4357/ab3c4d