A publishing partnership

The Peculiar X-Ray Transient Swift J0840.7−3516: An Unusual Low-mass X-Ray Binary or a Tidal Disruption Event?

, , , , , , , , , , , and

Published 2021 April 6 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Megumi Shidatsu et al 2021 ApJ 910 144 DOI 10.3847/1538-4357/abe6a1

Download Article PDF
DownloadArticle ePub

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

0004-637X/910/2/144

Abstract

We report on the X-ray properties of the new transient Swift J0840.7−3516, discovered with Swift/BAT in 2020 February, using extensive data from Swift, MAXI, NICER, and NuSTAR. The source flux increased for ∼103 s after the discovery, decayed rapidly over ∼5 orders of magnitude in five days, and then remained almost constant over nine months. Large-amplitude short-term variations on timescales of 1–104 s were observed throughout the decay. In the initial flux rise, the source showed a hard power-law-shaped spectrum with a photon index of ∼1.0 extending up to ∼30 keV, above which an exponential cutoff was present. The photon index increased in the following rapid decay and became ∼2 at the end of the decay. A spectral absorption feature at 3–4 keV was detected in the decay. It is not straightforward to explain all the observed properties by any known class of X-ray sources. We discuss the possible nature of the source, including a Galactic low-mass X-ray binary with multiple extreme properties and a tidal disruption event by a supermassive black hole or a Galactic neutron star.

Export citation and abstract BibTeX RIS

1. Introduction

Observations of X-ray transients have provided valuable opportunities to study a huge variety of Galactic and extragalactic variable sources, from stellar objects including flaring stars, cataclysmic variables, X-ray binaries, and gamma-ray bursts (GRBs), to supermassive black holes, including active galactic nuclei (AGNs) and tidal disruption events (TDEs) of stars. In particular, the combination of all-sky monitoring and follow-up observations has deepened our knowledge of their behavior in both early and late phases after their emergence, and helped us to achieve critical information on their nature and the physical mechanisms of energetic transient phenomena.

Swift J0840.7−3516 (also known as GRB 200205A) was discovered with the Neil Gehrels Swift Observatory (Swift; Gehrels et al. 2004) on 2020 February 5 UT 06:35 (Evans et al. 2020; Kennea et al. 2020; Osborne et al. 2020), located at ∼4° above the Galactic plane. The Monitor of All-sky X-ray Image (MAXI; Matsuoka et al. 2009) nova alert system (Negoro et al. 2016) also detected the source about 20 minutes after the Swift detection (Niwano et al. 2020). The MAXI/Gas Slit Camera (GSC) intensity increased from ∼70 mCrab to ∼210 mCrab in the 2–10 keV band from the first to second scan after the Swift detection, and then rapidly decreased to ∼30 mCrab in the third scan. Follow-up X-ray observations have been made with Swift and the Neutron Star Interior Composition Explorer (NICER; Gendreau et al. 2016) 59 and 16 times (as of 2020 December 10), respectively, and once with the Nuclear Spectroscopic Telescope Array (NuSTAR; Harrison et al. 2013) on 2020 February 8–9. A possible periodicity of 8.96 s was reported by using the Swift/XRT data (Kennea et al. 2020), but in subsequent NICER observations no periodic variations were significantly detected (Iwakiri et al. 2020).

Optical and UV counterparts were found in photometric observations (Lipunov et al. 2020; Melandri et al. 2020; Malesani et al. 2020; Mazaeva & IKI-FuN follow-up Collaboration 2020). Malesani et al. (2020) suggested that the source may be a Galactic source rather than a GRB because of the location and an unusually high optical/gamma-ray brightness ratio in early phases. No radio counterpart has been found so far; Borghese et al. (2020) posed a 3σ upper limit of 18 μJy at 7.25 GHz using the Australia Telescope Compact Array (ATCA) on 2020 February 11.

Despite the intensive follow-up observations, the nature of Swift J0840.7−3516 is still unknown. In this article, we analyze X-ray data from the source obtained with Swift, MAXI, NICER, and NuSTAR to understand its nature and radiation processes. We utilized HEAsoft version 6.26.1 for data reduction and analysis. Errors represent 90% confidence ranges of single parameters unless otherwise specified.

2. Observation and Data Reduction

2.1. Swift

Since the first BAT detection on February 5 UT 06:35, Swift observed the source almost every day in February and about once a week from the beginning of March to the end of November, with exposure times of 0.1–5 ks. We used all XRT data of the 59 observations. Figure 1 shows the Swift/XRT light curves and hardness ratio for the entire period, obtained via the web interface provided by the Swift team 16 (Evans et al. 2007, 2009), in which we used the latest CALDB (version 2019 September 10) when this source was first observed with Swift. The same interface was used to retrieve XRT spectra. We created time-averaged spectra for the individual observations until February 9. An exception is the second observation (OBSID = 00954304001) performed on February 5 UT 07–16, which was divided into intervals for each snapshot (i.e., each continuous exposure) because of strong variations, although the last three intervals in UT 14–16 were merged due to low statistics. We combined the data for every ∼40 days from February 10 to June 14, and merged all the data from June 15 and later, to improve statistics of the spectra after the decay.

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Swift/XRT light curves in the bands 1.5–10 keV, 0.3–1.5 keV, and 0.3–10 keV, and the hardness ratio between 1.5–10 keV and 0.3–1.5 keV, from top to bottom. The black open and red filled squares represent the windowed-timing (WT) mode and the photon-counting (PC) mode data, respectively. They are binned in each snapshot before 4 × 105 s from the BAT detection (T = 0) and in each observation for later times. Errors in the abscissa and ordinate represent the start/end times of the bin from the mean photon arrival time, and the 1σ errors of the count rates or the ratios, respectively. MAXI/GSC data are also plotted with blue stars (see Section 2.2 for the conversion to XRT count rates).

Standard image High-resolution image

A time-averaged hard X-ray spectrum was also made from the BAT event-mode data taken at the first detection on February 5 (OBSID = 00954304000). The BAT data were downloaded from the HEASARC archive 17 and processed with the script batgrbproduct included in HEAsoft, referring to the Swift/BAT Calibration Database (CALDB) released on 2017 October 16. We adopted the default binning (binned in every 2 keV) to produce the spectrum.

2.2. MAXI

MAXI/GSC detected the source ∼20 minutes after the BAT detection and the subsequent two scans. We retrieved the GSC data in these individual scans via the on-demand system. 18 The GSC data are plotted in the top panel of Figure 1, together with the XRT data. They were converted to the XRT count rates at 1.5–10 keV through the HEASARC online tool WebPIMMS (Mukai 1993), 19 using the best-fit parameters of the absorbed power-law model obtained from the time-averaged spectra in the corresponding scans (see Section 3.3.1). As noticed in the figure, MAXI uniquely observed the source around the flux peak of Swift J0840.7−3516.

2.3. NICER

NICER started the first observation ∼8 hours after the BAT detection, and revisited the source every day or every few days until February 27, with exposure times of 0.4–10 ks. We used all the data taken before February 9, which have sufficient statistics for spectral and timing analysis. The NICER/XTI data were downloaded from the HEASARC archive. They were reprocessed with the pipeline tool nicerl2 based on the NICER CALDB version 2020 February 2, before producing light curves and time-averaged spectra of the individual OBSIDs. The response matrix file nixtiref20170601v001.rmf and ancillary response file nixtiaveonaxis20170601v003.arf were adopted in spectral analysis. The background generator nibackgen3C50 version 4 was utilized to produce background spectra for each observation. The estimated background level was found to be overestimated in some OBSIDs, as shown in Figure 2. To absorb the deviation, we adjusted the normalizations of the background spectra of the individual OBSIDs so that the count rates at 10–15 keV (where background counts are dominant) are equal between the actual data and the generated background spectra. The renormalization factors are within the range 0.9–1.4. We have found that this treatment only slightly changes the results of the spectral fits in the following sections (by < 15% for the flux and <5% for the other parameters) from those obtained by using the original background spectra and that this does not affect the conclusions of the paper.

Figure 2. Refer to the following caption and surrounding text.

Figure 2. Black filled circles: time-averaged, folded spectrum produced from the first NICER observation (OBSID = 2201010101); the background is not subtracted. Red open squares: background spectrum generated with nibackgen3C50. Blue crosses: the same background spectrum renormalized so that its 10–15 keV count rate is equal to that of the data.

Standard image High-resolution image

2.4. NuSTAR

The NuSTAR observation was performed from February 8 UT 07:18 to February 9 UT 06:46 with an exposure time in the normal observing mode of 42 ks (after standard screening to produce cleaned event data was applied). The NuSTAR data were retrieved from the HEASARC archive and reprocessed through nupipeline and the latest NuSTAR CALDB as of 2020 February 5. NuSTAR light curves, the time-averaged spectrum, and its response files were created with nuproducts. The source and background regions were defined as circular regions with a radius of 50'' centered on the target position and a radius of 80'' in a blank sky area on the same chip, respectively.

3. Analysis and Results

3.1. Long-term Light Curve

As shown in Figure 1, the source initially increased its intensity and reached the peak at T ∼ 103 s (where T is the time since the BAT trigger). It then exhibited a very rapid decay, decreasing its 1.5–10 keV intensity by ∼5 orders of magnitude in <106 s. In the decay, the source showed a plateau phase from T ∼ 3 × 104 s to T ∼ 1 × 105 s. In the early decline phase before the plateau, the hardness ratio, calculated by dividing the 1.5–10 keV count rate by the 0.3–1.5 keV count rate, rapidly decreased, indicating spectral softening. After the end of the decay at T ∼ 5 × 105 s, the source showed flat light curves in all energy ranges, with a non-periodic variation by a factor of ∼2 (see the third panel of Figure 1).

3.2. Short-term Variability

Figure 3 displays examples of short-term light curves at different periods. Strong variation on timescales of 1–103 s can be seen throughout the decay. The X-ray intensity varied by a factor of several to even more than an order of magnitude (see Figure 3(d)), although the amplitude became somewhat weaker around the end of the initial decline, T ∼ 3 × 104 s (Figure 3(c)). We note that the strong variability was also clearly seen in the MAXI scan around the flux peak (T ∼ 103 s).

Figure 3. Refer to the following caption and surrounding text.

Figure 3. Short-term light curves in different periods. Note that different energy ranges are adopted for Swift, NICER, and NuSTAR (see the labels on the ordinate axes). The background component is not subtracted in the NICER light curves. For clarity, the NICER light curves only cover the first continuous scans of the observations, while the other two include the entire periods of the observations.

Standard image High-resolution image

We searched for periodicity using the available Swift/XRT, NICER, and NuSTAR data, after applying barycentric correction for the target position (R.A., decl.) (J2000) = (08h40m40fs84, $-35^\circ 16^{\prime} 24\buildrel{\prime\prime}\over{.} 8$) with the ftool barycorr, based on the JPL planetary ephemeris DE-200. Adopting the ${Z}_{1}^{2}$ (Buccheri et al. 1983) and epoch-folding techniques, we found a peak at 8.96 s in the periodogram of the first Swift/XRT WT-mode observations (OBSID = 00954304000 and 00954304001; Figure 4), which is consistent with the report by Kennea et al. (2020). The chance probability of this 8.96 s period was < 10−10 based on the ${Z}_{1}^{2}$ method. No significant periodicity was detected in the other data. In Figure 5 we plot the light curve in Figure 3 folded by 8.96 s using the ftool efold.

Figure 4. Refer to the following caption and surrounding text.

Figure 4. Periodogram of Swift/XRT WT-mode data taken on 2020 February 5, with an exposure time of 765 s. Analyses using ${Z}_{1}^{2}$ search (Buccheri et al. 1983) and e-folding techniques are overplotted (black and gray, respectively). A clear peak is seen at P = 8.96 s (red dotted line). This represents the only dataset in which this period is detected.

Standard image High-resolution image
Figure 5. Refer to the following caption and surrounding text.

Figure 5. Light curves in Figures 3(a)–(d) folded by 8.96 s, normalized by their averaged count rates. The Swift/XRT data of OBSID = 00954304001 were added to (a) and the whole observation periods were used in (b)–(d).

Standard image High-resolution image

Using the light curves, we also searched for possible Type-I X-ray bursts, which are characterized by a short rise time of ≲10 s and a subsequent longer decay for a few tens of seconds or more (e.g., Lewin et al. 1993). However, no evidence of the X-ray bursts was found in any of the Swift, NICER, and NuSTAR data. Specifically, the flare seen in the NuSTAR light curve (Figure 3(d)) at ∼2.48 × 105 s from the BAT detection was found to have a ∼200 s rise time, by using a light curve with shorter time bins, and hence is unlikely to be a Type-I X-ray burst.

3.3. Analysis of the X-Ray Spectrum

3.3.1. Overall Spectral Evolution

Figure 6 presents representative soft X-ray spectra obtained with the Swift/XRT, MAXI/GSC, and NICER/XTI at different epochs. The spectrum was hardest at the initial, brightest phase (T ≲ 104 s), then rapidly softened, and finally became almost flat in the EFE form, where E and FE are the photon energy and the energy flux per unit energy, respectively. A hint of an absorption feature is seen at 3–4 keV in the Swift and NICER spectra taken between T = 2 × 104 and 5 × 104 s (blue and magenta spectra in the left panel of Figure 6), which we look into in Section 3.3.2.

Figure 6. Refer to the following caption and surrounding text.

Figure 6. Representative spectra on 2020 February 5 (left) and afterward (right), unfolded by using a single power-law model with a photon index of 2. Left: the data of the first Swift/XRT observation at T = (1.4–2.3) × 102 s are presented with black open squares. The others are the MAXI/GSC spectrum obtained at T = (2.3–2.5) × 103 s (orange), Swift/XRT spectra at (4.5–4.8) × 103 s (red), (1.22–1.24) × 104 s (green), and (1.8–1.9) × 104 s (blue), and the NICER/XTI spectrum at (2–5) × 104 s (magenta), from top to bottom. Right: Swift/XRT spectra obtained at T = (1.0–1.3) × 105 s (black open squares), (2.3–2.5) × 105 s (red), (2.7–3.0) × 105 s (blue), (4–37) × 105 s (green), and (8–11) × 106 s (magenta open circles). Note that some of the spectra were obtained by averaging multiple intermittent observations (see Figure 1).

Standard image High-resolution image

To quantify the overall spectral behavior, we applied an absorbed power-law model to each spectrum. Throughout the spectral analysis, we used the 0.5–10 keV, 2–20 keV, and 1.0–10 keV ranges of the Swift, MAXI and NICER data, respectively. The C-statistic (Cash 1979) was employed because many of the spectra have relatively low statistics. The TBabs model was adopted for the absorption component, with the solar abundance table given by Wilms et al. (2000). The equivalent hydrogen column density NH was allowed to vary in Swift and NICER data, while it was fixed in the MAXI data at 3 × 1021 cm−2, which is the Galactic column in the direction of Swift J0840.7−3516 estimated with the tool nh in HEAsoft. This is because MAXI is not sensitive to the soft X-rays below 2 keV and it is difficult to constrain NH. We note that the free parameters obtained from the MAXI spectra remained consistent within their 90% confidence ranges when NH = 0 and 1.0 × 1022 cm−2 were adopted. When NH is allowed to vary, it is not constrained at all and only the lower limits of Γ and unabsorbed fluxes were obtained.

The model successfully reproduced all the spectra. Figure 7 plots the individual fit parameters in chronological order. The spectrum until T ∼ 104 s had a small photon index of Γ = 1.0–1.2. The value increased in the following rapid decline, and became Γ ∼ 2.0 after the decay. During the decay, the column density was comparable with the total Galactic column, except for the plateau phase at T ∼ (3 × 104)–105 s, where NH increased up to ∼1 × 1022 cm−2. At this phase, Γ was estimated to be slightly larger, Γ ∼ 2–3. These behaviors may suggest variations of the intrinsic absorption, although it could be affected by the degeneracy with Γ. At late times of and after the decay, the NH values are consistent with or slightly smaller than the Galactic column, but have large uncertainties due to the low statistics.

Figure 7. Refer to the following caption and surrounding text.

Figure 7. Time variations in the best-fit parameters of the absorbed power-law model: (a) unabsorbed 0.5–10 keV flux, (b) NH, (c) Γ, (d) C-statistic/(degree of freedom). Swift, MAXI, and NICER data are shown by black circles, blue squares, magenta triangles, respectively. The red dashed line in panel (b) shows the total Galactic column (NH = 3 × 1021 cm−2).

Standard image High-resolution image

3.3.2. Absorption Structure

In Figure 6, a possible absorption feature was found at 3–4 keV in the Swift and NICER spectra obtained between T = 2 × 104 and 5 × 104 s. To describe the profile, we tested (1) a Gaussian line model (gauss, with a negative normalization) and (2) a simplified edge model (edge). The same absorbed power-law model as in Section 3.3.1 was used as the continuum model, where all the parameters were allowed to vary. We also adopted the same energy ranges of the spectra as those in that section.

The best-fit values of the gauss and edge components are listed in Table 1, and the Swift/XRT and NICER data and their best-fit models with edge are shown in Figure 8. The quality of fit for the Swift/XRT spectrum was significantly improved from the case of the continuum model, whereas the improvement was not significant in the NICER data. Using the script simftest on XSPEC, we estimated the chance probability of improvement by the additional edge component as 0.001 and 0.2 (corresponding ∼4σ and ∼1σ) for the Swift and NICER data, respectively. The two spectra gave consistent edge/line energies. The Fe K line and/or edge around 6–7 keV are most prominently seen in many X-ray sources including accreting black holes and neutron stars. Assuming the observed feature originated in the neutral Fe K-shell edge, the redshift estimated from the edge model is z ∼ 1.1 (see Section 4.2 for discussion).

Figure 8. Refer to the following caption and surrounding text.

Figure 8. Swift/XRT (left) and NICER (right) spectra and the best-fit TBabsedgepowerlaw model. The data versus model ratios for the models excluding and including edge are plotted in the middle and bottom panels, respectively.

Standard image High-resolution image

Table 1. Best-fit Parameters of Absorption Components

ParameterUnitSwift NICER
(1) TBabsedgepowerlaw
Eedge keV3.3 ± 0.1 ${3.4}_{-0.2}^{+0.6}$
τ  1.43 ± 0.7 ${0.52}_{-0.51}^{+0.78}$
C-stat/d.o.f 204/263316/326
 
(2) TBabs(gauss+powerlaw)
Ecen keV ${3.7}_{-0.1}^{+0.2}$ ${3.9}_{-0.2}^{+0.1}$
σ keV ${0.2}_{-0.1}^{+0.2}$ <0.36
Ngau a   ${5}_{-3}^{+2}\times {10}^{-4}$ ${1.4}_{-1.1}^{+1.7}\times {10}^{-5}$
EWkeV ${0.5}_{-0.2}^{+0.3}$ ${0.24}_{-0.18}^{+0.29}$
C-stat/d.o.f 206/262314/325
 
TBabspowerlaw (continuum)
C-stat/d.o.f 276/265319/328

Note.

a Absolute value of the normalization of gauss, defined as the total photon flux in the absorption line, in units of photons cm−2 s−1.

Download table as:  ASCIITypeset image

If the feature is the neutral Fe K absorption edge of cold gas, Fe K emission or absorption lines would also be produced, depending on its geometry and temperature. We tested adding a negative/positive Gaussian component at z = 1.1 with a line center energy of 6.4 keV to the absorbed power-law continuum model, to consider the neutral Fe Kα absorption/emission line. The 90% upper limits of the line strengths for the Swift data were estimated to be 4.0 × 10−4 and 3.3 × 10−4 photons cm−2 s−1 (∼1.2 × 102 eV and ∼97 eV in the equivalent width), for the positive and negative Gaussian cases, respectively.

We also applied, to the Swift spectrum, the neutral Compton reflection model pexmon (Nandra et al. 2007), which accounts for both the Fe K emission lines and edge self-consistently, to test the possibility of a reflection component like those seen in X-ray binaries and AGNs. We replaced the power-law component in the original continuum model by pexmon and allowed the reflection scale factor (rel_refl) to vary; the latter is the solid angle Ω of the reflector visible from the illuminating source, in units of 2π. The abundance of Fe and other elements was assumed to be the same as the solar abundance, and the redshift and the cutoff energy were fixed at 1.1 and 0 keV (i.e., the high-energy cutoff was not included). We tested two cases of inclination angles: i = 0° and 60°. In both cases, the model gave somewhat better fits than the absorbed power-law model, yielding C-stat/d.o.f. = 217/263. We obtained only a weak upper limit of the reflection factor, Ω/2π < 2.0 and 2.6, for i = 0° and 60°, respectively.

3.3.3. Analysis of Wideband Spectra

Hard X-ray spectra with good statistics were obtained from the Swift/BAT observation at an early phase (T ∼ 102 s) and the NuSTAR observation in the last part of decay (T ∼ 2 × 105 s). Combining them with the Swift/XRT and NICER data taken (quasi-)simultaneously, we obtained wideband spectra at T ∼ 2 × 102 s (hereafter SP1) and at T ∼ 2 × 105 s (SP2), as shown in Figures 9(a) and (f), respectively. The SP1 spectrum was produced from the Swift/XRT and BAT data of OBSID = 00954304000, and the SP2 spectrum was from the Swift/XRT, NICER, and NuSTAR data of OBSID = 00954304004, 2201010105, and 90601304002, respectively. The former spectrum has a clear high-energy cutoff at ∼30 keV, while the latter shows a flat profile extending to at least ∼20 keV. Here, we adopted χ2 statistics for the Swift/BAT spectrum, because it was already grouped by channels to have a Gaussian distribution and hence C-statistics should not be used. 20 For consistency, the χ2 statistics are applied to the Swift/XRT, NICER, and NuSTAR spectra as well. To ensure that the distribution underlying the data can be approximated by the Gaussian distribution, the Swift/XRT data of SP1 were binned so that at least 40 counts are included in each spectral bin. The Swift/XRT, NICER, and NuSTAR spectra in SP2 were binned to have at least 20, 20, and 40 counts per bin, respectively.

Figure 9. Refer to the following caption and surrounding text.

Figure 9. Wideband X-ray spectra obtained at T ∼ 2 × 102 s (left) and ∼2 × 105 s (right), fitted with four different models. The Swift, NICER, and NuSTAR data are plotted in black, blue, and red, respectively. The best-fit TBabspcfabspowerlaw model and TBabs(bbodyrad+powerlaw) model are adopted in panels (a) and (f), respectively. The dashed line in panel (f) presents the contribution of bbodyrad. The ratios of data to model for TBabs*powerlaw ((b) and (g)), pcfabsTBabspowerlaw ((c) and (h)), TBabs*(bbodyrad+powerlaw) ((d) and (i)), and TBabspcfabspowerlaw ((e) and (j)) are also plotted.

Standard image High-resolution image

We first fit the two spectra with an absorbed cutoff power-law model: TBabscutoffpl in the XSPEC terminology. The ratios of data to model are shown in Figures 9(b) and (g) and the best-fit parameters are listed in Table 2. Although the model fit both spectra fairly well, small residual structures remained. In Figure 9(b) the discrepancy between the data and model can be seen at the lowest energies, and in Figure 9(g), a small hump at 0.8–2 keV and an excess above ∼10 keV are present.

Table 2. Best-fit Parameters for the Wideband Spectra

ID NH Γ Ecut Na kTin, kTbb Rin, Rbb b ${N}_{{\rm{H}}}^{\mathrm{pcf}}$ Cv c FX χ2/dof
 (1021 cm−2) (keV) (keV)(km)(1021 cm−2) (erg s−1 cm−2) 
TBabscutoffpl
SP12.4 ± 0.10.8 ± 0.1 ${24}_{-5}^{+7}$ 0.13 ± 0.029 × 10−9 119/99
SP22.5 ± 0.31.98 ± 0.09>196(7 ± 1) × 10−4 5 × 10−12 204/172
TBabspcfabscutoffpl
SP1<1.8 ${1.2}_{-0.1}^{+0.3}$ ${36}_{-9}^{+30}$ ${0.24}_{-0.04}^{+0.12}$ ${17}_{-5}^{+13}$ ${0.7}_{-0.15}^{+0.04}$ 8 × 10−9 96/97
SP22.6 ± 0.22.0 ± 0.1>193(7 ± 1) × 10−4 <0.25 e 5 × 10−12 203/170
TBabs(bbodyrad+cutoffpl)
SP16.8 ± 0.21.0 ± 0.2 ${30}_{-8}^{+14}$ ${0.19}_{-0.03}^{+0.04}$ ${0.07}_{-0.01}^{+0.02}$ ${6}_{-4}^{+12}\times {10}^{3}$ 8 × 10−9 97/97
SP22.1 ± 0.5 ${1.7}_{-0.2}^{+0.1}$ >49(4 ± 1) × 10−4 0.33 ± 0.05 ${2.0}_{-0.5}^{+0.8}$ 6 × 10−12 150/170
TBabs(diskbb+cutoffpl)
SP16.7 ± 0.21.0 ± 0.2 ${31}_{-8}^{14}$ ${0.19}_{-0.03}^{+0.04}$ 0.08 ± 0.02 ${8}_{-6}^{+19}\times {10}^{3}$ 8 × 10−9 97/97
SP2 ${2.6}_{-0.4}^{+0.5}$ ${1.7}_{-0.3}^{+0.1}$ >33(3 ± 1) × 10−4 0.5 ± 0.1 ${1.4}_{-0.4}^{+0.7}$ 6 × 10−12 150/170

Notes.

a Normalization of cutoffpl, defined as the photon flux in units of photons keV−1 cm−2 s−1 at 1 keV. b D = 8 kpc and i = 60° (for diskbb) were assumed. c Covering fraction of pcfabs. d Unabsorbed 1–100 keV flux. e Not constrained.

Download table as:  ASCIITypeset image

Next we investigated whether an additional, optically thick thermal component or partial covering absorption can better reproduce the soft X-ray profiles. We adopted pcfabs as the partial covering absorption model at z = 0. For the thermal component, we tested a blackbody component (bbodyrad) and a disk blackbody component (diskbb; Mitsuda et al. 1984). In SP1, these models significantly improved the quality of fit (Figures 9(c), (d), and (e)), and in any of the models the χ2 values decreased by ∼20 with the degrees of freedom reduced by 2 (Table 2) from the simple absorbed cutoff power-law model. The fit quality of SP2 was not improved by adding the pcfabs model (Figure 9(h)), but an additional thermal component (Figures 9(i) and (j)) reduced the χ2 value by ∼50. We obtained a low temperature and a large radius of (disk) blackbody in SP1, and a slightly higher temperature and a much smaller radius in SP2.

3.4. Profile of Flux Decay

To investigate the profile of the long-term flux decay, we analyzed the flux light curve in Figure 7(a), obtained in the analysis of the soft X-ray spectra based on the absorbed power-law model (see Section 3.3.1). First, to roughly characterize the decay profile, we fitted the data points with a single power-law function in terms of the time T from the BAT detection:

Equation (1)

where FX is the unabsorbed 0.5–10 keV flux and Tpeak is the time at the flux peak, which we assumed to be the second scan of MAXI at T ≈ 2 × 103 s. Here, we only considered the decay period from T ≈ 1 × 103 s to 2 × 104 s and ignored the initial flux rise, the final constant-flux phase, and the plateau phase at T ≈ (2–5) × 104 s during the decay. A decay index of α ≈ 1.7 was found to approximately reproduce the observed decay.

Next, we applied a multiply broken power-law model to the flux data over the entire period, to better describe the decay profile and estimate the total radiated energy accurately. We found that the flux profile can be well reproduced by using eight power-law segments with photon indices of α1, α2, ..., α8, and with seven breaks: Tbr1, Tbr2, ..., Tbr7, from earlier to later times. We described the initial flux rise until the second MAXI scan by two power-law segments. Because only three data points were available in this period, we fixed α1 and α2 at the values calculated from them and assumed Tbr1 and Tbr2 as the times of first and second MAXI scans. We note that this complex model is not physically motivated and is used just to obtain an accurate measurement of the fluence.

Figure 10 plots the light-curve data and the best-fit model and Table 3 gives the best-fit parameters. We also estimated the X-ray fluence at 0.5–10 keV for the total outburst period as F0.5–10 keV = 2.9 × 10−5 erg cm–2, by integrating the best-fit multiply broken power-law model in Figure 10 over the entire period from T = 0 s to 3 × 107 s.

Figure 10. Refer to the following caption and surrounding text.

Figure 10. Flux light curve (same as Figure 7(a)) and the best-fit multiply broken power-law model. The same data colors as Figure 7(a) are adopted. The dashed line indicates a single power law with an index of α = 1.7 in terms of the time from the flux peak (see Equation (1)).

Standard image High-resolution image

Table 3. Best-fit Parameters of the Multiply Broken Power-law Model for the Long-term Light Curve

ParameterUnitValue
α1  0.06 (fixed)
Tbr1 s1.1 × 102 (fixed)
α2  −1.2 (fixed)
Tbr2 s2.5 × 103 (fixed)
α3  1.4 ± 0.1
Tbr3 s(1.0 ± 0.1) × 104
α4   ${5.1}_{-0.7}^{+0.8}$
Tbr4 s ${2.9}_{-0.2}^{+0.3}\times {10}^{4}$
α5  −0.6 ± 0.1
Tbr5 s(1.30 ± 0.05) × 105
α6  1.8 ± 0.2
Tbr6 s >3.8 × 105
α7  >6
Tbr7 s ${4.8}_{-0.3}^{+0.5}\times {10}^{5}$
α8   ${0.1}_{-0.1}^{+0.2}$
F0.5–10 keV a erg cm−2 2.9 × 10−5

Notes.

a 0.5–10 keV fluence in the period T = 0–3 × 107 s.

Download table as:  ASCIITypeset image

In Section 3.3.3, we found that including a partial covering absorption component give a better fit to the broadband X-ray spectrum in the early phase. To investigate whether its inclusion affects the flux light curve, we tested adding pcfabs to the TBabspowerlaw model and fitting the soft X-ray spectra in Section 3.3.1. The TBabspcfabspowerlaw model improved the fits of the spectra obtained by the end of the first rapid decay (T < 105 s) but not in the later period. The unabsorbed 0.5–10 keV fluxes were found to change by only ≲10% from the TBabs*powerlaw model, and this does not affect the overall light-curve profile in Figure 10.

4. Discussion

4.1. Summary of X-Ray Properties

We have studied the X-ray properties of Swift J0840.7−3516 using Swift, MAXI, NICER, and NuSTAR data. Here we summarize the results, to discuss the nature of the source in the following sections.

  • 1.  
    A rapid decay of the outburst by ∼5 orders of magnitude within ∼5 days was observed. After the decay, the source flux was almost constant over nine months (Figure 1).
  • 2.  
    The overall decay profile can be approximated with a function ∝ t−1.7, where t is the time from the flux peak. More precisely, the profile is better described with a multiply broken power-law model (Figure 10 and Table 3). The total X-ray fluence until 3 × 107 s was estimated to be ∼3 × 10−5 erg cm−2.
  • 3.  
    We detected strong short-term variations on timescales of 1–103 s during the decay (Figure 3). A possible periodicity of 8.96 s was detected in the first Swift/XRT observation, but not in the other observations (Figures 4 and 5).
  • 4.  
    The soft X-ray spectrum was well described with an absorbed power-law model. The source showed the hardest spectrum with a photon index of ∼1, which gradually increased with decreasing flux and finally reached Γ ∼ 2 after the decay (Figure 7). The absorption column density was comparable to the total Galactic column, although some variations were present.
  • 5.  
    A clear spectral rollover was observed at ∼30 keV at the initial rise of the outburst. The wideband X-ray spectrum in this phase was better described by adding a partial covering absorption or a (disk) blackbody component to an absorbed cutoff power-law model. An additional thermal component also improved the fit of the wideband spectrum taken in the last period of the decay (Figure 9 and Table 2).
  • 6.  
    A possible absorption feature at ∼3.4 keV was detected between T ∼ 2 × 104 and 5 × 104 s, when the light curve exhibited a plateau following the first steep decay (Figures 6, 8, and Table 1). Assuming it to be the neutral iron K edge, the redshift is estimated to be z = 1.1.

4.2. Possible Nature of the Source

4.2.1. A Galactic Origin

Let us first investigate the possibility of a Galactic transient. Given that the source is located near the Galactic plane, it would be reasonable to take this possibility into account. Because the absorption column density is comparable to the total Galactic column for a large fraction of the outburst period, the source would not be very near. Assuming isotropic emission, the maximum and minimum luminosities at 0.5–10 keV are estimated as 2 × 1037(D/8 kpc)2 erg s−1 and 7 × 1032(D/8 kpc)2 erg s−1, respectively. The large luminosity range and the short timescale of the decay, and the relatively faint optical counterpart (with an apparent $r^{\prime} $ band magnitude of ∼20 mag; Mazaeva & IKI-FuN follow-up Collaboration 2020), rule out the possibility of a high-mass X-ray binary (e.g., Walter et al. 2015). Also the timescale of the X-ray decay is much shorter than magnetar outbursts (∼100 days; Coti Zelati et al. 2018), although a couple of them showed two-step decays reminiscent of that seen in our target. The observed hard, nonthermal spectrum without any prominent emission lines also makes unlikely the possibilities of a stellar flare (Güdel & Nazé 2009) and a cataclysmic variable (Mukai 2017).

We have confirmed the 8.96 s peak reported by Kennea et al. (2020) in the periodogram made with the first Swift/XRT observation. Given the fact that it did not appear in any other observations, the variation was either a transient phenomenon of the source or an artifact (such as red noise). If the former is the case, it is difficult, due to the limited exposure, to determine whether it is a pulsation or a quasi-periodic oscillation (QPO). Hence, the possibilities of an X-ray binary pulsar and a magnetar are neither proved nor ruled out.

Typical black hole (BH) and neutron star (NS) low-mass X-ray binaries (LMXBs) also make it difficult to explain some of the observed properties of Swift J0840.7−3516. The luminosity range of the decay is consistent with LMXBs, and the upper limit of the radio flux, 18 μJy at 7.5 Hz, obtained five days after the discovery (Borghese et al. 2020), also agrees with the radio/X-ray flux correlations of both BH and NS LMXBs (e.g., Corbel et al. 2013). However, the observed decay period of ∼5 days is much shorter than those of typical LMXBs: a few tens of days to more than a year. This means that Swift J0840.7−3516 accreted a much smaller mass than normal LMXBs in an unusually short period. Moreover, the spectrum in the rise of the outburst had a somewhat small photon index, Γ ∼ 1.0, compared with typical LMXBs in the hard state, Γ ∼ 1.5–1.9 (e.g., Done et al. 2007).

Yet these unusual properties cannot completely rule out an LMXB nature. There are actually some LMXBs that show similar properties. An LMXB candidate showing such a short outburst with a similar luminosity range has recently been found (MAXI J1957+032; Beri et al. 2019). There are also several known LMXBs with short outbursts but smaller peak luminosities of 1034–1036 erg s−1, which are called "very faint X-ray transient" (Wijnands et al. 2006; Degenaar & Wijnands 2010; Bahramian et al. 2021). In addition, very hard spectra in a bright hard state have been observed in a couple of NS LMXBs (Parikh et al. 2017) and BH LMXBs including V 404 Cyg (e.g., Kimura et al. 2016), V4641 Sgr (e.g., Maitra & Bailyn 2006), and Swift J1858.6−0814 (Hare et al. 2020). All these sources have a strong short-term flux variation by a more than order of magnitude (e.g., Kitamoto et al. 1989; Revnivtsev et al. 2002; Wijnands et al. 2017), also in agreement with the present case. In addition, partial covering absorption is sometimes required in these sources, especially in high-flux phases (Kimura et al. 2016), consistent with Swift J0840.7−3516 in a bright phase.

We note that the photon index of Swift J0840.7−3516 after the decay, ≲2, is reminiscent of a BH rather than an NS LMXB, which usually shows a larger value, Γ ∼ 3 (Wijnands et al. 2015), although the NS LMXB EXO 1745−248 showed a similar hard spectrum in quiescence (Rivera Sandoval et al. 2018). On the other hand, an NS LMXB would rather be favored, in that a very small radius of the thermal component, 1–2(D/8 kpc) km, consistent with emission from an NS surface, was obtained from the NuSTAR+Swift/XRT+NICER spectrum in a faint phase. This holds even if the observed power-law component was entirely produced by Comptonized disk photons. Assuming conservation of the number of photons and an isotropic scattered emission (Kubota & Makishima 2004), and using the intrinsic disk photons estimated from the best-fit diskbb+cutoffpl model, we obtain the inner disk radius as ∼3(D/8 kpc) km (where i = 60° is assumed). Similar thermal emission likely emitted by the NS surface can be seen in NS LMXBs at low X-ray luminosities below ∼1035 erg s−1 but not in a more luminous hard state, where the power-law component dominates the X-ray flux (e.g., Sakurai et al. 2014; Shidatsu et al. 2017). The thermal component in Swift J0840.7−3516 was observed at a consistent luminosity; the second rapid decay from 105 s since the discovery, in which this component was found, started at ∼1035(D/8 kpc)2 erg s−1.

Even if an LMXB scenario works for all the properties raised above, the observed absorption feature is difficult to explain. Although ionized Si and S lines from disk winds are sometimes seen at similar energies (e.g., Ueda et al. 2009), they are much weaker than the present case and always with strong iron Kα lines at ∼7 keV. Assuming the observed feature to be the redshifted Fe K edge, we obtained z = 1.1. If this is the gravitational redshift caused by a Schwarzschild stellar-mass BH, the feature should be produced at ∼1.3 RS, where RS is the Schwarzschild radius. It is unclear that an absorber with such a small size can exist in the vicinity of a BH. Another possibility is that the observed feature is a different blueshifted line produced by an outflow at a relativistic speed, like baryonic jets observed in SS 433. However, it is unclear whether such baryonic jets can exist at a low Eddington ratio and produce a single absorption feature.

Instead, the origin of the source might be explained by a tidal disruption of an asteroid by a neutron star (Newman & Cox 1980; Colgate & Petschek 1981), which was considered for the origin of GRBs 30–40 years ago. In this model, the collision of an asteroid and an NS surface produces strong MeV gamma rays and hot plasma. The plasma could be rapidly cooled by radiation and cause Compton downscattering, resulting in a power-law-shaped X-ray spectrum. The radiation energy of Swift J0840.7−3516 released in 107 s is estimated as 2 × 1039(D/8 kpc)2 erg, which is converted to a total accreted mass of ∼1020(η/0.01)(D/8 kpc)2 g (where η is the radiative efficiency), consistent with an asteroid. It is unclear, however, whether or not this scenario can explain the observed evolution of the photon index of the X-ray spectrum and the absorption feature.

4.2.2. An Extragalactic Origin

We next investigate the possibility of an extragalactic origin. If the source is an extragalactic X-ray binary located in a nearby galaxy, the maximum luminosity is far beyond the Eddington luminosity of a BH with a mass of 10 M. Even if the distance is as small as ∼1 Mpc, the peak luminosity is calculated as ≳ 1041 erg s−1. Such a bright X-ray binary is categorized as a hyperluminous X-ray source and considered to have an intermediate-mass BH (Farrell et al. 2009). In this case, it is even more difficult than in the case of a normal X-ray binary with a stellar-mass BH to explain the short duration of the outburst, because of the large size of the accretion disk. For the same reason, an AGN origin is unlikely.

Some of the observed properties of Swift J0840.7−3516 agree with those of GRBs. Absorption features have been detected in a few GRBs (Amati et al. 2000; Frontera et al. 2004; Bellm et al. 2014) and interpreted as the redshifted iron K line or edge, although they were detected mainly in the prompt phase, unlike the present case (but see Bellm et al. 2014). The absorption feature in our case was observed just after the first rapid decay, which may be considered to be the beginning of the afterglow, and could be explained by the interaction between the jets and the surrounding gas. If the estimated redshift z = 1.1 is the cosmological redshift, it is converted to a luminosity distance of 7 Gpc, assuming H0 = 70 km s−1 Mpc−1, ΩM = 0.3, and ΩΛ = 0.7. In this case, the maximum and minimum luminosities are calculated to be 1 × 1049 erg s−1 and 4 × 1044 erg s−1, respectively, where isotropic emission is assumed. The redshift and maximum luminosity are well within the distribution of known GRBs. The source showed a hard spectrum with Γ ∼ 1 and a spectral rollover at ∼30 keV in the early phase, which are compatible with soft GRBs in the prompt phase (e.g., Yonetoku et al. 2010). Using the best-fit cutoff power-law model of the Swift/XRT+BAT spectrum (see Section 3.3.3), we obtain the 2–10 keV luminosity of 5 × 1048 erg s−1 in the rest frame of the source. This is consistent with the correlation between the peak energy and X-ray luminosity of Swift-detected GRBs reported by D'Avanzo et al. (2012).

The decay profile for the first 106 s and the spectral evolution of Swift J0840.7−3516 also share similar properties with GRBs. The two-step decay could be attributed to the prompt and afterglow emission. However, the observed plateau phase started later and lasted an order of magnitude longer than the typical shallow decay phase of GRBs (Zhang et al. 2006). Moreover, the constant flux over nine months after the decay is very different from, and difficult to explain by, X-ray afterglows of GRBs.

Instead, the source may be interpreted as a TDE by a supermassive BH. The overall decay profile is approximated to t−1.7, which is consistent with what is theoretically expected for a TDE (t−5/3; Rees 1988; Phinney 1989) and the actual X-ray light curves of observed TDEs (e.g., Burrows et al. 2011; Cenko et al. 2012). The strong short-term variability is reminiscent of the jetted TDE Swift J164449.3+573451 (Burrows et al. 2011; Mangano et al. 2016, hereafter Swift J1644). X-ray absorption features have also been detected in a TDE at lower energies than that in the present case (Miller et al. 2015), and were interpreted as an outflow at a velocity of a few hundred km s−1. The total radiation energy at 1–20 keV is calculated from the fluence to be 9 × 1052 erg in the rest frame of Swift J0840.7−3516, assuming isotropic emission and a luminosity distance of 7 Gpc. This corresponds to a total accreted mass of ∼0.9 M, assuming the radiative efficiency of the standard accretion disk of a Schwarzschild BH. The peak luminosity, ∼1049 erg s−1, is more than one order of magnitude larger than that of the most luminous TDE, Swift J1644.

Swift J1644 exhibited a steep flux drop ∼500 days after the discovery and an almost constant flux for the following ∼1000 days, which can be explained by the end of accretion and scattering of X-rays by the surrounding gas, respectively (Cheng et al. 2016). Similar flux behaviors were observed in Swift J0840.7−3516, with a rapid decay until T ∼ 5 × 105 s and a subsequent constant-flux phase. If the end of the decay at ∼5 × 105 s actually reflects the end of accretion, the BH mass (MBH) of Swift J0840.7−3516 can be estimated by comparing the decay times, or the accretion periods, of the two sources, as described below. We note that the scattering echo scenario for the constant-flux phase requires spectral softening, because of the energy dependence of the scattering angle. However, no significant softening was observed and therefore a different explanation would be needed in the present case.

When a star is tidally disrupted by a BH, some of the disrupted matter orbits around it and falls onto the BH (Rees 1988). The kinetic energy of the disrupted matter in a unit mass is described as

Equation (2)

where a and G are the semimajor axis and the gravitational constant, respectively. Using Kepler's third law, $T\sim \sqrt{{a}^{3}/{{GM}}_{\mathrm{BH}}}$, we obtain

Equation (3)

Thus, the accretion time is proportional to MBH.

Using this relation, the BH mass of Swift J0840.7−3516 should be ∼102 times smaller than that of Swift J1644, which was previously estimated as 106–107 M (Burrows et al. 2011). In this case, we obtain a BH mass of Swift J0840.7−3516 of 104–105 M, and therefore we may have witnessed a rare case of a TDE in a dwarf galaxy. The smaller BH mass and higher peak luminosity than those of Swift J1644 imply that Swift J0840.7−3516 is a more highly beamed, jetted TDE. This is of course just a rough estimate, and different approaches such as observations of the host galaxy would be required for a more accurate measurement.

A QPO at 5 mHz was detected in Swift J1644 and associated with the Keplerian frequency at the innermost stable circular orbit of the BH (Reis et al. 2012). If the observed 8.96 s variation in Swift J0840.7−3516 is explained by a QPO produced in the same manner, its BH mass is calculated to be 104–105 M, depending on the BH spin. This is consistent with the mass estimated above from the decay period.

5. Summary and Future Prospects

Using X-ray data from MAXI, Swift, NICER, and NuSTAR, we have found unusual X-ray properties of Swift J0840.7−3516 that are difficult to explain by any known class of objects. Plausible interpretations may be an LMXB with multiple extreme properties, or the tidal disruption of a star by a supermassive BH or of an asteroid by a Galactic NS. In any case, this source has provided an opportunity to study an interesting and perhaps rare astrophysical phenomenon around a compact object.

To uncover the true nature of the source, observations of the optical/infrared counterpart are essential. So far, only a few results from photometric observations have been reported, all of which were performed in the first couple of days after the discovery. Future optical imaging with high sensitivity and high spatial resolution, using telescopes of the largest class, could provide morphology information on the optical counterpart and determine its host galaxy if it is an extragalactic object. In addition, high-sensitivity spectroscopy would enable an accurate determination of the redshift of the source.

This work made use of Swift data supplied by the UK Swift Science Data Centre at the University of Leicester, and MAXI data provided by RIKEN, JAXA, and the MAXI team. We thank the NuSTAR team for performing the ToO observation. M.S. thanks Yuichi Terashima, Tohru Nagao, and Kazuma Joh for discussion of optical and infrared counterparts. Part of this work was financially supported by Grants-in-Aid for Scientific Research 19K14762 (M.S.) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan. P.A.E. acknowledges UKSA support. J.K. was supported by NASA grant 80NSSC20K1103.

Facilities: MAXI (GSC) - , Swift (BAT and XRT) - , NICER (XTI) - , NuSTAR (FPMA and FPMB). -

Software: XSPEC (v12.9.0n; Arnaud 1996, HEAsoft v6.26.1; HEASARC 2014).

Footnotes

Please wait… references are loading.
10.3847/1538-4357/abe6a1