You are on page 1of 14

The Astrophysical Journal, 954:138 (14pp), 2023 September 10 https://doi.org/10.

3847/1538-4357/ace90c
© 2023. The Author(s). Published by the American Astronomical Society.

Constraining the White-dwarf Mass and Magnetic Field Strength of a New Intermediate
Polar through X-Ray Observations
Benjamin Vermette1 , Ciro Salcedo1 , Kaya Mori1 , Julian Gerber1 , Kyung Duk Yoon1 , Gabriel Bridges1 ,
1
Charles J. Hailey
1
, Frank Haberl2 , Jaesub Hong3 , Jonathan Grindlay3 , Gabriele Ponti4 , and Gavin Ramsay5
Columbia Astrophysics Laboratory, Columbia University, New York, NY 10027, USA; cgs2155@columbia.edu
2
Max-Planck-Institut für extraterrestrische Physik, Giessenbachstrasse 1, D-85748 Garching, Germany
3
Harvard-Smithsonian Center for Astrophysics, Harvard University, Cambridge, MA 02138, USA
4
INAF-Osservatorio Astronomico di Brera, Via E. Bianchi 46, I-23807 Merate (LC), Italy
5
Armagh Observatory and Planetarium, College Hill, Armagh, BT61 9DG, UK
Received 2022 May 4; revised 2023 July 14; accepted 2023 July 17; published 2023 August 31

Abstract
We report timing and broadband spectral analysis of a Galactic X-ray source, CXOGBS J174517.0−321356 (J1745),
with a 614 s periodicity. Chandra discovered the source in the direction of the Galactic Bulge. Gong proposed that
J1745 was either an intermediate polar (IP) with a mass of ∼1 Me, or an ultracompact X-ray binary (UCXB). To
confirm J1745ʼs nature, we jointly fit XMM-Newton and NuSTAR spectra, ruling out a UCXB origin. We have
developed a physically realistic model that considers a finite magnetosphere radius, X-ray absorption from the preshock
region, and reflection from the white-dwarf (WD) surface to properly determine the IP properties, especially its WD
mass. To assess systematic errors on the WD mass measurement, we consider a broad range of specific accretion rates
(m = 0.6–44 g cm−2 s−1) based on the uncertain source distance (d = 3–8 kpc) and fractional accretion area
( f = 0.001–0.025). Our model properly implements the fitted accretion column height in the X-ray reflection model and
accounts for the underestimated mass accretion rate due to the (unobserved) soft X-ray blackbody and cyclotron cooling
emissions. We found that the lowest accretion rate of m = 0.6 g cm−2 s−1, which corresponds to the nearest source
distance and maximum f value, yields a WD mass of (0.92 ± 0.08)Me. On the other hand, as long as the accretion rate
is m  3 g cm−2 s−1, the WD mass is robustly measured to be (0.81 ± 0.06)Me, nearly independent of m . The derived
WD mass range is consistent with the mean WD mass of nearby IPs. Assuming spin equilibrium between the WD and
accretion disk, we constrained the WD magnetic field to B  7 MG, indicating that it could be a highly magnetized IP.
Our analysis presents the most comprehensive methodology for constraining the WD mass and B field of an IP by
consolidating the effects of cyclotron cooling, finite magnetospheric radius, and accretion column height.
Unified Astronomy Thesaurus concepts: X-ray binary stars (1811); Cataclysmic variable stars (203); DQ Herculis
stars (407)

1. Introduction mCVs with stronger WD magnetic fields (B ∼ 10–240 MG)


whose interactions with companions cause the synchronization
Cataclysmic variables (CVs) are interacting white-dwarf
of their orbital and spin periods. Polars typically have softer
(WD) binary systems powered by accretion via Roche-lobe
X-ray spectra (kT ∼ 5–20 keV) than IPs due to faster cyclotron
overflow from a late-type main-sequence companion star.
cooling (Mukai 2017). The CV population in the solar
Thanks to the extensive optical and X-ray surveys of CVs, the
neighborhood (d < 150 pc) was recently determined from an
last 50 yr have seen enormous strides in our understanding of
unbiased, volume-limited sample based on the Gaia optical
the accretion physics, binary formation, and evolution of CVs.
survey, finding a surprisingly higher fraction of mCVs (36%)
CVs are the most common type of interacting compact binaries
and are also recognized as one of the candidate progenitor than previously predicted by conventional CV formation
classes for type Ia supernovae and gravitational wave sources models (Pala et al. 2020).
In the Galactic Center (GC), Bulge, and Ridge regions at
detectable by the Laser Interferometer Space Antenna in the
d ∼ 1–8 kpc, the optical/UV emission is extinct by significant
mHz frequency band (Hillebrandt & Niemeyer 2000; Meliani
absorption and dust scattering, making the X-ray observations
et al. 2000). The majority of hard X-ray emission in our galaxy
is represented by magnetic CVs (mCVs) of two subclasses. In all the more important. Dedicated hard X-ray surveys by the
mCVs, infalling material is funneled onto the WD poles along Rossi X-ray Timing Explorer, the International Gamma-Ray
magnetic field lines, heated at a standoff shock, and emits Astrophysics Laboratory (INTEGRAL), Suzaku, and Nuclear
thermal X-rays. Intermediate polars (IPs) are a type of mCV Spectroscopic Telescope Array (NuSTAR) revealed that the
that have nonsynchronized orbits and WD magnetic fields diffuse X-ray emission (which mostly represents an unresolved
(B ∼ 0.1–10 MG) strong enough to truncate the inner accretion population of CVs) from the GC, Bulge, and Ridge is distinct
disk; they are copious emitters of hard X-rays in each region. The NuSTAR discovery of the central hard
(kT ∼ 20–40 keV; Mukai 2017). On the other hand, polars are X-ray emission (CHXE with kT ∼ 35 keV; Perez et al. 2015),
which spatially coincides with the nuclear star cluster (NSC) in
the central 10 pc (Schödel et al. 2014), suggests that the CV
Original content from this work may be used under the terms
of the Creative Commons Attribution 4.0 licence. Any further population is dominated by IPs with mean WD mass
distribution of this work must maintain attribution to the author(s) and the title 〈M〉 ∼ 0.9Me (Hailey et al. 2016). Outside the CHXE/NSC
of the work, journal citation and DOI. region, the plasma temperatures of the Bulge and Ridge

1
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

measured by INTEGRAL, Suzaku, and NuSTAR (kT ∼ analysis results and the lack of an IR counterpart (Section 5),
8–15 keV; Türler et al. 2010; Yuasa et al. 2012; Perez et al. we identified that J1745−3213 is a new IP in the direction of
2019) are in stark contrast to the ones of the CHXE the Bulge (Section 6.1). By applying a new X-ray spectral
(kT ∼ 35 keV; Perez et al. 2015). While the CHXE is largely model developed for mCVs to the XMM-Newton + NuSTAR
characterized by IPs, the diffuse X-ray emission in the Bulge spectra, we determined the WD mass of J1745−3213
and Ridge may consist of a different CV population dominated (Section 6.2). In Section 7, we summarize our results and
by polars or nonmagnetic CVs (Hailey et al. 2016; Perez et al. discuss the significance of constraining WD masses of IPs
2019; Xu et al. 2019b). This is also supported by an Fe line using broadband X-ray spectral and timing data.
diagnostics study using Suzaku data of the Ridge X-ray
emission (Nobukawa et al. 2016; Xu et al. 2016). The different 2. X-Ray Observations and Data Reduction
but unresolved CV populations in these regions may have
emerged due to differences in star-forming history, binary J1745−3213 was discovered by Chandra (Jonker et al. 2014)
evolution (e.g., more interactions with stars in the NSC), or and was serendipitously observed by XMM-Newton in 2010
metallicities. However, it is nearly impossible to deconvolve while it targeted a nearby black-hole binary, H1743-322. J1745
the diffuse X-ray emission into different CV populations (with −3213 was not in the field of view of the 2010 XMM-Newton
varying LX and kT distributions) through analysis of diffuse European Photon Imaging Camera (EPIC)-PN detector. In the
hard X-ray continuum or Fe line emission only. metal oxide semiconductor (MOS) data, the point source
The only unambiguous way to explore CV populations in the appeared elongated due to the large off-axis angle; we thus
GC, Bulge, and Ridge is by analyzing the detected X-ray point used an elliptical region with semimajor and semiminor axes of
sources and determining their source types by analysis of their 40″ × 30″ for source extraction. The background region was
plasma temperatures or Fe emission lines. For example, in the extracted using an elliptical annulus with an inner ring the size
GC region, which has been extensively observed in the X-ray of the source extraction region and an outer ring with
band (Wang et al. 2002; Muno et al. 2009), Chandra ACIS semimajor and semiminor axes of 60″ × 45″, respectively.
analysis of X-ray point sources investigated the CV population To identify the source type, follow-up XMM-Newton (obser-
by confirming that low-luminosity IPs (LX  1033 erg s−1) ving between 0.2 and 15 keV) and NuSTAR (observing between
make up a large fraction of the CHXE (Xu et al. 2019a). 3 and 79 keV) observations were performed in 2021. For all of
However, in the Bulge region outside the GC region (r ∼ 100 the observations, the source extraction regions were centered
pc from Sgr A*), X-ray sources and their populations are poorly around the Chandra position (R.A. = 17:45:17.0 ± 0.11, decl. =
−32:13:56.5 ± 0.11; J2000). We reduced all XMM-Newton
understood, partially due to the lack of extensive X-ray surveys
EPIC data using SAS 19.1 and extracted source events using
and/or follow-up of deeper X-ray observations. The situation
evselect after applying GTI filters generated by emchain for MOS
will be improved by the ongoing Swift X-Ray Telescope,
and epchain for PN (Gabriel et al. 2004). The second XMM-
Chandra/XMM-Newton, and eROSITA X-ray surveys of the
Newton observation targeted J1745−3213 with 31 ks of exposure,
Bulge region (Bahramian et al. 2021; Mondal et al. 2022), and
and a circular region with a radius of 20″ for both MOS and PN
by follow-up observations. Indeed, follow-up NuSTAR and
was adopted for extracting source events. The background was
Chandra observations of unidentified INTEGRAL sources in extracted using a circular annulus of inner and outer radii of 20″
the Galactic disk detected a handful of IPs (Coughenour et al. and 30″, respectively.
2022). These INTEGRAL source investigations, as well as the NuSTAR observed J1745−3213 in 2021 March for 49 ks
NuSTAR legacy observations of 25 IPs (Shaw et al. 2020), exposure. All data were processed using the standard NuSTAR
have demonstrated the significant role of NuSTAR’s broadband processing procedures (Harrison et al. 2013). A r = 50″ circular
X-ray spectroscopy data for identifying mCVs and their WD region around the Chandra source position was used for
masses. extracting source events for both timing and spectral analysis.
Here, we report XMM-Newton and NuSTAR observations A circular annulus of inner and outer radii of 50″ and 80″ was
of a new IP, CXOGBS J174517.0-321356 (J1745−3213 used to extract the background. In the NuSTAR observations,
hereafter), located 3°. 2 away from the GC. J1745−3213 was the stray-light background contamination from H1743−322
discovered in the Galactic Bulge Survey with a high degree of did not overlap with our source nor with background extraction
X-ray absorption indicating a distance greater than 3 kpc regions.
(Jonker et al. 2014). The archived XMM-Newton observations
in 2010 (originally targeting a black-hole binary H1743−322)
showed a periodicity at P = 614 s (Gong 2020), which was 3. Timing Analysis
confirmed by a follow-up on-axis XMM-Newton observation Here, we present our timing analysis of the XMM-Newton
in 2021 (Gong 2022). The source was previously speculated to and NuSTAR observations using the Stingray software
be an IP or ultracompact X-ray binary (UCXB; Gong 2022). package (Bachetti et al. 2022). We applied the barycentric
The 614 s periodicity was interpreted as a spin period in the IP correction to the extracted source events using the barycen
case or an orbital period in the UCXB case. Our joint XMM- command in SAS. The photon events were then filtered to
Newton and NuSTAR observations in 2021 detected hard different energy bands (0.3–10, 0.3–5, and 5–10 keV) using
X-ray emission extending up to ∼50 keV. Combining all Hendrics tools in Stingray. For each XMM-Newton
available XMM-Newton and NuSTAR data supports the observation, we analyzed MOS1 and MOS2 data jointly, while
hypothesis that J1745−3213 is an IP with WD mass PN data (available only from the 2021 observation) was
M ∼ 0.75–1.0Me. In Section 2, we describe the X-ray analyzed separately. First, we produced periodograms over a
observations and data reduction of J1745−3213. In broad frequency band from f = 5 × 10−4 Hz to f = 2 Hz using
Sections 3 and 4, we present our X-ray timing and spectral Stingrayʼs weighted Z2n function z_n_search (Figure 1). In
analysis of J1745−3213, respectively. Based on the X-ray power density spectra, we found no evidence of red noise in

2
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Figure 1. Z 32 broadband (0.5 mHz–2 Hz) periodograms. From left to right: 0553950201 MOS, 0870990201 MOS, and 0870990201 PN. In each periodogram, 3σ and
5σ significance levels are marked by solid and faded lines, respectively. The most significant peak is in the region of P ∼ 614 seconds. The first observation
significantly detected a peak at 1227 s, which was less significantly detected in later observations.

Figure 2. Z 32 periodograms (top) and pulse profiles folded by their peak period (bottom) and reference time at T0 = 55197 MJD. From left to right: 0553950201
MOS1, 0870990201, MOS1 and 0870990201 PN. In each periodogram, 3σ and 5σ significance levels are marked by solid and faded lines, respectively. In the pulse
profile plots, the red horizontal lines indicate the mean counts per phase bin, and 1σ error bars are included at f = 0 − 1.

this frequency band thus validating the use of the Z2 test. In the method. By applying the Z2 test jointly on the 2010 and 2021
soft (0.3–5 keV) band alone, we detected peaks at f ≈ XMM-Newton observation data, we obtained an upper limit on
1.628 mHz (corresponding to P ≈ 614 s) with 3σ to 5σ the period derivative to ∣P ∣ < 4.2 ´ 10-11 [s s−1] (90% CL).
significance. Furthermore we detected peaks at f ≈ 0.815 mHz We performed similar Z2 tests over a wider frequency band
(corresponding to P ≈ 1227 s) with less significance (especially ( f = 0.5–2 mHz) to capture the ∼1227 s peak. We additionally
in the newest observation where it was recorded below 3σ saw a significant peak at ∼1841 s. However, analysis of pulse
significance). We found no other periodic signals above the profiles indicated that these peaks were not representative of
3σ level. additional physical phenomena, but rather of lower harmonics
To characterize the peak at P ∼ 614 s more accurately, of the fundamental ∼613 s signal. The pulsed profiles in the
we conducted Z2 tests over a narrow frequency band 0.3–5 keV band are generated by folding source events by
( f = 1–2 mHz) with a fine step size of Δf = 2 × 10−6 Hz (see the peak period in each XMM-Newton data set. They are
Figure 2 for the periodograms in the 0.3–5 keV band). We characterized by a single peak with ∼30%–40% pulsed
found that the 614 s peak is broad (with Δf  0.003 mHz) and, fractions (bottom panel in Figure 2). The light-curve peaks
in some cases, asymmetric, possibly due to timing noise between the different observations and instruments do not
intrinsic to the source. We calculated 90% confidence limit appear to be aligned in phase, but we attribute this to the
+1.13
(CL) error bars on the peak periods by use of the Z2 test tool in slightly different peak periods (e.g., P = 615.01- 0.61 [s] for
+0.97
Hendrics (Bachetti et al. 2022). We found that the peak MOS and 614.37-0.95 [s] for PN in the 2021 XMM-Newton
periods between the two XMM-Newton observations’ MOS observation). In the 5–10 keV band where we did not detect the
and PN data are consistent with each other within error. 614 s period, we calculated an upper limit on the pulsed
The signal was not detected in our identical analysis of fraction using the HENzsearch command to 31%, 34%, and
background events. We also confirmed the 614 s periodicity in 28% for the 2010 MOS, 2021 MOS, and PN data, respectively.
the same frequency and energy bands using the epoch-folding Combining the newest and highest-quality observations (2021

3
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Table 1 band. We jointly fit the NuSTAR spectra with the 2021 XMM-
Time Line of X-Ray Observations Newton observation data (with ∼3000 combined MOS1,
Date ObsId Target Telescope Exposure (ks) MOS2, and PN counts in the 0.5–10 keV band) and the 2010
XMM-Newton observation data (with ∼2800 combined MOS1
2010- 0553950201 H1743 XMM- 86
and MOS2 counts in the 1.5–10 keV band). The MOS spectra
10-09 −322 Newton
2021- 0870990201 J1745 XMM- 31
in the 2010 XMM-Newton observations are dominated by a
02-28 −3213 Newton background below ∼1.5 keV.
2021- 30601021002 J1745 NuSTAR 49 We first fitted a handful of phenomenological models to
03-06 −3213 characterize the overall X-ray spectral properties using XSPEC
(Arnaud 1996). Each spectral model was multiplied by tbabs
Note. In the 2010 XMM-Newton observations, PN was operated in the Small (for interstellar medium (ISM) absorption) using the Wilms
Window mode putting J1745−3213 outside of the FOV. abundance data (Wilms et al. 2000). Furthermore, we used a
cross-normalization factor between the XMM-Newton and
MOS + PN data) yields an upper limit of modulation of 23% NuSTAR spectra using constant in XSPEC. Figures 3 and 4
above 5 keV. The 1227 s signal shows two symmetric present the best model fits and residuals for (1) a power-law
nonoverlapping pulses and the 1841 s signal shows three model; (2) a power-law fit with a Gaussian line component at
symmetric nonoverlapping pulses (see the Appendix for more ∼6–7 keV to account for Fe emission lines; (3) an APEC model
information). Double-peaked modulation in UCXBs is char- fit; (4) an APEC model with a Gaussian line component fixed at
acterized by asymmetric rising and falling flanks and asym- E = 6.4 keV; and (5) an APEC model with a Gaussian line at
metric peaks (Norton et al. 2004; Annala & Poutanen 2010). 6.4 keV and a pcfabs partial absorption component. A single
Similarly, asymmetric peaks are also characteristic of double- power-law model yields a poor fit to the XMM-Newton and
peaked modulation in IPs, where the spin-folded light curve NuSTAR spectra with a cn2 = 1.41 (380 dof) largely due to the
demonstrates uneven peak amplitudes (Nasiroglu et al. 2012). residuals at E ∼ 6–7 keV. Adding a Gaussian line model
These asymmetries would indicate that two linked physical improved the fit to cn2 = 1.15 (377 dof). The best-fit line
features are being observed at a single fundamental frequency energy of 6.65 keV and large equivalent width of 940 eV
(e.g., the two poles of a rotating object). There would still be a indicates a complex Fe K emission line mostly originating from
significant signal at half of the fundamental frequency. H- and He-like ions. The X-ray source spectra are apparently
However, to continue the example of poles, this would characterized by thermal emission. Similarly, an absorbed
effectively fold the two poles together. Again, it is easiest to blackbody model yielded a poor fit with cn2 ~ 2.
check the multiples of any significant frequency to investigate We then fit an absorbed APEC model, i.e., a single-
whether a feature is a fundamental signal or not. Our data did temperature optically thin thermal plasma model (Smith et al.
not demonstrate asymmetry; the pulse profiles have peaks 2001). The best-fit plasma temperature and abundance are
+12 +1.5
symmetric in both height and shape at multiples of 614 s. Thus, kT = 33- 7 keV and Z = 2.2-0.8 Z , respectively. The fit is
we maintain that the 614 s signal represents a spin period. further improved by adding a narrow Gaussian line fixed at
Notably, this contrasts with the conclusion presented in Gong E = 6.4 keV indicating the presence of a neutral Fe K-α line.
(2020) where an asymmetry in the pulsation of the 1227 s Furthermore, we added a partial-covering absorption model
signal is visible. However, our data are supplemented by the (pcfabs), improving the fit of low-energy bins. Although
2021 observation which is on-axis and captures the source with unaccounted for in Gong (2022), pcfabs is needed to model
the PN detector (Table 1). the X-ray absorption due to the accretion curtain that IPs
For NuSTAR data, we applied the barycentric correction to exhibit (Norton & Watson 1989b; Hailey et al. 2016). The
the extracted photon events using the barycorr command line pcfabs ∗ (APEC+gauss) model, while freezing the
tool from HEASoft Version 6.25 (HEASARC 2014). We Gaussian line at E = 6.4 keV, represents the best-fit case with
searched for a periodic signal in NuSTAR data using the same cn2 = 1.00 (376 dof). The plasma temperature is still high at
Z 32 test and a step size of Δf = 2 × 10−6 between f = 1–2 mHz kT = 25 ± 6 keV. Our fit results are summarized in Table 3.
where we expect the 614 s signal and its higher harmonics. A Through the best-fit flux normalization factors, we found that
broadband 3–10 or 3–40 keV search in the frequency band of the X-ray flux decreased by ∼15% from 2010 to the 2021
f = 10−4–2 Hz yielded no significant detection except a peak at observations. The neutral hydrogen column density is
P ∼ 32.3 minutes (with 5σ significance). However, the 32.3 NH ∼ 4 × 1022 cm−2 for all the models except when we
minute peak is likely a harmonic of the NuSTAR’s 96.8 minute applied the partial-covering absorption model (which yielded a
orbital period as evidenced by the lack of a counterpart in the lower value of NH = 1.6 × 1022 cm−2).
XMM-Newton data. In Table 2, we present our 3–5 keV
NuSTAR results as the 614 s modulation was detected by 5. Optical and Infrared Counterparts
XMM-Newton below 5 keV. We found no significant
pulsations and the P ∼ 614 s signal appeared to be at 1.4σ We found no optical counterpart within ∼15″ of J1745
level. The insignificant detection of the 614 s periodicity, with a −3213 (Gaia Collaboration 2020) due to high ISM absorption.
pulsed fraction <32%, is likely due to the poor statistics at Greiss et al. (2014) identified a possible IR counterpart to
E < 5 keV where XMM-Newton observations demonstrated J1745−3213 by cross-referencing the Chandra Galactic Bulge
that the signal was most pronounced. Survey and the Visible and Infrared Survey Telescope for
Astronomy Variables in the Via Lactea (VVV) Survey (Greiss
et al. 2014). However, the IR counterpart (VVV J174517.24-
4. X-Ray Spectral Analysis
321357.67) is offset by 3″ from J1745−3213 and, conse-
The NuSTAR spectra of the source (ObsID 30601021002) quently, the match between the two sources has a large false-
held ∼2200 combined FPMA + FPMB counts in the 3–50 keV alarm probability of 0.6 (Greiss et al. 2014). We, therefore,

4
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Figure 3. XMM-Newton EPIC and NuSTAR spectra fit with four phenomenological models. From top left to top right: (1) an absorbed power-law model fit; (2) an
absorbed power law fitted with a Gaussian at 6.4 keV component; (3) an absorbed APEC model fit; and (4) an absorbed APEC model with a Gaussian at 6.4 keV
component. The NuSTAR observation used here is 30901021002 and is presented in the 3–50 keV band, with FPMA (1146 counts) in black and FPMB (1097 counts)
in red. The XMM-Newton observation 0553950201 is presented in the 1.5–10 keV band with MOS1 (1438 counts) in magenta and MOS2 (1342 counts) in yellow.
The more recent XMM-Newton observation 0870990201 is presented in the 0.5–10 keV band with MOS1 (667 counts) in green, MOS2 (671 counts) in navy blue,
and PN (1688 counts) in light blue.

Table 2
Summary of XMM-Newton EPIC and NuSTAR Timing Analysis

Instrument Peak Frequency (mHz) Peak Period (s) Counts Z 32 Significance (σ)b P (%)a
XMM-Newton 0.3–5 keV

0553950201
+0.0006 +0.08
MOS 1.6288-0.0002 613.94-0.22 2636 115.77 >7 30 ± 2
0870990201
+0.0016 +1.13
MOS 1.6260-0.0030 615.01-0.61 1156 78.56 7.0 37 ± 3
+0.0025 +0.97
PN 1.6277-0.0026 614.37- 0.95 1114 75.74 6.9 37 ± 3

NuSTAR 3–5 keV

30601021002
+0.0010 +0.58
FPMA + FPMB 1.6317-0.0025 612.84-0.36 768 23.09 1.4 <32

Notes. All errors shown are at 90% confidence intervals. The results are from the Z2 test in the narrow frequency band. For the 3–5 keV NuSTAR data from which the
614 s signal was not significantly detected, the peak frequencies/periods (without error bars) and the upper limits on the pulsed fraction are listed. The total counts for
each observation, instrument, and energy band are listed as well.
a
Pulsed fractions [%] and upper limits were calculated by HENzsearch (Bachetti et al. 2022). All of the hard-energy band Z 32 searches produced no significant peaks
above 1σ.

5
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Table 3
Phenomenological Model Fits to XMM-Newton EPIC and NuSTAR Spectra

Parameter pow pow + gauss APEC APEC+gauss pcfabs∗ (APEC+gauss)


+0.07 +0.07
C2021a 0.87- 0.06 0.87- 0.06 0.84 ± 0.06 0.85 ± 0.06 0.87 ± 0.06
+0.10 +0.09 +0.09
C2010a 1.11- 0.09 1.14 ± 0.09 1.09- 0.08 1.11- 0.08 1.18 ± 0.09
NH(i) (10 22 cm-2)b 4.6 ± 0.4 4.1 ± 0.4 4.2 ± 0.3 4.0 ± 0.3 1.6 ± 0.5
+2.4
NH( p) (10 22 cm-2)c ... ... ... ... 7.7- 1.6
+0.08
ccf ... ... ... ... 0.77- 0.09
Γ 1.48 ± 0.08 1.46 ± 0.08 ... ... ...
+11.6 +15.3 +5.8
kT (keV) ... ... 33.0- 6.5 37.3- 9.1 24.7- 5.6
+1.5 +1.7 +0.7
Zd(Ze) ... ... 2.2-0.8 2.6-1.1 1.1-0.5
+0.09
Eline (keV) ... 6.65- 0.10 ... 6.4f 6.4f
+0.26
σline (keV) ... 0.43- 0.13 ... 0.01f 0.01f
+66 +59 +35
EWline (eV) ... 937- 114 ... 188- 10 146- 37
FX(10−12 erg cm−2 s−1)e 3.91 ± 0.02 3.88 ± 0.02 3.34-+0.03
0.04 3.41 ± 0.04 3.33 ± 0.05
cn2 (dof) 1.41 (380) 1.15 (377) 1.26 (379) 1.18 (378) 1.00 (376)

Notes. All errors shown are 1σ intervals.


a
Cross-normalization factors of the XMM-Newton data in 2021 (C2021) and in 2010 (C2010) with respect to the NuSTAR observation in 2021.
b
The ISM hydrogen column density associated with tbabs which is multiplied to all the models.
c
NH( p) and cf are the hydrogen column density and covering fraction associated with pcfabs.
d
Abundance relative to solar.
e
Unabsorbed 3–50 keV flux of the NuSTAR data.
f
The parameter is frozen.

main-sequence F9/F9.5 spectral type or later (Pecaut &


Mamajek 2013). Using X-ray observations, Jonker et al.
(2014) established a lower bound of 3 kpc for the source
distance. If the true source distance were at this lower bound,
the extinction would be lower than E(B − V ) = 1.8. A lower
extinction would result in an even fainter spectral type than the
above results. Therefore, regardless of distance, we conclude
that the companion star of J1745−3213 is a main-sequence star
of type F9/F9.5 or later.

6. Discussion
In this section, we first discuss the source type of J1745
−3213 primarily based on the X-ray spectral properties
(Section 6.1). Then, after identifying J1745−3213 as an IP,
we measure its WD mass using our X-ray spectral model for
mCVs (Section 6.2) and discuss some implications of
characterizing the mass and B field of an IP at distances (d 
Figure 4. XMM-Newton and NuSTAR spectra and residuals of J1745−3213 3 kpc) where optical observations are hampered by significant
fit with a tbabs ∗ pcfabs ∗ (APEC+gauss) model, which yielded a extinction .
cn2 = 1.00 for 376 degrees of freedom and strongly supports the IP scenario.
The Gaussian line fitted at 6.4 keV and modeling the Fe K emission line
+68
yielded an equivalent width of 146-49 eV, which is typical of IPs.
6.1. Source Identification
consider no IR counterpart was detected for J1745−3213. The We present our source identification scheme for J1745
source is thus likely fainter than the sensitivity limits for the −3213 based on the X-ray analysis results and counterpart
VVV Survey. We used the sensitivity limits of the VVV searches in other wavelengths. The lack of a bright IR
Survey to constrain the spectral type of the companion star. The counterpart (Section 5) rules out the possibility that J1745
magnitude limits for the VVV Survey are J = 20.2, H = 18.2, −3213 is a high-mass X-ray binary. The strong Fe K emission
and Ks = 18.1. Using the relative extinctions from the VVV lines suggest that the source is not a low-mass X-ray binary
Survey (Saito et al. 2012) and assuming E(B − V ) = 1.8 (the whose X-ray emission is nonthermal (occasionally with weak
typical extinction value in the direction of the Galactic Bulge; Fe lines; Mori et al. 2021). The presence of Fe emission lines as
Greiss et al. 2014), we found the J, H, and Ks extinctions well as the long-term variability rules out a pulsar origin.
(1.56, 1.02, and 0.655, respectively) with an optical extinction Below we consider two possible source types suggested by
AV = 5.55. Assuming a source distance of 8 kpc (Wevers et al. Gong (2022) and conclude that the X-ray source is most likely
2016), we derived absolute magnitudes of MJ = 4.125, MH = an IP mainly based on the X-ray spectral properties obtained by
2.665, and MKs = 2.930 . Using the MKs limit value, we find a the XMM-Newton and NuSTAR observations.

6
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

6.1.1. UCXB accretion curtain or reflected from the WD surface (Norton &
Watson 1989b; Hailey et al. 2016). In addition, the 614 s period is
Solely based on the original 2010 XMM-Newton observation, it
in the typical spin period range of IPs (see Table 3 in Suleimanov
was proposed that J1745−3213 may be a UCXB with a short
et al. 2019), and the low-energy X-ray spin modulation is often
orbital period of 614 s (Gong 2022). UCXBs are low-mass X-ray
observed among IPs due to the accretion curtain absorption
binaries usually containing a WD donor and a neutron star accretor
(Anzolin et al. 2008). While double-peaked modulation is a
(or very rarely a stellar-mass black hole) and with an ultrashort
common feature of IPs, the nonoverlapping, highly symmetric
(<1 hr) orbital period. X-ray spectra of known NS-UCXBs are
pulses in the ∼1227 and ∼1841 s regions indicate that the ∼614 s
well characterized by a kT ∼ 1–3 keV blackbody component
signal is a single-peaked modulation. While this is typically
originating from the NS boundary layers (Koliopanos 2015). We
associated with slower-spinning IPs, this feature has been
found a UCXB origin implausible because an absorbed blackbody
previously detected in faster-spinning IPs (Pspin = 564 s, 450 s;
model did not fit the 0.5–50 keV XMM-Newton + NuSTAR
Bonnet-Bidaud et al. 2007; Gorgone et al. 2021). This indicates
spectra well. Furthermore, other combinations with a blackbody
that the emissions of only one of the two WD poles are visible in
model (e.g., blackbody + power law) did not fit the X-ray spectra
the observations. Lastly, radio observations of the local position of
due to the prominent Fe lines at E = 6–7 keV. In addition, the
J1745 with the Australia Telescope Compact Array in Australia in
known UCXBs exhibit no or weak Fe K emission lines in their
late 2021 provided inconclusive detection of the source, while
X-ray spectra (Koliopanos et al. 2021). UCXBs with He-rich
later Very Large Array observations yielded no detection in the
donors show weak Fe lines with equivalent widths (EWs) ≈ 50–80
radio band at all (Gong 2022). The lack of radio emission is also
eV (except for X-ray transients, which show stronger Fe K
congruent with the IP hypothesis as Barrett et al. (2020) reported
features), whereas those with C/O or O/Ne/Mg-rich donors show
that only a small fraction of IPs had been detected in the
no traces of Fe emission lines (Koliopanos et al. 2021). Note that
radio band.
these weak or no Fe line features were observed from a dozen very
bright or outbursting UCXBs with LX > 1036 erg s−1. For a much
6.2. Determining the White-dwarf Mass of J1745
fainter UCXB similar to J1745−3213 (L  1034 erg s−1), Fe line
emission should be negligible due to a lack of bright illuminating In this section, we outline our methodology for determining
X-rays. Thus, the strong Fe emission lines detected with a total the WD mass of J1745−3213. In general, X-ray emission from
EW ∼ 0.9 keV also rule out the UCXB hypothesis. IPs originates primarily from the accretion column. Some
X-rays are reprocessed via absorption in the accretion curtain
and reflection by the WD surface (Norton & Watson 1989b;
6.1.2. Intermediate Polar Hailey et al. 2016).We utilize the broadband X-ray spectra to
constrain parameters such as the magnetospheric radius,
The variable, thermal X-ray spectra with atomic emission lines
accretion rate, and WD mass.
indicate that J1745−3213 is a CV. In particular, the hard X-ray
To model X-ray emission from the accretion column, we
spectra extending up to E ∼ 50 keV and hot plasma temperature
developed an XSPEC model based on the accretion flow equations
of ∼30–40 keV suggest an IP origin. IPs are copious emitters of
presented in Saxton et al. (2005). The XSPEC model, called
hard X-rays above 20 keV as a large fraction of the hard X-ray
MCVSPEC, assumes a steady, planar accretion flow channeled
sources detected by INTEGRAL have been identified as IPs
along B field lines from the shock height to the WD surface
(Masetti et al. 2010). The typical IP plasma temperature is
(K. Mori et al. 2023, in preparation). In IPs, the accretion flow
kT ∼ 20–40 keV (Mukai 2017). On the other hand, the other types
begins freefalling along the magnetic field lines from the
of CVs, namely, polars and nonmagnetic CVs (nmCVs or dwarf
magnetospheric radius (Rm) defined by the balance between
novae), exhibit much softer and fainter X-ray spectra. Polars, with
magnetic and gas pressures. For a given set of M, Rm, and specific
typically higher magnetic fields (B ∼ 10–200 MG) than those of
accretion rate m [g cm−2 s−1] taken as input parameters, coupled
IPs, have lower plasma temperatures (kT ∼ 5–20 keV) due to
differential equations for the flow continuity and momentum/
faster cyclotron cooling (Warner 2003; Mukai 2017). On the other
energy conservation are numerically solved along the accretion
hand, nmCVs are characterized by even fainter and softer X-ray
column to determine the density and temperature profiles. The
emission with kT  10 keV (Byckling et al. 2010). Even a few
model considers both thermal Bremsstrahlung and cyclotron
outliers with harder X-ray spectra (e.g., SS Cyg with kT ∼ 20 keV
cooling, the latter of which is prevalent in polars or highly
during quiescence (Ponman et al. 1995)) do not reach the plasma
magnetized IPs with B  30 MG (Wu et al. 1994). An X-ray
temperature as high as those of J1745−3213. Furthermore,
model spectrum is calculated by integrating the emissivity (with
nmCVs tend to be far less luminous than IPs, with typical
different kT and ρ values) throughout the accretion column by
0.25–10 keV luminosity ranging anywhere from 1030 erg s−1 to
implementing the most updated atomic database for collisionally
1032 erg s−1 (Wada et al. 2017). The measured, unabsorbed X-ray
ionized plasma.6 A significant improvement in our modeling
flux of ∼1.4 × 10−12 erg cm−2 s−1 converts to LX = 2 ×
relevant to IPs, and missing from the analysis of Gong (2022),
1033–1 × 1034 erg s−1 assuming the source distance of 3–8 kpc.
is the incorporation of the recent observation (Suleimanov et al.
This X-ray luminosity is much higher than those of nmCVs.
2016) that the accreting gas begins its infall from the disk
Moreover, the measured EW for the Fe K-α line at 6.4 keV is
magnetospheric radius, not from infinity, as was previously
188 eV and 146 eV for APEC+gauss and pcfabs ∗ (APEC
assumed in all models. As pointed out by Suleimanov et al.
+gauss) models, respectively (Table 2). These results are
(2016), if the finite magnetospheric radius effects for IPs are
consistent with the observed mean EW of the neutral Fe K-α line
not taken into account, it will lead to an overestimation
of 115 eV for IPs and are markedly different from the polars’ and
of the freefall velocity at the shock height as vff =
nmCVs’ of 68 and 62 eV, respectively (Xu et al. 2016). The best- 1 1
fit model (Figure 3) requires a partial-covering absorption term, 2GM ( R + h - R ) . Furthermore, as vff is directly linked to
m
and this is also consistent with the IP hypothesis because X-ray
6
emission from the accretion column is partially shielded by the http://atomdb.org

7
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

6.2.2. Mass Accretion Rate and Source Distance


The total mass accretion rate M [g s−1] can be estimated
(
from the bolometric luminosity (L): L = GMM R - R . As
1 1
m
the source was not detected below ∼1 keV, we set L ≈ LX, for
)
now, where LX is the absorption-corrected X-ray luminosity,
assuming that most of the radiation is emitted in the X-ray band
through thermal Bremsstrahlung emission. We used
LX = 4πd2FX where FX is the unabsorbed X-ray flux in
0.8–50 keV and assumed source distance (d). However, in
reality, L > LX because some X-rays emitted from the accretion
column can be reprocessed in the lower energy band
(E  1 keV). For example, a soft X-ray blackbody component
of kT = 86 eV constitutes ∼30% of the whole X-ray emission
for IP 1RXSJ154814.5-452845 (Haberl et al. 2002). Further-
more, when the WD B field is high, magnetic CVs can have a
nonnegligible fraction of the radiation through cyclotron
Figure 5. MCVSPEC model spectrum with M = 1Me and m = 1 g cm−2 s−1 emission in the optical and UV bands. As no optical or UV
and Rm/R = 20. emission from the source at a large distance (d > 3 kpc) is
observable, we need to rely on modeling the ratio of thermal
bremsstrahlung and cyclotron cooling rates (ò = Lcyc/Lbr). ò is a
the shock temperature, WD masses in IPs will be under- function of the accretion column height (h) varying with the
estimated if the effect of the finite magnetospheric radius is plasma density and temperature. In general, ò(h) is the highest
ignored. Furthermore, MCVSPEC accounts for the effects of at the shock height (òs) where the plasma density and
finite shock height by a self-correcting shooting method that temperature are lowest and highest, respectively (Wu et al.
adjusts the freefall velocity and other appropriate parameters to 1994). Based on our spectral fits and the range of the specific
the model’s derived shock height until it converges. For accretion rate, as described below, we found that òs ∼ 0.1–0.5
example, Figure 5 shows an MCVSPEC model spectrum with and that ò rapidly decreases toward the WD surface. To
WD mass M = 1Me and m = 1 g cm−2 s−1 and Rm/R = 20. estimate the overall contribution of the cyclotron cooling
In the following subsections, we outline how we constrain Rm component in the bolometric luminosity, we calculated the
(Section 6.2.1) and m (Sections 6.2.2 and 6.2.3) prior to the mean ratio (̄ ) by integrating ò over the accretion height (from
spectral fitting and eventually measure the WD mass h = 0 to hs), using the best-fit temperature and density profile
(Section 6.2.4). The process begins by deriving Rm/R from the obtained by MCVSPEC, and dividing by hs. For each case of the
spin period, assuming that the IP is in a spin equilibrium. We then model fitting, ̄ allows us to estimate how much luminosity we
bound the specific accretion rate by considering a range of the are missing by considering the X-ray flux measurements. If the
cyclotron emission is not negligible, the accretion rate is
source distance (d) and fractional accretion column area ( f ).
underestimated by assuming L ≈ LX, especially for higher M
Following these parameter constraints, we proceed to X-ray
and M values, which yield greater ̄ (see Wu et al. (1994) for
spectral fitting with the MCVSPEC and X-ray reflection model, how òs depends on M and M ). In the subsequent sections, we
where some of the input parameters (e.g., a reflection fraction) are find that ̄ is at most ∼20% and describe how the cyclotron
properly linked with each other or frozen to the predetermined emission component affects our IP parameter determination.
values (e.g., Rm/R). As described below, our physically motivated Hereafter, we adopt a range of M = (2–10) ´ 1016 [g s−1],
model allows us to constrain the IP parameters. To clarify our corresponding to d = 3–8 kpc, respectively, assuming L ≈ LX.
method, we provided a flowchart in the Appendix. We will later discuss how the underestimation of the mass
accretion rate due to the “missing” luminosity in the soft X-ray
band and below will affect the WD mass measurements.
6.2.1. Magnetospheric Radius
Suleimanov et al. (2019) suggested that the frequency of a 6.2.3. Specific Mass Accretion Rate
spectral break in the power density spectrum could be used to
estimate Rm. However, besides the spin period of 614 s and its Given that Rm/R and M are constrained, we proceed to
harmonics, we found no other feature or break in the power density bound the specific accretion rate (m ) by considering the
spectra obtained from the XMM-Newton and NuSTAR data. fractional accretion column area ( f ) in two extreme cases. Note
M
Therefore, following Shaw et al. (2020), we estimated Rm that the specific accretion rate is defined as m = 4pR2f , where f
assuming that the WD is in spin equilibrium with the accretion is the fractional area of the accretion column base with respect
disk. We derived Rm/R = 20–22 for M = (1–1.4)Me using to the WD surface. In general, f is unknown observationally
GMP 2 1 3 except for an eclipsing IP XY Ari with f < 2 × 10−3
( )
Rm = 4p 2 where P is the WD spin period (614 s). Note (Hellier 1997). We first set an upper limit of f to
fmax = 2R = 2.5 ´ 10-2 . This is the maximum fractional area
R
that the dependence on M is small compared to P, so we adopt the
mass estimate of Gong (2022). The ratio of Rm/R ∼ 20 is within m
of the accretion column for a dipole B field geometry assuming
the range of Rm/R = 8–25 obtained by a systematic study of 45 that the infalling gas originates from the entire accretion disk
IPs using NuSTAR and Swift-BAT data (Suleimanov et al. 2019). (Frank et al. 2002). A combination of d = 3 kpc and fmax leads
Hereafter we adopt Rm/R = 20 for MCVSPEC. to a minimum specific accretion rate of m = 0.6 g cm−2 s−1.

8
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Table 4
MCVSPEC Fitting Results

Parameter Case A Case B Case C Case D


Rm/R 20 20 20 20
m a [g cm−2 s−1] 0.6 3.0 10 44
+0.07
M [Me] 0.94 ± 0.09 0.82 ± 0.06 0.81 ± 0.06 0.81- 0.06
+0.25 +0.29 +0.28 +0.26
Z [Ze] 1.17- 0.27 1.03- 0.23 1.02- 0.23 1.00-0.22
+0.04
C2021 0.84 ± 0.04 0.85 ± 0.04 0.85- 0.03 0.86 ± 0.04
+0.04
C2010 1.13 ± 0.04 0.97 ± 0.04 0.97- 0.03 0.97 ± 0.04
NH(i) (10 22 cm-2) 2.5 ± 0.5 2.5 ± 0.5 2.5 ± 0.5 2.5 ± 0.05
+1.7 +1.6 +1.57 +1.6
NH( p) (10 22 cm-2) 11.2- 1.4 11.0- 1.3 11.0-1.27 10.9- 1.3
+0.05
c bf 0.79- 0.06 0.80 ± 0.05 0.80 ± 0.05 0.80 ± 0.05
rrefl 0.47 0.76 0.86 0.96
+0.00 +0.00 +0.00 +0.00
cos (i ) 0.95- 0.39 0.95- 0.36 0.95- 0.36 0.95- 0.36
cn2 (dof) 1.12 (371) 1.11 (371) 1.11 (371) 1.11 (371)
hs/R 18% 3% 1% <1%
òsc 0.11 0.19 0.24 0.50
̄ d 0.05 0.07 0.10 0.19
B (MG)e 7 15 27 57

Notes. All errors shown are 1σ intervals.


a
The specific accretion rate is presented without error bars as it was frozen for fits.
b
cf is the covering fraction associated with pcfabs, not to be confused with the fractional accretion area.
c
The ratio of the Bremsstrahlung cooling timescale to the cyclotron cooling timescale at the shock (see Wu et al. 1994). This is derived from the fit parameters.
d
The mean ratio of the Bremmstrahlung cooling timescale to the cyclotron cooling timescale along the accretion column. This is derived from the profiles of the
model.
e
An estimate of the magnetic field strength derived from the fit parameters.

This value sets a lower limit of m as, in reality, the mass m = 3 g cm−2 s−1 (Case B) and m = 10 g cm−2 s−1 (Case C),
accretion originates from the so-called threading region where and the maximum value of m = 44 g cm−2 s−1 corresponding
the gas in the inner accretion disk is connected with the to d = 8 kpc and f = 0.001 (Case D). We note that the shock
magnetic field lines toward the polar regions (Belloni et al. height (hs) is calculated and output from the MCVSPEC model
2021), in contrast to emission coming from the entire disk fit as it affects the reflection scaling factor (rrefl) in the
(corresponding to f = fmax ). For a lower limit of f, we adopted reflect model. For instance, in the tall accretion column,
the theoretical range of 0.001  f  0.02 from Rosen (1992) X-rays emitted near the shock are less likely to be reflected off
who considered various IP parameters and geometrical the WD surface. Initially, we fit the model to the XMM-
configurations. Using the theoretical lower bound of Newton and NuSTAR spectra without a reflection (reflect)
f = 0.001 and d = 8 kpc, we derive the maximum specific component. This procedure is necessary to bound hs that can be
accretion rate of m = 44 g cm−2 s−1. Hereafter, we present our used for calculating the reflection scaling factor (rrefl),
fit results for the range of m = 0.6–44 g cm−2 s−1. following the formula given by Tsujimoto et al. (2018):
rrefl = W 2p = 1 - 1 - 1 (1 + hs R)2 . As hs is calculated
6.2.4. Spectral Fit Results implicitly in the MCVSPEC model using a shooting method and
updating vff iteratively in the accretion column profile
Given that two of the input model parameters (Rm/R and m ) computation, we cannot link hs and rrefl directly in XSPEC.
are constrained as above, we are able to fit MCVSPEC to the
Thus, we need several iterations to fit the X-ray spectra by
XMM-Newton and NuSTAR spectra with less parameter
updating hs and adjusting rrefl in the reflection model. We then
degeneracy. In contrast to Gong (2022), we also considered
continued to fit the model, deriving new M and hs values with
additional photon reprocessing effects associated with X-ray
emission from IPs. Some X-rays emitted from the accretion each iteration and adjusting (and freezing) rrefl accordingly
column can be absorbed in the preshock region or reflected by until hs did not change by more than 1% of the WD radius. In
the WD surface as manifested in the neutral Fe fluorescence the end, we found that the reflection scaling factors are 0.47,
lines (Hayashi et al. 2021). Following previous X-ray studies 0.76, 0.86, and 0.96 for the m = 0.6, m = 3, m = 10, and
(Hailey et al. 2016), we add a Gaussian line component at m = 44 g cm−2 s−1 cases, respectively. Note that a higher
E = 6.4 keV for the neutral Fe K-α fluorescence line, pcfabs accretion rate reduces the shock height and thus increases the
in XSPEC to account for X-ray absorption at the preshock degree of X-ray reflection.
region, and reflect to account for X-ray reflection off the Table 4 presents our X-ray spectral fit results for the four
WD surface. Our physically motivated model then consists of accretion rates considered. For all the cases, the shock heights
tbabs ∗ pcfabs ∗ (MCVSPEC ∗ reflect+gauss). derived from the model remained below 20% of the WD radius,
Furthermore, the constant model was added to account for ensuring the validity of the MCVSPEC model, which assumes a
the flux variability between the different X-ray observations. planar geometry of the accretion column (compared to the dipole
We considered four representative accretion rates: the field configuration studied by Canalle et al. 2005 and Hayashi &
minimal value of m = 0.6 g cm−2 s−1 corresponding to Ishida 2014). We found that all the cases yielded similarly good
d = 3 kpc and f = fmax (Case A), two intermediate cases of fits to the XMM-Newton and NuSTAR spectra with cn2 ~ 1.1

9
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Figure 6. XMM-Newton and NuSTAR spectra fit by tbabs ∗ pcfabs ∗ (reflect ∗ MCVSPEC+gauss) and residuals for RM/R = 20. Top left:
m = 0.6 g cm−2 s−1; top right: m = 3.0 g cm−2 s−1; bottom left: m = 10 g cm−2 s−1; bottom right: m = 44 g cm−2 s−1.

(Figure 6). Fitting the atomic lines by MCVSPEC also allows us to m is the largest. This is because, for the fixed Rm/R = 20, the WD
determine the Fe abundance to Z = (0.8–1.5)Ze. B field is the largest at B = 57 MG as a result of the highest m
The best-fit WD mass depends on the assumed accretion rate value. From the best-fit WD mass and assumed m value, we
more significantly for the lowest accretion rate (Case A) than in calculated the B field using the formula for Rm (Norton &
the other cases with higher m values (Case B–D) for two Watson 1989a). However, even in this case with a higher B field,
reasons. The low-accretion-rate case yields a much higher we found that òs = 0.5 and ̄ = 0.19, indicating that a fraction of
shock height of hs/R = 0.18. First, it will reduce the freefall the cyclotron emission within the overall radiative cooling is 50%
velocity and make the model X-ray spectrum softer as a result and 19% at the shock height and on average throughout the
of lower shock temperature, thus requiring to increase the WD accretion column, respectively. Our results are consistent with the
mass to fit the X-ray spectral data. This effect was investigated analytical solution of Wu et al. (1994) who found that the X-ray
by Hayashi & Ishida (2014) who found that mass accretion spectral shape does not vary significantly due to the cyclotron
rates in the range of m ~ 1–100 [g cm−2 s−1] did not alter the cooling effect when òs < 20 (Wu et al. 1994). To take the cyclotron
X-ray spectral shape. Second, the smaller shock height will emission in the bolometric luminosity into account, we increased
increase the degree of X-ray reflection (i.e., larger rrefl) and the mass accretion rates by 5%–19% based on the derived ̄
make the model X-ray spectrum harder (Hailey et al. 2016). values. In addition, a fraction of the X-ray emission could be
Therefore, we need to decrease the WD mass in order to fit the reprocessed through soft X-ray blackbody radiation. By taking the
X-ray spectra. Above m ~ 3.0 g cm-2 s-1, the spectral para- aforementioned IP 1RXSJ154814.5-452845 observed by Haberl
meters (including the WD mass) become insensitive to these et al. (2002) as a representative case, we estimate ∼30% of the
two effects associated with hs. This is because the small hs/R X-ray luminosity is missing through soft X-ray blackbody
ratio no longer impacts the freefall velocity, and also, the radiation as X-ray emission from our source is heavily absorbed
reflection fraction is close to unity. In this m range, the WD below 1 keV. Overall, m may be underestimated by ∼35%–50%
mass measurement converges to ∼0.81 Me and is almost for Cases A–D. Nevertheless, such increases in the m values do
independent of the assumed m value. not change the WD mass significantly for Cases B, C, and D
As described earlier, we should take into account the fact that because they are already in the high accretion regime where M is
the mass accretion rate is underestimated due to the (unobserved) insensitive to m variation as described above. For Case A where m
cyclotron and soft X-ray blackbody emissions. First, we found that increases from 0.6 to 0.8 g cm−2 s−1, the WD mass goes down
the cyclotron cooling effect is most pronounced for Case D when from 0.94 ± 0.09Me to 0.92 ± 0.08Me. We adopt the latter WD

10
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

mass value as a more realistic solution. However, we note that òs paper presents the first application of a comprehensive X-ray
and ̄ increase rapidly with the WD mass and exceed unity spectral model to a newly discovered IP, incorporating the
(meaning that cyclotron cooling is more dominant over thermal finite magnetosphere radius, the impact of the shock height,
Bremsstrahlung cooling), when it is above ∼1Me. We caution and X-ray reflection in a self-consistent manner. By accounting
that, for massive IPs with M  1Me and without optical or UV for uncertainties associated with source distance and fractional
detection, the “missing” cyclotron cooling component should be accretion column area, we have determined the systematic
estimated as shown above and reflected in the mass accretion rates. errors related to the WD mass. Our methodology could be
applied to other IPs to measure their WD masses more
accurately. Notably, recent dynamical measurements of WD
6.2.5. IP Parameter Determination
masses in GK Per and XY Ari have presented challenges to the
While the statistical errors for M are small (<10%) due to the previous (and differing) WD mass measurements using X-ray
high-quality X-ray data, the systematic errors arising from the spectroscopy data (Álvarez-Hernández et al. 2021, 2023). To
unknown mass accretion rate are more predominant as reduce systematic errors, we stress the significance of
manifested by the discrepancy between Case A (M = measuring source distances, detecting or estimating cyclotron
0.94Me) and the other cases (M = (0.81–0.82)Me). As emission in the optical/UV band (especially for highly
discussed earlier, m depends on both the source distance and magnetized IPs), and obtaining broadband X-ray spectra
fractional accretion column area. Therefore, if we take the beyond 10 keV. In the future, the HEX-P X-ray probe mission,
broadest parameter space for d and f, the WD mass was which covers a broad X-ray band of 0.1–80 keV, will be ideal
constrained to M = (0.75–1.03)Me. If we assume that the for determining the WD masses accurately from a substantial
source is in the Galactic Center (at d = 8 kpc), the minimum number of mCVs (Madsen et al. 2019).
possible accretion rate would be m = 1.5 g cm-2 s-1, corresp-
onding to hs/R ∼ 7% and rrefl = 0.64. In this case, Acknowledgments
M = (0.85 ± 0.06)Me, and thus the WD mass is in the range of
M = (0.75–0.91)Me, consistent with the mean WD mass of IPs When we were preparing our manuscript for resubmission,
(Zorotovic et al. 2011; Shaw et al. 2020). we became aware of Gong (2022) who analyzed the same
Based on the best-fit WD masses, Rm/R = 20, and assumed XMM-Newton and NuSTAR data of J1745−3213. We thank
mass accretion rates, we estimated the WD B field using the the referee for carefully reading our manuscript and making
magnetic radius formula (Norton et al. 2004)—B = 7, 15, 27, valuable comments. We thank Dr. Matteo Bachetti for helping
and 57 MG for Cases A, B, C, and D, respectively. The loose us with the X-ray timing analysis. K.M. acknowledges support
constraints on the WD B field in the range of B = 7–57 MG are from NuSTAR Cycle 6 Guest Observer Program grant
on the higher end compared to the typical range for IPs NNH19ZDA001N. G.P. acknowledges funding from the
(B ∼ 0.1–10 MG). The B fields in Cases C and D are European Research Council (ERC) under the European
comparable to or even exceed the highest IP B field ever Unionʼs Horizon 2020 research and innovation program (grant
measured from V405 Aur (B = 32 MG; Piirola et al. 2008). agreement No. 865637). C.G.S. acknowledges continuous
support from Cristina “Yaya” Salcedo during this project and
the development of MCVSPEC.
7. Conclusion Software: HEASoft Version 6.25 (HEASARC 2014), FTools
We have both provided further analysis on J1745 and presented Version 6.25 (HEASARC 2014).
a novel methodology for constraining the IP parameters, including
the WD mass and B field. We presented further evidence, using Appendix
X-rays below ∼5 keV, that there is a statistically significant 614 s
period. With our enlarged XMM-Newton data set, we do not
confirm the asymmetry in the pulse shape of the 1227 s X-ray A.1. XMM-Newton–EPIC Folded Light Curves
periodicity seen by Gong (2022). Consequently, we argue that the We present phase profiles folded with 614, 1228, and 1841 s
614 s period is the spin period, the 1227 s period is a lower periods with the newest and highest-quality available data to justify
harmonic, and the emission is from one pole only. Using those that the 614 s signal represents a fundamental period. The
results and a wide range of accretion rates and our novel X-ray symmetry between two peaks that appear in a single-phase profile
spectroscopic model MCVSPEC, we measured the WD mass of indicates each peak is the same signal repeated twice. All higher
J1745. Furthermore, our physically motivated model encapsulates harmonics of the 614 s periodicity show peaks within error of each
many features and corrections that are not found together in other other. Furthermore, the dip between the peaks goes below the
IP models, such as the effect of the shock height and magneto- mean counts (red horizontal lines), indicating that these are
sphere radius. Its calculation of the shock height also serves as a nonoverlapping signals. It must be noted, however, that the
means to constrain the effects of X-ray reflection of the WD original observation was excluded as it anomalously only detected
surface of an IP. As demonstrated in our study of J1745-3213, it is our source with the MOS1 instrument. In summary, based on the
essential to follow up X-ray sources (which may show periodic new XMM-Newton observation of the source on-axis, for which
signals and/or large hardness ratios as potential signatures of we provide a joint pulse profile from all three modules, indicates
magnetic CVs) with NuSTAR for firmly determining the source that the signal at 1228 s is a repeated 614 s signal (Figure 7).
types and WD masses. Our results demonstrate promising
outcomes from mCV populations in the X-ray band as a
A.2. Methodology for Determining WD Mass and Magnetic
complementary approach to the optical/UV studies (Pala et al.
Field Strength
2022).
Furthermore, our methodology may be applied to its greatest For additional visual clarity, we present a flowchart outlining
effect on sources of known or well-constrained distance. This the different steps for determining WD mass (Figure 8).

11
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Figure 7. Pulse profiles at the peak frequency and its harmonics for 0870990201 MOS (top), 0870990201 PN (middle), and 0870990201 joint (bottom). The red
horizontal lines indicate the mean counts per phase bin and 1σ error bars are included at f = 0 − 1.

12
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Figure 8. The blue box indicates the step related to X-ray timing analysis. We derived variables from the known or fit parameters in the steps colored in green.
Magenta boxes indicate the steps involved in spectral fittings.

ORCID iDs Bachetti, M., Huppenkothen, D., Khan, U., et al. 2022, StingraySoftware/
stingray: Version 1.0, v1.0, Zenodo, doi:10.5281/zenodo.6394742
Benjamin Vermette https://orcid.org/0000-0003-4292-1281 Bahramian, A., Heinke, C. O., Kennea, J. A., et al. 2021, MNRAS, 501, 2790
Ciro Salcedo https://orcid.org/0000-0001-8586-7233 Barrett, P., Dieck, C., Beasley, A. J., Mason, P. A., & Singh, K. P. 2020,
Kaya Mori https://orcid.org/0000-0002-9709-5389 AdSpR, 66, 1226
Belloni, D., Rodrigues, C. V., Schreiber, M. R., et al. 2021, ApJS, 256, 45
Julian Gerber https://orcid.org/0000-0001-6470-6553 Bonnet-Bidaud, J. M., de Martino, D., Falanga, M., Mouchet, M., &
Kyung Duk Yoon https://orcid.org/0000-0001-8532-1273 Masetti, N. 2007, A&A, 473, 185
Gabriel Bridges https://orcid.org/0000-0002-6653-4975 Byckling, K., Mukai, K., Thorstensen, J. R., & Osborne, J. P. 2010, MNRAS,
Charles J. Hailey https://orcid.org/0000-0002-3681-145X 408, 2298
Frank Haberl https://orcid.org/0000-0002-0107-5237 Canalle, J. B. G., Saxton, C. J., Wu, K., Cropper, M., & Ramsay, G. 2005,
A&A, 440, 185
Jaesub Hong https://orcid.org/0000-0002-6089-5390 Coughenour, B. M., Tomsick, J. A., Shaw, A. W., et al. 2022, MNRAS, 511, 4582
Jonathan Grindlay https://orcid.org/0000-0002-1323-5314 Frank, J., King, A., & Raine, D. J. 2002, Accretion Power in Astrophysics
Gabriele Ponti https://orcid.org/0000-0003-0293-3608 (3rd ed.; Cambridge: Cambridge Univ. Press)
Gavin Ramsay https://orcid.org/0000-0001-8722-9710 Gabriel, C., Denby, M., Fyfe, D. J., et al. 2004, in ASP Conf. Ser. 314,
Astronomical Data Analysis Software and Systems (ADASS) XIII, ed.
F. Ochsenbein, M. G. Allen, & D. Egret (San Francisco, CA: ASP), 759
Gaia Collaboration 2020, yCat, I/350
References Gong, H. 2020, AAS Meeting, 235, 439.05
Gong, H. 2022, ApJ, 933, 240
Álvarez-Hernández, A., Torres, M. A. P., Rodríguez-Gil, P., et al. 2021, Gorgone, N. M., Woudt, P. A., Buckley, D., et al. 2021, ApJ, 923, 243
MNRAS, 507, 5805 Greiss, S., Steeghs, D., Jonker, P. G., et al. 2014, MNRAS, 438, 2839
Álvarez-Hernández, A., Torres, M. A. P., Rodríguez-Gil, P., et al. 2023, Haberl, F., Motch, C., & Zickgraf, F. J. 2002, A&A, 387, 201
MNRAS, 524, 3314 Hailey, C. J., Mori, K., Perez, K., et al. 2016, ApJ, 826, 160
Annala, M., & Poutanen, J. 2010, A&A, 520, A76 Harrison, F. A., Craig, W. W., Christensen, F. E., et al. 2013, ApJ, 770, 103
Anzolin, G., de Martino, D., Bonnet-Bidaud, J.-M., et al. 2008, A&A, Hayashi, T., & Ishida, M. 2014, MNRAS, 438, 2267
489, 1243 Hayashi, T., Kitaguchi, T., & Ishida, M. 2021, MNRAS, 504, 3651
Arnaud, K. A. 1996, in ASP Conf. Ser. 101, Astronomical Data Analysis Software HEASARC 2014, HEAsoft: Unified Release of FTOOLS and XANADU,
and Systems V, ed. G. H. Jacoby & J. Barnes (San Francisco, CA: ASP), 17 Astrophysics Source Code Library, ascl:1408.004

13
The Astrophysical Journal, 954:138 (14pp), 2023 September 10 Vermette et al.

Hellier, C. 1997, MNRAS, 291, 71 Ponman, T. J., Belloni, T., Duck, S. R., et al. 1995, MNRAS, 276, 495
Hillebrandt, W., & Niemeyer, J. C. 2000, ARA&A, 38, 191 Rosen, S. R. 1992, MNRAS, 254, 493
Jonker, P. G., Torres, M. A. P., Hynes, R. I., et al. 2014, ApJS, 210, 18 Saito, R. K., Hempel, M., Minniti, D., et al. 2012, A&A, 537, A107
Koliopanos, F. 2015, PhD thesis, Ludwig-Maximilians-Universität München Saxton, C. J., Wu, K., Cropper, M., & Ramsay, G. 2005, MNRAS, 360, 1091
Koliopanos, F., Peault, M., Vasilopoulos, G., & Webb, N. 2021, MNRAS, 501, 548 Schödel, R., Feldmeier, A., Kunneriath, D., et al. 2014, A&A, 566, A47
Madsen, K., Hickox, R., Bachetti, M., et al. 2019, BAAS, 51, 166 Shaw, A. W., Heinke, C. O., Mukai, K., et al. 2020, MNRAS, 498, 3457
Masetti, N., Parisi, P., Palazzi, E., et al. 2010, A&A, 519, A96 Smith, R. K., Brickhouse, N. S., Liedahl, D. A., & Raymond, J. C. 2001, ApJL,
Meliani, M. T., de Araujo, J. C. N., & Aguiar, O. D. 2000, A&A, 358, 417 556, L91
Mondal, S., Ponti, G., Haberl, F., et al. 2022, A&A, 666, A150 Suleimanov, V., Doroshenko, V., Ducci, L., Zhukov, G. V., & Werner, K.
Mori, K., Hailey, C. J., Schutt, T. Y. E., et al. 2021, ApJ, 921, 148 2016, A&A, 591, A35
Mukai, K. 2017, PASP, 129, 062001 Suleimanov, V. F., Doroshenko, V., & Werner, K. 2019, MNRAS, 482, 3622
Muno, M. P., Bauer, F. E., Baganoff, F. K., et al. 2009, ApJS, 181, 110 Tsujimoto, M., Morihana, K., Hayashi, T., & Kitaguchi, T. 2018, PASJ,
Nasiroglu, I., Słowikowska, A., Kanbach, G., & Haberl, F. 2012, MNRAS, 70, 109
420, 3350 Türler, M., Chernyakova, M., Courvoisier, T. J. L., et al. 2010, A&A, 512, A49
Nobukawa, M., Uchiyama, H., Nobukawa, K. K., Yamauchi, S., & Wada, Q., Tsujimoto, M., Ebisawa, K., & Hayashi, T. 2017, PASJ, 69, 10
Koyama, K. 2016, ApJ, 833, 268 Wang, Q. D., Gotthelf, E. V., & Lang, C. C. 2002, Natur, 415, 148
Norton, A. J., & Watson, M. G. 1989a, MNRAS, 237, 715 Warner, B. 2003, Cataclysmic Variable Stars (Cambridge: Cambridge Univ.
Norton, A. J., & Watson, M. G. 1989b, MNRAS, 237, 853 Press)
Norton, A. J., Wynn, G. A., & Somerscales, R. V. 2004, ApJ, 614, 349 Wevers, T., Hodgkin, S. T., Jonker, P. G., et al. 2016, MNRAS, 458, 4530
Pala, A. F., Gänsicke, B. T., Belloni, D., et al. 2022, MNRAS, 510, 6110 Wilms, J., Allen, A., & McCray, R. 2000, ApJ, 542, 914
Pala, A. F., Gänsicke, B. T., Breedt, E., et al. 2020, MNRAS, 494, 3799 Wu, K., Chanmugam, G., & Shaviv, G. 1994, ApJ, 426, 664
Pecaut, M. J., & Mamajek, E. E. 2013, ApJS, 208, 9 Xu, X.-j., Li, Z., Zhu, Z., et al. 2019a, ApJ, 882, 164
Perez, K., Hailey, C. J., Bauer, F. E., et al. 2015, Natur, 520, 646 Xu, X.-j., Wang, Q. D., & Li, X.-D. 2016, ApJ, 818, 136
Perez, K., Krivonos, R., & Wik, D. R. 2019, ApJ, 884, 153 Xu, X.-j., Yu, Z.-l., & Li, X.-d 2019b, ApJ, 878, 53
Piirola, V., Vornanen, T., Berdyugin, A., & Coyne, S. J. G. V. 2008, ApJ, Yuasa, T., Makishima, K., & Nakazawa, K. 2012, ApJ, 753, 129
684, 558 Zorotovic, M., Schreiber, M. R., & Gänsicke, B. T. 2011, A&A, 536, A42

14

You might also like