ArticlePDF Available

Discovery of a Double-detonation Thermonuclear Supernova Progenitor

Authors:

Abstract

We present the discovery of a new double detonation progenitor system consisting of a hot subdwarf B (sdB) binary with a white dwarf companion with an P=76.34179(2) min orbital period. Spectroscopic observations are consistent with an sdB star during helium core burning residing on the extreme horizontal branch. Chimera light curves are dominated by ellipsoidal deformation of the sdB star and a weak eclipse of the companion white dwarf. Combining spectroscopic and light curve fits we find a low mass sdB star, M_(sdB) = 0.383 ± 0.028 M_⊙ with a massive white dwarf companion, M_(WD) = 0.725 ± 0.026 M_⊙. From the eclipses we find a blackbody temperature for the white dwarf of 26,800 K resulting in a cooling age of ≈25 Myrs whereas our MESA model predicts an sdB age of ≈170 Myrs. We conclude that the sdB formed first through stable mass transfer followed by a common envelope which led to the formation of the white dwarf companion ≈25 Myrs ago. Using the MESA stellar evolutionary code we find that the sdB star will start mass transfer in ≈6 Myrs and in ≈60 Myrs the white dwarf will reach a total mass of 0.92 M_⊙ with a thick helium layer of 0.17 M_⊙. This will lead to a detonation that will likely destroy the white dwarf in a peculiar thermonuclear supernova. PTF1 2238+7430 is only the second confirmed candidate for a double detonation thermonuclear supernova. Using both systems we estimate that at least ≈1% of white dwarf thermonuclear supernovae originate from sdB+WD binaries with thick helium layers, consistent with the small number of observed peculiar thermonuclear explosions.
Draft version October 26, 2021
Typeset using L
A
T
E
X default style in AASTeX63
Discovery of a double detonation thermonuclear supernova progenitor
Thomas Kupfer ,1Evan B. Bauer ,2Jan van Roestel ,3Eric C. Bellm ,4Lars Bildsten,5, 6 Jim Fuller,3
Thomas A. Prince,3Ulrich Heber ,7Stephan Geier,8Matthew J. Green,9Shrinivas R. Kulkarni ,3
Steven Bloemen,10 Russ R. Laher ,11 Ben Rusholme ,11 and David Schneider7
1Department of Physics and Astronomy, Texas Tech University, PO Box 41051, Lubbock, TX 79409, USA
2Center for Astrophysics |Harvard & Smithsonian, 60 Garden St, Cambridge, MA 02138, USA
3Division of Physics, Mathematics and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA
4DIRAC Institute, Department of Astronomy, University of Washington, 3910 15th Avenue NE, Seattle, WA 98195, USA
5Kavli Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA
6Department of Physics, University of California, Santa Barbara, CA 93106, USA
7Dr. Karl Remeis-Observatory & ECAP, Astronomical Institute, Friedrich-Alexander University Erlangen-Nuremberg (FAU),
Sternwartstr. 7, 96049 Bamberg, Germany
8Institut ur Physik und Astronomie, Universit¨at Potsdam, Haus 28, Karl-Liebknecht-Str. 24/25, D-14476 Potsdam-Golm, Germany
9Department of Astrophysics, School of Physics and Astronomy, Tel Aviv University, Tel Aviv 6997801, Israel
10Department of Astrophysics/IMAPP, Radboud University Nijmegen, P.O. Box 9010, 6500 GL Nijmegen, The Netherlands
11IPAC, California Institute of Technology, 1200 E. California Blvd, Pasadena, CA 91125, USA
(Received June 1, 2019; Revised January 10, 2019; Accepted October 26, 2021)
Submitted to ApJL
ABSTRACT
We present the discovery of a new double detonation progenitor system consisting of a hot sub-
dwarf B (sdB) binary with a white dwarf companion with an Porb=76.34179(2) min orbital period.
Spectroscopic observations are consistent with an sdB star during helium core burning residing on
the extreme horizontal branch. Chimera light curves are dominated by ellipsoidal deformation of the
sdB star and a weak eclipse of the companion white dwarf. Combining spectroscopic and light curve
fits we find a low mass sdB star, MsdB = 0.383 ±0.028 Mwith a massive white dwarf companion,
MWD = 0.725 ±0.026 M. From the eclipses we find a blackbody temperature for the white dwarf
of 26,800 K resulting in a cooling age of 25 Myrs whereas our MESA model predicts an sdB age of
170 Myrs. We conclude that the sdB formed first through stable mass transfer followed by a common
envelope which led to the formation of the white dwarf companion 25 Myrs ago.
Using the MESA stellar evolutionary code we find that the sdB star will start mass transfer in 6 Myrs
and in 60 Myrs the white dwarf will reach a total mass of 0.92 Mwith a thick helium layer of 0.17 M.
This will lead to a detonation that will likely destroy the white dwarf in a peculiar thermonuclear super-
nova. PTF1 J2238+7430 is only the second confirmed candidate for a double detonation thermonuclear
supernova. Using both systems we estimate that at least 1 % of white dwarf thermonuclear super-
novae originate from sdB+WD binaries with thick helium layers, consistent with the small number of
observed peculiar thermonuclear explosions.
Keywords: editorials, notices miscellaneous catalogs surveys
1. INTRODUCTION
Most hot subdwarf B stars (sdBs) are core helium burning stars with masses around 0.5 Mand thin hydrogen
envelopes (Heber 1986,2009,2016). A large number of sdB stars are in close orbits with orbital periods of Porb <10 days
Corresponding author: Thomas Kupfer
tkupfer@ttu.edu
arXiv:2110.11974v1 [astro-ph.SR] 22 Oct 2021
2Kupfer et al.
(Napiwotzki et al. 2004;Maxted et al. 2001), with the most compact systems reaching orbital periods of .1 hour (e.g.
Vennes et al. 2012;Geier et al. 2013;Kupfer et al. 2017a,b,2020a,b). The only way to form such tight binaries is
orbital shrinkage through a common envelope phase followed by the loss of angular momentum due to the radiation
of gravitational waves (Han et al. 2002,2003;Nelemans 2010).
SdB binaries with white dwarf (WD) companions which exit the common envelope phase at Porb.2 hours will reach
contact while the sdB is still burning helium (Bauer & Kupfer 2021). Due to the emission of gravitational waves the
orbit of the binary will shrink until the sdB fills its Roche Lobe at a period of 30 100 min, depending on the
evolutionary stage and envelope thickness of the hot subwarf (e.g. Savonije et al. 1986;Tutukov & Fedorova 1989;
Tutukov & Yungelson 1990;Iben & Tutukov 1991;Yungelson 2008;Piersanti et al. 2014;Brooks et al. 2015;Neunteufel
et al. 2019;Bauer & Kupfer 2021).
The known population of sdB + WD binaries consists mostly of systems with orbital periods too large to start
accretion before the sdB turns into a WD (Kupfer et al. 2015). Currently only four detached systems with a WD
companion are known to have Porb<2 hours (Vennes et al. 2012;Geier et al. 2013;Kupfer et al. 2017a,b;Pelisoli
et al. 2021). Just recently Kupfer et al. (2020a,b) discovered the first two Roche lobe filling hot subdwarfs as part
of a high-cadence Galactic Plane survey using the Zwicky Transient Facility (Kupfer et al. 2021). Both systems can
be best explained as Roche Lobe filling sdOB stars which have started mass transfer to a WD companion. The light
curves in both systems show deep eclipses from an accretion disc. Due to their high effective temperatures, both sdOB
stars are predicted to be in a short lived phase where the sdOB undergoes residual hydrogen shell burning.
The most compact known sdB binary where the sdB is still undergoing core-helium burning is CD–3011223. The
binary has an orbital period Porb =70.5 min and a high mass WD companion (MWD 0.75 M;Vennes et al. 2012;
Geier et al. 2013). The sdB in CD–3011223 will begin transferring helium to its WD companion in 40 Myr when
the system has shrunk to an orbital period Porb 40 min. After the WD accretes 0.1 M, helium burning is predicted
to be ignited unstably in the accreted helium layer on the WD surface (Brooks et al. 2015;Bauer et al. 2017). This
could either disrupt the WD even when the mass is significantly below the Chandrasekhar mass, a so-called double
detonation supernova (e.g. Livne 1990;Livne & Arnett 1995;Fink et al. 2010;Woosley & Kasen 2011;Wang & Han
2012;Shen & Bildsten 2014;Wang 2018) or just detonate the He-shell without disrupting the WD which results in a
faint and fast .Ia supernova with subsequent weaker He-flashes (Bildsten et al. 2007;Brooks et al. 2015). Therefore,
systems like CD–3011223 are predicted to be either the progenitors for double detonation thermonuclear supernovae
or perhaps faint and fast .Ia supernovae that do not disrupt the WD.
De et al. (2019,2020) presented the discovery of a sample of calcium-rich transients consistent with a thick helium
shell double detonation on a sub-Chandrasekhar-mass WD (Polin et al. 2019,2021). The majority of these transients
are located in old stellar populations with only a small sub-sample found in in star forming environments.
The question remains just how common systems like CD–3011223 are. To address this question we have conducted a
search for (ultra-)compact post-common envelope systems using the Palomar Transient Factory (PTF; Law et al. 2009;
Rau et al. 2009) and subsequently the Zwicky Transient Facility (ZTF; Graham et al. 2019;Masci et al. 2019) based
on a color selected sample from Pan-STARRS data release 1. The PTF used the Palomar 4800 Samuel Oschin Schmidt
telescope to image up to 2000 deg2of the sky per night to a depth of Rmould 20.6 mag or g021.3 mag. PTF was
succeeded by the Zwicky Transient Facility which started science operation in March 2018 using the same telescope but
a new camera with a field-of-view of 47 deg2. Here we report the discovery of a new thermonuclear supernova double
detonation progenitor system consisting of an sdB with a WD companion: PTF1 J223857.11+743015.1 (hereafter
PTF1 J2238+7430) with orbital period of 76 min showing similar properties to CD–3011223.
2. OBSERVATIONS
2.1. Photometry
As part of the Palomar Transient Factory (PTF), the Palomar 48-inch (P48) telescope imaged the sky every night.
The reduction pipeline for PTF applies standard de-biasing, flat-fielding, and astrometric calibration to raw images
(Laher et al. 2014). Relative photometry correction is applied and absolute photometric calibration to the few per-
cent level is performed using a fit to SDSS fields observed in the same night (Ofek et al. 2012). The lightcurve of
PTF1 J2238+7430 has 144 epochs, with good photometry in the Rmould band with a typical uncertainty of 0.01-
0.02 mag. The majority of observations were conducted during the summer months June - August 2013 and 2014 and
the cadence is highly irregular, ranging from a few minutes to years. The object was also observed as part of the
Zwicky Transient Facility (ZTF) public survey (Graham et al. 2019;Bellm et al. 2019). Image processing of ZTF data
A new double detonation progenitor system 3
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Phase
15.35
15.40
15.45
15.50
15.55
Magnitude [mag]
ZTF-
r
PTF-
R
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
400
300
200
100
0
100
200
300
400
RV (km\,s 1)
DBSP
ISIS
HIRES
ESI
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Phase
25
0
25
RV (km\,s 1)
Figure 1. Left panel: Phase folded at Porb =76.341750 min ZTF and PTF light curve for PTF1 J2238+7430. Right panel:
Radial velocity plotted against orbital phase for PTF1 J2238+7430. The RV data were phase folded with the orbital period and
are plotted twice for better visualization. The residuals are plotted below.
is described in full detail in Masci et al. (2019). We extracted the light curve from ZTF data release 6 which consists
of 34 observations in ZTF-rtaken randomly over 1.5 years between August 2018 and November 2019.
High-cadence observations were conducted using the Palomar 200-inch telescope with the high-speed photometer
CHIMERA (Harding et al. 2016) which is a 2-band photometer which uses frame-transfer, electron-multiplying CCDs
to achieve 15 ms dead time covering a 5×5 arcmin field of view. Simultaneous optical imaging in two bands is enabled
by a dichroic beam splitter centered at 567nm. Data reduction was carried out with the ULTRACAM pipeline (Dhillon
et al. 2007) customized for CHIMERA. All frames were bias-subtracted and flat-fielded. 1300 observations in g0and
r0with a 5 sec exposure time were obtained on 2017-07-26 and 2700 observations in g0and i0with a 4sec exposure
time were obtained on 2017-12-14.
2.2. Spectroscopy
Optical spectra were obtained with the Palomar 200-inch telescope and the Double-Beam Spectrograph (DBSP; Oke
& Gunn 1982) using a low resolution mode (R1500). 31 consecutive exposures were obtained on 2017-05-25 and
2017-05-29 and 15 consecutive exposures were obtained on 2017-05-25 using a 180 sec exposure time. Each night an
average bias and normalized flat-field frame was made out of 10 individual bias and 10 individual lamp flat-fields. To
account for telescope flexure, an arc lamp was taken at the position of the target after each observing sequence. For the
blue arm, FeAr and for the red arm, HeNeAr arc exposures were taken. Both arms of the spectrograph were reduced
using a custom PyRAF-based pipeline 1(Bellm & Sesar 2016). The pipeline performs standard image processing and
spectral reduction procedures, including bias subtraction, flat-field correction, wavelength calibration, optimal spectral
extraction, and flux calibration.
Additionally PTF1 J2238+7430 was also observed with the William Herschel Telescope (WHT) and the ISIS spec-
trograph (Carter et al. 1993) using a medium resolution mode (R600B grating, R2500). 10 consecutive exposures
with an exposure time of 180 sec were obtained on 2017-07-26. 10 bias frames were obtained to construct an average
bias frame and 10 individual lamp flat-fields were obtained to construct a normalized flat-field. CuNeAr arc exposures
were taken before and after the observing sequence to correct for instrumental flexure. One dimensional spectra were
extracted using optimal extraction and were subsequently wavelength and flux calibrated.
To obtain high-resolution spectra, PTF1 J2238+7430 was observed with Keck/HIRES and Keck/ESI. We obtained
5 consecutive exposures with Keck/HIRES on 2017-08-14 and 2017-08-30 as well as 14 consecutive exposures with
Keck/ESI on 2018-07-20. ThAr arc exposures were taken at the beginning of the night. The spectra were reduced
1https://github.com/ebellm/pyraf-dbsp
4Kupfer et al.
3
2.8
2.6
2.4
2.2
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
normalized flux
H10
H9
H8
Hǫ
Hβ
60
40
20
0
-20
-40
-60
3.6
3.4
3.2
λ(˚
A)
H12
H11
2.4
2.2
2
1.8
1.6
1.4
1.2
ESI Hei4471
0.8
0.6
0.4
1
ESI Hei4921
ESI Hei6678
HIRES Hei4026
HIRES Hei4471
HIRES Hei4921
10
5
0
-5
-10
-15
2.8
2.6
normalized flux
ESI Hei4026
λ(˚
A)
Figure 2. Left panel: Fit of synthetic LTE models to the hydrogen Balmer lines of a coadded DBSP spectrum. The normalized
fluxes of the single lines are shifted for better visualisation. Right panel: Fits of vrot sin ito the helium lines seen in the HIRES
and ESI spectra. The atmospheric parameters were fixed to the values derived from the WHT and DBSP spectra.
using the MAKEE2pipeline following the standard procedure: bias subtraction, flat fielding, sky subtraction, order
extraction, and wavelength calibration.
3. ORBITAL AND ATMOSPHERIC PARAMETERS AND LIGHT CURVE FITTING
As evident in Fig. 1PTF1 J2238+7430 shows strong periodic ellipsoidal variability in its light curve at Porb =
76.341750(1) min. This variability is caused by the tidal deformation of the sdB primary under the influence of the
gravitational force of the companion. We use the PTF and the ZTF lightcurve with its multi-year baseline and the
Chimera light curves to derive the orbital period of the systems. The analysis was done with the Gatspy module for
time series analysis which uses the Lomb-Scargle periodogram3(VanderPlas & Ivezi´c 2015). The error was derived
from bootstrapping.
Radial velocities were measured by fitting Gaussians, Lorentzians, and polynomials to the hydrogen and helium
lines to cover continuum, line, and line core of the individual lines using the FITSB2 routine (Napiwotzki et al. 2004).
The procedure is described in full detail in Geier et al. (2011). We fitted the wavelength shifts compared to the rest
wavelengths using a χ2-minimization. Assuming circular orbits, a sine curve was fitted to the folded radial velocity
(RV) data points (Fig. 1).
Atmospheric parameters such as effective temperature, Teff , surface gravity, log g, helium abundance, log y=
log n(He)
n(H), and projected rotational velocity, vrot sin i, were determined by fitting the rest-wavelength corrected av-
erage DBSP, ISIS and HIRES spectra with metal-line-blanketed LTE model spectra (Heber et al. 2000). Teff and
log gwere derived from the Balmer and helium lines from the ISIS and DBSP spectra whereas log yand vrot sin i
were measured with the HIRES spectra. High-resolution echelle spectra are not well suited to measure Teff and log g
because the broad hydrogen absorption lines span several orders and merging of the echelle spectra could introduce
2http://www.astro.caltech.edu/tb/ipac staff/tab/makee/
3http://dx.doi.org/10.5281/zenodo.14833
A new double detonation progenitor system 5
Table 1. Overview of the measured and derived parameters for PTF1 J2238+7430
Right ascension RA [hrs] 22:38:57.11
Declination Dec [] +74:30:15.1
Magnitudebg[mag] 15.244±0.023
Parallaxa$[mas] 1.0001 ±0.0225
Distance d[kpc] 1.00 ±0.03
Absolute Magnitude Mg[mag] 4.40 ±0.20
(reddening corrected)
Proper motiona(RA) µαcos(δ) [mas yr1] 0.344 ±0.056
Proper motiona(Dec) µδ[mas yr1]1.833 ±0.051
Atmospheric parameters of the sdB
Effective temperaturecTeff [K] 23 600±400
Surface gravityclog g5.42±0.06
Helium abundancedlog y2.11±0.03
Projected rotational velocitydvrot sin i[km s1] 185±5
Orbital parameters
T0[BMJD UTC] 57960.47584170(3)
Orbital period Porb [min] 76.341750(1)
RV semi-amplitude K[km s1] 378.0±3.7
System velocity γ[km s1]6.2±2.14
Binary mass function fm[M] 0.0597±0.0020
Derived parameters
Mass ratio q=MWD
MsdB 0.528 ±0.020
sdB mass MsdB [M] 0.383 ±0.028
sdB radius RsdB [R] 0.190 ±0.003
WD mass MWD [M] 0.725 ±0.026
WD radius RWD [M] 0.0109+0.0002
0.0003
WD blackbody temperature Teff [K] 26,800 ±4600
Orbital inclination i[] 88.4+1.6
3.3
Separation a[R] 0.615 ±0.010
Roche filling factor RsdB/RRochelobe 0.951 ±0.010
afrom Gaia eDR3 (Gaia Collaboration et al. 2016,2021)
bfrom PanSTARRS DR1 (Chambers et al. 2016)
cadopted from from DBSP and ISIS
dadopted from ESI and HIRES
systematic errors. The full procedure is described in detail in Kupfer et al. (2017a,b). PTF1 J2238+7430 shows typical
Teff , log g, and log yand vrot sin i=185±5 km s1. The rotational velocity is consistent with a tidally locked sdOB star
(see Sec. 4.1). Table 1summarizes the atmospheric and orbital parameters.
To model the lightcurves obtained with CHIMERA we used the LCURVE code (Copperwheat et al. 2010). We use
a Roche geometry, and the free parameters in our fit are: the phase (t0), the scaled radii (r1,2), the mass ratio
q, the inclination i, secondary temperature TWD, and the velocity scale ([K+KWD ]/sin i). We use a passband-
dependent gravity-darkening law and use a gravity darkening value (yg,r) from Claret & Bloemen (2011) and find
β= 0.425 for g0,β= 0.395 for r0, and β= 0.37 for i0. We assume an uncertainty of 0.03 on the value and use a
Gaussian prior. We use fixed limb darkening coefficients (a1,a2,a3,a4) taken from Claret & Bloemen (2011). We
use a1= 0.82, a2=0.65, a3= 0.55, and a4=0.19 for g0,a1= 0.81, a2=0.89, a3= 0.79, and a4=0.27
for r0, and a1= 0.78, a2=1.01, a3= 0.91, and a4=0.31 for i0. We also model the relativistic beaming (F) as
in Bloemen et al. (2011). We calculate the beaming parameters by assuming a blackbody spectrum and using the
effective wavelength of the g0,r0, and i0filters. We find F= 1.80 for g0,F= 1.57 for r0, and F= 1.46 for i0. The full
6Kupfer et al.
0.47
3.6
3.4
3.2
3
0.46
0.45
0.44
0.43
0.42
0.1
0
-0.1
rel. Flux
∆Flux
g, 2017-07-26
MJD - 57960 [days]
0.49
0.48
3.8
0.08
3.8
3.6
3.4
3.2
3
0.2
0
-0.2
0.2
0.18
0.16
g, 2017-12-14
0.14
0.12
0.1
4
∆Flux
MJD - 58101 [days]
rel. Flux
1.6
1.5
1.4
1.3
0.42
0.05
0
r, 2017-07-26
-0.05
MJD - 57960 [days]
rel. Flux
∆Flux
0.49
0.48
0.47
0.46
0.45
0.44
0.43
0.1
1.6
1.5
1.4
1.3
0
-0.1
0.2
0.18
0.16
0.14
0.12
0.1
0.08
1.7
rel. Flux
i, 2017-12-14
MJD - 58101 [days]
∆Flux
3.05
3
2.95
0.472
MJD -57960 [days]
3.2
3.15
3.1
0.48
0.478
rel. Flux
0.476
0.474
rel. Flux
1.33
MJD -57960 [days]
0.48
0.478
0.476
0.474
0.472
1.43
1.42
1.41
1.4
1.39
1.38
1.37
1.36
1.35
1.34
Figure 3. Chimera light curves un-binned (grey) and binned (black) shown together with the LCURVE fits (red) observed optical
SDSS bandpasses. The lower two panels show the region when the WD is being eclipsed by the sdB. The blue solid curve marks
the same model without eclipses of the WD. The lower panels show the region when the white dwarf is being eclipsed. Lower
left panel: g0light curve, Lower right panel: r0light curve
approach is also described in Kupfer et al. (2017a,b,2020a,b) and Ratzloff et al. (2019). In addition, we add a 2nd
order polynomial to correct for any long timescale trends which are the result of a changing airmass over the course of
the observations. The best value of χ2for this model was 1350 for 1300 data points for the g-band light curve which
includes also a weak eclipse of the hot WD. Although the eclipse is weak (1 %), the χ2for the non-eclipsing solution
is 1400 which is statistically significantly worse compared to the solution with the weak eclipse. We use the MCMC
sampler emcee (Foreman-Mackey et al. 2013) to determine the best-fit values and uncertainty on the parameters.
4. RESULTS
4.1. System parameters
Although, PTF1 J2238+7430 is a single-lined binary we can derive system parameters using the combined results
from the light curve analysis with results from the spectroscopic fitting. Parameters derived in this way by a si-
A new double detonation progenitor system 7
multaneous fit to the Chimera light curves are summarized in Table 1. The given errors are all 95 % confidence
limits.
We find that PTF1 J2238+7430 consists of a low mass sdB with a high-mass WD companion. We derive a mass
ratio q=MsdB/MWD = 0.528 ±0.020, a mass for the sdB MsdB = 0.383 ±0.028 M, and a WD companion mass
MWD = 0.725 ±0.026 M. PTF1 J2238+7430 is found to be eclipsing at an inclination angle of i= 88.4+1.6
3.3
which
allows us to measure the radius and the black-body temperature of the WD companion. We determine a black-body
temperature of 26,800 ±4600 K for the WD and a radius of RWD = 0.0109+0.0002
0.0003 R. The radius was found to be
<5% above the zero-temperature value and is fully consistent with predictions of at most a few percent above the
zero-temperature value for the radius.
We calculate the absolute magnitude (Mg) of PTF1 J2238+7430 using the visual PanSTARRS g-band magnitude
g=15.244±0.023 mag and the parallax from Gaia eDR3 (Gaia Collaboration et al. 2016,2021). Because the object is
located near the Galactic Plane, significant reddening can occur. Green et al. (2019) present updated 3D extinction
maps based on Gaia parallaxes and stellar photometry from Pan-STARRS 1 and 2MASS4and find towards the
direction of PTF1 J2238+7430 an extinction of E(gr) = 0.24 ±0.03 at a distance of 1.00 kpc; this results in a total
extinction in the g-band of Ag= 0.84 ±0.11 mag, and with the corrected magnitude, we find an absolute magnitude
of Mg= 4.40 ±0.20 mag consistent with a hot subdwarf star (Geier et al. 2019).
4.2. Comparison with Gaia parallax
To test whether our derived system parameters are consistent with the parallax provided by Gaia eDR3, we compared
the measured parameters from the light curve fit to the predictions using the Gaia parallax. The approach follows
a similar strategy as described in Ratzloff et al. (2019) and Kupfer et al. (2020a). Using the absolute magnitude
Mg= 4.40 ±0.20 mag, we find a luminosity of L= 11.5±3.0 Lusing a bolometric correction BCg=2.30 mag
derived for our stellar parameters from the MESA Isochrones & Stellar Tracks (MIST; Dotter 2016;Choi et al. 2016;
Paxton et al. 2011,2013,2015,2018). Using the Stefan-Boltzmann law applied to a black body (L= 4σπR2
sdBT4
eff ),
we can solve for the radius of the sdBs, and combined with R2
sdB =GMsdB/g, we can solve for mass of the sdBs:
MsdB =LsdB10log(g)
4πσGT 4
eff
(1)
Using these equations we find MsdB = 0.39 ±0.10 Mand RsdB = 0.17 ±0.03 R. Although the error bars are rather
large, this result is in agreement with the results from the light curve and spectroscopic fits.
4.3. Kinematics of the binary systems
We found that PTF1 J2238+7430 has evolved from a 2.1 Mstar and we expect the system is a member of a
young stellar population (see Sec. 5). Using the proper motion from Gaia eDR3 (Gaia Collaboration et al. 2016,2018,
2021), the distance and the systemic velocities (see Tab. 1) we calculate the Galactic motion for PTF1 J2238+7430.
We employed the approach described in Odenkirchen & Brosche (1992) and Pauli et al. (2006). As in Kupfer et al.
(2020a), we use the Galactic potential of Allen & Santillan (1991) as revised by Irrgang et al. (2013). The orbit was
integrated from the present to 3 Gyr into the past. We find that the binary moves within a height of 200 parsec of the
Galactic equator and with very little eccentricity between 9 and 10 kpc from the Galactic center. From the Galactic
orbit we conclude that PTF1 J2238+7430 is a member of the Galactic thin disk population consistent with being
member of a young stellar population.
5. PREDICTED EVOLUTION OF THE BINARY SYSTEM
5.1. Formation of the sdB + WD system
Ruiter et al. (2010) found that the dominant way to form compact double carbon-oxygen core WDs is through stable
mass transfer which forms the sdB followed by a phase of unstable mass transfer which forms the white dwarf com-
panion. They present a specific example which starts with a 2.88 Mand 2.45 Mbinary pair. In PTF1 J2238+7430
weak eclipses of the WD companion imply a blackbody temperature of 26 800 ±4600 K, implying a cooling time of
25 million years, significantly shorter than the predicted current age of the sdB of 170 million years (see Sec. 5.2).
4http://argonaut.skymaps.info/
8Kupfer et al.
1
2
3
5
MS MS
MS
post-MS
MS
sdB
RLOF
CEE
4
sdB WD
GWs
GWs
MS - main-sequence star RLOF - stable Roche-lobe overflow
CEE - common envelope evolution GW - gravitational waves
Peculiar SN Ia
6
7
8
Runaway
sdB star
RLOF
Current phase
Figure 4. Visualization of the proposed evolutionary pathway for PTF1 J2238+7430. The red box marks the current evolu-
tionary phase. Each evolutionary phase is numbered according to their order in the evolution and the direction of the sequence
is marked with arrows.
Therefore, we predict that the sdB was formed first, and we propose the following evolutionary scenario (illustrated
in Fig. 4) for PTF1 J2238+7430 which explains all observational properties and is similar to the scenario discussed in
Ruiter et al. (2010).
The system started as a 2.14 Mmain sequence star (see Sect. 5.2) which will become the sdB, and a slightly
lower mass companion with an orbital period of a few weeks. The sdB progenitor evolves first and starts stable mass
transfer to the companion star. At the end of that phase the sdB has formed with the observed mass of 0.4 Mand
the orbital periods has substantially widened. The companion star has accreted 1.7 Mof material from the sdB
progenitor and turned into a 3.5–4 Mstar which will then evolve off the main sequence and overflow its Roche Lobe
while the sdB star is still burning helium. Due to the large mass ratio at this point, mass transfer will be unstable
and initiate a common envelope. The CE phase could happen either during the RGB or AGB phase of the secondary
depending on the binary separation at that point. In either case it would leave a compact binary with a massive WD
and an sdB at an orbital period of 86 minutes. The observed high WD mass of 0.725 ±0.026 Mis consistent with
the evolution from an intermediate mass main sequence star (Cummings et al. 2018). The final phase of unstable
mass transfer happened 25 million years ago, after which the WD cooled to its currently observed temperature
while gravitational wave radiation decreased the orbital period to the currently observed period of 76 minutes. As also
discussed in Ruiter et al. (2010), there could exist a substantial fraction of compact sdB+WD binaries where the sdB
was formed first through stable mass transfer.
5.2. Future evolution
To understand the future evolution of the system we employed MESA version 12115 (Paxton et al. 2011,2013,2015,
2018,2019). Bauer & Kupfer (2021) use MESA models to show that sdB stars with mass M.0.47 Mcan descend
A new double detonation progenitor system 9
220002400026000280003000032000 Teff [K]
5.2
5.4
5.6
5.8
log(g/cms2)
Binary model
Single-star model
10.0
7.5
log(˙
M/Myr1)
0.2
0.4
MsdB [M]
0.8
0.9
MWD [M]
He-shell detonation
0 20 40 60
Myr from now
25
50
75
Porb [min]
Figure 5. Left panel: Predicted evolution based on the MESA model for the PTF1 J2238+7430 system. The current observed
log gand Teff and error bars for the system are shown in red. The dashed curve shows the evolution the star would follow in
isolation, while the solid curve shows the trajectory it follows due to encountering the Roche limit, depicted by the gray shaded
region. Right panel: Future evolution of the system until the helium ignites.
from either lower-mass main sequence progenitors that ignite central He burning via an off-center degenerate He flash
(MZAMS .2.3 M), or they can descend from higher-mass main sequence progenitors that ignite He at the center under
non-degenerate conditions (MZAMS &2.3 M). They show that these scenarios lead to different H envelope structures
that influence the subsequent radius evolution of the sdB star, with stars descended from higher mass progenitors
having more compact envelopes and correspondingly higher log gvalues. The measured log gfor PTF1 J2238+7430
requires a relatively extended envelope with a radius that requires that sdB star descend from the lower-mass channel
with a progenitor mass around 2 M. We find that our best matching MESA model for the measured log gand Teff
of this system is a 0.41 MsdB model descended from a 2.14 Mmain sequence star that ignited the He core via a
degenerate He-core flash. This model has a sharp transition from the He core to an H envelope with solar composition.
When He ignites, we remove most of the envelope, leaving a thin H envelope layer of 103Mso that the subsequent
sdB evolution track matches the observed log gand Teff of PTF1 J2238+7430. Figure 5shows the log g–Teff evolution
of this MESA model, where it approaches the current observed state of PTF1J2238+7430 after 170 Myr of evolution,
and will encounter its Roche lobe and begin transferring mass soon after.
We model the future binary evolution of this system with a 0.75 MWD companion using the MESA binary capa-
bilities. The sdB is currently observed at 95% Roche Lobe filling and will continue to spiral in due to gravitational
wave radiation. In our model the sdB will soon fill its Roche lobe and start to donate its hydrogen rich envelope in
six million years at a low rate of .1010 Myr1(see Bauer & Kupfer 2021, for a detailed oberview). Because of the
large initial radius of the H envelope, mass transfer will proceed at this low rate for 50 Myr before the H envelope
is exhausted and the He core is finally exposed at a much more compact radius. While the sdB is still helium core
burning 60 Myr from today, the sdB will begin to donate helium rich material onto the WD at the expected rate of
108Myr1, as shown in Figure 5. A helium rich layer will slowly build up for 10million years, reaching a critical
mass of 0.17 M. At this point the binary has an orbital period of 10 min. The sdB has been stripped down to a
mass of 0.25 M, and the WD has a total mass of 0.92 M.
10 Kupfer et al.
Our MESA model predicts that at this point the accreting WD will experience a thermonuclear instability that will
lead to a detonation that will likely destroy the WD in a thermonuclear supernova (Woosley & Kasen 2011;Bauer
et al. 2017). The structure of our MESA model at the point of detonation is very similar to the model for CD–3011223
in Bauer et al. (2017), and we have used similar modeling assumptions as those described in that work. At the time
of the thermonuclear supernova the sdB has an orbital velocity of 911 km s1and will be released as a hyper-runaway
star exceeding the escape velocity of the Galaxy (Bauer et al. 2019;Neunteufel 2020;Neunteufel et al. 2021). Fig. 4
illustrates the evolutionary sequence proposed for PTF1 J2238+7430.
6. SUPERNOVA RATE ESTIMATE
Models of thermonuclear supernovae in WDs with thick (&0.1 M) helium shells indicate that they will yield
transients classified as peculiar Type I supernovae (Polin et al. 2019;De et al. 2019). PTF1 J2238+7430 together
with CD-3011223 therefore mark a small sample of double detonation peculiar thermonuclear supernova progenitors.
Using both systems we can estimate a lower limit of thermonuclear supernovae originating in compact hot subdwarf
+ WD binaries where the sdB donates helium rich material during helium core burning. Both systems will have an
age of 500 Myrs at the time of the helium shell detonation and are located within 1 kpc. Because of their young age,
we compare the rate of these double detonation progenitors to the supernova Ia rate as a function of star formation.
Under the assumption that these systems typically have an age of 500 Myrs at time of explosion we find a lower limit
of double detonation explosions of 2
500 kpc2Myr1from the two known systems. We can compare that to the local
star formation rate of 103Mkpc2yr1which leads to a double detonation rate of 4×106yr1.Sullivan et al.
(2006) found a supernova Ia rate of 3.9±0.7×104SNe yr1(Myr1)1of star formation. With a Galactic star
formation rate of 1Myr1, we find that the rate at which peculiar thermonuclear supernovae with thick 0.15 M
helium shells occur in star forming galaxies could be at least 1 % of the type Ia supernova rate. This is in reasonable
agreement with the presently observed low rate of thick helium shell detonations.
De et al. (2019) presented the discovery of peculiar Type I supernova consistent with a thick helium shell double
detonation on a sub-Chandrasekhar-mass WD (Polin et al. 2019,2021). However, one of the distinct differences is
that the transient occurred in the outskirts of an elliptical galaxy which points to an old stellar population which
is in disagreement with our observed systems which represent a young population. More recently, De et al. (2020)
present a sample of calcium rich transients originating from double-detonations with helium shells. They find that the
majority of transients are located in old stellar populations. However, De et al. (2020) note that a small subsample
(iPTF16hgs, SN2016hnk and SN 2019ofm) were found in star forming environments, suggesting that there is a small
but likely non-zero contribution from young systems which could potentially be related to systems like CD-3011223
and PTF1 J2238+7430.
7. SUMMARY AND CONCLUSION
As part of our search for short period sdB binaries we discovered PTF1J2238+7430 using PTF and subsequently
ZTF light curves. We find a period of Porb=76.34179(2) min. Follow-up observations confirmed the system as an
sdB with MsdB = 0.383 ±0.028 Mand a WD companion with MWD = 0.725 ±0.026 M. High-speed photometry
observations with Chimera revealed a weak WD eclipse which allows us to measure the blackbody temperature and
radius of the WD. We find a temperature of 26,800 ±4600 K and a radius of RWD = 0.0109+0.0002
0.0003 Rfully consistent
with cooling models for carbon-oxygen core WDs. We find a cooling age of 25 Myrs for the WD which is significantly
shorter than our age estimate for the sdB which is 170 Myrs. This can be explained by the sdB forming first through
stable mass transfer, followed by the WD forming 25 Myrs ago through a common envelope phase. This shows that
evolutionary scenarios where the sdB is formed first through stable mass transfer must be considered for compact sdB
binaries with WD companions.
We employed MESA to calculate the future evolution of the system and find that the sdB in PTF1J2238+7430 will
start mass transfer of the hydrogen rich envelope in 6 Myr. In 60 Myr, after a phase of hydrogen and helium
mass transfer, the WD will build up a helium layer of 0.17Mleading to a total WD mass of 0.92 M. Our models
predict that at this point the WD likely detonates in a peculiar thermonuclear supernova making PTF1 J2238+7430
the second known progenitor for a supernova with a thick helium layer. Using both systems we estimate that at least
1 % of type Ia supernova originate from compact sdB+WD binaries in young populations of galaxies with similar star
formation rates compared to the Milky Way. Although this is only a lower limit the estimate is broadly consistent
with the low number of observed peculiar thermonuclear supernovae.
A new double detonation progenitor system 11
ACKNOWLEDGMENTS
This research benefited from interactions that were funded by the Gordon and Betty Moore Foundation through
grant GBMF5076. This work was supported by the National Science Foundation through grants PHY-1748958 and
ACI- 1663688. TK would like to thank Ylva otberg for providing the template for Fig. 4. TK acknowledge support
from the National Science Foundation through grant AST #2107982. DS was supported by the Deutsche Forschungs-
gemeinschaft (DFG) under grants HE 1356/70-1 and IR 190/1-1.
Observations were obtained with the Samuel Oschin Telescope at the Palomar Observatory as part of the PTF
project, a scientific collaboration between the California Institute of Technology, Columbia University, Las Cumbres
Observatory, the Lawrence Berkeley National Laboratory, the National Energy Research Scientific Computing Center,
the University of Oxford, and the Weizmann Institute of Science.
Based on observations obtained with the Samuel Oschin 48-inch Telescope at the Palomar Observatory as part of
the Zwicky Transient Facility project. ZTF is supported by the National Science Foundation under Grant No. AST-
1440341 and a collaboration including Caltech, IPAC, the Weizmann Institute for Science, the Oskar Klein Center at
Stockholm University, the University of Maryland, the University of Washington, Deutsches Elektronen-Synchrotron
and Humboldt University, Los Alamos National Laboratories, the TANGO Consortium of Taiwan, the University of
Wisconsin at Milwaukee, and Lawrence Berkeley National Laboratories. Operations are conducted by COO, IPAC,
and UW.
Some of the data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific
partnership among the California Institute of Technology, the University of California and the National Aeronautics
and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck
Foundation. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the
summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the
opportunity to conduct observations from this mountain.
Some results presented in this paper are based on observations made with the WHT operated on the island of
La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Institutio de
Astrofisica de Canarias.
This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.
int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www.cosmos.esa.int/
web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the
institutions participating in the Gaia Multilateral Agreement.
Facilities: PO:1.2m (PTF), PO:1.2m (ZTF), Hale (DBSP), ING:Herschel (ISIS), Keck:I (HIRES), Keck:II (ESI),
Hale (Chimera)
Software: Gatspy (VanderPlas & Ivezi´c 2015;Vanderplas 2015), FITSB2 (Napiwotzki et al. 2004), LCURVE (Copper-
wheat et al. 2010), emcee (Foreman-Mackey et al. 2013), MESA (Paxton et al. 2011,2013,2015,2018,2019), Matplotlib
(Hunter 2007), Astropy (Astropy Collaboration et al. 2013,2018), Numpy (Oliphant 2015), ISIS (Houck & Denicola
2000), MAKEE (http://www.astro.caltech.edu/tb/ipac staff/tab/makee/)
REFERENCES
Allen, C., & Santillan, A. 1991, RMxAA, 22, 255
Astropy Collaboration, Robitaille, T. P., Tollerud, E. J.,
et al. 2013, A&A, 558, A33,
doi: 10.1051/0004-6361/201322068
Astropy Collaboration, Price-Whelan, A. M., Sip˝ocz, B. M.,
et al. 2018, AJ, 156, 123, doi: 10.3847/1538-3881/aabc4f
Bauer, E. B., & Kupfer, T. 2021, arXiv e-prints,
arXiv:2106.13297. https://arxiv.org/abs/2106.13297
Bauer, E. B., Schwab, J., & Bildsten, L. 2017, ApJ, 845, 97,
doi: 10.3847/1538-4357/aa7ffa
Bauer, E. B., White, C. J., & Bildsten, L. 2019, ApJ, 887,
68, doi: 10.3847/1538-4357/ab4ea4
Bellm, E. C., & Sesar, B. 2016, pyraf-dbsp: Reduction
pipeline for the Palomar Double Beam Spectrograph,
Astrophysics Source Code Library.
http://ascl.net/1602.002
Bellm, E. C., Kulkarni, S. R., Graham, M. J., et al. 2019,
PASP, 131, 018002, doi: 10.1088/1538-3873/aaecbe
Bildsten, L., Shen, K. J., Weinberg, N. N., & Nelemans, G.
2007, ApJL, 662, L95, doi: 10.1086/519489
12 Kupfer et al.
Bloemen, S., Marsh, T. R., Østensen, R. H., et al. 2011,
MNRAS, 410, 1787,
doi: 10.1111/j.1365-2966.2010.17559.x
Brooks, J., Bildsten, L., Marchant, P., & Paxton, B. 2015,
ApJ, 807, 74, doi: 10.1088/0004-637X/807/1/74
Carter, D., et al. 1993
Chambers, K. C., Magnier, E. A., Metcalfe, N., et al. 2016,
arXiv e-prints. https://arxiv.org/abs/1612.05560
Choi, J., Dotter, A., Conroy, C., et al. 2016, ApJ, 823, 102,
doi: 10.3847/0004-637X/823/2/102
Claret, A., & Bloemen, S. 2011, A&A, 529, A75,
doi: 10.1051/0004-6361/201116451
Copperwheat, C. M., Marsh, T. R., Dhillon, V. S., et al.
2010, MNRAS, 402, 1824,
doi: 10.1111/j.1365-2966.2009.16010.x
Cummings, J. D., Kalirai, J. S., Tremblay, P. E.,
Ramirez-Ruiz, E., & Choi, J. 2018, ApJ, 866, 21,
doi: 10.3847/1538-4357/aadfd6
De, K., Kasliwal, M. M., Polin, A., et al. 2019, ApJL, 873,
L18, doi: 10.3847/2041-8213/ab0aec
De, K., Kasliwal, M. M., Tzanidakis, A., et al. 2020, ApJ,
905, 58, doi: 10.3847/1538-4357/abb45c
Dhillon, V. S., Marsh, T. R., Stevenson, M. J., et al. 2007,
MNRAS, 378, 825, doi: 10.1111/j.1365-2966.2007.11881.x
Dotter, A. 2016, ApJS, 222, 8,
doi: 10.3847/0067-0049/222/1/8
Fink, M., opke, F. K., Hillebrandt, W., et al. 2010, A&A,
514, A53, doi: 10.1051/0004-6361/200913892
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman,
J. 2013, PASP, 125, 306, doi: 10.1086/670067
Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al.
2016, A&A, 595, A1, doi: 10.1051/0004-6361/201629272
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al.
2018, A&A, 616, A1, doi: 10.1051/0004-6361/201833051
—. 2021, A&A, 649, A1, doi: 10.1051/0004-6361/202039657
Geier, S., Raddi, R., Gentile Fusillo, N. P., & Marsh, T. R.
2019, A&A, 621, A38, doi: 10.1051/0004-6361/201834236
Geier, S., Hirsch, H., Tillich, A., et al. 2011, A&A, 530,
A28, doi: 10.1051/0004-6361/201015316
Geier, S., Marsh, T. R., Wang, B., et al. 2013, A&A, 554,
A54, doi: 10.1051/0004-6361/201321395
Graham, M. J., Kulkarni, S. R., Bellm, E. C., et al. 2019,
arXiv e-prints. https://arxiv.org/abs/1902.01945
Green, G. M., Schlafly, E., Zucker, C., Speagle, J. S., &
Finkbeiner, D. 2019, ApJ, 887, 93,
doi: 10.3847/1538-4357/ab5362
Han, Z., Podsiadlowski, P., Maxted, P. F. L., & Marsh,
T. R. 2003, MNRAS, 341, 669,
doi: 10.1046/j.1365-8711.2003.06451.x
Han, Z., Podsiadlowski, P., Maxted, P. F. L., Marsh, T. R.,
& Ivanova, N. 2002, MNRAS, 336, 449,
doi: 10.1046/j.1365-8711.2002.05752.x
Harding, L. K., Hallinan, G., Milburn, J., et al. 2016,
MNRAS, 457, 3036, doi: 10.1093/mnras/stw094
Heber, U. 1986, A&A, 155, 33
—. 2009, ARA&A, 47, 211,
doi: 10.1146/annurev-astro-082708- 101836
—. 2016, PASP, 128, 082001,
doi: 10.1088/1538-3873/128/966/082001
Heber, U., Reid, I. N., & Werner, K. 2000, A&A, 363, 198
Houck, J. C., & Denicola, L. A. 2000, in Astronomical
Society of the Pacific Conference Series, Vol. 216,
Astronomical Data Analysis Software and Systems IX,
ed. N. Manset, C. Veillet, & D. Crabtree, 591
Hunter, J. D. 2007, Computing In Science & Engineering,
9, 90, doi: 10.1109/MCSE.2007.55
Iben, Jr., I., & Tutukov, A. V. 1991, ApJ, 370, 615,
doi: 10.1086/169848
Irrgang, A., Wilcox, B., Tucker, E., & Schiefelbein, L. 2013,
A&A, 549, A137, doi: 10.1051/0004-6361/201220540
Kupfer, T., Geier, S., Heber, U., et al. 2015, A&A, 576,
A44, doi: 10.1051/0004-6361/201425213
Kupfer, T., van Roestel, J., Brooks, J., et al. 2017a, ApJ,
835, 131, doi: 10.3847/1538-4357/835/2/131
Kupfer, T., Ramsay, G., van Roestel, J., et al. 2017b, ApJ,
851, 28, doi: 10.3847/1538-4357/aa9522
Kupfer, T., Bauer, E. B., Marsh, T. R., et al. 2020a, ApJ,
891, 45, doi: 10.3847/1538-4357/ab72ff
Kupfer, T., Bauer, E. B., Burdge, K. B., et al. 2020b,
ApJL, 898, L25, doi: 10.3847/2041-8213/aba3c2
Kupfer, T., Prince, T. A., van Roestel, J., et al. 2021,
MNRAS, 505, 1254, doi: 10.1093/mnras/stab1344
Laher, R. R., Surace, J., Grillmair, C. J., et al. 2014, PASP,
126, 674, doi: 10.1086/677351
Law, N. M., Kulkarni, S. R., Dekany, R. G., et al. 2009,
PASP, 121, 1395, doi: 10.1086/648598
Livne, E. 1990, ApJl, 354, L53, doi: 10.1086/185721
Livne, E., & Arnett, D. 1995, ApJ, 452, 62,
doi: 10.1086/176279
Masci, F. J., Laher, R. R., Rusholme, B., et al. 2019,
PASP, 131, 018003, doi: 10.1088/1538-3873/aae8ac
Maxted, P. f. L., Heber, U., Marsh, T. R., & North, R. C.
2001, MNRAS, 326, 1391,
doi: 10.1111/j.1365-8711.2001.04714.x
Napiwotzki, R., Karl, C. A., Lisker, T., et al. 2004,
Astrophysics and Space Science, 291, 321,
doi: 10.1023/B:ASTR.0000044362.07416.6c
Nelemans, G. 2010, Ap&SS, 329, 25,
doi: 10.1007/s10509-010-0392-0
A new double detonation progenitor system 13
Neunteufel, P. 2020, A&A, 641, A52,
doi: 10.1051/0004-6361/202037792
Neunteufel, P., Kruckow, M., Geier, S., & Hamers, A. S.
2021, A&A, 646, L8, doi: 10.1051/0004-6361/202040022
Neunteufel, P., Yoon, S. C., & Langer, N. 2019, A&A, 627,
A14, doi: 10.1051/0004-6361/201935322
Odenkirchen, M., & Brosche, P. 1992, Astronomische
Nachrichten, 313, 69, doi: 10.1002/asna.2113130204
Ofek, E. O., Laher, R., Law, N., et al. 2012, PASP, 124, 62,
doi: 10.1086/664065
Oke, J. B., & Gunn, J. E. 1982, PASP, 94, 586,
doi: 10.1086/131027
Oliphant, T. E. 2015, Guide to NumPy, 2nd edn. (USA:
CreateSpace Independent Publishing Platform)
Pauli, E.-M., Napiwotzki, R., Heber, U., Altmann, M., &
Odenkirchen, M. 2006, A&A, 447, 173,
doi: 10.1051/0004-6361:20052730
Paxton, B., Bildsten, L., Dotter, A., et al. 2011, ApJs, 192,
3, doi: 10.1088/0067-0049/192/1/3
Paxton, B., Cantiello, M., Arras, P., et al. 2013, ApJs, 208,
4, doi: 10.1088/0067-0049/208/1/4
Paxton, B., Marchant, P., Schwab, J., et al. 2015, ApJs,
220, 15, doi: 10.1088/0067-0049/220/1/15
Paxton, B., Schwab, J., Bauer, E. B., et al. 2018, ApJS,
234, 34, doi: 10.3847/1538-4365/aaa5a8
Paxton, B., Smolec, R., Schwab, J., et al. 2019, ApJS, 243,
10, doi: 10.3847/1538-4365/ab2241
Pelisoli, I., Neunteufel, P., Geier, S., et al. 2021, Nature
Astronomy, 5, 1052, doi: 10.1038/s41550- 021-01413-0
Piersanti, L., Tornamb´e, A., & Yungelson, L. R. 2014,
MNRAS, 445, 3239, doi: 10.1093/mnras/stu1885
Polin, A., Nugent, P., & Kasen, D. 2019, ApJ, 873, 84,
doi: 10.3847/1538-4357/aafb6a
—. 2021, ApJ, 906, 65, doi: 10.3847/1538-4357/abcccc
Ratzloff, J. K., Barlow, B. N., Kupfer, T., et al. 2019, ApJ,
883, 51, doi: 10.3847/1538-4357/ab3727
Rau, A., Kulkarni, S. R., Law, N. M., et al. 2009, PASP,
121, 1334, doi: 10.1086/605911
Ruiter, A. J., Belczynski, K., Benacquista, M., Larson,
S. L., & Williams, G. 2010, ApJ, 717, 1006,
doi: 10.1088/0004-637X/717/2/1006
Savonije, G. J., de Kool, M., & van den Heuvel, E. P. J.
1986, A&A, 155, 51
Shen, K. J., & Bildsten, L. 2014, ApJ, 785, 61,
doi: 10.1088/0004-637X/785/1/61
Sullivan, M., Le Borgne, D., Pritchet, C. J., et al. 2006,
ApJ, 648, 868, doi: 10.1086/506137
Tutukov, A. V., & Fedorova, A. V. 1989, Soviet Astronomy,
33, 606
Tutukov, A. V., & Yungelson, L. R. 1990, Soviet Ast., 34,
57
Vanderplas, J. 2015, gatspy: General tools for Astronomical
Time Series in Python, v0.3.0, Zenodo,
doi: 10.5281/zenodo.14833
VanderPlas, J. T., & Ivezi´c, v. 2015, ApJ, 812, 18,
doi: 10.1088/0004-637X/812/1/18
Vennes, S., Kawka, A., O’Toole, S. J., emeth, P., &
Burton, D. 2012, ApJL, 759, L25,
doi: 10.1088/2041-8205/759/1/L25
Wang, B. 2018, ArXiv e-prints.
https://arxiv.org/abs/1801.04031
Wang, B., & Han, Z. 2012, NewAR, 56, 122,
doi: 10.1016/j.newar.2012.04.001
Woosley, S. E., & Kasen, D. 2011, ApJ, 734, 38,
doi: 10.1088/0004-637X/734/1/38
Yungelson, L. R. 2008, Astronomy Letters, 34, 620,
doi: 10.1134/S1063773708090053
... CD-30 Binary Evolution 3 tions obtained observationally (Geier et al. 2013;Brooks et al. 2015;Bauer et al. 2017). More recently, modelling compact sdB-WD systems in close accordance with observed atmospheric parameters has proved to be effective at yielding constraints that have implications on their formation and evolution (Kupfer et al. 2022). Furthermore, such modelling can also be used to investigate the two solutions from G13 in light of the updated parameters from Section 2. This provides strong motivation to model CD-30 in detail, especially owing to it being a double detonation supernova progenitor. ...
... Previous modelling of CD-30 by G13 (see also Brooks et al. 2015;Bauer et al. 2017) was also based on the classical formation channel, assuming that the WD formed first, followed by the sdB. However, a recent study of a compact sdB-WD binary by Kupfer et al. (2022) revealed an sdB older than the WD for the first time, making it necessary to explore an additional formation channel. ...
... As described in Kupfer et al. (2022), and shown here in Figure 6 as the alternate channel, the distinguishing feature of this younger WD channel is the formation of the sdB first via stable mass transfer, followed by the formation of the WD via common envelope ejection. The resulting system is a compact sdB-WD binary with the sdB age more than the WD age. ...
Article
Full-text available
We present a detailed modelling study of CD-30○11223 (CD-30), a hot subdwarf (sdB)-white dwarf (WD) binary identified as a double detonation supernova progenitor, using the open-source stellar evolution software MESA. We focus on implementing binary evolution models carefully tuned to match the observed characteristics of the system including log g and Teff. For the first time, we account for the structure of the hydrogen envelope throughout the modelling, and find that the inclusion of element diffusion is important for matching the observed radius and temperature. We investigate the two sdB mass solutions (0.47 and 0.54 M⊙) previously proposed for this system, strongly favouring the 0.47 M⊙ solution. The WD cooling age is compared against the sdB age using our models, which suggest an sdB likely older than the WD, contrary to the standard assumption for compact sdB-WD binaries. Subsequently, we propose a possible alternate formation channel for CD-30. We also perform binary evolution modelling of the system to study various aspects such as mass transfer, orbital period evolution and luminosity evolution. Our models confirm CD-30 as a double detonation supernova progenitor, expected to explode ≈55 Myr from now. The WD accretes a ≈0.17 M⊙ thick helium shell that causes a detonation, leaving a 0.30 M⊙ sdB ejected at ≈750 km s−1. The final 15 Myr of the system are characterised by helium accretion which dominates the system luminosity, possibly resembling an AM CVn-type system.
... Some alternative progenitor systems with a double detonation include a hot subdwarf B binary with a WD companion (Geier et al. 2013;Kupfer et al. 2022), a WD in a dynamically unstable system where the secondary is either a He WD or a hybrid between He/CO (Guillochon et al. 2010;Pakmor et al. 2013), and a potential outcome is a dynamically driven double degenerate double detonation (D 6 ) where the companion WD survives the explosion and is flung away (Shen et al. 2018b), and a WD accreting mass from a He star (Neunteufel et al. 2016;Polin et al. 2019). ...
Article
Full-text available
We present photometric and spectroscopic data for SN 2022joj, a nearby peculiar Type Ia supernova (SN Ia) with a fast decline rate (Δ m 15,B = 1.4 mag). SN 2022joj shows exceedingly red colors, with a value of approximately B − V ≈ 1.1 mag during its initial stages, beginning from 11 days before maximum brightness. As it evolves, the flux shifts toward the blue end of the spectrum, approaching B − V ≈ 0 mag around maximum light. Furthermore, at maximum light and beyond, the photometry is consistent with that of typical SNe Ia. This unusual behavior extends to its spectral characteristics, which initially displayed a red spectrum and later evolved to exhibit greater consistency with typical SNe Ia. Spectroscopically, we find strong agreement between SN 2022joj and double detonation models with white dwarf masses of around 1 M ⊙ and a thin He shell between 0.01 and 0.05 M ⊙ . Moreover, the early red colors are explained by line-blanketing absorption from iron peak elements created by the double detonation scenario in similar mass ranges. The nebular spectra in SN 2022joj deviate from expectations for double detonation, as we observe strong [Fe iii ] emission instead of [Ca ii ] lines as anticipated, though this is not as robust a prediction as early red colors and spectra. The fact that as He shells get thinner these SNe start to look more like normal SNe Ia raises the possibility that this is the triggering mechanism for the majority of SNe Ia, though evidence would be missed if the SNe are not observed early enough.
... At present the catalog of candidate verification binaries includes detached (Brown et al. 2016a) and semidetached double white dwarfs (DWDs; the latter called AM CVn type binaries; see Solheim (2010) for a recent review), hot subdwarf stars with a white dwarf companion (see Geier et al. 2013;Pelisoli et al. 2021;Kupfer et al. 2022 for recent discoveries), semidetached white dwarf-neutron star binaries (so-called ultracompact X-ray binaries; Nelemans & Jonker 2010), double neutron stars (Lyne et al. 2004) and cataclysmic variables (CVs; Scaringi et al. 2023). In Kupfer et al. (2018), we analyzed ∼50 known candidates using distances derived from parallaxes provided in the Gaia DR2 catalog (Gaia Collaboration et al. 2018). ...
Article
Full-text available
Galactic compact binaries with orbital periods shorter than a few hours emit detectable gravitational waves (GWs) at low frequencies. Their GW signals can be detected with the future Laser Interferometer Space Antenna (LISA). Crucially, they may be useful in the early months of the mission operation in helping to validate LISA's performance in comparison to prelaunch expectations. We present an updated list of 55 candidate LISA-detectable binaries with measured properties, for which we derive distances based on Gaia Data Release 3 astrometry. Based on the known properties from electromagnetic observations, we predict the LISA detectability after 1, 3, 6, and 48 months using Bayesian analysis methods. We distinguish between verification and detectable binaries as being detectable after 3 and 48 months, respectively. We find 18 verification binaries and 22 detectable sources, which triples the number of known LISA binaries over the last few years. These include detached double white dwarfs, AM CVn binaries, one ultracompact X-ray binary, and two hot subdwarf binaries. We find that across this sample the GW amplitude is expected to be measured to ≈10% on average, while the inclination is expected to be determined with ≈15° precision. For detectable binaries, these average errors increase to ≈50% and ≈40°, respectively.
... One explanation for these Ca-rich transients (theorized to be single detonations of WDs) is that they originate from high-velocity, kicked systems, and explode at considerable distances from their original location within the host galaxy prior to their occurrences (Lyman et al. 2014). Some alternative progenitor system with a double detonation include: a hot subdwarf B (sdB) binary with a WD companion (Geier et al. 2013;Kupfer et al. 2022), a WD in a dynamically unstable system where the secondary is a either He WD or a hybrid between He/CO (Guillochon et al. 2010;Pakmor et al. 2013), and a potential outcome is a dynamically driven double degenerate double-detonation (D 6 ) where the companion WD survives explosion and is flung away (Shen et al. 2018b), and a WD accreting mass from a He star (Neunteufel et al. 2016;Polin et al. 2019). ...
Preprint
We present photometric and spectroscopic data for SN 2022joj, a nearby peculiar Type Ia supernova (SN Ia) with a fast decline rate ($\rm{\Delta m_{15,B}=1.4}$ mag). SN 2022joj shows exceedingly red colors, with a value of approximately ${B-V \approx 1.1}$ mag during its initial stages, beginning from $11$ days before maximum brightness. As it evolves the flux shifts towards the blue end of the spectrum, approaching ${B-V \approx 0}$ mag around maximum light. Furthermore, at maximum light and beyond, the photometry is consistent with that of typical SNe Ia. This unusual behavior extends to its spectral characteristics, which initially displayed a red spectrum and later evolved to exhibit greater consistency with typical SNe Ia. We consider two potential explanations for this behavior: double detonation from a helium shell on a sub-Chandrasekhar-mass white dwarf and Chandrasekhar-mass models with a shallow distribution of $\rm{^{56}Ni}$. The shallow nickel models could not reproduce the red colors in the early light curves. Spectroscopically, we find strong agreement between SN 2022joj and double-detonation models with white dwarf masses around 1 $\rm{M_{\odot}}$ and thin He-shell between 0.01 and 0.02 $\rm{M_{\odot}}$. Moreover, the early red colors are explained by line-blanketing absorption from iron-peak elements created by the double detonation scenario in similar mass ranges. However, the nebular spectra composition in SN 2022joj deviates from expectations for double detonation, as we observe strong [Fe III] emission instead of [Ca II] lines as anticipated from double detonation models. More detailed modeling, e.g., including viewing angle effects, is required to test if double detonation models can explain the nebular spectra.
... There are three main proposed progenitor channels for helium CVs: the interaction of double WDs (Iben & Tutukov 1987;Marsh et al. 2004), the interaction of a WD accretor with a helium-burning star (Kupfer et al. 2020a), and transitional cataclysmic variables, in which a WD interacts with an evolved main-sequence donor (Augusteijn et al. 1993; Thorstensen et al. 2002;Podsiadlowski et al. 2003;Breedt et al. 2012;Carter et al. 2013;Littlefield et al. 2013;Kato & Osaki 2014;Ramsay et al. 2014;Burdge et al. 2022a). Several mass-transferring WD plus helium star pairs have been discovered in recent years (Burdge et al. 2020a;Kupfer et al. 2020aKupfer et al. , 2020bKupfer et al. , 2022, though it is unclear whether these systems will evolve into the post-period minimum helium CVs we see or undergo a thermonuclear supernova well before reaching such orbital periods. Finally, we know of many detached short orbital period double WD systems, the shortest of which are the eclipsing binaries ZTF J1539+5027 (Burdge et al. 2019a) and ZTF J2243+5242 ; with periods of ∼6.9 and ∼8.8 minutes, respectively), and at least two interacting WD systems that are undergoing direct impact accretion (Marsh et al. 2004), HM Cancri (Ramsay & Hakala 2002) and V407 Vul (Marsh & Steeghs 2002; with orbital periods of ∼5.4 and ∼9.5 minutes, respectively). ...
Article
Full-text available
We report the discovery of ZTF J0127+5258, a compact mass-transferring binary with an orbital period of 13.7 minutes. The system contains a white dwarf accretor, which likely originated as a post–common envelope carbon–oxygen (CO) white dwarf, and a warm donor ( T eff,donor = 16,400 ± 1000 K). The donor probably formed during a common envelope phase between the CO white dwarf and an evolving giant that left behind a helium star or white dwarf in a close orbit with the CO white dwarf. We measure gravitational wave–driven orbital inspiral with ∼51 σ significance, which yields a joint constraint on the component masses and mass transfer rate. While the accretion disk in the system is dominated by ionized helium emission, the donor exhibits a mixture of hydrogen and helium absorption lines. Phase-resolved spectroscopy yields a donor radial velocity semiamplitude of 771 ± 27 km s ⁻¹ , and high-speed photometry reveals that the system is eclipsing. We detect a Chandra X-ray counterpart with L X ∼ 3 × 10 ³¹ erg s ⁻¹ . Depending on the mass transfer rate, the system will likely either evolve into a stably mass-transferring helium cataclysmic variable, merge to become an R CrB star, or explode as a Type Ia supernova in the next million years. We predict that the Laser Space Interferometer Antenna (LISA) will detect the source with a signal-to-noise ratio of 24 ± 6 after 4 yr of observations. The system is the first LISA-loud mass-transferring binary with an intrinsically luminous donor, a class of sources that provide the opportunity to leverage the synergy between optical and infrared time domain surveys, X-ray facilities, and gravitational-wave observatories to probe general relativity, accretion physics, and binary evolution.
... Compact sdB binaries with WD companions can show ellipsoidal deformation in their light curves (e.g. Bloemen et al. 2011 ;Geier et al. 2013 ;Kupfer et al. 2020a , b ;Pelisoli et al. 2021 ;Kupfer et al. 2022 ). Compact sdB binaries with low-mass main-sequence star companions show quasi-sinusoidal variability due to the reflection effect, resulting from the extreme temperature difference and small separation distance between the cool low-mass companion and the hot sdB star. ...
Article
We conduct a systematic search for periodic variables in the hot subdwarf catalogue using data from the Zwicky Transient Facility. We present the classification of 67 HW Vir binaries, 496 reflection effect, pulsation or rotation sinusoids, 11 eclipsing signals, and 4 ellipsoidally modulated binaries. Of these, 486 are new discoveries that have not been previously published including a new mass-transferring hot subdwarf binary candidate. These sources were determined by applying the Lomb-Scargle and Box Least Squares periodograms along with manual inspection. We calculated variability statistics on all periodic sources, and compared our results to traditional methods of determining astrophysical variability. We find that ≈60% percent of variable targets, mostly sinusoidal variability, would have been missed using a traditional varindex cut. Most HW Virs, eclipsing systems and all ellipsoidal variables were recovered with a varindex >0.02. We also find a significant reddening effect, with some variable hot subdwarfs meshing with the main sequence stripe in the Hertzprung-Russell Diagram. Examining the positions of the variable stars in Galactic coordinates, we discover a higher proportion of variable stars within |b| < 25○ of the Galactic Plane, suggesting that the Galactic Plane may be fertile grounds for future discoveries if photometric surveys can effectively process the clustered field.
... The diversities in host environments indicate multiple formation channels in the He-shell DDet SN population. Those in star-forming galaxies, SN 2020jgb being the most unambiguous example, could originate from some analogues of the two subdwarf B binaries with WD companions (Iben et al. 1987;Geier et al. 2013;Kupfer et al. 2022) discovered in young stellar populations. On the other hand, those with large host offsets could not be easily formed in situ. ...
Article
Full-text available
The detonation of a thin (≲0.03 M ⊙ ) helium shell (He-shell) atop a ∼1 M ⊙ white dwarf (WD) is a promising mechanism to explain normal Type Ia supernovae (SNe Ia), while thicker He-shells and less massive WDs may explain some recently observed peculiar SNe Ia. We present observations of SN 2020jgb, a peculiar SN Ia discovered by the Zwicky Transient Facility (ZTF). Near maximum brightness, SN 2020jgb is slightly subluminous (ZTF g -band absolute magnitude −18.7 mag ≲ M g ≲ −18.2 mag depending on the amount of host-galaxy extinction) and shows an unusually red color (0.2 mag ≲ g ZTF − r ZTF ≲ 0.4 mag) due to strong line-blanketing blueward of ∼5000 Å. These properties resemble those of SN 2018byg, a peculiar SN Ia consistent with an He-shell double detonation (DDet) SN. Using detailed radiative transfer models, we show that the optical spectroscopic and photometric evolution of SN 2020jgb is broadly consistent with a ∼0.95–1.00 M ⊙ (C/O core + He-shell) progenitor ignited by a ≳0.1 M ⊙ He-shell. However, one-dimensional radiative transfer models without non-local-thermodynamic-equilibrium treatment cannot accurately characterize the line-blanketing features, making the actual shell mass uncertain. We detect a prominent absorption feature at ∼1 μ m in the near-infrared (NIR) spectrum of SN 2020jgb, which might originate from unburnt helium in the outermost ejecta. While the sample size is limited, we find similar 1 μ m features in all the peculiar He-shell DDet candidates with NIR spectra obtained to date. SN 2020jgb is also the first peculiar He-shell DDet SN discovered in a star-forming dwarf galaxy, indisputably showing that He-shell DDet SNe occur in both star-forming and passive galaxies, consistent with the normal SN Ia population.
Article
Full-text available
We report the discovery of the first hot subdwarf B (sdB) star with a massive compact companion in a wide ( P = 892.5 ± 60.2 d) binary system. It was discovered based on an astrometric binary solution provided by the Gaia mission Data Release 3. We performed detailed analyses of the spectral energy distribution (SED) as well as spectroscopic follow-up observations and confirm the nature of the visible component as a sdB star. The companion is invisible despite of its high mass of M comp = 1.50 −0.45 +0.37 M ⊙ . A main sequence star of this mass would significantly contribute to the SED and can be excluded. The companion must be a compact object, either a massive white dwarf or a neutron star. Stable Roche lobe overflow to the companion likely led to the stripping of a red giant and the formation of the sdB, the hot and exposed helium core of the giant. Based on very preliminary data, we estimate that ∼9% of the sdBs might be formed through this new channel. This binary might also be the prototype for a new progenitor class of supernovae type Ia, which has been predicted by theory.
Article
Type Ia supernovae (SNe Ia) play a key role in the fields of astrophysics and cosmology. It is widely accepted that SNe Ia arise from thermonuclear explosions of white dwarfs (WDs) in binary systems. However, there is no consensus on the fundamental aspects of the nature of SN Ia progenitors and their actual explosion mechanism. This fundamentally flaws our understanding of these important astrophysical objects. In this review, we outline the diversity of SNe Ia and the proposed progenitor models and explosion mechanisms. We discuss the recent theoretical and observational progress in addressing the SN Ia progenitor and explosion mechanism in terms of the observables at various stages of the explosion, including rates and delay times, pre-explosion companion stars, ejecta-companion interaction, early excess emission, early radio/X-ray emission from circumstellar material (CSM) interaction, surviving companion stars, late-time spectra and photometry, polarization signals, and supernova remnant properties, etc. Despite the efforts from both the theoretical and observational side, the questions of how the WDs reach an explosive state and what progenitor systems are more likely to produce SNe Ia remain open. No single published model is able to consistently explain all observational features and the full diversity of SNe Ia. This may indicate that either a new progenitor paradigm or the improvement of current models is needed if all SNe Ia arise from the same origin. An alternative scenario is that different progenitor channels and explosion mechanisms contribute to SNe Ia. In the next decade, the ongoing campaigns with the James Webb Space Telescope (JWST), Gaia and the Zwicky Transient Facility (ZTF), and upcoming extensive projects with the Vera C. Rubin Observatory Legacy Survey of Space and Time (LSST) and the Square Kilometre Array (SKA) will allow us to conduct not only studies of individual SNe Ia in unprecedented detail but also systematic investigations for different subclasses of SNe Ia. This will advance theory and observations of SNe Ia sufficiently far to gain a deeper understanding of their origin and explosion mechanism.
Article
Full-text available
Context. Hot subdwarfs in close binaries with either M dwarf, brown dwarf, or white dwarf companions show unique light variations. In hot subdwarf binaries with M dwarf or brown dwarf companions, we can observe the so-called reflection effect, while in hot subdwarfs with close white dwarf companions, we find ellipsoidal modulation and/or Doppler beaming. Aims. Analyses of these light variations can be used to derive the mass and radius of the companion and determine its nature. Thereby, we can assume the most probable sdB mass and the radius of the sdB derived by the fit of the spectral energy distribution and the Gaia parallax. Methods. In the high signal-to-noise space-based light curves from the Transiting Exoplanet Survey Satellite and the K2 space mission, several reflection effect binaries and ellipsoidal modulation binaries have been observed with much better quality than with ground-based observations. The high quality of the light curves allowed us to analyze a large sample of sdB binaries with M dwarf or white dwarf companions using LCURVE . Results. For the first time, we can constrain the absolute parameters of 19 companions of reflection effect systems, covering periods from 2.5 to 19 h and with companion masses from the hydrogen-burning limit to early M dwarfs. Moreover, we were able to determine the mass of eight white dwarf companion to hot subdwarf binaries showing ellipsoidal modulations, covering the as-yet unexplored period range of 7 to 19 h. The derived masses of the white dwarf companions show that all but two of the white dwarf companions are most likely helium-core white dwarfs. Combining our results with previously measured rotation velocities allowed us to derive the rotation period of seven sdBs in short-period binaries. In four of those systems, the rotation period of the sdB agrees with a tidally locked orbit, whereas in the other three systems, the sdB rotates significantly more slowly.
Article
Full-text available
Binary systems of a hot subdwarf B (sdB) star + a white dwarf (WD) with orbital periods less than 2–3 hr can come into contact due to gravitational waves and transfer mass from the sdB star to the WD before the sdB star ceases nuclear burning and contracts to become a WD. Motivated by the growing class of observed systems in this category, we study the phases of mass transfer in these systems. We find that because the residual outer hydrogen envelope accounts for a large fraction of an sdB star’s radius, sdB stars can spend a significant amount of time (∼tens of megayears) transferring this small amount of material at low rates (∼10 ⁻¹⁰ –10 ⁻⁹ M ⊙ yr ⁻¹ ) before transitioning to a phase where the bulk of their He transfers at much faster rates ( ≳10 ⁻⁸ M ⊙ yr ⁻¹ ). These systems therefore spend a surprising amount of time with Roche-filling sdB donors at orbital periods longer than the range associated with He star models without an envelope. We predict that the envelope transfer phase should be detectable by searching for ellipsoidal modulation of Roche-filling objects with P orb = 30–100 minutes and T eff = 20,000–30,000 K, and that many (≥10) such systems may be found in the Galactic plane after accounting for reddening. We also argue that many of these systems may go through a phase of He transfer that matches the signatures of AM CVn systems, and that some AM CVn systems associated with young stellar populations likely descend from this channel.
Article
Full-text available
Supernovae Ia are bright explosive events that can be used to estimate cosmological distances, allowing us to study the expansion of the Universe. They are understood to result from a thermonuclear detonation in a white dwarf that formed from the exhausted core of a star more massive than the Sun. However, the possible progenitor channels leading to an explosion are a long-standing debate, limiting the precision and accuracy of supernovae Ia as distance indicators. Here we present HD 265435, a binary system with an orbital period of less than a hundred minutes that consists of a white dwarf and a hot subdwarf, which is a stripped core-helium-burning star. The total mass of the system is 1.65 ± 0.25 solar masses, exceeding the Chandrasekhar limit (the maximum mass of a stable white dwarf). The system will merge owing to gravitational wave emission in 70 million years, likely triggering a supernova Ia event. We use this detection to place constraints on the contribution of hot subdwarf–white dwarf binaries to supernova Ia progenitors.
Article
Full-text available
Context. Thermonuclear supernovae (SNe), a subset of which are the highly important SNe Type Ia, remain one of the more poorly understood phenomena known to modern astrophysics. In recent years, the single degenerate helium (He) donor channel, where a white dwarf star accretes He-rich matter from a hydrogen-depleted companion, has emerged as a promising candidate progenitor scenario for these events. An unresolved question in this scenario is the fate of the companion star, which would be evident as a runaway hot subdwarf O/B stars (He sdO/B) in the aftermath of the SN event. Aims. Previous studies have shown that the kinematic properties of an ejected companion provide an opportunity to closer examine the properties of an SN progenitor system. However, with the number of observed objects not matching predictions by theory, the viability of this mechanism is called into question. In this study, we first synthesize a population of companion stars ejected by the aforementioned mechanism, taking into account predicted ejection velocities, the inferred population density in the Galactic mass distribution, and subsequent kinematics in the Galactic potential. We then discuss the astrometric properties of this population. Methods. We present 10 ⁶ individual ejection trajectories, which were numerically computed with a newly developed, lightweight simulation framework. Initial conditions were randomly generated, but weighted according to the Galactic mass density and ejection velocity data. We then discuss the bulk properties (Galactic distribution and observational parameters) of our sample. Results. Our synthetic population reflects the Galactic mass distribution. A peak in the density distribution for close objects is expected in the direction of the Galactic centre. Higher mass runaways should outnumber lower mass ones. If the entire considered mass range is realised, the radial velocity distribution should show a peak at 500 km s ⁻¹ . If only close US 708 analogues are considered, there should be a peak at (∼750 − 850) km s ⁻¹ . In either case, US 708 should be a member of the high-velocity tail of the distribution. Conclusions. We show that the puzzling lack of confirmed surviving companion stars of thermonuclear SNe, though possibly an observation-related selection effect, may indicate a selection against high mass donors in the SD He donor channel.
Article
Full-text available
Context. We present the early installment of the third Gaia data release, Gaia EDR3, consisting of astrometry and photometry for 1.8 billion sources brighter than magnitude 21, complemented with the list of radial velocities from Gaia DR2. Aims. A summary of the contents of Gaia EDR3 is presented, accompanied by a discussion on the differences with respect to Gaia DR2 and an overview of the main limitations which are present in the survey. Recommendations are made on the responsible use of Gaia EDR3 results. Methods. The raw data collected with the Gaia instruments during the first 34 months of the mission have been processed by the Gaia Data Processing and Analysis Consortium and turned into this early third data release, which represents a major advance with respect to Gaia DR2 in terms of astrometric and photometric precision, accuracy, and homogeneity. Results. Gaia EDR3 contains celestial positions and the apparent brightness in G for approximately 1.8 billion sources. For 1.5 billion of those sources, parallaxes, proper motions, and the ( G BP − G RP ) colour are also available. The passbands for G , G BP , and G RP are provided as part of the release. For ease of use, the 7 million radial velocities from Gaia DR2 are included in this release, after the removal of a small number of spurious values. New radial velocities will appear as part of Gaia DR3. Finally, Gaia EDR3 represents an updated materialisation of the celestial reference frame (CRF) in the optical, the Gaia -CRF3, which is based solely on extragalactic sources. The creation of the source list for Gaia EDR3 includes enhancements that make it more robust with respect to high proper motion stars, and the disturbing effects of spurious and partially resolved sources. The source list is largely the same as that for Gaia DR2, but it does feature new sources and there are some notable changes. The source list will not change for Gaia DR3. Conclusions. Gaia EDR3 represents a significant advance over Gaia DR2, with parallax precisions increased by 30 per cent, proper motion precisions increased by a factor of 2, and the systematic errors in the astrometry suppressed by 30–40% for the parallaxes and by a factor ~2.5 for the proper motions. The photometry also features increased precision, but above all much better homogeneity across colour, magnitude, and celestial position. A single passband for G , G BP , and G RP is valid over the entire magnitude and colour range, with no systematics above the 1% level
Article
We present a new three-dimensional map of dust reddening, based on Gaia parallaxes and stellar photometry from Pan-STARRS 1 and 2MASS. This map covers the sky north of a decl. of −30°, out to a distance of a few kiloparsecs. This new map contains three major improvements over our previous work. First, the inclusion of Gaia parallaxes dramatically improves distance estimates to nearby stars. Second, we incorporate a spatial prior that correlates the dust density across nearby sightlines. This produces a smoother map, with more isotropic clouds and smaller distance uncertainties, particularly to clouds within the nearest kiloparsec. Third, we infer the dust density with a distance resolution that is four times finer than in our previous work, to accommodate the improvements in signal-to-noise enabled by the other improvements. As part of this work, we infer the distances, reddenings, and types of 799 million stars. (Our 3D dust map can be accessed at doi: 10.7910/DVN/2EJ9TX or through the Python package dustmaps , while our catalog of stellar parameters can be accessed at doi: 10.7910/DVN/AV9GXO . More information about the map, as well as an interactive viewer, can be found at argonaut.skymaps.info .) We obtain typical reddening uncertainties that are ∼30% smaller than those reported in the Gaia DR2 catalog, reflecting the greater number of photometric passbands that enter into our analysis.
Article
The sub-Chandrasekhar-mass double-detonation (DDet) scenario is a contemporary model for Type Ia supernovae (SNe Ia). The donor star in the DDet scenario is expected to survive the explosion and to be ejected at the high orbital velocity of a compact binary system. For the first time, we consistently perform 3D hydrodynamical simulations of the interaction of supernova ejecta with a helium (He) star companion within the DDet scenario. We map the outcomes of 3D impact simulations into 1D stellar evolution codes and follow the long-term evolution of the surviving He-star companions. Our main goal is to provide the post-impact observable signatures of surviving He-star companions of DDet SNe Ia, which will support the search for such companions in future observations. Such surviving companions are ejected with high velocities of up to about 930 km s ⁻¹ . We find that our surviving He-star companions become significantly overluminous for about 10 ⁶ yr during the thermal re-equilibration phase. After the star re-establishes thermal equilibrium, its observational properties are not sensitive to the details of the ejecta-donor interaction. We apply our results to the hypervelocity star US 708, which is one of the fastest unbound stars in our Galaxy; it travels with a velocity of about 1200 km s ⁻¹ , making it a natural candidate for an ejected donor remnant of a DDet SN Ia. We find that a He-star donor with an initial mass of ≳0.5 M ⊙ is needed to explain the observed properties of US 708. Based on our detailed binary evolution calculations, however, a progenitor system with such a massive He-star donor cannot get close enough at the moment of the SN explosion to explain the high velocity of US 708. Instead, if US 708 is indeed the surviving He-star donor of a DDet SN Ia, it would require the entire pre-supernova progenitor binary to travel at a velocity of about 400 km s ⁻¹ . It could, for example, have been ejected from a globular cluster in the direction of the current motion of the surviving donor star.
Article
We present the goals, strategy and first results of the high-cadence Galactic plane survey using the Zwicky Transient Facility (ZTF). The goal of the survey is to unveil the Galactic population of short-period variable stars, including short period binaries and stellar pulsators with periods less than a few hours. Between June 2018 and January 2019, we observed 64 ZTF fields resulting in 2990 deg2 of high stellar density in ZTF-r band along the Galactic Plane. Each field was observed continuously for 1.5 to 6 hrs with a cadence of 40 sec. Most fields have between 200 and 400 observations obtained over 2 − 3 continuous nights. As part of this survey we extract a total of ≈230 million individual objects with at least 80 epochs obtained during the high-cadence Galactic Plane survey reaching an average depth of ZTF-r ≈ 20.5 mag. For four selected fields with 2 million to 10 million individual objects per field we calculate different variability statistics and find that ≈1-2 % of the objects are astrophysically variable over the observed period. We present a progress report on recent discoveries, including a new class of compact pulsators, the first members of a new class of Roche Lobe filling hot subdwarf binaries as well as new ultracompact double white dwarfs and flaring stars. Finally we present a sample of 12 new single-mode hot subdwarf B-star pulsators with pulsation amplitudes between ZTF-r = 20 − 76 mmag and pulsation periods between P = 5.8 − 16 min with a strong cluster of systems with periods ≈6 min. All of the data have now been released in either ZTF Data Release 3 or data release 4.
Article
The results of the evolutionary modelling of subdwarf B stars are presented. For the first time, we explore the core and near-core mixing in subdwarf B stars using new algorithms available in the mesa code: the predictive mixing scheme and the convective pre-mixing scheme. We show how both methods handle problems related to the determination of the convective boundary and the discrepancy between the core masses obtained from asteroseismology and evolutionary models, and long-standing problems related to the core-helium-burning phase, such as the splitting of the convective core and the occurrence of breathing pulses. We find that the convective pre-mixing scheme is the preferable algorithm. The masses of the convective core in the case of the predictive mixing and the combined convective and semiconvective regions in the case of the convective pre-mixing scheme are higher than in the models with only the Ledoux criterion, but they are still lower than the seismic-derived values. Both algorithms are promising and alternative methods of studying models of subdwarf B stars.
Article
We use non-local thermal equilibrium radiative transport modeling to examine observational signatures of sub-Chandrasekhar mass double detonation explosions in the nebular phase. Results range from spectra that look like typical and subluminous Type Ia supernovae (SNe) for higher mass progenitors to spectra that look like Ca-rich transients for lower mass progenitors. This ignition mechanism produces an inherent relationship between emission features and the progenitor mass as the ratio of the nebular [Ca ii ]/[Fe iii ] emission lines increases with decreasing white dwarf mass. Examining the [Ca ii ]/[Fe iii ] nebular line ratio in a sample of observed SNe we find further evidence for the two distinct classes of SNe Ia identified in Polin et al. by their relationship between Si ii velocity and B -band magnitude, both at time of peak brightness. This suggests that SNe Ia arise from more than one progenitor channel, and provides an empirical method for classifying events based on their physical origin. Furthermore, we provide insight to the mysterious origin of Ca-rich transients. Low-mass double detonation models with only a small mass fraction of Ca (1%) produce nebular spectra that cool primarily through forbidden [Ca ii ] emission.
Article
Using the Zwicky Transient Facility alert stream, we are conducting a large spectroscopic campaign to construct a complete, volume-limited sample of transients brighter than 20 mag, and coincident within 100″ of galaxies in the Census of the Local Universe catalog. We describe the experiment design and spectroscopic completeness from the first 16 months of operations, which have classified 754 supernovae. We present results from a systematic search for calcium-rich gap transients in the sample of 22 low-luminosity (peak absolute magnitude M > −17), hydrogen-poor events found in the experiment. We report the detection of eight new events, and constrain their volumetric rate to ≳15% ± 5% of the SN Ia rate. Combining this sample with 10 previously known events, we find a likely continuum of spectroscopic properties ranging from events with SN Ia–like features (Ca-Ia objects) to those with SN Ib/c–like features (Ca-Ib/c objects) at peak light. Within the Ca-Ib/c events, we find two populations distinguished by their red ( g − r ≈ 1.5 mag) or green ( mag) colors at the r -band peak, wherein redder events show strong line blanketing features and slower light curves (similar to Ca-Ia objects), weaker He lines, and lower [Ca ii ]/[O i ] in the nebular phase. We find that all together the spectroscopic continuum, volumetric rates, and striking old environments are consistent with the explosive burning of He shells on low-mass white dwarfs. We suggest that Ca-Ia and red Ca-Ib/c objects arise from the double detonation of He shells, while green Ca-Ib/c objects are consistent with low-efficiency burning scenarios like detonations in low-density shells or deflagrations.