MULTI-WAVELENGTH OBSERVATIONS OF THE BLACK WIDOW PULSAR 2FGL J2339.6-0532 WITH OISTER AND SUZAKU

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

Published 2015 March 25 © 2015. The American Astronomical Society. All rights reserved.
, , Citation Yoichi Yatsu et al 2015 ApJ 802 84 DOI 10.1088/0004-637X/802/2/84

0004-637X/802/2/84

ABSTRACT

Multi-wavelength observations of the black widow binary system 2FGL J2339.6-0532 are reported. The Fermi gamma-ray source 2FGL J2339.6-0532 was recently categorized as a black widow in which a recycled millisecond pulsar (MSP) is evaporating the companion star with its powerful pulsar wind. Our optical observations show clear sinusoidal light curves due to the asymmetric temperature distribution of the companion star. Assuming a simple geometry, we constrained the range of the inclination angle of the binary system to $52{}^\circ \lt i\lt 59{}^\circ $, which enables us to discuss the interaction between the pulsar wind and the companion in detail. The X-ray spectrum consists of two components: a soft, steady component that seems to originate from the surface of the MSP, and a hard, variable component from the wind-termination shock near the companion star. The measured X-ray luminosity is comparable to the bolometric luminosity of the companion, meaning that the heating efficiency is less than 0.5. In the companion orbit, 1011 cm from the pulsar, the pulsar wind is already in the particle-dominant stage with a magnetization parameter of $\sigma \lt 0.1$. In addition, we precisely investigated the time variations of the X-ray periodograms and detected a weakening of the orbital modulation. The observed phenomenon may be related to unstable pulsar wind activity or weak mass accretion, both of which can result in the temporal extinction of the radio pulse.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

A millisecond pulsar (MSP) is believed to evolve from a dead pulsar in a binary system via mass accretion from a companion star that has spun up the pulsar for billions of years (Alpar et al. 1982). Indeed, most MSPs are discovered in binary systems; however, isolated MSPs also exist. The missing link between the isolated MSPs and binary MSPs is black widow pulsars in which a sufficiently spun up pulsar is evaporating its companion star with its powerful pulsar wind. The prototype of the black widow is PSR B1957+20 with a 1.61 ms radio pulse (Fruchter et al. 1988). So far, less than 10 black widow-class objects have been discovered; however, the evolution process from an accreting MSP to a rotation-powered MSP is still unclear (Roberts 2011, 2013).

Moreover, a black widow may be a good probe for a deep investigation of pulsar wind. Currently, young rotation-powered pulsars are believed to have pulsar wind nebulae (PWNe) made up of relativistic electron–positron plasma, as seen in the Crab Nebula (Weisskopf et al. 2000; Hester et al. 2002); however, the physical mechanism that generates the pulsar wind is not yet understood. Although we do not know whether the descriptions of the Crab pulsar can be applied to a recycled MSP with a magnetic field four orders of magnitude weaker than that of typical radio pulsars, if this is the case, then the black widows can provide a crucial opportunity for us to probe pulsar wind via the direct interaction with the companion nearby the light cylinder. Theoretically, the pulsar wind is dominated by Poynting flux at the light cylinder, and therefore the magnetic energy in the wind must be converted into kinetic energy just after the wind flies out of the light cylinder, as reported by Aharonian et al. (2012), and we can approach the origin of the pulsar wind much deeper with black widows to investigate the unknown physical mechanism which can explain the σ paradox. In order to clarify the history of MSP formation and also to constrain the physical mechanism of particle acceleration just around the pulsar, a black widow pulsar is an intriguing target.

The Large Area Telescope on the Fermi satellite, with unprecedented sensitivity and angular resolution in the energy range from 100 MeV to greater than 300 GeV, has discovered more than 2000 gamma-ray sources since its launch, and 30% of these sources are still unidentified. A bright gamma-ray source discovered at high galactic latitude, 2FGL J2339.6-0532 (1FGL J2339.7-0531), was also listed in the first Fermi source catalog as an unidentified source (Abdo et al. 2010; Ackermann et al. 2012). The gamma-ray flux amounts to $3.0\pm 0.2\times {{10}^{-11}}$ erg s−1 cm−2 with a variable index of 15.7, indicating that the gamma-ray flux seems to be steady at the month timescale. While the gamma-ray spectrum has a cutoff structure at 3 GeV. These characteristics lead us to believe that the object is a pulsar. A follow-up X-ray observation conducted with Chandra discovered an X-ray point source within the error circle expected from the gamma-ray image (Kong et al. 2012). However, a radio pulse was not detected at the position at that moment. On the other hand, ground-based optical observations discovered clear sinusoidal variability with a period of 4.63 hours, which implies that the object is in a binary system and the observed optical variability is likely related to the orbital motion. The intensity of the optical counterpart varies from 20 to 17 mag in the R band (Romani & Shaw 2011; Kong et al. 2012). Moreover, the phase-resolved spectroscopy indicated that the companion may be a K-class star with a mass of 0.075 ${{M}_{}}$. This means that the pulsar-side hemisphere of the companion star is drastically heated and the energy might be supplied from an unknown recycled MSP via pulsar wind, as seen in PSR B1957+20.

In this paper, we report multi-wavelength observations of the newly discovered black widow binary system 2FGL J2339.6-0532, covering the near-infrared, optical, and X-ray energy bands. In Sections 2 and 3, the optical and X-ray observations and the obtained results are described, respectively. In Section 4, we discuss the orbital parameters based on the obtained phase-resolved spectral energy distribution (SED) and the properties of the pulsar wind just around the pulsar as well as an interpretation of the X-ray light curve showing intriguing irregular variability.

2. OPTICAL OBSERVATIONS

2.1. Observation

2FGL J2339.6-0532 was observed from 22 September to 2011 October 7 utilizing the global telescope network Optical and Infrared Synergetic Telescopes for Education and Research (OISTER).22 OISTER consists of 14 independent observatories that are funded by Japanese universities and research associations. For this work, we also asked for photometry observations from the Kottamia Astronomical observatory, which has a 188 cm reflector. Thanks to the locations of the observatories distributed across the globe, we can continuously trace the variability all day long. Moreover, the telescopes cover a wide range of wavelengths from Ks to B band which can provide the phase-resolved SED of the target. The photometric observations conducted are summarized in Table 1. The locations of the reference stars and the target are shown in Figure 1.

Figure 1.

Figure 1. Hα image of the vicinity of 2FGL J2339.6-0532 observed with the Kiso-105 cm Schmidt telescope. The circle region with a radius of 10'' indicates the location of the target. Two rectangular regions indicate the photometric reference stars shown in Table 2.

Standard image High-resolution image

Table 1.  Summary of Optical Observations with OISTER

Telescope (Instrument) Diameter 2011 September 2011 October
    22 23 24 27 28 29 30 1 4 6 7
Pirka telescope (MSI),a 160 cm BVRI BVRI V V ... ... ... ... ... ...
Gunma Astronomical Observatory 150 cm reflector (GIRCS)b 150 cm ... ... ... ... J ... ... ... ... ... ...
MITSuME-Akeno telescope (Tricolor Camera)c 50 cm gRI gRI gRI gRI gRI gRI gRI ... ... ... ...
Kiso Schmidt (2kCCD)d 105 cm ... ... ... ... ... R/Hα ... ... ... ... ...
Kyoto Sangyo University 1.3 m telescope (Two-color Camera)e 130 cm ... ... ... ... gz ... ... ... ... ... ...
Nayuta telescope (NIC)f 200 cm (JHKS) ... ... ... ... ... ... ... ... ... ...
MITSuME-OAO telescope(Tricolor Camera)c 50 cm ... gRI ... gRI gRI ... ... gRI ... gRI gRI
OAO 188 cm telescope (ISLE)g 188 cm ... ... ... J ... ... ... ... ... ... ...
Bisei Spaceguard Center 1 m telescope (Optical CCD)h 100 cm ... ... ... (r) (r) ... ... ... ... ... ...
Kanata telescope (HOWPol)j 150 cm ... ... ... R R ... ... ... V ... B
Kagoshima University 1 m telescope (IR CCD)j 100 cm ... ... ... (J) (J) ... ... (J) ... ... ...
Murikabushi telescope (Tricolor Camera)c 105 cm ... ... ... gRI ... gRI gRI ... ... ... ...
Kottamia Astronomical Observatory 188 cm telescope (Optical CCD)k 188 cm ... ... ... R R R B ... ... ... ...
IRSF 1.4 m telescope (SIRIUS)l 140 cm ... ... ... ... JHKS JHKS JHKS ... ... ... ...
miniTAO (ANIR)l 104 cm ... ... ... ... ... RJ BJ ... ... ... ...

Note. The letters on the table describe the observed energy bands. Lower case and upper case characters correspond to the SDSS system and the Johnson–Cousins system, respectively. The observations shown in parentheses did not produce usable data due to the weather condition.

Calibration observation for the field photometry. aNayoro, Hokkaido Pref., Japan (Watanabe et al. 2012). bTakayama, Gunma Pref., Japan (Takahashi et al. 2009). cHokuto, Yamanashi Pref., Japan (Kotani et al. 2005). dKiso, Nagano Pref., Japan (Itoh et al. 2001). eKyoyo, Kyoto Pref., Japan (http://www.kyoto-su.ac.jp/kao/) fSayo, Hyogo Pref., Japan (http://www.nhao.jp/en/). gAsaguchi, Okayama Pref., Japan (Yanagisawa et al. 2010). hAsaguchi, Okayama Pref., Japan (Yoshida 2005). iBisei, Okayama Pref., Japan (http://spaceguard.or.jp/bsgc_jsf/pamphlet/index.htm). jHigashi-hiroshima, Hiroshima Pref., Japan (Kawabata et al. 2008). kSatsuma-sendai, Kagoshima Pref., Japan (http://milkyway.sci.kagoshima- u.ac.jp/1/m/sys/index.html). lIshigaki, Okinawa Pref., Japan (Kotani et al. 2005). mKottamia, Egypt (Azzam et al. 2014). nSutherland, South Africa (Nagayama et al. 2003). oAtacama, Chile (Motohara et al. 2010; Konishi et al. 2015).

Download table as:  ASCIITypeset image

2.2. Hα Imaging

In the case of PSR B1957+20, the binary system is surrounded by an Hα bow shock due to its super-sonic proper motion with respect to the surrounding ISM (Kulkarni & Hester 1988). To search for a similar feature around 2FGL J2339.6-0532, we took Hα images with the Kiso Schmidt telescope at the minimum phase to prevent contamination from the companion star. Figure 1 shows the obtained Hα image in the vicinity of 2FGL J2339.6-0532 utilizing an Hα narrow-band filter (${\Delta }\lambda =9$ nm) in which we could not find the evidence of bow-shock nebula.

To evaluate the detection limit, we calculated the standard deviation of the 33'' $\;\times \;67^{\prime\prime} $ (50 × 100 pixel) rectangular region near the target and obtained STDEV = 7.82 ADU. The sky background of the CCD image shows a weak striped pattern running in the southeast-to-northwest direction with a peak-to-peak value of ∼8 ADU, which seems to be the dominant noise component restricting the detection limit. Regardless, if we adopt a conservative threshold with a 3σ confidence level, the detection limit becomes 23.5 ADU. Based on the flux calibrations of the reference stars listed in Table 2, the conversion coefficient of ADU to energy flux, $(8.3\pm 0.1)\times {{10}^{-18}}$ erg s−1 cm−2 ADU−1, was observed for the Hα filter with a bandwidth of 9 nm. Finally, we obtained a 3σ detection limit of $\lt 1.9\times {{10}^{-16}}$ erg s−1 cm−2 pixel−1, which corresponds to a surface brightness of $\lt 8.7\times {{10}^{-17}}$ erg s−1 cm−2 arcsec−2.23

Table 2.  Summary of Field Photometry

  Reference-1 Reference-2 Aλ
Name(USNO-2.0A) U0825_19993817 U0825_19993871  
Coordinate(J2000.0) (23:39:39.487, −05:32:40.56) (23:39:40.366, −05:31:51.89)  
B 18.435 ± 0.014 17.603 ± 0.008 0.120
V 17.834 ± 0.018 16.828 ± 0.008 0.091
R 17.484 ± 0.016 16.388 ± 0.007 0.072
I 17.296 ± 0.020 16.105 ± 0.008 0.050
g' 18.098 ± 0.011a 17.177 ± 0.006a 0.110
J 16.47 ± 0.10b 15.19 ± 0.04b 0.024
H 16.45 ± 0.23b 14.87 ± 0.05b 0.015
Ks 16.12b,c 14.55 ± 0.08b 0.011
Hα (656 nm) 314 ± 46 μJyd 854 ± 170 μJyd ...

Note. Errors are with 1σ confidence level.

Galactic extinction at the target coordinate (Cardelli et al. 1989; Schlafly & Finkbeiner 2011). aSDSS magnitudes were calculated from B-band and V-band magnitudes based on Smith et al. (2002). bFor the flux calibrations of J, H, and Ks, we referred to the 2MASS catalog (Skrutskie et al. 2006). cThe catalog error of the Reference-1 for the Ks band was not available. dThe flux density was evaluated from the absorption-corrected SED assuming a simple balckbody model. The estimated temperature of the reference stars were ${{T}_{{\rm BB},{\rm ref}1}}=7320\pm 160$ K, ${{T}_{{\rm BB},{\rm ref}1}}=6300\pm 180$ K.

Download table as:  ASCIITypeset image

The prototype black widow PSR B1957+20, residing at a distance of 1.6 kpc from the earth, has a bow shock nebula extending over ∼30'' × 30'' with an Hα flux of $3.3\;\times \;{{10}^{-14}}$ erg s−1 integrated over the entire nebula (Kulkarni & Hester 1988). If our target is accompanied by an equivalent bow-shock nebula, then the Hα flux and the size should be $6.1\times {{10}^{-14}}$ erg s−1 cm−2 and $\sim 41^{\prime\prime} \times 41^{\prime\prime} $, respectively. Therefore, the expected surface brightness is $3.6\times {{10}^{-17}}$ erg s−1 cm−2 arcsec−2, which is slightly lower than the detection limit.

Compared with the other bow shock PWNe, the size and intensity both seem to be scattered across a range of two orders of magnitude, possibly reflecting their shock conditions (Chatterjee & Cordes 2002). Therefore, we conclude that this observation provides only a weak upper limit of surface brightness in the Hα band.

2.3. Photometry

The other telescopes, except for the Kiso observatory, carried out photometric observations covering the Ks– B bands from 2011 September 9th to 2011 September 30th. For flux calibration, we chose two reference stars near the target object, U0825_19993817 (R = 17.5) and U0825_19993871 (R = 16.4) in the USNO-2.0 A catalog, so that all of the observatories can employ common references without saturation. The photon fluxes of the references stars were measured and compared to the standard stars in the Landolt catalog (Landolt 1992) with the MSI of the Pirka telescope in Hokkaido prefecture (Japan). Since the Pirka employs the Johnson–Cousins filter system, we estimated the photon flux in the SDSS g' band based on the observed B and V magnitudes using a conversion equation proposed by Smith et al. (2002),

Equation (1)

For J, H, and Ks, we employed the photometric data in the Two Micron All-Sky Survey (2MASS) catalog for flux calibration (Skrutskie et al. 2006). The obtained magnitudes of the reference stars are summarized in Table 2.

Figure 2 shows the light curves of 2FGL J2339.6-0532 obtained by OISTER. We re-confirmed the clear sinusoidal modulation as reported by Kong et al. (2012) and Romani & Shaw (2011). The absolute flux seems consistent with past observations: the modulation amplitude is about 4.5 magnitude in the R band and is larger at shorter wavelengths. In the R band, we successfully observed the maximum phases of the target eight times during the observation campaign. The photon flux at the maximum phase is ${{R}_{{\rm max} }}=17.7$ mag on average and varies among peaks within a range of ±0.1 mag. There was no irregular activity like the flares observed in 2FGL J1311.6-3429 (Kataoka et al. 2012).

Figure 2.

Figure 2. Multi-color light curve of 2FGL J2339.6-0532 obtained via OISTER.

Standard image High-resolution image

2.4. Phase-resolved SED

To discuss the energetics quantitatively, we estimated the energy flux based on the photometric data shown in Figure 2. First, we corrected the galactic extinction at the coordinate of the target (Schlafly & Finkbeiner 2011; Cardelli et al. 1989);24 the employed extinction at various wavelength are listed in Table 2. Then, we converted the absorption corrected magnitudes to the energy flux using conversion equations presented in Fukugita et al. (1996) and Tokunaga & Vacca (2005) for optical (B, V, R, I, g') and IR(J, H, Ks) energy bands, respectively.

Figure 3(a) shows the energy fluxes yielded as functions of the binary orbital phase. Phase = 0 was set to be MJD = 55500.0. While the optical light curves show symmetric structures, the IR (H and Ks) light curves possess somewhat asymmetric shapes. This may be due to the geometry of the companion star or the evaporating stellar gas. Panel (b) shows the SED during the brightening phase from the orbital phase = 0–0.5 in panel (a). Clearly the peak frequency increases as the orbital phase increases from orbital phase 0–0.5, implying that the temperature of the companion increases with the orbital phase. At the maximum phase, the SED is fitted with a blackbody model with an effective temperature of ${{T}_{{\rm eff}}}=7540\pm 130$ K, which is consistent with the past study (Romani & Shaw 2011).

Figure 3.

Figure 3. (a) Phase-folded multi-color light curves. The assumed orbital period is 4.63435 hr and Phase = 0 is set to be MJD = 55500 days. Marker colors represent the energy bands: red = R, green = V, blue = B, magenta = H, and black = Ks. (b)Spectral energy distributions at increasing phase from minimum to maximum. Colors correspond to the orbital phase: black = 0–0.1, magenta = 0.1–0.2, red = 0.2–0.3, green = 0.3–0.4, and blue = 0.4–0.5.

Standard image High-resolution image

In this paper, we assumed a simple emission model from the surface of the companion star to explain the observed optical emission. As described in Figure 4, we supposed that a companion star with a radius ${{r}_{{\rm comp}}}$ is orbiting around a pulsar at an orbital radius ${{R}_{{\rm orb}}}$, and the hemisphere of the companion that faces the pulsar is heated by the pulsar wind while the opposite side of the companion has ordinal temperature ${{T}_{{\rm cool}}}$. Note that this model does not take into account the effect of energy transfer via convection or advection on the surface of the companion, and therefore the temperature distribution of the companion is simply described by energy injection via the pulsar wind, although the heating efficiency is unclear. In this paper, we assumed that the spin-down energy is perfectly converted into the isotropic pulsar wind and the injected energy via the pulsar wind heats the companion up with a heating efficiency of f for simplicity. We also adopted a spin-down luminosity of $2.3\times {{10}^{34}}$ erg s−1 based on the timing analysis in the radio band recently reported by Ray et al. (2014).

Figure 4.

Figure 4. Schematic image of the assumed emission model in which a companion star with a radius ${{r}_{{\rm comp}}}$ is orbiting around a pulsar at an orbital radius ${{R}_{{\rm orb}}}$. The hemisphere of the companion star facing the pulsar is heated. In this paper, the inclination angle i is defined as the angle between the line of sight and the normal line to the orbital plane.

Standard image High-resolution image

We then fitted the obtained phase-resolved SED with the model function described above (Figure 5). Since the model seems to correlate weakly with the inclination angle, we fixed the inclination angle i. The resultant model parameters are summarized in Figure 6 as functions of the inclination angle. Although the fit seems rather poor—mainly because of the large residual in the IR band—we found that an inclination angle of $i=59{}^\circ $ minimizes the ${{\chi }^{2}}$. For comparison, we also attempted this analysis without IR data and obtained a best-fit parameter of $i=52{}^\circ $. These results seem consistent with the inclination angle reported by Romani & Shaw (2011). The obtained best-fit parameters are summarized in Table 3. Additionally, the obtained temperature distributions of the companion surface are plotted in Figure 7 as functions of the zenith angle of the pulsar from the measured point on the companion surface.

Figure 5.

Figure 5. Phase-resolved SED fitted with a model function.

Standard image High-resolution image
Figure 6.

Figure 6. Obtained parameters from model-fitting of the phase-resolved SED as functions of inclination angle i. Panels (a)–(d) show the orbital radius ${{R}_{{\rm orb}}}$ and companion radius ${{r}_{{\rm comp}}}$, the companion temperature at the cool side, the heating efficiency f, and the resultant ${{\chi }^{2}}$ (dof = 901), respectively. The bold gray line in panel (a) represents the semimajor axis of the companion calculated from the radial velocity of 350 km s−1 with respect to the mass centroid of the system (Romani & Shaw 2011). In this calculation, we employed a spin-down luminosity of the pulsar, ${{L}_{{\rm SD}}}=2.3\times {{10}^{34}}$, based on Ray et al. (2014).

Standard image High-resolution image
Figure 7.

Figure 7. Simulated temperature of the companion surface as a function of the zenith angle of the pulsar from the position for which the calculation is performed. The simulations were performed for the best-fit parameters listed in Table 3.

Standard image High-resolution image

Table 3.  Obtained Parameters of the Binary System from the Phase-resolved SED

Data set Optical+IRa Opticala
Orbital radius (${{R}_{{\rm orb}}}$)b $1.08\pm 0.02\;\times {{10}^{11}}$ cm $1.18\pm 0.02\;\times {{10}^{11}}$ cm
Companion radius (${{r}_{{\rm comp}}}$) $1.26\pm 0.02\;\times {{10}^{10}}\ {{d}_{1.1}}$ cmc $1.53\pm 0.02\;\times {{10}^{10}}\ {{d}_{1.1}}$ cmc
Companion Temperature (${{T}_{{\rm cool}}}$) $3170\pm 70\;$ K $2740\pm 120\;$ K
Heating Efficiency (f)d 1.42 ± 0.05 1.16 ± 0.04
Inclination angle (i) $59\buildrel{\circ}\over{.} 0\pm 1\buildrel{\circ}\over{.} 5$ $52\buildrel{\circ}\over{.} 1\pm 1\buildrel{\circ}\over{.} 0$
Bolometric Luminositye $1.09\pm 0.17\;\times {{10}^{32}}\ d_{1.1}^{2}$ erg s−1 $1.12\pm 0.16\;\times {{10}^{32}}\ d_{1.1}^{2}$ erg s−1
${{\chi }^{2}}$ 1392.6 637.6

Note. Errors are with 1σ confidence level.

a"Optical" and "IR" correspond to the [B, V, R, I, g'] data set and the [J, H, Ks] data set, respectively. b ${{R}_{{\rm orb}}}$ are calculated based on the radial velocity of the companion star reported by Romani & Shaw (2011). c ${{d}_{1.1}}$ is the distance in units of 1.1 kpc. dAssuming an isotropic pulsar wind with a luminosity of $2.3\times {{10}^{34}}$ erg s−1 based on Ray et al. (2014). eBolometric luminosity of the companion's heated hemisphere.

Download table as:  ASCIITypeset image

3. X-RAY OBSERVATION

3.1. Observation

2FGL J2339.6-05312 was observed with the Suzaku satellite on 2011 June 29th. The target was observed at the XIS nominal point with the normal imaging mode. The net exposure time after the standard data reduction process was 96 ks. The accumulated photons are 2555 counts with the front-illuminated CCDs (XIS0 + XIS3) and 1874 counts with the back-illuminated CCD (XIS1) in the 0.4–8 keV energy band. During the observation, the event rates were $\sim 8.25\pm 0.28\times {{10}^{-3}}$ (XIS0, XIS3) photons s−1 and $8.89\pm 0.50\times {{10}^{-3}}$ photons s−1 (XIS1) in the energy band.

3.2. Spectral Analysis

Figure 8 shows the X-ray spectrum obtained with Suzaku. At first, we tried to fit the X-ray spectra with a single power-law function, however, the resultant fit was poor (${{\chi }^{2}}/\nu $= 67.71/55) due to residuals below 1 keV implying the presence of a soft excess component. We therefore added a blackbody to the model function for the soft excess, and obtained a better fit (${{\chi }^{2}}/\nu =51.82/55$). The dotted lines in Figure 8 describe the best-fit model functions. The resultant parameters of the model fitting are summarized in Table 4. In both cases, the column density converged to zero. The resultant upper limit is $\lt 1.2\times {{10}^{21}}$ cm−2 with 90% confidence level, and is consistent with the Galactic absorption of $3.23\times {{10}^{20}}$ cm2 for this direction (Dickey & Lockman 1990; Kalberla et al. 2005).

Figure 8.

Figure 8. X-ray spectra observed by the XIS on board the Suzaku satellite. Black and red data points corresponds to the spectra observed with the front-illuminated CCDs (XIS0 and XIS3) and with the back illuminated CCD (XIS1), respectively. Dotted lines show the best-fit model functions consisting of a low-temperature blackbody and a power-law function. The model parameters are found in Table 4.

Standard image High-resolution image

Table 4.  Results of Model Fitting of the Averaged X-ray Spectra

Model wabs*(bbody + power law) wabs*power law
${{N}_{{\rm H}}}$ (${{10}^{21}}$ cm−2) $\lt 1.2$ (0) $\lt 0.13$ (0)
Blackbody Temperature (keV) 0.15 ± 0.06
Blackbody Radius (km) $0.28_{-0.16}^{+0.93}\;$
Power-law Photon Index $1.14_{-0.15}^{+0.14}$ 1.32 ± 0.08
Power-law Flux $_{0.5-10{\rm keV}}$ (${{10}^{-13}}$ erg cm−2 s−1) 2.50 ± 0.16 2.50 ± 0.15
Total flux $_{0.5-10{\rm keV}}$ (${{10}^{-13}}$ erg cm−2 s−1) $2.67_{-1.25}^{+0.01}$ $2.46_{-0.73}^{+0.04}$
${{\chi }^{2}}$ (dof) 51.82 (51) 67.71(53)

Note. Errors are with 90% confidence level.

Blackbody radius assuming a distance of 1.1 kpc.

Download table as:  ASCIITypeset image

In order to clarify the origin of these two components, phase-folded X-ray light curves were generated as shown in Figure 9. In this figure, Phase = 0 was set to be MJD = 55500.0 in the same way as the optical light curves shown in Figure 3(a). Panels (a)–(c) in Figure 9 represent the energy bands 0.4–1.0 keV, 2.0–4.0 keV, and 4.0–8.0 keV, respectively. Clearly, the hard X-ray light curves show clear modulation coinciding with the orbital motion. This timing coincidence may imply that the non-thermal emission originates from the companion surface. However, the shape of the light curves is somewhat different from the sinusoidal shapes observed in the optical, i.e., a double-peaked shape at the 2.0–4.0 keV band and a flat top shape at 4.0–8.0 keV. On the other hand, the soft X-rays seem to be steady. This may imply that the soft component has a different emitting region, namely, the pulsar.

Figure 9.

Figure 9. Phase-folded X-ray light curves for the (a) 0.4–1.0 keV, (b) 2.0–4.0 keV, and (c) 4.0–8.0 keV energy bands. Phase = 0 is set to be MJD = 55500.0 in the same way as the optical light curves. Dotted line in panel (a) is the best-fit sinusoidal curve for evaluating the modulation factor. Dashed line in panel (c) shows a model function for synchrotron emission from a thin layer covering the companion hemisphere.

Standard image High-resolution image

To evaluate the upper limit of the modulation factor in the soft X-ray energy band, we tried to fit the X-ray light curve with a sinusoidal curve and a constant component, $A{\rm sin} (2\pi x)+B$, where x is the orbital phase. The obtained amplitude of the sinusoidal curve is $A=(0.80\pm 0.34)\times {{10}^{-3}}$ counts s−1 against the constant component of $B=(3.95\pm 0.23)\times {{10}^{-3}}$ counts s−1. Therefore, the upper limit of the modulation factor is $A/B\lt 28.6$% with 1σ confidence level. Note that the maximum point was not fixed to be phase = 0.5 in this estimation; this helped to maximize the modulation factor. The best-fit model curve is shown in Figure 9(a).

If we adopt the two-component model consisting of a blackbody and a power-law function from the X-ray spectroscopy, then we can roughly evaluate the contributions from these two components to the constant emission. The phase-averaged absorbed-photon flux for the blackbody and the power law in the 0.4–1.0 keV energy band are both $2.0\times {{10}^{-5}}$ photons s−1. Assuming that the constant component originates in the blackbody, the modulation factor is expected to be (PL)/(BB + PL)∼50%, which is about twice the measured value. Taking into account the large errors arising from the uncertainties of the blackbody components (see Table 4), the measured modulation factor should be larger because of the tight upper limits to the blackbody component. This may indicate that the constant emission may consist of the blackbody from the neutron star (NS) surface and the non-thermal component originating in the pulsar magnetosphere.

3.3. Temporal Analysis

The upper panel of Figure 10 shows the obtained X-ray light curve with a bin size of 600 s for the energy range of 0.4–8.0 keV. Note that the whole length of the light curve is 190 ks, containing occultation periods of Earth and corresponding to twice the net exposure time of 96 ks. Although the photon statistics seem poor, the X-rays show periodic variability coincident with an orbital period of 16,683 s. We expected to observe the flare-like features reported in previous studies (Kataoka et al. 2012; Papitto et al. 2013; Ferrigno et al. 2014; Linares et al. 2014), which may be related to the accretion activity; however, no significant evidence for this was discovered. Nonetheless, the shape of the light curve seems to change over time.

Figure 10.

Figure 10. Upper panel shows the X-ray light curve for the energy band of 0.4–8.0 keV with a bin size of 600 s. Lower panel shows the running periodogram generated from the X-ray light curve shown in the upper panel with the Lomb-Scargle algorithm. We employed a 50 ks rectangular window function that scans over the entire X-ray light curve of 190 ks. The horizontal line in the lower panel corresponds to the center time of the sampled time span of 50 ks. The gray scale shows the arbitrary amplitude of modulation. Note that the amplitude of 6.3 correspond to 90% confidence level.

Standard image High-resolution image

In order to clarify the temporal variation in the waveform of the X-ray light curve, we traced the temporal changes of the power spectrum with the Lomb-Scargle algorithm, which can be applied to an unevenly sampled data set. In this analysis, we employed a rectangular function as a window function to detect small changes from a short data sequence. The width of the window function was set to 50 ks in order to cover three orbital laps and to detect the light curve variations for the total exposure time of 190 ks. We then scanned the light curve by sweeping the window function and generated a power spectrum for every timespan in the light curve. Note that this method cannot detect rapid variations of the power spectra (faster than a few tens of kiloseconds) because the width of the time window has a finite timespan.

The lower panel of Figure 10 shows the obtained periodogram as a function of time, which represents the central time of the sampled time window. Therefore, the periodogram at t = 25 ks consists of the X-ray light curve in the time span of $0\leqslant t\leqslant 50$ ks. Both ends of the grayscale map are omitted because the time window can sweep in the range of the obtained light curve. The sweeping step size was set to 600 s. Therefore, the time resolution of the running periodogram is about 230, which is much greater than the timescale that can be detected by this method. Nevertheless, a clear peak structure was detected at a frequency of around ${{f}_{1}}=0.6\times {{10}^{-5}}$ Hz, which corresponds to the orbital frequency. On the other hand, ${{f}_{2}}=1.2\times {{10}^{-5}}$ Hz appears to denote the second harmonic of the orbital frequency, which may be related to the dip structure of the double-peaked light curve. Interestingly, these two components in the power spectra decrease sequentially after t = 1.2$\times {{10}^{4}}$ s.

4. DISCUSSION

4.1. Constraints on the Inclination Angle of the Binary Orbit

The model fitting of the phase-resolved SED resulted in inclination angles of $i=59{}^\circ $ from the entire data set (covering optical + IR bands) and $i=52{}^\circ $ from the optical-only data. This discrepancy seems to arise from the systematic errors of the reference stars in the IR band (J, H, Ks), resulting in an underestimate of the target luminosity in the IR band. Therefore, the SED of the optical+IR data set tended to return higher companion temperatures than the optical-only data. On the other hand, it is also difficult to constrain the temperature with the optical-only data (B, V, g', R, and I), which may then result in an underestimate of the companion temperature. Therefore, the true inclination angle must be in the range of $52{}^\circ \lt i\lt 59{}^\circ $. Hereafter, we take $i\sim 55{}^\circ \pm 5{}^\circ $.

We should also summarize the pulsar parameters in preparation for the discussion below. Ray et al. (2014) recently reported the detection of radio and gamma-ray pulse emission from the central MSP with a period of 2.88 ms with a spin-down luminosity of ${{L}_{{\rm SD}}}=2.3\times {{10}^{34}}$ erg s−1. Based on these parameters, the pulsar's surface magnetic field can be estimated, assuming that the pulsar's rotation energy is dissipated by dipole radiation (Longair 1994),

Equation (2)

where ${{I}_{45}}$ is the pulsar's moment of inertia in units of 1045 g cm2, which is a typical magnetic field for a standard MSP.

4.2. Condition for Mass Accretion

In the case of NS binary systems, the mass accretion process is thought to be controlled by the propeller effect (Stella et al. 1986), i.e., the pressure balance between the ram pressure of accreting material ${{p}_{{\rm ram}}}$ and the magnetic pressure of the pulsar ${{p}_{{\rm B}}}$. Here we define ${{r}_{{\rm m}}}$ where ${{p}_{{\rm B}}}$ and ${{p}_{{\rm ram}}}$ are in balance:

Equation (3)

where G is the gravitational constant, ${{r}_{*}}$ is the NS radius, ${{M}_{*}}$ is the NS mass, and $\dot{m}$ is the mass accretion rate. Near the pulsar, charged particles are trapped by the pulsar magnetosphere and are corotating with the pulsar. If the accreting gas approaches the NS within a radius,

Equation (4)

at which the centrifugal force and the gravitational attraction are equal, then the gas will accrete onto the NS (where ${{\omega }_{*}}$ is the angular velocity of the NS).

Utilizing ${{B}_{*}}$ from the pulse profile, the lower limit of the accretion rate $\dot{m}$ that fulfills the condition of ${{r}_{{\rm c}}}\gt {{r}_{{\rm m}}}$ can be calculated,25

Equation (5)

Recently, Ray et al. (2014) reported temporal extinctions of the pulse emission in the radio implying that the accretion on the MSP is ongoing. If that is the case, the mass accretion rate must be higher than Equation (5) and the expected luminosity during the mass accretion will amount to $2.4\times {{10}^{36}}$ erg s−1, which is about five orders of magnitude higher than the X-ray luminosity at the 0.5–10.0 keV energy band. Papitto et al. (2013) reported intriguing activities of IGR J18245-2452 in the intermediate stage between the rotation and accretion power. Linares et al. (2014) and Ferrigno et al. (2014) reported weak accretion flows (below $\dot{m}\lt {{10}^{6}}$ g s−1) in IGR J18245-2452; these exhibit a variety of activities possibly reflecting accretion conditions. In case of ${{r}_{{\rm m}}}\approx {{r}_{{\rm c}}}$, namely, the weak-propeller regime, the accretion flow is only partly inhibited and the X-ray luminosity is in the range of ${{10}^{35}}\sim {{10}^{37}}$ erg s−1. For cases of ${{r}_{{\rm m}}}\gt {{r}_{{\rm c}}}$, namely, the strong propeller regime, the X-ray luminosity is in the range of ${{10}^{33}}\sim {{10}^{34}}$ with a very fast and striking variability. Regardless, the X-ray luminosity of 2FGL J2339.6-0532 is still lower than that of IGR J18245-2452 in the strong-propeller regime by an order of magnitude.

In addition, IGR J18245-2452 exhibited another quiescent state in 2002 with an X-ray luminosity of $\sim {{10}^{32}}$ erg s−1, which is comparable to the observed X-ray luminosity of 2FGL J2339.6-0532. Ferrigno et al. (2014) claimed that ${{r}_{{\rm m}}}$ reached the light-cylinder radius in this regime. This condition would allow for the existence of pulsar wind and would also prevent mass accretion. Thus, we conclude that the binary system can be in the same condition seen in IGR J18245-2452 in 2002 or in the late stage of the evolution process without mass accretion. In the latter case, where the mass accretion already ended, the radio-pulse extinction can also be explained by the unstable activity of the pulsar's magnetosphere itself, as with "nulling" pulsars. (Gajjar et al. 2012 and references therein.)

4.3. Origin of the X-ray Emission

As shown in the spectral analysis, the soft X-rays below 1 keV seem to originate from a blackbody. The obtained radius of 0.28 km for d = 1.1 kpc is unusually small for an accretion disk and thus the radiation can be interpreted as the X-ray radiation from the NS surface. A radius of 0.28 km is still small for a NS, and therefore the emitting region could be a hot spot around the polar cap area. In order to constrain the geometry of the polar cap, a high time resolution X-ray observation is required. In addition, the small variability and the lack of eclipse in the X-ray light curve are consistent with the estimated inclination angle of $i\sim 55{}^\circ $.

In contrast, the hard X-ray light curves vary with the orbital motion. The X-ray luminosity of the non-thermal component from the averaged X-ray spectra amounts to ${{L}_{{\rm pow}}}=3.6\times {{10}^{31}}$ erg s−1 for a distance of 1.1 kpc at 0.5–10.0 keV. Therefore, the ratio of ${{L}_{{\rm pow}}}$ to the spin-down luminosity is about 0.15%, which is the typical value for a standard radio pulsar (Becker & Truemper 1997; Kargaltsev & Pavlov 2008).

To explain the flat-top modulation of the 4–8 keV light curve, two possible emitting regions can be proposed. One is an extended emitting region just around the pulsar like a halo. In this hypothesis, the minimum phase at the inferior conjunction can be explained by an eclipse of the halo. The observed smooth modulation requires the emitting region to be larger than the companion, 1010cm∼, which is still small for a standard pulsar wind nebula. In many cases, the non-thermal X-rays from a pulsar wind nebula are from downstream of the termination shock and the radius of the termination shock is constrained by the pressure balance. In this case, the upper limit of the emitter size is the orbital radius of ${{R}_{{\rm orb}}}\sim {{10}^{11}}$ cm. At this distance, however, the wind pressure is still too strong to be terminated.

The other hypothesis for the origin of the hard X-rays is that of a shocked pulsar wind interacting with the companion star (Bogdanov et al. 2005, 2014a, 2014b). The companion star orbiting nearby the pulsar can partially terminate the strong pulsar wind. In this case, the flat-top light curve can be explained by the Lorentz boost of the shocked plasma. As shown in Figure 11, we assumed a thin emitting layer covering the hemisphere of the companion star. The surface of the emitting region is the wind termination shock and the injected electron plasma decelerated at the shock to a bulk velocity ${{\beta }_{2}}$ reflecting the magnetization parameter upstream of the shock. At the superior conjunction phase, the shocked plasma is flowing in a direction away from us, and therefore the synchrotron luminosity may be weakened (Pelling et al. 1987). The dashed line in the bottom panel of Figure 9 shows the calculated photon flux for the thin layer of the shocked pulsar wind on the surface of the companion with a photon index of p = 1.14 and an inclination angle of $i=55{}^\circ $. The best-fit bulk velocity of the shocked wind is ${{\beta }_{2}}=0.43\pm 0.15$ in units of the light speed. If that is the case, the X-ray flux is attenuated by a factor of

Equation (6)

Here, we employed a magnetization parameter σ for the pulsar wind upstream of the shock, $\sigma =B_{1}^{2}/(4\pi {{n}_{1}}{{u}_{1}}{{\gamma }_{1}},{{m}_{e}}{{c}^{2}})$ (Kennel & Coroniti 1984), where B1 is the magnetic field just upstream of the shock, n1 is the number density of the pulsar wind, u1 is the flow velocity defined by $1+u_{1}^{2}={\Gamma }_{1}^{2}$ (${{{\Gamma }}_{1}}$ is the bulk Lorentz factor of the wind upstream of the shock), and c is the light velocity. Assuming that half of the wind energy is carried as Poynting flux ($\sigma =1$), B1 amounts to

Equation (7)

Furthermore, the pulsar wind is compressed at the shock, and therefore the downstream magnetic field B2 is higher than B1. Adopting ${{B}_{2}}\gt 5.5$ G as a lower limit, we can constrain the gyration radius of the synchrotron electrons emitting 1 keV X-rays,

Equation (8)

where ${{\epsilon }_{{\rm X}}}$ is the synchrotron photon energy. The obtained gyration radius is far smaller than the companion, ${{r}_{{\rm comp}}}\sim {{10}^{10}}$ cm, and therefore the shock-heated electrons cannot escape from the downstream immediately. In terms of the enegetics, the phase averaged X-ray luminosity of the power-law component was ${{L}_{{\rm pow},0.5-10\,{\rm keV}}}=3.6\times {{10}^{31}}$ erg s−1 from the spectral fitting. Taking into account the Lorentz boost effect of 1/0.42 and the ratio of the maximum flux against the phase averaged flux of ∼3.0, the intrinsic synchrotron luminosity from the companion becomes $L_{{\rm pow},0.5-10\,{\rm keV}}^{\prime }=1.1\times {{10}^{32}}$, which is comparable to the bolometric luminosity of the heated hemisphere of the companion, ${{L}_{{\rm BB}}}\sim 1.1\times {{10}^{32}}$ ergs s−1. This brief estimation also indicates that the heating efficiency should be smaller than 0.5. This result is obviously different from the heating efficiency obtained from the phase-resolved SED (see Figure 5) in which an isotropic pulsar wind is assumed. This discrepancy may indicate that the pulsar wind from the central MSP is not isotropic as expected in the crab nebula (Bogovalov & Khangoulyan 2002). The wind pressure at the companion is then ${{p}_{{\rm wind}}}={{L}_{{\rm SD}}}/4\pi R_{{\rm orb}}^{2}\;c\sim 4.7$ dyn which may be greater than that of the coronal gas ($T={{10}^{6}}$ K, ${{n}_{{\rm p}}}={{10}^{8}}$ cm−3 for the Sun) but much smaller than the photosphere. Therefore the thickness of the emitting region should not be higher than the corona of the companion. On the other hand, the radiation length of the relativistic electrons is quite long: $716A/(Z(Z+1){\rm ln} (278/\sqrt{(Z)})\sim 4\times {{10}^{17}}$ cm for the coronal gas of ${{n}_{{\rm p}}}\sim {{10}^{8}}$ cm−3. Therefore, these electrons can radiate synchrotron X-rays even in the immediate area of the photosphere.

Figure 11.

Figure 11. Schematic view of the Hard X-ray emission from the surface of the companion.

Standard image High-resolution image

If the hard X-rays originated in the shocked pulsar wind on the surface of the companion star as is discussed, then the composition of the pulsar wind can be constrained based on the downstream flow velocity. Considering the Rankin–Hugoniot relation for magnetized electron–positron plasma for small σ, we can find the downstream flow velocity as a function of the magnetization parameter σ (Kennel & Coroniti 1984):

Equation (9)

where ${{u}_{2}}={{{\Gamma }}_{2}}{{\beta }_{2}}$. Finally, we find the downstream flow velocity, ${{\beta }_{2}}=\sqrt{(1+9\sigma )/(7-9\sigma )}$, in units of the light speed. Note that the equation predicts the lower limit of the shocked pulsar wind, ${{\beta }_{2}}=1/\sqrt{7}\sim 0.38$, at $\sigma =0$. On the other hand, the flow velocity of ${{\beta }_{2}}=0.43\pm 0.15$ from the X-ray light curve indicates that the pulsar wind is already in the particle-dominant state, i.e., $\sigma \sim 0.028$, at the surface of the companion, only $1.1\times {{10}^{11}}$ cm apart from the pulsar. Note that the post-shock flow velocity has an inverse-correlation with the inclination angle that affects the resultant magnetization parameter. If we adopt an error range of inclination angle, $50{}^\circ \sim 60{}^\circ $, the magnetization parameter can range from 0.008 (for $i=60{}^\circ $) to 0.1 (for $i=50{}^\circ $). The upper limit of $\sigma \lt 0.1$ is still much smaller than that calculated for the similar system of PSR J1023+0038 reported by Bogdanov et al. (2011), in which a thick post-shock emitting region was assumed. Past theoretical studies claimed that the pulsar wind is highly magnetized at the beginning just around the light cylinder. However, the observed σ is far smaller than 1, and this discrepancy has been called the σ paradox. Recently, Aharonian et al. (2012) reported evidence of particle acceleration just around the pulsar based on gamma-ray pulse analysis of the Crab pulsar. The above argument of σ in this work is consistent with their result and may be one of the σ parameters measured at the nearest distance from a pulsar. Furthermore, this may imply that the old MSPs and the young radio pulsars have the same particle acceleration mechanism.

One of the intriguing features of this object is the double-peak structure in the 2–4 keV X-ray light curve. Since the dip is observed only in the softer band (Figure 9), this feature can be interpreted as absorption of soft X-rays. If we accept that the non-thermal X-rays originate in the shocked pulsar wind on the surface of the companion as discussed above, the absorber must be lying around the pulsar because the dip appears at the optical maximum phase (= the superior conjunction). In addition, the orbital inclination angle $i\sim 55{}^\circ $ requires that the absorber must be extended far from the orbital plane ($\sim 8\times {{10}^{10}}$ cm).

As Ray et al. (2014) claimed, mass accretion onto the pulsar seems to continue intermittently, however, the argument of the propeller effect rules out a powerful accretion. This is consistent with the lack of a bright disk component in X-ray spectra. In addition to the propeller effects, the measured companion radius is about 1/2 of the Roche lobe, and therefore the stellar material cannot overcome the Lagrange point L1 to accrete on the NS. To feed mass to the pulsar, explosive ejection faster than ∼1000 km s−1 may be required for the mass to escape from the gravity potential.

In addition, as inferred from the discussion on the heating efficiency, the pulsar wind may be concentrated on the orbital plane, which would prevent the accretion flow. If the stellar material that is blown away by the equatorial pulsar wind is drifting around the pulsar's polar regions, and it may accrete from the poloidal direction. On the other hand, the running periodogram shown in Figure 10 indicates that the dip structure related to the second harmonic component is unstable, and appears to have disappeared ∼120 ks from the start time of the observation. To confirm the scenario in which the stellar gas is drifting around the pulsar and accretes intermittently, simultaneous X-ray and radio observations are required.

5. CONCLUSION

We presented multi-wavelength observations of a newly found black widow binary system 2FGL J2339.6-0532 covering near-infrared to X-ray regimes. Thanks to a wide wave waveband and long coverage, we successfully obtained a phase-resolved SED that enabled us to constrain the orbital parameters more precisely. The obtained SED seems consistent with past studies, and we calculated an inclination angle of $i\sim 55{}^\circ $, while taking into account the results of recent radio observations. Based on the argument of the propeller-effect argument, the lower limit of the acrretion rate was estimated to be $\dot{m}\gt 1.3\times {{10}^{16}}$ g s−1, which is about five orders of magnitude larger than that expected from the observed X-ray luminosity. In addition, the estimated orbital parameters imply that the companion's radius is only half of the Roche lobe, making it difficult to feed the pulsar continuously. We also obtained an Hα image of the vicinity of the target and could not detect any diffuse structure with a 3σ detection limit of $\lt 8.7\times {{10}^{-17}}$ erg s−1 cm−2 arcsec−2. We therefore conclude that the target does not have a bow-shock nebula brighter than the Hα nebula around PSR B1957+20.

In the X-ray regime, we discovered a steady, soft X-ray component below 1 keV, which seemed to originate from the neutron-star surface, and which showed no evidence of an accretion disk. On the other hand, the hard component above 2 keV showed periodic modulation synchronized with the orbital motion, implying that the hard X-rays originate in the shocked pulsar wind near the companion surface. The observed X-ray luminosity is comparable to the bolometric luminosity of the heated hemisphere of the companion. This means that the heating efficiency should be smaller than 0.5 and the pulsar-wind distribution should be anisotropic. Adopting the above scenario, we estimated the magnetization parameter of the pulsar wind to be $\sigma \sim 0.03$, based on the Rankin–Hugoniot relation. This implies that the pulsar wind is already in the particle-dominant state at a distance of 1.1 × 1011 cm from the pulsar. Moreover, we also investigated the time variability of the modulation pattern by using the running periodogram of the X-ray light curve, wherein we detected a weakening of the modulation pattern. This may be related to an intermittent weak accretion or an unstable pulsar-wind activity, both of which can cause temporal extinction of the radio-pulse emission.

This research has made use of optical and near-infrared data obtained from the Optical and Near-Infrared Astronomy Inter-University Cooperation Program, supported by the MEXT of Japan and Kottamia Astronomical Observatory supported by the NRIAG of Egypt. This research has also made use of X-ray data obtained from the Suzaku satellite, a collaborative mission between the space agencies of Japan (JAXA) and the USA (NASA). Operation of ANIR on the miniTAO telescope is supported by Grant-in-Aid for Scientific Research (21684006, 22253002, and 22540258) and the Institutional Program for Young Researcher Overseas Visits, operated by Japan Society for the Promotion of Science (JSPS). Part of this work has been supported by NAOJ Research Grant for Universities. The authors thank the referee for useful suggestions that helped to improve the original manuscrip. Y.Y. is deeply grateful to Suguru Saito for generous support and encouragement.

Facilities: Suzaku - Suzaku (ASTRO-EII)

Footnotes

Please wait… references are loading.
10.1088/0004-637X/802/2/84