Abstract

We present Very Large Telescope optical spectra of the black hole candidate X-ray binary MAXI J1836−194 at the onset of its 2011 outburst. Although the spectrum was taken at the beginning of the outburst and contains a significant contribution from the optically thin synchrotron emission that originates in the radio jet, we find that the accretion disc was already large and bright. Single-peaked, narrow Hα and He ii λ4686 lines imply the most face-on accretion disc observed in a black hole low-mass X-ray binary to date, with an inclination angle between 4° and 15°, assuming a black hole mass of between 5 and 12 M, for distances of between 4 and 10 kpc. We use New Technology Telescope observations of the system in quiescence to place strong upper limits on the mass and radius of the donor star and the orbital period. The donor is a main-sequence star with a mass <0.65 M and a radius <0.59 R with an orbital period of <4.9 h. From those values and Roche lobe geometry constraints we find that the compact object must be >1.9 M if the system is located 4 kpc away and >7.0 M at 10 kpc.

INTRODUCTION

Low-mass X-ray binaries (LMXBs) spend the majority of their lifetimes in a state of quiescence with faint or undetectable X-ray emission and an optical spectrum dominated by the companion star. Occasionally, they show dramatic X-ray brightening (outburst), the consequence of an increased accretion rate on to the central compact object; this may be due to thermal-viscous disc instabilities (Lasota 2001) or an increased mass transfer rate from the donor star. An X-ray binary may go through a number of X-ray spectral states during an outburst (Fender, Belloni & Gallo 2004; Remillard & McClintock 2006; Belloni 2010), before returning to quiescence after a few weeks or months. In the simplest classification, the two main outburst states are a hard (non-thermal) state, where most of the accretion power is liberated in a hot corona or a relativistic outflow, and a soft (thermal) state, when most of the accretion power is efficiently radiated by an accretion disc.

The main sources of energy in the disc are viscous heating and the interception and thermalization of X-ray photons from regions close to the compact object (van Paradijs & McClintock 1995). The latter generally dominates at large disc radii, where most of the optical/ultraviolet (UV) emission comes from. In the viscous-dominated part of the disc, the emitted spectrum is a multicolour blackbody with an effective temperature, Teff, scaling with radius, R, as, Teff(R) ∼ R−3/4 (Shakura & Sunyaev 1973), while in the irradiation-dominated outer disc, Teff(R) ∼ R−1/2 (Frank, King & Raine 2002). A typical signature of external X-ray irradiation in the outer disc is the formation of a temperature-inversion layer in the disc chromosphere; as a result, He ii λ4686 and Hα are detected in emission, and higher lines of the Balmer series generally have narrow emission cores superposed on broader absorption formed in deeper layers (van Paradijs & McClintock 1995).

In the soft state, the optical/near-infrared (NIR) continuum comes almost entirely from the accretion disc, which becomes much brighter than the donor star (van Paradijs & McClintock 1995). In the hard state there may also be a significant NIR/optical/UV contribution from the optically thin synchrotron emission of the jet (Jain et al. 2001; Corbel & Fender 2002; Russell et al. 2006) as well as the donor star. While X-ray studies of an X-ray binary probe the accretion processes in the region immediately surrounding the compact object, optical/UV studies provide important constraints on the size and orbital period of the binary system, and on the mass of the donor star (see van Paradijs & McClintock 1995; Charles & Coe 2006 for a full review of optical and UV emission from X-ray binaries).

In this paper we present the results of an optical spectroscopic study of the X-ray transient MAXI J1836−194. This LMXB was discovered in the early stages of its outburst on 2011 August 30 (Negoro et al. 2011) by Monitor of All-sky X-ray Image (MAXI)/Gas Slit Camera (GSC) on the International Space Station (Matsuoka et al. 2009). The system was classified as a rapidly spinning (a = 0.88 ± 0.03; Reis et al. 2012) black hole (BH) candidate by virtue of its radio/X-ray properties (Miller-Jones et al. 2011; Russell et al. 2011; Strohmayer & Smith 2011). MAXI J1836−194 underwent a state transition from the hard state to the hard-intermediate state (HIMS) on 2011 September 10 (inferred from INTEGRAL, Swift and RXTE X-ray spectral and timing properties; Ferrigno et al. 2012). Instead of softening further to a full soft state (as is usually the case for other BH LMXBs), the outburst ‘failed’ (Brocksopp, Bandyopadhyay & Fender 2004) and the source transitioned back to the hard state on 2011 September 28 and faded towards quiescence (Ferrigno et al. 2012) over several months while the jet remained active (Russell et al. 2013). A re-brightening of the source was detected with Swift/Burst Alert Telescope (BAT) in 2012 March which lasted for several months (Krimm et al. 2012; Yang, Wijnands & Kennea 2012a) before a final return to quiescence by 2012 July (Yang et al. 2012b). From the X-ray spectral state evolution and inner-disc temperature, we argue (in our companion paper Russell et al. 2014, hereafter TDR14) that the luminosity in the early outburst phase was maximally a few per cent of the Eddington luminosity (typical for a HIMS to hard state transition; Maccarone 2003; Dunn et al. 2010). This suggests a plausible range of distances between 4 and 10 kpc for this source. Here, we expand on the preliminary results presented by Pakull & Motch (2011), based on their Very Large Telescope (VLT) observations. We show that the optical/UV signatures of an accretion disc were already present in the first few days after the discovery of the outburst although the optical/NIR band contained a significant contribution from the jet (Russell et al. 2013). We use these signatures to infer the inclination of the accretion disc. Swift Ultraviolet/Optical Telescope (UVOT) and European Southern Observatory New Technology Telescope (NTT) measurements of the optical luminosity in quiescence also allow us to constrain the size of the donor star and hence the orbital period of the binary system, as well as place a lower limit on the BH mass.

OBSERVATIONS AND REDUCTIONS

VLT

We observed MAXI J1836−194 with the Focal Reducer and Low Dispersion Spectrograph 2 (FORS2) on the VLT as a target-of-opportunity during the implementation of Programme 087.D-0870 (PI: M. Pakull). The target was observed on 2011 August 31, between Modified Julian Date (MJD) 55804.99–55805.05, and 2011 September 01 (MJD 55805.99–55806.02). We used the 1200B grism, covering the wavelength range from 3729 to 5200 Å (‘blue spectra’), and the 1200R grism, from 5869 to 7308 Å (‘red spectra’). The first night consisted of four 900 s observations for the blue spectra and four 450 s observations for the red spectra, while the second night consisted of two 900 s observations and two 450 s observations for the blue and red spectra respectively.

Calibration data (bias, flat-field and wavelength calibrators) were acquired following the standard calibration plan, implemented at the beginning or the end of the observing nights. We used the same slit width (1.0 arcsec) for all observations, giving a spectral resolution of 2.3 Å for the red spectra and 1.9 Å for the blue spectra on the first night and on the second night 1.6 Å for both the red and the blue spectra. The orientation of the slit was at a parallactic angle of 112| $_{.}^{\circ}$|5 the first night, and 141| $_{.}^{\circ}$|3 the second night. We took B-band sky and through-slit acquisition images each night before starting our spectroscopic observations to confirm that the target was well centred in the slit. The average full width at half-maximum (FWHM) seeing, estimated at 5000 Å, was 1.18 arcsec for the blue spectra and 1.51 arcsec for the red spectra on the first night and 0.81 and 0.68 arcsec, respectively, on the second night.

The data were reduced with ESOREX version 3.9.6 and the FORS2 kit recipes version 4.8.7 (Appenzeller et al. 1998) was used to correct raw frames for bias, apply flat-fields and to apply wavelength calibration. Arc images of He, Hg and Cd as well as Hg, Cd, Ar and Ne lamps provided wavelength solutions for the 1200B and 1200R grisms, respectively, yielding a wavelength accuracy of ∼0.05 Å for both the blue and red spectra (taken from the scatter of the fit to all the arc lines). One-dimensional spectra were extracted from the wavelength calibrated image using a MIDAS (Banse et al. 1988) procedure optimized for providing the best signal-to-noise ratio. We then used the IRAF (Tody 1986) tasks rvcorrect and dopcorrect to correct the spectra for the relative radial velocity of the Earth, shifting the wavelength scale to the heliocentric system. We also used standard IRAF tasks to combine and/or compare the spectra from the two nights. We then normalized the spectrum to the continuum with an eighth-order cubic spline, where we masked the Balmer lines (including their broad absorption components), the interstellar lines, He ii λ4686 and the Bowen blend. We then measured the main properties of the lines [central position, FWHM, equivalent width (EW)] with the IRAF task splot, where the errors are taken as the statistical error in the Gaussian fit and verified the results with the fitting package QDP (Tennant 1991).

Optical / UV Swift

Swift monitored MAXI J1836−194 every few days during its 2011 outburst (Ferrigno et al. 2012) and 2012 re-flaring. Observations in six optical/UV filters were made with the UVOT (Roming et al. 2005), of which we use the three optical filters (u, b and v) that provide the strictest constraints on the luminosity of the low-mass donor star. UVOT data were pre-processed at the Swift Data Centre (Breeveld et al. 2010) and reduced using standard Swift pipelines. The FTOOLS1 (Blackburn 1995) task, uvotimsum, was used to combine all images from the same filter. We used uvotsource and a suitable background region to extract an aperture photometry magnitude from a source region of 2.5 arcsec radius, centred on the UVOT position. Magnitudes are based on the UVOT photometric system, which differs from the Bessell system by Vv < 0.04 mag (for all colours) and Bb < 0.04 mag (for colours bv < 1.5 mag; Poole et al. 2008). The differences between the UVOT and Bessell photometric magnitudes are small compared to the uncertainty in our photometry (see Section 3.1) and we therefore treat the UVOT v and b magnitudes as standard V and B magnitudes.

NTT

We obtained optical images during quiescence using the ESO Faint Object Spectrograph and Camera (EFOSC2; Buzzoni et al. 1984) mounted on the NTT at La Silla Observatory, Chile. On 2013 June 11 three observations were acquired in the Gunn i band with an exposure time of 300 s and a FWHM seeing of 1.0–1.1 arcsec. On 2013 June 12 and 13 single exposures of 300 s were obtained in the Gunn g-, r- and i-band filters with a seeing in the i band of 1.1 and 0.7 arcsec, respectively. Sky conditions were not photometric on all three nights, so additional 10, 30 and 60 s exposures of the field were obtained on June 12 and 13 to enable absolute flux calibration of the frames relative to data from the AAVSO Photometric All-Sky Survey (APASS) Data Release 7 (Henden et al. 2009). We used 2 × 2 on-chip binning during the observations, providing a scale of approximately 0.24 arcsec pixel−1. Frames were bias subtracted and then flat-fielded using dome flat-field observations. Given the faintness of the optical sources under study (see Fig. 1, bottom panel), we excluded data with image quality worse than 1.0 arcsec FWHM seeing.

Figure 1.

Top panel: B-band image of MAXIJ1836−194 (marked by an arrow) and surrounding field, taken with VLT/FORS2 on 2011 August 31 as an acquisition image before our spectral observations. Middle panel: i-band NTT image of the same field taken on 2013 June 13, after the system had returned to quiescence, where the red circle of radius 0.5 arcsec indicates the outburst position of the system. Bottom panel: another i-band NTT image of the field taken on 2013 June 11 of the system in quiescence. Here the yellow circle of radius 0.08 arcsec shows the unrelated field star at a J2000 position of RA = 18:35:43.474 and Dec. = −19:19:10.25, while the red circle of radius 0.08 arcsec shows the fitted location of the optical counterpart detected during outburst, at the J2000 position of RA = 18:35:43.451 and Dec. = −19:19:10.43. The optical counterpart of MAXI J1836−194 is no longer detected.

We applied point spread function (PSF) fitting on the images using daophot in IRAF to compute the instrumental magnitudes of the detected stars. Flux calibration of the field was performed using APASS objects in the MAXI J1836−194 field, achieving zero-point errors <0.1 mag. Differential photometry was performed to derive the fluxes of the sources of interest. The atmospheric calibration of the frames was defined by fitting for the reference point position, the scale and the position angle. As a first step, we obtained an astrometric solution on a 10 s i-band exposure. By comparing the positions of stars against entries from the fourth US Naval Observatory CCD Astrograph Catalogue (UCAC4; Zacharias et al. 2013), we found nine suitable UCAC4 sources well distributed on the image that were not saturated or blended and appear stellar. We obtained a solution with 0.0788 arcsec rms residuals. We then used the sources from this 10 s i-band image as secondary astrometric calibrators for our 300 s images. Taking into account uncertainties (including the systematic 0.05 arcsec uncertainty that is inherent to the UCAC4 catalogue) we estimated a positional accuracy of 0.13 arcsec. Additionally a publicly available 30 s i-band image of the field obtained during outburst with the Faulkes telescope on 2011 September 15 was calibrated using the same secondary reference tiers as for the 300 s NTT images. This frame was acquired with a 0.3 arcsec pixel−1 scale under a seeing of 1.2 arcsec FWHM and shows the optical counterpart at a J2000 position of RA = 18:35:43.451, Dec. = −19:19:10.43.

RESULTS

Optical photometry

During the two nights of VLT observations (2011 August 31, 2011 September 1), MAXI J1836−194 was detected as one of the brightest optical sources in the field (Fig. 1, top panel). Quasi-simultaneous Swift/UVOT observations on the same nights provide an accurate measurement of its optical/UV brightness and colours: we obtain u = 16.43 ± 0.08 mag, b = 17.12 ± 0.08 mag, v = 16.37 ± 0.08 mag on the first night, and u = 16.39 ± 0.07 mag, b = 16.97 ± 0.07 mag, v = 16.40 ± 0.07 mag on the second night (Fig. 2).

Figure 2.

Swift/UVOT light curve, showing the 2011 outburst and decline towards quiescence. The date of the first VLT observation is marked by the green arrow. The state transitions are marked by the vertical dashed lines, HS denotes the hard state and HIMS is the hard-intermediate state (the states are defined by X-ray spectral and timing properties presented by Ferrigno et al. 2012).

Instead of brightening as the source moved into the HIMS, the optical/UV luminosity decreased; on 2011 September 17, we measured u = 16.79 ± 0.08 mag, b = 17.33 ± 0.07 mag, v = 17.18 ± 0.07 mag from Swift/UVOT (Fig. 2). The optical/UV dimming is discussed further in Section 4. We used the Swift/X-Ray Telescope (XRT) data where the soft disc-like component to the XRT spectra was largest (mid-September; TDR14) to infer the most important disc parameters; in particular, in the context of this work, the outer disc radius and X-ray reprocessing fraction.

MAXI J1836−194 underwent a re-brightening during 2012 March, followed by a decline to quiescence (Krimm et al.

2012; Yang et al. 2012a). In the last three Swift/UVOT observations of 2012 June 29, July 5 and July 10, the optical counterpart was not visible, nor was it found in the stacked image from the three data sets. We take the detection limit for that combined image as the best-available 3σ upper limit to the quiescent optical flux in the u, b and v bands: u > 20.5 mag, b > 20.8 mag, v > 19.5 mag (Table 1).

Table 1.

Upper limits on the magnitude and magnitude minus the distance modulus (dm) of MAXI J1836−194 during quiescence; u, b and v upper limits are determined from the summed Swift/UVOT images of 2012 June 29, July 05 and July 10, while the g, r and i upper limits are from the 2013 June 12 and 13 NTT images of the system in quiescence. The distance ranges presented cover the extent of plausible distances to the source (derived from the temperature and flux of the disc; TDR14) and magnitudes presented are not corrected for extinction.

FilterMagnitudeMag − dm
4 kpc6 kpc8 kpc10 kpc
u>20.5>7.5>6.6>6.0>5.5
b>20.8>7.8>6.9>6.3>5.8
v>19.5>6.5>5.6>5.0>4.5
g>24.0>11.0>10.1>9.5>9.0
r>23.7>10.7>9.8>9.2>8.7
i>23.5>10.5>9.6>9.0>8.5
FilterMagnitudeMag − dm
4 kpc6 kpc8 kpc10 kpc
u>20.5>7.5>6.6>6.0>5.5
b>20.8>7.8>6.9>6.3>5.8
v>19.5>6.5>5.6>5.0>4.5
g>24.0>11.0>10.1>9.5>9.0
r>23.7>10.7>9.8>9.2>8.7
i>23.5>10.5>9.6>9.0>8.5
Table 1.

Upper limits on the magnitude and magnitude minus the distance modulus (dm) of MAXI J1836−194 during quiescence; u, b and v upper limits are determined from the summed Swift/UVOT images of 2012 June 29, July 05 and July 10, while the g, r and i upper limits are from the 2013 June 12 and 13 NTT images of the system in quiescence. The distance ranges presented cover the extent of plausible distances to the source (derived from the temperature and flux of the disc; TDR14) and magnitudes presented are not corrected for extinction.

FilterMagnitudeMag − dm
4 kpc6 kpc8 kpc10 kpc
u>20.5>7.5>6.6>6.0>5.5
b>20.8>7.8>6.9>6.3>5.8
v>19.5>6.5>5.6>5.0>4.5
g>24.0>11.0>10.1>9.5>9.0
r>23.7>10.7>9.8>9.2>8.7
i>23.5>10.5>9.6>9.0>8.5
FilterMagnitudeMag − dm
4 kpc6 kpc8 kpc10 kpc
u>20.5>7.5>6.6>6.0>5.5
b>20.8>7.8>6.9>6.3>5.8
v>19.5>6.5>5.6>5.0>4.5
g>24.0>11.0>10.1>9.5>9.0
r>23.7>10.7>9.8>9.2>8.7
i>23.5>10.5>9.6>9.0>8.5

Our more recent NTT images of the system in quiescence (Fig. 1, bottom panel) show a number of blended sources near the location of MAXI J1836−194. The brightest of these sources is a point-like object at a J2000 position of RA = 18:35:43.474, Dec. = −19:19:10.25. PSF photometry yields i = 21.4 ± 0.1 mag (2013 June 11) and g = 23.7 ± 0.2 mag, r = 22.4 ± 0.1 mag and i = 21.6 ± 0.2 mag (2013 June 13). This source is 0.37 arcsec away (0.32 arcsec in RA and 0.18 arcsec in Dec.) from the position derived for the optical counterpart to MAXI J1836−194 during outburst (RA = 18:35:43.451 and Dec. = −19:19:10.43, J2000; Section 2.3). Given that the images were astrometrically calibrated using the same reference stars, the uncertainty in the separation between the two objects is dominated by the uncertainty in the determination of the target profile centre. Using different algorithms implemented in IRAF to calculate these positions (centroid, Gaussian and PSF fitting), we derive positional uncertainties of 0.06 and 0.08 arcsec in RA and Dec., respectively, for the source found in the NTT images. The positional uncertainties are negligible for the bright optical counterpart detected in outburst. Therefore, the separation between the two objects is significant (5σ in RA), showing that the object in the NTT images is a field star unrelated with the transient. South-west from this source there is evidence for unresolved emission overlapping the expected position for the transient. We derive 3σ upper limits of g > 24.0 mag, r > 23.7 mag and i > 23.5 mag at this position for the best quality images (Table 1).

Optical spectroscopy

Our early outburst VLT spectra (Fig. 3) show clear signs of an optically thick, X-ray irradiated accretion disc; rotationally broadened He ii λ4686, Hα, Hβ and Hγ emission in a broad absorption trough, the Bowen region at λ ∼ 4630–4640 Å, as well as other rotationally broadened H i absorption lines from the Balmer series (identifiable up to H10) and He i absorption lines, consistent with findings presented by Pakull & Motch (2011).

Figure 3.

Top panel: averaged blue spectrum of MAXI J1836−194 obtained 2011 August 31 and 2011 September 01, with a zoomed portion showing the He ii λ4686, Hβ and Bowen region. Bottom panel: averaged red spectrum from August 31 and September 01, with a zoom of the Hα peak. Continuum intensities have been normalized at one.

We fitted Gaussian profiles to a selection of the de-blended Balmer emission and absorption components (Table 2). We found EWs of the emission components, measured from the two-night combined spectrum, to be Hα = −0.83 ± 0.02 Å, Hβ emission core = −0.25 ± 0.03 Å and Hγ emission core = −0.15 ± 0.03 Å (where negative EWs represent emission and positive EWs represent absorption). We find EWs for the Hβ and Hγ absorption components of 2.54 ± 0.07 and 2.5 ± 0.1 Å, respectively. For the broad and shallow Hα absorption trough, the FWHM is not well constrained due to the strong emission line, but we estimate the FWHM to be ≈50 Å and the full width at zero intensity (FWZI) to be ≈90 Å. The de-reddened fluxes of the Balmer series and He ii λ4686 line are presented in Table 2 and will be discussed further in Section 4. We determined the reddening, |$E(B-V) = 0.53^{+0.03}_{-0.02}$| mag, by fitting an irradiated disc model to the broad-band optical/UV/X-ray spectrum with XSPEC (TDR14). Diffuse interstellar bands provided an alternative estimate of the extinction. In particular, the λ4430 interstellar band is clearly visible with a central absorption of 6.5 per cent corresponding to E(BV) = |$0.6^{+0.2}_{-0.1}$| mag (Krelowski et al. 1987). We also get similar results from the λ6284 interstellar band using the calibration from Jenniskens & Desert (1994). Results from Marshall et al. (2006) indicate that there is an absorption layer of E(BV) ≈ 0.5 mag located within 2 kpc in the direction of the source, providing a lower limit to the distance to the source.

Table 2.

The intrinsic full width at half-maximum (FWHM), equivalent width (EW) and the de-reddened flux within the line for some notable lines at the onset of the 2011 outburst. Negative values represent emission and positive values represent absorption. Both the emission core and broad absorption of Hβ were recorded.

LineFWHMEWEmitted flux
×10−15
(Å)(Å)(erg s−1 cm−2)
Hα (emis.)6.0 ± 0.2−0.83 ± 0.023.15 ± 0.09
Hα (abs.)  ≈501.20 ± 0.05
Hβ (emis.)4.0 ± 0.4−0.25 ± 0.031.5 ± 0.2
Hβ (abs.)35.1 ± 0.92.54 ± 0.07
Hγ (emis.)4.2 ± 0.7−0.15 ± 0.031.0 ± 0.2
Hγ (abs.)23 ± 22.5 ± 0.1
He II λ46865.5 ± 0.3−0.52 ± 0.033.2 ± 0.2
Bowen blend−0.58 ± 0.063.9 ± 0.4
LineFWHMEWEmitted flux
×10−15
(Å)(Å)(erg s−1 cm−2)
Hα (emis.)6.0 ± 0.2−0.83 ± 0.023.15 ± 0.09
Hα (abs.)  ≈501.20 ± 0.05
Hβ (emis.)4.0 ± 0.4−0.25 ± 0.031.5 ± 0.2
Hβ (abs.)35.1 ± 0.92.54 ± 0.07
Hγ (emis.)4.2 ± 0.7−0.15 ± 0.031.0 ± 0.2
Hγ (abs.)23 ± 22.5 ± 0.1
He II λ46865.5 ± 0.3−0.52 ± 0.033.2 ± 0.2
Bowen blend−0.58 ± 0.063.9 ± 0.4
Table 2.

The intrinsic full width at half-maximum (FWHM), equivalent width (EW) and the de-reddened flux within the line for some notable lines at the onset of the 2011 outburst. Negative values represent emission and positive values represent absorption. Both the emission core and broad absorption of Hβ were recorded.

LineFWHMEWEmitted flux
×10−15
(Å)(Å)(erg s−1 cm−2)
Hα (emis.)6.0 ± 0.2−0.83 ± 0.023.15 ± 0.09
Hα (abs.)  ≈501.20 ± 0.05
Hβ (emis.)4.0 ± 0.4−0.25 ± 0.031.5 ± 0.2
Hβ (abs.)35.1 ± 0.92.54 ± 0.07
Hγ (emis.)4.2 ± 0.7−0.15 ± 0.031.0 ± 0.2
Hγ (abs.)23 ± 22.5 ± 0.1
He II λ46865.5 ± 0.3−0.52 ± 0.033.2 ± 0.2
Bowen blend−0.58 ± 0.063.9 ± 0.4
LineFWHMEWEmitted flux
×10−15
(Å)(Å)(erg s−1 cm−2)
Hα (emis.)6.0 ± 0.2−0.83 ± 0.023.15 ± 0.09
Hα (abs.)  ≈501.20 ± 0.05
Hβ (emis.)4.0 ± 0.4−0.25 ± 0.031.5 ± 0.2
Hβ (abs.)35.1 ± 0.92.54 ± 0.07
Hγ (emis.)4.2 ± 0.7−0.15 ± 0.031.0 ± 0.2
Hγ (abs.)23 ± 22.5 ± 0.1
He II λ46865.5 ± 0.3−0.52 ± 0.033.2 ± 0.2
Bowen blend−0.58 ± 0.063.9 ± 0.4

We determine the heliocentric velocity of the disc on both nights by fitting Gaussian profiles to a selection of suitable emission and absorption lines (Table 3) and measuring their average velocity shifts. The average recessional velocity is 68 ± 20 km s−1 for the first night of VLT observations, 55 ± 17 km s−1 for the second night and 61 ± 15 km s−1 for the combined two-night spectrum. We also obtain consistent velocity shifts when we use only the sample of emission lines (59 ± 9 km s−1) or the sample of absorption lines (60 ± 16 km s−1); we thus conclude that both emission and absorption lines in Table 3 come from the disc surface. In principle the recessional velocity may be different on each night due to the orbital phase of the system because the disc follows the orbital motion of the BH. However, from the derived system parameters (presented in Section 4) we verify that the amplitude of the projected radial velocity of the BH around the centre of mass is ∼9 km s−1 for a 5 M BH and ∼5 km s−1 for a 12 M BH, so this effect is well within the uncertainties from our averaged velocities.

Table 3.

Line positions and calculated systemic velocity for each individual night and for the two-night combined spectrum. The 1σ uncertainties in the stronger lines (Hα, Hβ and He ii λ4686) are ±0.2 Å, and ±0.3 Å for all other lines.

LineNight 1Night 2Combined
(Å)(Å)(Å)(Å)
Hα 6562.86564.16563.76563.8
Hβ 4861.34862.44861.94862.3
Hγ 4340.54341.24341.34341.3
Hδ 4101.74102.04102.14102.1
H8 3889.13890.33889.73889.7
H9 3835.43836.43836.23836.3
H10 37993800.13799.73800.0
He i 4026.24027.04027.34027.0
He i 4143.84145.24145.14145.0
He i 4387.94389.04388.94389.0
He i 4471.54472.54472.04472.5
Mg ii 4481.34482.14482.24482.1
He ii 46864686.94686.74686.8
Velocity (km s−1)68 ± 2055 ± 1761 ± 15
LineNight 1Night 2Combined
(Å)(Å)(Å)(Å)
Hα 6562.86564.16563.76563.8
Hβ 4861.34862.44861.94862.3
Hγ 4340.54341.24341.34341.3
Hδ 4101.74102.04102.14102.1
H8 3889.13890.33889.73889.7
H9 3835.43836.43836.23836.3
H10 37993800.13799.73800.0
He i 4026.24027.04027.34027.0
He i 4143.84145.24145.14145.0
He i 4387.94389.04388.94389.0
He i 4471.54472.54472.04472.5
Mg ii 4481.34482.14482.24482.1
He ii 46864686.94686.74686.8
Velocity (km s−1)68 ± 2055 ± 1761 ± 15
Table 3.

Line positions and calculated systemic velocity for each individual night and for the two-night combined spectrum. The 1σ uncertainties in the stronger lines (Hα, Hβ and He ii λ4686) are ±0.2 Å, and ±0.3 Å for all other lines.

LineNight 1Night 2Combined
(Å)(Å)(Å)(Å)
Hα 6562.86564.16563.76563.8
Hβ 4861.34862.44861.94862.3
Hγ 4340.54341.24341.34341.3
Hδ 4101.74102.04102.14102.1
H8 3889.13890.33889.73889.7
H9 3835.43836.43836.23836.3
H10 37993800.13799.73800.0
He i 4026.24027.04027.34027.0
He i 4143.84145.24145.14145.0
He i 4387.94389.04388.94389.0
He i 4471.54472.54472.04472.5
Mg ii 4481.34482.14482.24482.1
He ii 46864686.94686.74686.8
Velocity (km s−1)68 ± 2055 ± 1761 ± 15
LineNight 1Night 2Combined
(Å)(Å)(Å)(Å)
Hα 6562.86564.16563.76563.8
Hβ 4861.34862.44861.94862.3
Hγ 4340.54341.24341.34341.3
Hδ 4101.74102.04102.14102.1
H8 3889.13890.33889.73889.7
H9 3835.43836.43836.23836.3
H10 37993800.13799.73800.0
He i 4026.24027.04027.34027.0
He i 4143.84145.24145.14145.0
He i 4387.94389.04388.94389.0
He i 4471.54472.54472.04472.5
Mg ii 4481.34482.14482.24482.1
He ii 46864686.94686.74686.8
Velocity (km s−1)68 ± 2055 ± 1761 ± 15

The Hα and He ii λ4686 lines are single peaked and narrower than typically observed in Galactic LMXBs. Therefore, instead of using the peak-to-peak separation, we use their FWHM to provide an approximation of twice the projected Keplerian velocity of the disc annulus of the emitting region. For Hα we measure a Gaussian FWHM of 6.40 ± 0.20 Å for night 1 and 5.75 ± 0.20 Å for night 2. The He ii λ4686 line profile produces FWHM = 5.57 ± 0.30 Å on night 1 and FWHM = 5.43 ± 0.30 Å on night 2. We deconvolved the instrumental resolution for both the red and the blue spectra on the second night (seeing limited). Averaging the intrinsic FWHMs from both nights, we obtain a velocity of 260 ± 6 km s−1 for Hα and 333 ± 14 km s−1 for He ii λ4686. We assume that this corresponds to approximately twice the Keplerian velocity of the disc rings that give the highest contribution to those emission lines, and we use these results to determine the disc inclination (Section 4.2).

ANALYSIS AND DISCUSSION

During the first few days after the discovery of the outburst, MAXI J1836−194 was in the hard spectral state with a significant contribution to the optical/UV emission from the optically thin synchrotron emission from the radio jet (a spectral index of α = −0.6 ± 0.1, where flux Sν∝να; Russell et al. 2013). In addition to this power-law synchrotron component, the optical emission on 2011 August 31 (night 1) and September 01 (night 2) contained a multicolour thermal component from the irradiated accretion disc. From multiwavelength spectral fitting, we estimate that ≈55 per cent of the B-band flux, and ≈70 per cent of the V-band flux came from the synchrotron component on those dates (TDR14). Two weeks after the VLT observations, the disc became the dominant source of optical emission, not because of a dramatic increase in its flux, but because of the temporary disappearance of the synchrotron component due to jet quenching (Russell et al. 2013), resulting in a reduction in the total optical/UV emission (Fig. 2).

Emission lines

The emission lines or cores of the Balmer series are seen clearly at Hα, Hβ and Hγ, with Hβ and Hγ being dominated by pressure-broadened absorption troughs (Mayo, Wickramasinghe & Whelan 1980). In particular, the FWHM of the Hα and Hβ absorption troughs are ≈50 and ≈35 Å, respectively, while the FWZI of the Hα and Hβ absorption troughs are ≈90 and ≈75 Å, similar to other Galactic LMXBs (e.g. GRO 1655−40 and XTE J1118+480; Soria, Wu & Hunstead 2000; Torres et al. 2002).

While the spectra look remarkably similar, the EW of the Hα emission line (−0.83 ± 0.02 Å) is much smaller (in absolute value) than what is usually found in Galactic BH transients early in an outburst, when typical magnitudes of the EWs are ∼3–6 Å (Fender et al. 2009). The reason for this significantly lower EW is the optically thin synchrotron component from the jet dominating the red part of the optical spectrum, which we shall argue in Section 4.2 is due to beaming because of the face-on nature of the system. Accounting for the jet component (model presented in TDR14), we find that the EW of Hα with respect to the disc continuum alone is in fact −4.2 ± 0.1 Å, similar to other Galactic XRBs (Fender et al. 2009). Similarly, the EW of He ii λ4686 with respect to the total blue continuum is −0.52 ± 0.03 Å, but with respect to the disc continuum alone is −1.6 ± 0.1 Å. The comparable EWs with respect to the disc component imply that the conditions in the disc in terms of optical depth and irradiation are similar to other LMXBs in the early stages of outburst.

From the de-reddened fluxes we find a Balmer decrement of f/f = 2.1 ± 0.3, lower than the canonical value of ≈2.87 valid for Case-B recombination of optically thin gas at T ≈ 104 K (Osterbrock & Ferland 2006); also, the ratio of f/f = 0.71 ± 0.2 is slightly higher than the expected Case-B recombination value of ≈0.466 (Osterbrock & Ferland 2006). The most likely explanation of such a discrepancy is that the lines are emitted in a layer with high electron density and optical depth, τ > 1, where the collisional excitation and de-excitation become dominant and case-B recombination is no longer valid. From Drake & Ulrich (1980), the measured Balmer ratios from our spectra suggest an electron density of ∼1012–1014 cm−3, consistent with their expectations for the upper layers of an optically thick accretion disc.

For He ii λ4686 we assume instead that collisional processes are not significant, at least in the outer disc. In this case the de-reddened flux of this line, |$f_{\rm {He\,{\scriptscriptstyle {II}}}} = (3.2\pm 0.2) \times 10^{-15}$| erg cm−2 s−1 =(7.5 ± 0.4) × 10−4 λ4686 photons cm−2 s−1 can be used to estimate the fraction of X-ray luminosity intercepted by the disc. Following Pakull & Angebault (1986), for X-ray photoionized gas, the He ii λ4686 photon flux is proportional to the illuminating flux of ionizing photons above 54 eV; for an optically thick region, the scaling factor is simply the ratio between the He+ recombination coefficient (summed over all levels above the ground state) and the effective recombination coefficient for the n = 4 → 3 transition. This ratio varies between ≈4.2 and ≈6.3 for temperatures T ≈ 10 000–40 000 K (Osterbrock & Ferland 2006), and implies that approximately one He ii λ4686 photon is emitted for every ∼4–6 UV/X-ray photons that are intercepted by the disc. The unabsorbed, isotropic photon flux above 54 eV measured from Swift XRT data taken on August 31 is ≈3.9 photons cm−2 s−1 (TDR14), meaning that we should observe approximately (0.65–0.98) fout λ4686 photons cm−2 s−1, where fout is the disc interception and reprocessing fraction. As  ≈7.5 × 10−4 λ4686 photons cm−2 s−1 are observed this means that the outer accretion disc must be intercepting ∼10−3 of the UV/X-ray photons that are emitted from closer to the BH; this reprocessing fraction is in good agreement with theoretical models (Vrtilek et al. 1990; de Jong, van Paradijs & Augusteijn 1996; Hynes et al. 2002; Gierliński, Done & Page 2009), demonstrating that, even at this early stage of the outburst, the accretion disc is already relatively large, bright and optically thick.

Inclination angle of the disc

The profiles and widths of the Hα and He ii λ4686 emission lines are good indicators of the Keplerian rotational velocity of the disc annuli responsible for their emission. Instead of the usual flat-topped or double-peaked profile (Paczyński 1977; Smak 1981), we observe single-peaked and relatively narrow Hα and He ii λ4686 profiles (Fig. 3), which are indicative of a face-on accretion disc (Smak 1981). Using the FWHM provides a good approximation of twice the Keplerian velocity of the disc annulus that contributes most to the line emission.2

First, we determined whether the peak of the Hα emission originates in the outermost disc annulus, that is, whether RRout. To determine Rout, we modelled the broad-band optical/UV/X-ray spectrum (TDR14) with the irradiated disc model diskir (Gierliński, Done & Page 2008) within XSPEC (Arnaud 1996); the main free parameters of this model are the colour temperature of the inner disc Tin, the disc blackbody normalization constant K, the fraction of X-ray flux intercepted and reprocessed in the outer disc fout and Rout expressed as a function of the inner disc (fitting) radius rin. The X-ray data provide strong constraints on Tin and K (and therefore also rin; Soria 2007), while the optical/UV data constrain Rout/rin and the irradiation fraction. In particular, we find typical irradiation fractions f ≈ 10−3 (in agreement with what we estimated from the X-ray/λ4686 flux ratio, Section 4.1), and typical values of y ≡ log (Rout/rin) ≈ 4.1 (TDR14). Since rinD10 kpc [K/cos (i)]1/2, we obtain Rout = 10yrin ≈ 1.4 × 1011cos (i)−1/2D10 kpc cm. The Rayleigh–Jeans tail of the irradiated outer disc spectrum suggests that T(Rout) ≈ 15 000 K. This characteristic temperature confirms that the outermost disc annulus must be a strong Balmer emitter (Osterbrock & Ferland 2006). Therefore, we are justified in assuming that the projected rotational velocity of the outer disc |$v_{R_{\rm out}} \approx v_{\rm H\alpha } \approx 130 \pm 4$| km s−1.

Using other well-known analytic models of irradiated discs (Vrtilek et al. 1990; King, Kolb & Szuszkiewicz 1997; Dubus et al. 1999), we obtain a temperature range ≈10 000–20 000 K for a radius ≈1011 cm and an X-ray luminosity ≈1037 erg s−1; typical of this system and consistent with the Hα emission. Even if we did not have any X-ray information available, we could still estimate Rout and T(Rout) simply by fitting a blackbody curve to the characteristic spectral peak in the optical/UV region that is typical of irradiated discs. This approximation can be qualitatively understood from the fact that the outer disc annuli dominate the optical emission by virtue of their large emitting area and moderately flat radial temperature gradient. It can also be quantitatively verified by re-fitting a simple blackbody to the optical/UV peak of a full irradiated disc spectrum. With this simpler method, we obtain a characteristic value of Rout ≈ 1.4 × 1011cos (i)−1/2D10 kpc cm at a temperature T ≈ 15 000 K, in agreement with what we obtained from our full optical to X-ray model. We can now write
\begin{eqnarray} v_{\rm H\alpha } & \approx &v_{\rm out} \displaystyle \approx \sqrt{\frac{G M_{\rm BH}}{R_{\rm out}}} \sin (i)\nonumber \\ &\approx &\displaystyle\frac{\left( G M_{\rm BH} \right)^{1/2}}{\left( 10^y D_{\rm 10\,kpc}\right)^{1/2} K^{1/4}} \sin (i)[\cos (i)]^{1/4}, \end{eqnarray}
(1)
where v ≈ 130 ± 4 km s−1. The rotational broadening of the emission cores in higher Balmer lines (v ≈ 123 km s−1; v ≈ 146 km s−1; Table 2) is comparable to that of Hα; however, the uncertainty is larger because Hβ and Hγ are dominated by the broader absorption trough. The rotational broadening of He ii λ4686 is slightly higher (⁠|$v_{{\rm He }\, {\small II}} = 166 \pm 7$| km s−1), consistent with peak emission at a characteristic radius |$R_{{\rm He}\,{\small II}} \approx 0.6 R_{\rm {H}\alpha }$|⁠, at higher temperatures. We numerically solve equation (1) for i as a function of distance and BH mass. We find that MAXI J1836−194 has a face-on disc (Table 4), with i between 4° and 15° (at a 90 per cent confidence limit) for a distance of 4–10 kpc and a BH mass of between 5 and 12 M (the typical range of Galactic BH masses; Kreidberg et al. 2012).
Table 4.

Outer disc radii and inclination angles for a range of plausible BH masses (Kreidberg et al. 2012) and source distances.

BH massDistance to sourceRoutInclination angle
(kpc)(×1010) cm(i°)
4|$5.3^{+1.3}_{-1.6}$||$7.0^{+2.2}_{-1.0}$|
6|$7.9^{+2.0}_{-2.3}$||$8.5^{+2.7}_{-1.2}$|
5 M
8|$10.6^{+2.6}_{-3.1}$||$9.8^{+3.1}_{-1.4}$|
10|$13.2^{+3.3}_{-3.9}$||$11.0^{+3.5}_{-1.5}$|
4|$5.3^{+1.3}_{-1.6}$||$5.5^{+1.8}_{-0.8}$|
6|$7.9^{+2.0}_{-2.3}$||$6.7^{+2.1}_{-1.0}$|
8 M
8|$10.5^{+2.6}_{-3.1}$||$7.8^{+2.5}_{-1.1}$|
10|$13.2^{+3.3}_{-3.9}$||$8.7^{+2.8}_{-1.2}$|
4|$5.3^{+1.3}_{-1.6}$||$4.5^{+1.4}_{-0.6}$|
6|$7.9^{+2.0}_{-2.3}$||$5.5^{+1.7}_{-0.8}$|
12 M
8|$10.5^{+2.6}_{-3.1}$||$6.3^{+2.0}_{-0.9}$|
10|$13.2^{+3.3}_{-3.9}$||$7.1^{+2.3}_{-1.0}$|
BH massDistance to sourceRoutInclination angle
(kpc)(×1010) cm(i°)
4|$5.3^{+1.3}_{-1.6}$||$7.0^{+2.2}_{-1.0}$|
6|$7.9^{+2.0}_{-2.3}$||$8.5^{+2.7}_{-1.2}$|
5 M
8|$10.6^{+2.6}_{-3.1}$||$9.8^{+3.1}_{-1.4}$|
10|$13.2^{+3.3}_{-3.9}$||$11.0^{+3.5}_{-1.5}$|
4|$5.3^{+1.3}_{-1.6}$||$5.5^{+1.8}_{-0.8}$|
6|$7.9^{+2.0}_{-2.3}$||$6.7^{+2.1}_{-1.0}$|
8 M
8|$10.5^{+2.6}_{-3.1}$||$7.8^{+2.5}_{-1.1}$|
10|$13.2^{+3.3}_{-3.9}$||$8.7^{+2.8}_{-1.2}$|
4|$5.3^{+1.3}_{-1.6}$||$4.5^{+1.4}_{-0.6}$|
6|$7.9^{+2.0}_{-2.3}$||$5.5^{+1.7}_{-0.8}$|
12 M
8|$10.5^{+2.6}_{-3.1}$||$6.3^{+2.0}_{-0.9}$|
10|$13.2^{+3.3}_{-3.9}$||$7.1^{+2.3}_{-1.0}$|
Table 4.

Outer disc radii and inclination angles for a range of plausible BH masses (Kreidberg et al. 2012) and source distances.

BH massDistance to sourceRoutInclination angle
(kpc)(×1010) cm(i°)
4|$5.3^{+1.3}_{-1.6}$||$7.0^{+2.2}_{-1.0}$|
6|$7.9^{+2.0}_{-2.3}$||$8.5^{+2.7}_{-1.2}$|
5 M
8|$10.6^{+2.6}_{-3.1}$||$9.8^{+3.1}_{-1.4}$|
10|$13.2^{+3.3}_{-3.9}$||$11.0^{+3.5}_{-1.5}$|
4|$5.3^{+1.3}_{-1.6}$||$5.5^{+1.8}_{-0.8}$|
6|$7.9^{+2.0}_{-2.3}$||$6.7^{+2.1}_{-1.0}$|
8 M
8|$10.5^{+2.6}_{-3.1}$||$7.8^{+2.5}_{-1.1}$|
10|$13.2^{+3.3}_{-3.9}$||$8.7^{+2.8}_{-1.2}$|
4|$5.3^{+1.3}_{-1.6}$||$4.5^{+1.4}_{-0.6}$|
6|$7.9^{+2.0}_{-2.3}$||$5.5^{+1.7}_{-0.8}$|
12 M
8|$10.5^{+2.6}_{-3.1}$||$6.3^{+2.0}_{-0.9}$|
10|$13.2^{+3.3}_{-3.9}$||$7.1^{+2.3}_{-1.0}$|
BH massDistance to sourceRoutInclination angle
(kpc)(×1010) cm(i°)
4|$5.3^{+1.3}_{-1.6}$||$7.0^{+2.2}_{-1.0}$|
6|$7.9^{+2.0}_{-2.3}$||$8.5^{+2.7}_{-1.2}$|
5 M
8|$10.6^{+2.6}_{-3.1}$||$9.8^{+3.1}_{-1.4}$|
10|$13.2^{+3.3}_{-3.9}$||$11.0^{+3.5}_{-1.5}$|
4|$5.3^{+1.3}_{-1.6}$||$5.5^{+1.8}_{-0.8}$|
6|$7.9^{+2.0}_{-2.3}$||$6.7^{+2.1}_{-1.0}$|
8 M
8|$10.5^{+2.6}_{-3.1}$||$7.8^{+2.5}_{-1.1}$|
10|$13.2^{+3.3}_{-3.9}$||$8.7^{+2.8}_{-1.2}$|
4|$5.3^{+1.3}_{-1.6}$||$4.5^{+1.4}_{-0.6}$|
6|$7.9^{+2.0}_{-2.3}$||$5.5^{+1.7}_{-0.8}$|
12 M
8|$10.5^{+2.6}_{-3.1}$||$6.3^{+2.0}_{-0.9}$|
10|$13.2^{+3.3}_{-3.9}$||$7.1^{+2.3}_{-1.0}$|

The inclination angle we determined is much lower than the ∼20°–∼80° range seen in other LMXBs (Narayan & McClintock 2005; Muñoz-Darias et al. 2013) and could be the reason why this system is so suitable for jet and jet break studies (Russell et al. 2013), as Doppler boosting would make the jet appear much brighter and may shift the jet break to higher frequencies than in similar systems with a less favourable orientation. The lack of phase-dependent modulation in the Swift/UVOT light curve (Fig. 2) supports our conclusion of a face-on accretion disc and the flat-topped shape of the hardness–intensity diagram (Ferrigno et al. 2012, fig. 5) is also consistent with other relatively low-inclination LMXBs (Muñoz-Darias et al. 2013), even though the system never settled into a full soft state. We also note that the widths of the Balmer absorption troughs are similar between this face-on system and other higher inclination systems suggesting that the broadening is mostly due to pressure effects, rather than Keplerian rotation.

Donor star size and orbital period

LMXBs are usually associated with older stellar populations (Zhang, Gilfanov & Bogdán 2012). Using solar-abundance isochrones based on the Padova stellar evolutionary models3 (Bressan et al. 2012) for stellar populations of age 1, 5 and 10 Gyr, we place mass and radius constraints (Fig. 4) on the companion star from the deepest quiescent luminosity limits of MAXI J1836−194 (Table 1). Generating a theoretical stellar population based on the Swift/UVOT and NTT photometric systems we find that the donor must be a main-sequence star with a mass <0.65 M and a radius <0.59 R (this occurs for a 1 Gyr population at a distance of 10 kpc). The binary period, P, can be determined solely from the mean density (⁠|$\bar{\rho }$|⁠) of the Roche lobe-filling companion star (Frank et al. 2002),
\begin{equation} \bar{\rho } = \frac{3 M_{2}}{4 \pi R_{2}^{3}} \cong \frac{3^5 \pi }{8\,G P^2} \cong 110 P^{-2}_{\rm h} \, {\rm \, g \, cm^{-3}}. \end{equation}
(2)
As the companion is on the lower main sequence, we know that its radius and mass are approximately equal in solar units (Kippenhahn & Weigert 1990; Frank et al. 2002), hence we place an upper limit on the binary period of <4.9 h. Table 5 presents the maximum allowable companion mass and radius, as well as the binary period for all population ages at distances of 4, 6, 8 and 10 kpc. Our derived mass, radius and orbital period values depend on the extinction. If we consider the lowest plausible reddening of E(BV) = 0.5 mag (see Section 3.2) we find that, for all plausible distances, the donor star mass and radius would only reduce by <0.05 (in M and R, respectively), and the orbital period would decrease by a maximum of ∼0.05 h. If the reddening is increased to E(BV) = 0.7 mag, the limits on the mass and radius increase by <0.05 (in M and R, respectively) and the orbital period will only increase by a maximum of 0.5 h, for all plausible distances. This demonstrates that uncertainty in the extinction does not have a significant effect on our companion star results. The mass and radius limits restrict the donor star to being a K- or M-type main-sequence star with a mass and radius that are consistent with other transient BH LMXB systems such as GRO J0422+32 and XTE J1118+480 (Remillard & McClintock 2006).
Figure 4.

Padova stellar evolutionary models for stellar populations of age 1, 5 and 10 Gyr (Bressan et al. 2012), corrected for reddening (E(BV) = 0.53 mag) to compare with the magnitudes presented in Table 1. Coloured lines represent the stellar population of given ages and black dashed lines are the upper limits on the magnitude from quiescent photometric observations.

Table 5.

Largest possible mass, size and period estimates derived from quiescent Swift/UVOT and NTT observations and Padova stellar evolutionary models of different population ages. Presented are the maximum limits for all population ages.

Distance4 kpc6 kpc8 kpc10 kpc
M2 (M)<0.33<0.45<0.55<0.65
R2 (R)<0.31<0.41<0.50<0.59
Period (h)<2.7<3.4<4.1<4.9
Distance4 kpc6 kpc8 kpc10 kpc
M2 (M)<0.33<0.45<0.55<0.65
R2 (R)<0.31<0.41<0.50<0.59
Period (h)<2.7<3.4<4.1<4.9
Table 5.

Largest possible mass, size and period estimates derived from quiescent Swift/UVOT and NTT observations and Padova stellar evolutionary models of different population ages. Presented are the maximum limits for all population ages.

Distance4 kpc6 kpc8 kpc10 kpc
M2 (M)<0.33<0.45<0.55<0.65
R2 (R)<0.31<0.41<0.50<0.59
Period (h)<2.7<3.4<4.1<4.9
Distance4 kpc6 kpc8 kpc10 kpc
M2 (M)<0.33<0.45<0.55<0.65
R2 (R)<0.31<0.41<0.50<0.59
Period (h)<2.7<3.4<4.1<4.9

Black hole mass

From the mass and radius limits we place on the companion star, and the outer disc radius from X-ray analysis (Table 4), we are able to infer a lower limit to the BH mass. The size of the Roche lobe is dependent on the mass ratio (q = M2/M1) and the binary separation a, approximated by the form (Eggleton 1983; Frank et al. 2002)
\begin{equation} \frac{R_{\rm L2}}{{\it a}} \approx \frac{0.49{\it q}^{2/3}}{0.6{\it q}^{2/3}+\ln \left(1+{\it q}^{1/3}\right)}. \end{equation}
(3)
Similarly, we obtain the average Roche lobe radius of the primary by replacing q with q−1 in equation (3). The ratio of the two radii is
\begin{equation} \frac{R_{\rm L2}}{R_{\rm L1}} \approx \frac{q^{4/3}\left[0.6q^{-2/3}+\ln \left(1+q^{-1/3}\right)\right]}{0.6q^{2/3}+\ln \left(1+q^{1/3}\right)} \approx q^{0.45}, \end{equation}
(4)
where the second approximation holds for 0.03 < q < 1. In equation (4), M2 and R2 (=RL2) are constrained by the photometric observations in quiescence, and RL1 ≈ 1.3Rout (Paczyński 1977; Whitehurst 1988) is obtained from the X-ray spectrum of the disc in outburst (as a function of distance). Thus, we can use equation (4) to place a lower limit on the mass of the primary. We find that the compact object must have a mass M1 > 1.9 M if the system is located at 4 kpc, M1 > 3.8 M at 6 kpc, M1 > 5.5 M at 8 kpc and M1 > 7.0 M at 10 kpc. This confirms that the compact object is a stellar mass BH, unless at the lowest projected distance (4 kpc), with a mass comparable to other transient BH LMXBs (Kreidberg et al. 2012). With this same method, if we could detect and measure the quiescent magnitude of the donor star (rather than just the upper limit) with deeper photometric observations, we could infer the BH mass without the need for phase-resolved spectroscopic measurements of radial velocity shifts (which are difficult to do in face-on systems).

CONCLUSIONS

We have used optical spectroscopic and photometric observations to determine or constrain the binary system parameters of MAXI J1836−194. From the optical luminosity of the irradiated disc, combined with the relative narrowness of the disc emission-line profiles, we estimate that the inclination angle of the disc is between ≈4° and ≈15° for a range of distances suggested by the X-ray luminosity. This is the lowest inclination angle found in any transient BH LMXB, and is likely to produce noticeable beaming effects in our measurements of the jet emission properties, making this system ideal for studies of the spectral break evolution during outburst (Russell et al. 2013; TDR14).

The donor star was not detected in our observations of the source after its return to quiescence; however, we can still place strong upper limits on its luminosity and determine that it must be a main sequence K or M star. For such stars, an upper limit to the luminosity also gives upper limits to the mass and radius (M2 < 0.65 M and R2 < 0.59 R), and the binary period (P < 4.7 h), because the star must fill its Roche lobe for mass transfer to occur. Combining our photometric constraints of donor star mass and radius with the outer radius of the accretion disc in outburst from X-ray analysis (Rout ≈ 1 × 1011 cm at 8 kpc), and therefore also the average radius of the BH Roche lobe, we obtain lower limits to the primary mass. This confirms that, except possibly at the lowest distance range, the primary is a typical stellar mass BH with similar mass to other Galactic LMXB transients.

We would like to thank the anonymous referee for comments that have improved this paper. We also thank Gary Da Costa and Daniel Bayliss for their contribution and help with optical monitoring. This research was supported under the Australian Research Council's Discovery Projects funding scheme (project number DP 120102393). This research has made use of the APASS data base, located at the AAVSO web site. Funding for APASS has been provided by the Robert Martin Ayers Sciences Fund.

Based on observations collected at the European Southern Observatory, Chile, under ESO Programme IDs 087.D-0914 and 089.D-0970.

2

In fact, the FWHM is a slight overestimate of the projected rotational velocity (Smak 1981), but the difference is ≲ 10 per cent for the moderately flat radial temperature profile expected in the outer region of an irradiated disc. In the absence of more direct measurements, the FWHM of the disc lines is a robust and conservative approximation to the projected rotational velocity of the emitting gas; if anything, the true projected rotational velocity may be slightly lower, which would further reduce the inclination angle estimated here.

REFERENCES

Appenzeller
I.
, et al. 
Messenger
1998
, vol. 
94
 pg. 
1
 
Arnaud
K. A.
Jacoby
G. H.
Barnes
J.
ASP Conf. Ser. Vol. 101, Astronomical Data Analysis Software and Systems V
1996
San Francisco
Astron. Soc. Pac.
pg. 
17
 
Banse
K.
Ponz
D.
Ounnas
C.
Grosbol
P.
Warmels
R.
Robinson
L. B.
Instrumentation for Ground-Based Optical Astronomy, Present and Future
1988
New York
Springer-Verlag
pg. 
431
 
Belloni
T.
Lecture Notes in Physics, Vol. 794, The Jet Paradigm: From Microquasars to Quasars
2010
Berlin
Springer-Verlag
Blackburn
J. K.
Shaw
R. A.
Payne
H. E.
Hayes
J. J. E.
ASP Conf. Ser. Vol. 77, Astronomical Data Analysis Software and Systems IV
1995
San Francisco
Astron. Soc. Pac.
pg. 
367
 
Breeveld
A. A.
, et al. 
MNRAS
2010
, vol. 
406
 pg. 
1687
 
Bressan
A.
Marigo
P.
Girardi
L.
Salasnich
B.
Dal Cero
C.
Rubele
S.
Nanni
A.
MNRAS
2012
, vol. 
427
 pg. 
127
 
Brocksopp
C.
Bandyopadhyay
R. M.
Fender
R. P.
New Astron.
2004
, vol. 
9
 pg. 
249
 
Buzzoni
B.
, et al. 
Messenger
1984
, vol. 
38
 pg. 
9
 
Charles
P. A.
Coe
M. J.
Lewin
W. H. G.
van der Klis
M.
Compact Stellar X-Ray Sources
2006
Cambridge
Cambridge Univ. Press
 
preprint (astro-ph/0308020)
Corbel
S.
Fender
R. P.
ApJ
2002
, vol. 
573
 pg. 
L35
 
de Jong
J. A.
van Paradijs
J.
Augusteijn
T.
A&A
1996
, vol. 
314
 pg. 
484
 
Drake
S. A.
Ulrich
R. K.
ApJS
1980
, vol. 
42
 pg. 
351
 
Dubus
G.
Lasota
J.-P.
Hameury
J.-M.
Charles
P.
MNRAS
1999
, vol. 
303
 pg. 
139
 
Dunn
R. J. H.
Fender
R. P.
Körding
E. G.
Belloni
T.
Cabanac
C.
MNRAS
2010
, vol. 
403
 pg. 
61
 
Eggleton
P. P.
ApJ
1983
, vol. 
268
 pg. 
368
 
Fender
R. P.
Belloni
T. M.
Gallo
E.
MNRAS
2004
, vol. 
355
 pg. 
1105
 
Fender
R. P.
Russell
D. M.
Knigge
C.
Soria
R.
Hynes
R. I.
Goad
M.
MNRAS
2009
, vol. 
393
 pg. 
1608
 
Ferrigno
C.
Bozzo
E.
Del Santo
M.
Capitanio
F.
A&A
2012
, vol. 
537
 pg. 
L7
 
Frank
J.
King
A.
Raine
D. J.
Accretion Power in Astrophysics
2002
3rd edn
Cambridge
Cambridge Univ. Press
Gierliński
M.
Done
C.
Page
K.
MNRAS
2008
, vol. 
388
 pg. 
753
 
Gierliński
M.
Done
C.
Page
K.
MNRAS
2009
, vol. 
392
 pg. 
1106
 
Henden
A. A.
Welch
D. L.
Terrell
D.
Levine
S. E.
Am. Astron. Soc. Meeting Abstr.
2009
, vol. 
214
 pg. 
407.02
 
Hynes
R. I.
Haswell
C. A.
Chaty
S.
Shrader
C. R.
Cui
W.
MNRAS
2002
, vol. 
331
 pg. 
169
 
Jain
R. K.
Bailyn
C. D.
Orosz
J. A.
McClintock
J. E.
Remillard
R. A.
ApJ
2001
, vol. 
554
 pg. 
L181
 
Jenniskens
P.
Desert
F.-X.
A&AS
1994
, vol. 
106
 pg. 
39
 
King
A. R.
Kolb
U.
Szuszkiewicz
E.
ApJ
1997
, vol. 
488
 pg. 
89
 
Kippenhahn
R.
Weigert
A.
Stellar Structure and Evolution
1990
Berlin
Springer-Verlag
Kreidberg
L.
Bailyn
C. D.
Farr
W. M.
Kalogera
V.
ApJ
2012
, vol. 
757
 pg. 
36
 
Krelowski
J.
Walker
G. A. H.
Grieve
G. R.
Hill
G. M.
ApJ
1987
, vol. 
316
 pg. 
449
 
Krimm
H. A.
, et al. 
Astron. Telegram
2012
, vol. 
3966
 pg. 
1
 
Lasota
J.-P.
New Astron. Rev.
2001
, vol. 
45
 pg. 
449
 
Maccarone
T. J.
A&A
2003
, vol. 
409
 pg. 
697
 
Marshall
D. J.
Robin
A. C.
Reylé
C.
Schultheis
M.
Picaud
S.
A&A
2006
, vol. 
453
 pg. 
635
 
Matsuoka
M.
, et al. 
PASJ
2009
, vol. 
61
 pg. 
999
 
Mayo
S. K.
Wickramasinghe
D. T.
Whelan
J. A. J.
MNRAS
1980
, vol. 
193
 pg. 
793
 
Miller-Jones
J. C. A.
Sivakoff
G. R.
Rupen
M.
Altamirano
D.
Astron. Telegram
2011
, vol. 
3628
 pg. 
1
 
Muñoz-Darias
T.
Coriat
M.
Plant
D. S.
Ponti
G.
Fender
R. P.
Dunn
R. J. H.
MNRAS
2013
, vol. 
432
 pg. 
1330
 
Narayan
R.
McClintock
J. E.
ApJ
2005
, vol. 
623
 pg. 
1017
 
Negoro
H.
, et al. 
Astron. Telegram
2011
, vol. 
3611
 pg. 
1
 
Osterbrock
D. E.
Ferland
G. J.
Astrophysics of Gaseous Nebulae and Active Galactic Nuclei
2006
Sausalito, CA
University Science Books
Paczyński
B.
ApJ
1977
, vol. 
216
 pg. 
822
 
Pakull
M. W.
Angebault
L. P.
Nature
1986
, vol. 
322
 pg. 
511
 
Pakull
M. W.
Motch
C.
Astron. Telegram
2011
, vol. 
3640
 pg. 
1
 
Poole
T. S.
, et al. 
MNRAS
2008
, vol. 
383
 pg. 
627
 
Reis
R. C.
Miller
J. M.
Reynolds
M. T.
Fabian
A. C.
Walton
D. J.
ApJ
2012
, vol. 
751
 pg. 
34
 
Remillard
R. A.
McClintock
J. E.
ARA&A
2006
, vol. 
44
 pg. 
49
 
Roming
P. W. A.
, et al. 
Space Sci. Rev.
2005
, vol. 
120
 pg. 
95
 
Russell
D. M.
Fender
R. P.
Hynes
R. I.
Brocksopp
C.
Homan
J.
Jonker
P. G.
Buxton
M. M.
MNRAS
2006
, vol. 
371
 pg. 
1334
 
Russell
D. M.
, et al. 
Astron. Telegram
2011
, vol. 
3689
 pg. 
1
 
Russell
D. M.
, et al. 
ApJ
2013
, vol. 
768
 pg. 
L35
 
Russell
T. D.
Soria
R.
Miller-Jones
J. C. A.
Curran
P. A.
Markoff
S.
Russell
D. M.
Sivakoff
G. R.
MNRAS
2014
, vol. 
439
 pg. 
1390
  
(TDR14)
Shakura
N. I.
Sunyaev
R. A.
A&A
1973
, vol. 
24
 pg. 
337
 
Smak
J.
Acta Astron.
1981
, vol. 
31
 pg. 
395
 
Soria
R.
Ap&SS
2007
, vol. 
311
 pg. 
213
 
Soria
R.
Wu
K.
Hunstead
R. W.
ApJ
2000
, vol. 
539
 pg. 
445
 
Strohmayer
T. E.
Smith
E. A.
Astron. Telegram
2011
, vol. 
3618
 pg. 
1
 
Tennant
A. F.
NASA Technical Memorandum
1991
, vol. 
4301
 
Tody
D.
Proc. SPIE
1986
, vol. 
627
 pg. 
733
 
Torres
M. A. P.
, et al. 
ApJ
2002
, vol. 
569
 pg. 
423
 
van Paradijs
J.
McClintock
J. E.
X-Ray Binaries
1995
Heidelberg
Astronomisches Rechen-Institut
pg. 
58
 
Vrtilek
S. D.
Raymond
J. C.
Garcia
M. R.
Verbunt
F.
Hasinger
G.
Kurster
M.
A&A
1990
, vol. 
235
 pg. 
162
 
Whitehurst
R.
MNRAS
1988
, vol. 
232
 pg. 
35
 
Yang
Y. J.
Wijnands
R.
Kennea
J. A.
Astron. Telegram
2012a
, vol. 
3975
 pg. 
1
 
Yang
Y. J.
Curran
P.
Kennea
J. A.
Kong
A. K. H.
Astron. Telegram
2012b
, vol. 
4255
 pg. 
1
 
Zacharias
N.
Finch
C. T.
Girard
T. M.
Henden
A.
Bartlett
J. L.
Monet
D. G.
Zacharias
M. I.
AJ
2013
, vol. 
145
 pg. 
44
 
Zhang
Z.
Gilfanov
M.
Bogdán
Á.
A&A
2012
, vol. 
546
 pg. 
A36