1: \documentclass[12pt,preprint,longabstract]{aastex}
2: \usepackage{lscape}
3:
4:
5: \def\msun{$M_{\odot}$}
6: \def\rsun{$R_{\odot}$}
7: \def\mdot{$\dot M$}
8: \def\eddmdot{$\dot m$}
9: \def\RS{$R_{\rm S}$}
10: \def\Rg{$R_{\rm g}$}
11: \def\fdg{\hbox{$.\!\!^\circ$}}
12: \def\fas{\hbox{$.\!\!\*(U$}}
13: \def\fss{\hbox{$.\!\!^s$}}
14: \def\ergsec{\hbox{erg s$^{-1}$ }}
15: \def\ergseccm{\hbox{erg s$^{-1} $cm$^{2}$ }}
16: \def\ergcm{\hbox{erg cm$^{-2}$ s$^{-1}$ }}
17: \def\erga{\hbox{erg cm$^{-2}$ s$^{-1}$ \AA$^{-1}$ }}
18: \def\ergh{\hbox{erg cm$^{-2}$ s$^{-1}$ Hz$^{-1}$ }}
19: \def\fcol{$f_{\rm col}$}
20: \def\Tin{$T_{\rm in}$}
21: \def\Teff{$T_{\rm eff}$}
22:
23: \begin{document}
24:
25: \title{A New Dynamical Model for the Black Hole Binary LMC X-1$^{\dagger}$}
26:
27: \author{Jerome A. Orosz}
28: \affil{Department of Astronomy, San Diego State University,
29: 5500 Campanile Drive, San Diego, CA 92182-1221}
30: \email{orosz@sciences.sdsu.edu}
31:
32: \author{Danny Steeghs}
33: \affil{Department of Physics, University of Warwick, Coventry, CV4 7AL, UK
34: and
35: Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA
36: 02138}
37: \email{D.T.H.Steeghs@warwick.ac.uk}
38:
39:
40: \author{Jeffrey E. McClintock, Manuel A. P. Torres, Ivan Bochkov,
41: Lijun Gou, Ramesh Narayan}
42: \affil{Harvard-Smithsonian Center for Astrophysics, 60 Garden Street,
43: Cambridge, MA 02138}
44: \email{jem@cfa.harvard.edu, mtorres@cfa.harvard.edu,
45: ibochkov@cfa.harvard.edu,lgou@cfa.harvard.edu,
46: narayan@cfa.harvard.edu}
47:
48: \author{Michael Blaschak}
49: \affil{Physics Department, California Polytechnic University, Pomona,
50: 3801 W.\ Temple Avenue
51: Pomona, CA 91768}
52: \email{mgblaschak@csupomona.edu}
53:
54:
55: \author{Alan M. Levine, Ronald A. Remillard}
56: \affil{Kavli Institute for Astrophysics and Space Research,
57: Massachusetts Institute of Technology,
58: Cambridge, MA 02139-4307}
59: \email{aml@space.mit.edu, rr@space.mit.edu}
60:
61:
62: \author{Charles D. Bailyn, Morgan M. Dwyer, Michelle Buxton}
63: \affil{Department of Astronomy, Yale University, P.O. Box 208101,
64: New Haven, CT 06520}
65: \email{charles.bailyn@yale.edu, morgan.dwyer@aya.yale.edu,
66: michelle.buxton@yale.edu}
67:
68: \altaffiltext{$\dagger$}{Based on observations made with
69: the Magellan 6.5m Clay
70: telescope at Las Campanas Observatory of the Carnegie Institution}
71:
72:
73:
74: \begin{abstract}
75: We present a dynamical model of the high mass X-ray binary LMC X-1 based
76: on high-resolution optical spectroscopy and extensive optical and
77: near-infrared photometry. From our new optical data we find an orbital
78: period of $P=3.90917 \pm 0.00005$ days. We present a refined analysis
79: of the All Sky Monitor data from {\em RXTE} and find an X-ray period of
80: $P=3.9094 \pm 0.0008$ days, which is consistent with the optical period.
81: A simple model of Thomson scattering in the stellar wind can
82: account for
83: the modulation seen in the X-ray light curves. The $V-K$ color of the
84: star ($1.17\pm 0.05$) implies $A_V=2.28\pm 0.06$, which is much larger
85: than previously assumed. For the secondary star, we measure a radius of
86: $R_2=17.0 \pm 0.8 \,R_{\odot}$ and a projected rotational velocity of
87: $V_{\rm rot}\sin i= 129.9 \pm 2.2$ km s$^{-1}$.
88: Using these measured properties to constrain the dynamical model, we
89: find an inclination of $i=36.38
90: \pm 1.92^{\circ}$, a secondary star mass of $M_2=31.79\pm
91: 3.48\,M_{\odot}$, and a black hole mass of $10.91\pm 1.41\,M_{\odot}$.
92: The present location of the secondary star in a temperature-luminosity
93: diagram is consistent with that of a star with an initial mass of
94: $35\,M_{\odot}$ that is 5 Myr past the zero-age main sequence. The star
95: nearly fills its Roche lobe ($\approx 90\%$ or more), and owing to the
96: rapid change in radius with time in its present evolutionary state, it
97: will encounter its Roche lobe and begin rapid and
98: possibly unstable mass transfer on a
99: timescale of a few hundred thousand years.
100: \end{abstract}
101:
102:
103: \section{Introduction}
104:
105: The first X-ray source to be discovered in the Magellanic Clouds,
106: LMC~X-1 (Mark et al.\ 1969), is a persistently luminous ($L_{\rm x} >
107: 10^{38}$~\ergsec) X-ray binary that has been observed by nearly all
108: X-ray missions during the past 37 years (e.g., Leong et al.\ 1971; Cui
109: et al.\ 2002). Spectroscopic studies of its optical counterpart
110: revealed an orbital period of $\approx 4$ days and a probable mass for
111: the compact star ``near $M \approx 6$~\msun'' (Hutchings et al.\ 1983,
112: 1987), making it the fourth dynamical black-hole candidate to be
113: established.
114:
115: The candidacy of this black hole has had a checkered history. Even
116: the most fundamental property of this binary system -- its orbital
117: period -- was established only very recently. For nearly
118: 20 years, the
119: accepted orbital period was 4.2288 d, the value adopted in the dynamical
120: study of Hutchings et al.\ (1987). The correct orbital period, $P =
121: 3.9081 \pm 0.0015$~d, was determined by Levine \& Corbet (2006) using
122: {\it RXTE} All-Sky Monitor (ASM)
123: X-ray data. The optical results reported in
124: this paper and further analysis of the ASM data
125: amply confirm that the Levine \& Corbet (2006)
126: period is correct. As an
127: interesting side note, the correct period is noted in the earlier
128: Hutchings et al.\ (1983) work as one of several candidate orbital
129: periods.
130:
131: Even the identity of the optical counterpart was not firmly established
132: until fairly recently. Initially, based on a rather uncertain X-ray
133: position, a B5 supergiant known as
134: R148 was favored over what is now known to be
135: the counterpart, an O7/O8 giant identified as ``star 32'' by Cowley,
136: Crampton \& Hutchings
137: (1978).
138: The counterpart was finally established through an analysis of
139: multiple ${\it ROSAT}$ HRI observations by Cowley et al.\ (1995). A
140: precise {\it Chandra} X-ray position (Cui et al.\ 2002) and the
141: agreement between the X-ray and optical periods (mentioned above) leave
142: no doubt that star 32 of Cowley et al.\ is the counterpart of LMC X-1.
143: We present high resolution
144: $V$- and $K$-band finding charts of the field in Figure
145: \ref{fc}.
146:
147:
148: In this paper, we confirm the basic model and the principal conclusions
149: presented by Hutchings et al.\ (1987) while greatly improving upon their
150: pioneering work. Thirty high resolution and
151: 14 medium-resolution spectra have allowed us to
152: reduce the uncertainty in the radial velocity amplitude $K$ by a factor
153: of 6 and to obtain a secure value for the rotational line broadening.
154: Of equal importance, we present the first optical light
155: curves and infrared magnitudes and colors
156: of LMC X-1, the analysis of which allows us to strongly constrain
157: the orbital inclination angle and hence our dynamical model of the
158: binary.
159: In \S\ref{photsec} and \S\ref{specsec} we present our new photometric
160: and spectroscopic observations. In \S\ref{resultsec}
161: we present improved measurements of the orbital period, the
162: temperature of the secondary star, and the radius of the secondary star.
163: These new observations are used to construct a dynamical model
164: of the system, which is presented in
165: \S\ref{discuss}; a summary of our results is in \S\ref{summary}.
166:
167: \section{Photometry}\label{photsec}
168:
169: \subsection{Observations and Reductions}
170:
171:
172: Optical and near infrared observations of LMC X-1 were obtained on 66
173: nights between 2007 January 6 and 2007 April 2 using the
174: ANDICAM\footnote{See http://www.astronomy.ohio-state.edu/ANDICAM.}
175: instrument on the 1.3m telescope at CTIO, which is operated by the
176: Small and Moderate Aperture Research Telescope System (SMARTS)
177: consortium\footnote{See http://www.astro.yale.edu/smarts.}. During
178: each visit to the source, a 120 second exposure in the $B$ filter and
179: a 120 second exposure in the $V$ filter were obtained simultaneously
180: with an exposure in $J$ that consisted of 10 dithered subexposures of
181: 12 seconds each. Flat field images for all filters were taken
182: nightly, and IRAF tasks were used to perform the standard
183: image reductions (i.e.\ bias subtraction and flat-fielding for the
184: optical and flat-fielding, sky subtraction, and image combination for
185: the infrared). All of the images were inspected visually, and a few
186: of them were discarded owing to poor signal-to-noise. We
187: retained 58 images in $B$, 57 images in $V$, and 57 images in $J$.
188:
189: We also observed the field of LMC X-1 on the nights of 2006 December 6
190: and December 8 using the 6.5m Magellan Baade telescope at Las Campanas
191: Observatory (LCO)
192: and the Persson's Auxiliary Nasmyth Infrared Camera
193: (PANIC; Martini et al.\ 2004). On December 6 three composite frames
194: were obtained in the $J$ band and four composite frames were obtained
195: in the $K$ band in 0\farcs 7 seeing. Each composite frame consists of
196: five 7-second dithered subexposures. On December 8, ten sequences in
197: $J$ with 3-second subexposures and eight sequences in $K$ with
198: 7-second subexposures were obtained in 0\farcs 5 seeing. The
199: conditions were photometric on both nights. The PANIC data were
200: reduced and processed with IRAF and custom PANIC software to produce
201: mosaic frames of the field (13 mosaic frames in $J$ and 12 mosaic
202: frames in $K$).
203:
204:
205: \subsection{Derivation of the Photometric Light Curves}\label{photanal}
206:
207: PSF-fitting photometry was used on the PANIC mosaic images to obtain
208: instrumental magnitudes of LMC X-1 and two nearby stars. The absolute
209: calibration was done with respect to stars from the 2MASS catalog
210: (Skrutskie et al.\ 2006). Comparison stars having less than 8000
211: detector counts in the PANIC frames were selected and checked for
212: variability. Weighted differential photometry of LMC X-1 was performed
213: with respect to seven and six comparison stars for the $J$ and
214: $K$-band mosaic images, respectively. For LMC X-1, we found a mean
215: $J$ magnitude of $J=13.76$ with an rms of 0.02 mag and a mean $K$
216: magnitude of $K=13.43$ with an rms of 0.01 mag. Optimal aperture
217: photometry was also performed for LMC X-1 and the comparison stars,
218: and we obtained similar values for the mean magnitudes. For
219: comparison, the magnitudes for LMC X-1 given in the 2MASS catalog are
220: $J=13.695\pm 0.063$ and $K=13.293\pm 0.063$, although these values
221: should be treated with caution owing to the presence of the bright
222: star R148 $\approx 6^{\prime\prime}$ to the east
223: (Figure \ref{fc}).
224:
225:
226:
227:
228: The photometric time series for the $B$ and $V$ filters were obtained
229: using the programs DAOPHOT IIe, ALLSTAR, and DAOMASTER (Stetson 1987,
230: 1992a, 1992b; Stetson, Davis, \& Crabtree 1991). The instrumental
231: magnitudes were placed on the standard scales using observations of
232: the Landolt (1992) fields RU 149, PG 1047+003, and PG 1657+078.
233: Aperture photometry was done on the standard stars and on several
234: bright and isolated stars in the LMC X-1 field and the DAOGROW
235: algorithm (Stetson 1990) was used to obtain optimal magnitudes. We
236: find for LMC X-1 an average $V$ magnitude of $V=14.60\pm 0.02$, and a
237: $B-V$ color of $B-V=0.17\pm 0.08$, marginally consistent with the
238: values of $V=14.52 \pm 0.05$ and $B-V=0.29\pm 0.02$ given in Bianchi
239: \& Pakull (1985).
240:
241: The photometric time series for the SMARTS $J$ band was obtained using
242: aperture photometry. A relatively small aperture radius ($\approx
243: 1''$) was used to exclude light from nearby sources. Stars from the
244: 2MASS catalog were used to place the instrumental magnitudes onto the
245: standard system. The $J$-band magnitudes from SMARTS are in agreement
246: with the mean $J$-band magnitude derived from the PANIC data.
247:
248: Figure \ref{lcfig1} shows the light curves phased on the photometric
249: ephemeris determined below. The light curves show the
250: double-wave modulation characteristic of ellipsoidal variations, with
251: maxima at the quadrature phases and minima at the conjunction phases.
252: The amplitude of the modulation ($\approx 0.06$ mag,
253: maximum to minimum) is not especially
254: large, which usually indicates a relatively low inclination, or a small
255: Roche lobe filling factor for the mass donor, or both.
256:
257:
258:
259: \section{Spectroscopy}\label{specsec}
260:
261: \subsection{Observations and Reductions}
262:
263: Thirty
264: spectra of LMC X-1 were obtained on the nights of 2005
265: January 19--24 using the Magellan Inamori Kyocera Echelle (MIKE)
266: spectrograph (Bernstein et al.\ 2002) and the 6.5 m Magellan Clay
267: telescope at LCO. The instrument was used in the
268: standard dual-beam mode with a $1\farcs 0\times5\farcs 0$ slit.
269: Exposure times ranged between 1200 and 2700 seconds and observing
270: conditions were good. The seeing was well below $1^{\prime\prime}$
271: most of the time with excursions as low as 0\farcs 5 and as high as
272: 1\farcs 2.
273: The pair of $2048\times4096$ pixel CCD detectors were
274: operated in the $2\times2$ on-chip binning mode.
275: The blue arm had a
276: wavelength coverage of
277: 3340--5065 \AA\ and the spectral dispersion on
278: the MIT detector was $\lambda/\Delta\lambda=100,000$ (0.03-0.05 \AA\
279: pixel$^{-1}$), whereas the red arm covered 4855--9420 \AA\ and the
280: dispersion on its SITe detector was $\lambda/\Delta\lambda=71,000$
281: (0.07-0.13 \AA\ pixel$^{-1}$). Our $1^{\prime\prime}$ slit delivered a
282: spectral resolution of $R=33,000$ and 28,000 in the blue and red arms
283: respectively. ThAr lamp exposures were obtained before and after each
284: pair of observations of the object, and several flux standards and
285: spectral-comparison stars were observed.
286:
287: MIKE was located at a Nasmyth focus, which minimizes flexure and
288: calibration problems. No instrument rotator or dispersion
289: compensation optics were available, and the position angle of the
290: spectrograph was set to the parallactic angle for an object at air
291: mass 1.3. Because all of our observations were taken below air mass
292: $\approx 1.5$, the effects of light loss due to atmospheric dispersion
293: are small and unlikely to significantly affect our results.
294:
295: The MIKE
296: data were reduced using a pipeline written by
297: Dan Kelson\footnote{http://www.ociw.edu/Code}.
298: The
299: pipeline performs all of the detector calibrations in the standard way
300: using bias frames and tungsten flat fields which were obtained each
301: afternoon. The arc spectra taken at the position of each target were
302: used to correct for the non-orthogonality between the dispersion and
303: spatial axes and construct a 2D wavelength solution for all object
304: frames. All object orders were then sky subtracted
305: using the technique discussed in Kelson (2003)
306: and optimally
307: extracted.
308: The signal-to-noise ratio per pixel in
309: the orders with useful lines was generally
310: in the range of $\approx 50-100$ for most of the spectra.
311:
312: The extracted spectra were inspected visually and artifacts due to
313: cosmic rays were removed manually using simple interpolation. In
314: addition to cosmic rays, there were some artifacts in the cores of the
315: higher Balmer lines owing to imperfect subtraction of nebular lines.
316: These artifacts were also removed by interpolation.
317:
318: Finally,
319: the individual orders in the cleaned spectra
320: were trimmed to remove the low signal-to-noise regions
321: at each end
322: and normalized using cubic spline fits.
323: The trimmed and normalized
324: orders were then merged into a pair of spectra (e.g.\
325: red arm and blue arm).
326: Figure \ref{newspecfig} shows the average
327: spectrum (in the restframe of the secondary) for much of the blue arm
328: and part of the red arm. The model spectrum shown in
329: Figure \ref{newspecfig} is discussed in \S\ref{tempsec}.
330:
331:
332: An additional fourteen spectra of LMC X-1 were obtained on the nights
333: of 2008 August 2-3 using
334: the Magellan Echellette Spectrograph (MagE; Marshall et al.\ 2008)
335: and the Clay telescope at LCO. We used the 0\farcs 7 slit
336: which yields a resolving power of $R=6000$.
337: The exposure times were 600 seconds and the seeing ranged from
338: 0\farcs 8 to 1\farcs 1 for the first night and
339: 1\farcs 4 to 2\farcs 7 for the second night. ThAr lamp exposures
340: were obtained before and after each pair of observations
341: of LMC X-1.
342:
343: The images were reduced with tasks in the IRAF `ccdproc' and `echelle'
344: packages. After the bias was subtracted from each image, pairs of
345: images were combined using a clipping algorithm to remove cosmic rays.
346: The resulting seven images were flat-fielded using a normalized master
347: flat and then rotated in order to align the background night sky
348: emission lines along the columns. After the background emission lines
349: were rectified, the spectra from individual orders were optimally
350: subtracted. Unfortunately, the rectification of the background lines
351: was not perfect, which produced artifacts due to nebular lines in the
352: cores of the higher Balmer lines. These imperfections were removed by
353: simple interpolation. The process used to merge the orders in the MIKE
354: spectra described above was also used to merge the MagE orders. The
355: resulting seven spectra cover a useful wavelength range of
356: 3200-9250~\AA.
357:
358: \subsection{Radial Velocities}\label{radvel}
359:
360: In a recent study of three high mass X-ray binaries with OB-star
361: companions, van der Meer et al.\ (2007) measured radial velocity curves
362: using single lines. They found that X-ray heating produced distortions
363: in some of the line profiles, thereby yielding for some cases different
364: $K$-velocities for different lines.
365:
366: In order to assess the effects of X-ray heating on our spectra, we
367: selected four bandpasses with strong lines that have a wide range of
368: excitation energies over which to derive radial velocities from the MIKE
369: spectra: 3780--3880~\AA, dominated by H9 \& H10 and including He I
370: $\lambda3820$; 3990-4062~\AA, He I $\lambda4026$; 4150--4250~\AA, He II
371: $\lambda4200$; and 4425--4525~\AA, He I $\lambda$ 4471. These lines
372: were selected because of their prominence in both object and template
373: spectra and their isolation from other spectral features.
374:
375: The radial velocities were determined using the {\sc fxcor}
376: task within IRAF, which is an implementation of the cross-correlation
377: analysis developed by Tonry \& Davis (1979).
378: For each bandpass, the spectra were low-frequency filtered either by
379: fitting a Legendre polynomial or by Fourier filtering. To optimize
380: the removal of any order blaze remnants, the Legendre polynomial
381: orders and the Fourier ramp filter components were determined
382: individually for each bandpass. For example, the region
383: containing H9 and H10
384: required a Legendere polynomial order of 8 (i.e.\ nine terms) or a
385: Fourier cut-on wavenumber of 4 and a full-on wavenumber of 8, whereas
386: the region containing the line He I $\lambda4471$ needed a 28-piece
387: polynomial or a cut-on of 9 and a full-on of 18.
388: The spectra were then continuum-subtracted
389: prior to computing the cross-correlation. No other filtering was
390: applied.
391:
392: The correlations were performed using three different templates: a
393: model spectrum from the OSTAR2002 grid (Lanz \& Hubeny 2003;
394: see
395: the discussion below) and the spectra of
396: two of our comparison stars: HD~93843, O6~III(f) and HD~101205,
397: O7~IIIn((f)) (Walborn 1972, 1973). Typical values of the Tonry \&
398: Davis (1979) $r$-value, a measure of signal-to-noise ratio, ranged
399: from about 10 to 70 and the median value was 26.
400:
401:
402: The radial velocity data for each bandpass were fitted to a circular
403: orbit model which returns the systemic velocity $V_{\rm 0}$, the time of
404: maximum velocity $T_{\rm max}$, and the velocity semiamplitude of the
405: secondary $K_2$. The orbital period was fixed at our adopted
406: value, which is given in Table 3. The model provided good fits to
407: the data. For each fit, the statistical errors on the velocities
408: returned by the fit were scaled by factors ranging from 0.3 to 2.1 and
409: the data were refitted in order to give a reduced chi-squared
410: ($\chi_{\nu}^{2}$) of unity.
411:
412:
413: As an aid to visualize the systematic effects, Figure \ref{Kfig}
414: summarizes the fitted values of $K_2$ and $T_{\rm max}$ obtained for
415: different combinations of
416: (i)
417: the four bandpasses, (ii) the three template stars and (iii) the two
418: modes of low-frequency filtering, Legendre and Fourier. The open
419: circles show, for example, the results obtained by fitting the velocity
420: data using the spectrum of HD 93843 as the template and employing
421: Legendre filtering. The other open symbols show for comparison the
422: results obtained with the other template/filter combinations. The
423: results shown as filled circles and bold error bars were derived using
424: the weighted mean velocities for each spectrum: $K_2 = 69.19 \pm 0.88$
425: km s$^{-1}$ and $T_{\rm max} = {\rm HJD~} 2453392.3043 \pm 0.0081$.
426: These mean values are indicated by the dashed lines.
427:
428: Figure \ref{Kfig} shows that our results for $K$ and $T_{\rm max}$ are
429: quite insensitive to the choice of bandpass, template or mode of
430: filtering, although there are some systematic effects at play. For
431: example, the weighted mean $K$-velocity of the He II line, $K
432: $($\lambda4200$)$ =70.87 \pm 0.81$ km s$^{-1}$, is greater than our
433: adopted mean value, albeit still within 2$\sigma$.
434: This modest shift in $K$ is far
435: less than was observed in the recent van der Meer
436: et al.\ (2007) study of three O-star X-ray
437: binaries that contain neutron star primaries.
438: Because we have considered lines with a wide
439: range of excitation energies (Balmer, He I and He II) and have found
440: quite consistent results, we conclude that our dynamical results are
441: little affected by X-ray heating, tidal distortion of the secondary,
442: stellar wind, and other commonly-observed sources of systematic
443: effects.
444:
445: Finally, we measured the equivalent widths of the H10, He I
446: $\lambda 4471$, and He II $\lambda 4200$ lines in the individual
447: spectra. There is no apparent trend with orbital phase
448: (see Figure \ref{ewfig}), which
449: further confirms that the effects of X-ray heating are minimal.
450:
451: Knowing that the velocities are little effected by X-ray heating, we
452: derived the final adopted velocities from the MIKE and MagE spectra
453: using {\sc fxcor}, a synthetic template,
454: and a cross correlation region that covers
455: nine He I and He II lines between 4000~\AA\ and 5020~\AA.
456: The mean velocity curve is shown plotted in Figure
457: \ref{rvfig1}.
458:
459:
460:
461: \section{Results}\label{resultsec}
462:
463:
464: \subsection{Refined Orbital Period}\label{periodogram}
465:
466: Levine \& Corbet (2006) analyzed {\em RXTE} ASM data of LMC X-1 taken
467: between 1996 March and 2005 November. They reported a periodicity of
468: $3.9081 \pm 0.0015$ days, which differs from the period of $P=4.2288$
469: days given in Hutchings et al.\ (1987). On the other hand, the X-ray
470: period is consistent with one of the possible periods given in
471: Hutchings et al.\ (1983), namely $P=3.909 \pm 0.001$ days.
472:
473: We derived a refined period from our optical data. The optical light
474: curves completely rule out the 4.2288 day period and strongly favor
475: the period near 3.909 days. We made a periodogram from the
476: Magellan radial velocities by fitting a three-parameter sinusoid
477: to
478: the data at various trial periods and recording the $\chi^2$ of
479: the fit. The results are shown in the top of Figure \ref{plotperiodogram}.
480: A unique period cannot be found from these data alone, leaving us with
481: possible periods near 3.884, 3.898 and 3.909 days.
482: Although the radial velocities
483: given in Hutchings et al.\ (1983, 1987) have relatively large errors
484: and the times given are accurate to only about one minute,
485: they can be used to rule out possible alias periods. When these
486: additional velocities are included in the analysis, only the period
487: near 3.909 days remains viable (see
488: the center panel in Figure \ref{plotperiodogram}).
489: Finally,
490: we used the ELC code (Orosz \& Hauschildt
491: 2000) to compute periodograms by modeling both our light and velocity
492: curves together. The details of the model fitting are discussed
493: thoroughly in \S\ref{ELC}.
494: For the purposes of computing a periodogram, the
495: period is held fixed at a given trial period and several other fitting
496: parameters are varied (here the orbit is assumed to be circular) until
497: the fit is optimized. The graph of $\chi^2$ vs.\ $P$ (see the bottom
498: panel of Figure
499: \ref{plotperiodogram}) then serves as a periodogram. The combination
500: of the SMARTS photometry and the Magellan radial velocities yields
501: a unique period of $3.90917\pm 0.00005$ days.
502: This period is consistent with the period found from the combination
503: of the Hutchings et al.\ (1983, 1987) and Magellan radial velocities.
504: This period is also consistent with the X-ray period found by
505: Levine \& Corbet (2006), and with the refined X-ray period that
506: we now derive.
507:
508:
509:
510: The ASM consists of
511: three Scanning Shadow Cameras (SSCs) mounted on a rotating Drive
512: Assembly \citep{asm96}. Approximately 53,700 measurements of the
513: intensity of LMC X-1, each from a 90-s exposure with a single SSC,
514: were obtained from the beginning of the {\it RXTE} mission in early
515: 1996 through 2008 June. A single exposure yields intensity estimates
516: in each of three spectral bands which nominally correspond to photon
517: energy ranges of 1.5-3, 3-5, and 5-12 keV with a sensitivity of a few
518: SSC counts s$^{-1}$ (the Crab Nebula produces intensities of 27, 23,
519: and 25 SSC counts s$^{-1}$ in the 3 bands, respectively). The X-ray
520: intensity of LMC X-1, as seen over more than 12 years in the ASM light
521: curves, has been more or less steady near 20 mCrab (1.5-12 keV) with
522: variations, when 10-day averages are considered, of $\approx \pm 10$\%.
523:
524: A periodicity in the X-ray intensity of LMC X-1 was found during a
525: search for periodicities in ASM data using advanced analysis
526: techniques. Two strategies were applied in this search to improve the
527: sensitivity. The first involves the use of appropriate weights such
528: as the reciprocals of the variances,
529: since the individual ASM measurements have a wide range of
530: associated uncertainties. The second strategy stems from the fact
531: that the observations of the source are obtained with a low duty
532: cycle, i.e., the window function is sparse (and complex). The
533: properties of the window function, in combination with the presence of
534: slow variations of the source intensity act to hinder the detection
535: of variations on short time scales. The power density
536: spectrum (PDS) of the
537: window function has substantial power at high frequencies, e.g., 1
538: cycle d$^{-1}$ and 1 cycle per spacecraft orbit ($\approx 95$ minute
539: period). Since the data may be regarded as the product of a
540: (hypothetical) continuous set of source intensity measurements with
541: the window function, a Fourier transform of the data is equivalent to
542: the convolution of a transform of a continuous set of intensity
543: measurements with the window function transform. The high frequency
544: structure in the window function transform acts to spread power at low
545: frequencies in the source intensity to high frequencies in the
546: calculated transform (or, equivalently, the PDS). This effectively
547: raises the noise level at high frequencies.
548:
549: In our analysis, the sensitivity to high frequency variations is
550: enhanced by subtracting a smoothed version of the light curve from the
551: unsmoothed light curve. To perform the smoothing, we do not simply
552: convolve a box function with the binned light curve data since that
553: would not yield any improvement in the noise level at high
554: frequencies. Rather, we ignore bins which do not contain any actual
555: measurements and we use weights based on estimates of the
556: uncertainties in the individual measurements to compute the smoothed
557: light curve.
558: The kernel used in the smoothing was a Gaussian with a
559: full width at half maximum of 30.0 days,
560: so the smoothed light curve
561: displayed only that variability with Fourier components at frequencies
562: below $\approx 0.03$ day$^{-1}$. The smoothed light curve was
563: subtracted from the unsmoothed light curve, and the difference light
564: curve was Fourier transformed. The results are illustrated in Figure
565: \ref{ASM_PDS}. The center frequency of the peak and the formal
566: uncertainty thereof are $0.25580 \pm 0.00005 $ d$^{-1}$. The
567: corresponding period is $3.9094 \pm 0.0008$. This period differs from
568: the optical period by only $1.3\times 10^{-4}$ days.
569:
570: In Figure \ref{ASMfold} we show folded X-ray light curves for the
571: 1.5-3, 3-5, and 5-12 keV photon-energy bands as well as for the
572: overall 1.5-12 keV band. There is a clear orbital modulation with the
573: minimum intensity occurring at phase 0, which corresponds to the time
574: of the inferior conjunction of the secondary star. The X-ray light
575: curves can be fitted to a function of the form $f(\phi) = a_0 +
576: a_1\cos(2\pi\phi) + a_2\sin(2\pi\phi)$. Table \ref{ASMfits} gives the
577: best-fitting coefficients for the three individual bands and for the
578: sum. In all cases, the $a_2$ terms are consistent with zero. We
579: discuss this modulation in more detail in
580: \S\ref{wind}.
581:
582:
583:
584: \subsection{Temperature and Rotational Velocity}\label{tempsec}
585:
586:
587: The effective temperature of the secondary star and its (projected)
588: rotational velocity are needed as inputs and constraints on the
589: dynamical model discussed below. Previously, Hutchings et al.\ (1983)
590: derived a spectral type of O7 for the secondary star based on the line
591: strengths of He I, He II, Si IV, and Mg II and a rotational velocity
592: of $V_{\rm rot}\sin i \approx 150$ km s$^{-1}$. Negueruela \& Coe
593: (2002) derived a spectral type of O8III based on the ratio of the He
594: II $\lambda4541$ and He I $\lambda4471$ lines and on the
595: strength of the Mg II $\lambda4481$ line. A spectral type of
596: O8III corresponds to an effective temperature of around 33,000~K
597: (Heap, Lanz, \& Hubeny 2006).
598:
599: We used model spectra from the OSTAR2002 grid (Lanz \& Hubeny 2003) to
600: derive improved values of the temperature and rotational velocity of
601: the secondary. The temperature and gravity sampling of the grid was
602: extended to 50~K and 0.05 dex, respectively, by first resampling the
603: high dispersion model spectra to a fine wavelength spacing common to
604: all models and then by interpolation of the resampled models. Hubeny's
605: program ROTINS3 was used to broaden model spectra with instrumental
606: and rotational broadening profiles, and to resample the resulting
607: models to match the wavelength sampling of the observed spectra.
608:
609:
610:
611:
612: The ``surface'' gravity of the secondary star is tightly constrained by
613: the dynamical solution discussed below and is found to be $\log g=3.50$
614: with a formal $1\sigma$ error less than 0.02 dex.
615: For the purposes of finding the temperature we allow $\log g$ to fall
616: within the generous range of $3.4\le \log g\le 3.6$.
617: With our large grid of model spectra over a large range of
618: temperatures, gravities, and rotational velocities, we performed
619: $\chi^2$ tests using nine helium lines between 4000~\AA\ and 5020~\AA.
620: The line ratios of He II $\lambda4541$/He I $\lambda 4471$ and He II
621: $\lambda 4200$/He I(+II) $\lambda4026$, which are the main
622: classification criteria for O-stars (e.g.\ Walborn \& Fitzpatrick 1990),
623: are included in this range. We find a temperature of $T_2=33,225\pm
624: 75$~K from the average MIKE spectrum and $T_2=33,200 \pm 150$~K from the
625: average MagE
626: spectrum. Figure \ref{newspecfig} shows the best-fitting model with the
627: MIKE spectrum. Generally the lines in the model spectrum match up with
628: the lines in the observed spectrum. There are a few exceptions such as
629: the O II + C III blend near 4035~\AA, the Bowen blend near 4640~\AA, the
630: He II line near 4686~\AA, and the C III line near 5690~\AA. The wings
631: and the cores of the Balmer lines are not well fit. However, in the
632: case of the MIKE spectra it is difficult to normalize the profiles of
633: the Balmer lines owing to their large widths which can take up a
634: substantial part of an individual order. The Balmer lines are matched
635: better in the MagE spectrum where it is much easier to normalize the
636: individual orders owing to the lower spectral resolution.
637:
638:
639: The errors on the temperature quoted above are the formal
640: $1\sigma$ errors. The true uncertainty on the effective
641: temperature is no doubt larger owing to the fact that the model
642: spectrum is not a perfect fit to the data. Also, there
643: are almost certainly systematic errors in the model atmospheres
644: as well. Accordingly, we will adopt a temperature of
645: $T_2=33,200$~K with an uncertainty of 500~K
646: for the analysis presented below.
647:
648:
649:
650: To refine our initial measurement of the rotational velocity of the
651: star, we used the He I lines near 4026.0~\AA\ and 4471.0~\AA\ and
652: the He II lines near 4541.4~\AA\ and 5411.4~\AA.
653: We computed large grids of models
654: with various temperatures, gravities, and rotational velocities
655: with the appropriate resolving power for the MIKE
656: spectrograph and
657: performed $\chi^2$ tests to the composite line profiles (here we
658: assume that the main broadening mechanism in the line profiles is
659: rotation). Figure \ref{vrotfig} shows the results.
660: The model profiles fit the observed profiles fairly well.
661: The rotational velocities found from the individual lines range from
662: $127.9\pm 1.0$ km s$^{-1}$ to $133.2\pm 1.2$
663: km s$^{-1}$, where the uncertainties
664: are the formal $1\sigma$ errors. The average value is
665: 129.9 km s$^{-1}$, and the standard deviation of the four measurements
666: is 2.18 km s$^{-1}$. Hence for the analysis below, we adopt a value of
667: $V_{\rm rot}\sin i=129.9\pm 2.2$ km s$^{-1}$.
668:
669:
670: \subsection{Radius of the Secondary Star and Reddening}
671:
672: Unlike the case for most Galactic X-ray binaries, the distance to LMC
673: X-1 is well-determined.
674: Given the distance, we can find the
675: radius of the secondary star if we can obtain good measurements of the
676: apparent magnitude of the star, its effective temperature, the
677: extinction, and the bolometric correction. If the radius of the
678: secondary star can be found independently, then the dynamical model of
679: the system is further constrained.
680:
681: The apparent magnitude of the system is easy to measure. The spectral
682: type of the star yields its effective temperature, and the bolometric
683: corrections can be estimated using detailed model atmosphere
684: computations (e.g.\ Lanz \& Hubeny 2003; Martins \& Plez 2006).
685: The remaining quantity, the extinction to the source, is unfortunately
686: subject to the largest uncertainties and difficult to measure.
687:
688:
689: Using spectra from the {\em International Ultraviolet Explorer} (IUE),
690: Bianchi \& Pakull (1985) determined a color excess of $E(B-V)=0.37$
691: (0.32 from extinction in the LMC and 0.05 Galactic foreground
692: extinction). Bianchi \& Pakull (1985) noted that the color excess
693: derived from the IUE data is lower than the value derived from the
694: optical color of the star. Their value of $B-V=0.29$ and
695: $(B-V)_0=-0.27$ (Martins \& Plez 2006) gives $E(B-V)=0.56$ Using our
696: measured value of $B-V=0.17$ we find $E(B-V)=0.44$. Thus we have a
697: range of color excess values of $0.37 \le E(B-V) \le 0.56$. Using
698: $A_V=R_VE(B-V)$ with $R_V=3.1$ (the mean value for the LMC, Cardelli,
699: Clayton, \& Mathis 1989), the $V$-band extinction would be in the
700: range of $1.15 \le A_V \le 1.74$.
701:
702:
703: These values of $A_V \le 1.74$ mag are at odds with the extinction
704: derived from two other techniques. In the first case, from our photometry
705: we find a
706: $V-K$ color of $V-K=1.17 \pm 0.05$.
707: O-stars, including dwarfs, giants, and supergiants of O-subtypes
708: 6 through 9, have
709: $V-K$ colors in a relatively narrow range of
710: $-0.87\le V-K \le -0.83$
711: (Martins \& Plez 2006).
712: Assuming $A_K=0.11A_V$ (Cardelli, Clayton, \&
713: Mathis 1989), we need $A_V=2.28\pm 0.06$ to match the observed $V-K$
714: color.
715: In the second case, the $V$-band extinction can be inferred
716: from the hydrogen column densities derived from X-ray spectra. Table
717: \ref{xray} gives seven determinations of the column density $N_H$ and
718: the inferred $V$-band extinction, assuming $A_V=N_H/(1.79\pm
719: 0.03)\times 10^{21}$, where $N_H$ is given in units of cm$^{-2}$
720: (Predehl \& Schmitt 1995). The values of the extinction cover a
721: considerable range: $2.57 \pm 0.12 \le A_V \le 4.53\pm 0.29$. We
722: believe the first measurement given in the table ($A_V=2.57\pm 0.12$
723: from Cui et al.\ 2002) is the most reliable given the model used (the
724: mult-temperature disk blackbody model) and the data (high resolution
725: grating spectra
726: from {\em Chandra} HETG). This measurement differs by the extinction
727: derived from the $V-K$ color by only $2.5\sigma$.
728: Furthermore, we note that the X-ray
729: column density may systematically exceed the value inferred
730: from optical reddening (i.e.\ dust)
731: considerations, if even a fraction of the dense stellar wind from
732: the O-star is not completely ionized.
733:
734: Adopting $A_V=2.28$ mag, the range of color excess values quoted above
735: ($0.37 \le E(B-V) \le 0.56$) gives $4.07 \le R_V \le 6.16$. Our
736: measured value of color excess of $E(B-V)=0.44$ gives $R_V=5.18$. The
737: nominal average value in the LMC is $R_V=3.1$, although certain
738: lines-of-sight can have $R_V$ values that differ significantly from
739: the average (Cardelli, Clayton, \& Mathis 1989).
740:
741: To compute the radius of the star, we assume a distance modulus to the
742: LMC of $18.41\pm 0.10$ (see Orosz et al.\ 2007). Although the
743: uncertainty of the distance modulus to the center of the LMC itself is
744: somewhat smaller than 0.1 mag, we adopt an uncertainty of 0.1 mag to
745: account for the unknown relative position of LMC X-1 with respect to the
746: LMC center. We use bolometric corrections for the $V$ and $K$ bands
747: derived from the OSTAR2002 models with the LMC mean metalicity (Lanz \&
748: Hubeny 2003).\footnote{The OSTAR2002 bolometric corrections are in good
749: agreement with the bolometric corrections given in Martins \& Plez
750: (2006). We prefer the former since they are tabulated in terms of the
751: effective temperature and gravity, although the bolometric corrections
752: are insensitive to the gravity.}
753: We use apparent magnitudes of $V=14.6\pm 0.05$ and $K=13.43\pm 0.03$
754: (where we have increased the uncertainties to be conservative), and
755: assume the star is the only source of optical light in the system
756: (see the discussion in the Appendix regarding the likely faintness of
757: the accretion disk).
758: We use an
759: effective temperature of $T_2=33,200\pm 500$~K, and set the gravity to
760: $\log g=3.50\pm 0.02$, as determined from the dynamical model described
761: below. We computed the stellar radius using the $V$-band magnitudes and
762: $K$-band magnitudes using a wide range of values of $A_V$, and the
763: results are shown in Figure \ref{radfig}. The lines cross near
764: $A_V\approx 2.2$ mag (where $R_2\approx17\,R_{\odot}$), which is not at
765: all surprising since we derived $A_V=2.28 \pm 0.06$ mag from the $V-K$
766: color above.
767: Note that the radius derived from the $K$ band is relatively insensitive
768: to
769: the assumed value of $A_V$, owing to the fact that $A_K=0.11 A_V$. The
770: radii found using the $K$-band are in the range $15.5\lesssim R_2
771: \lesssim 18.3\,R_{\odot}$ (this includes the $1\sigma$ errors) for
772: extinction values in the generous range of $1.4\le A_V \le 2.8$ mag.
773: For a given value of $A_V$, the uncertainty in the radius derived from
774: the $V$ band is on the order of $1.0\,R_{\odot}$ when the generous
775: uncertainties in the apparent magnitudes and the large range in
776: effective temperatures are used. For the $K$ band, the corresponding
777: uncertainty is on the order of $0.8\,R_{\odot}$. For the purposes of
778: the dynamical model below we adopt $R_2=17.0\pm 0.8\,R_{\odot}$, which
779: is the value we find using the $K$-band and $A_V=2.28 \pm 0.06$. The
780: luminosity of the star is then $\log L/L_{\odot}=5.50 \pm 0.05$.
781:
782:
783:
784: As noted above, the extinction derived from the IUE spectra is
785: unusually low. In order to see if the IUE spectra could be explained
786: by a different extinction law, we obtained the IUE spectra of LMC X-1
787: from the MAST archive. The spectra have been reprocessed using the
788: latest calibrations, and we combined all of the observations to
789: increase the signal-to-noise. The mean IUE spectrum was dereddened
790: using $A_V=2.28$ and $R_V=5.18$. The results are shown in Figure
791: \ref{plotiue}. We also show a model spectrum from the OSTAR2002 grid
792: with $T_{\rm eff}=32,500$~K and $\log g=3.5$, scaled using a distance
793: modulus of 18.41 and a stellar radius of $17.0 \,R_{\odot}$. The
794: agreement between the data and the model is good redward of the
795: 2200~\AA\ bump. However, the 2200~\AA\ bump itself seems
796: over-corrected, and the slope of the dereddened spectrum blueward of
797: the bump is much too flat. However, the following caveats should be
798: noted regarding the IUE data. First, the signal-to-noise of the
799: spectra is not terribly high (on the order of 10 per pixel or less).
800: Second, LMC X-1 was observed in the IUE large aperture ($10''$), and
801: the source was not centered in the aperture in order to exclude light
802: from the much brighter star R148. Finally, Pakull \& Angebault (1986)
803: showed that LMC X-1 is in a compact He III nebula, which is in turn
804: embedded in a larger H II region known as N159 (Bianchi \& Pakull
805: 1985). Considering the very large aperture of the IUE spectra, the UV
806: spectrum of LMC X-1 might be contaminated by scattered light from nearby
807: bright stars. It would be worthwhile to obtain additional UV spectra
808: of LMC X-1 to further investigate the UV extinction.
809:
810:
811:
812: \subsection{System Parameters}\label{parmsec}
813:
814:
815: \subsubsection{Simultaneous Fits to Light and Velocity Curves}\label{ELC}
816:
817: We have several ``observables'' of the LMC X-1 binary system (e.g.\
818: the radial velocity curve of the secondary, its $B$ and $V$ light
819: curves, etc.), and we seek the physical model whose observed
820: properties best match the observed data. While the ``forward''
821: problem of computing the observable properties of a binary system is
822: relatively straightforward, the ``inverse'' problem of deriving a
823: physical model from observed data is much more challenging. The ELC
824: code (Orosz \& Hauschildt 2000) is a comprehensive code for computing
825: the forward problem using a model based on Roche geometry.
826: ELC can also solve the inverse problem using its associated optimizing codes
827: based on various numerical techniques
828: (e.g.\ a ``grid search'', an ``amoeba'', a Levenberg-Marquardt scheme,
829: a genetic code, and a Monte Carlo Markov Chain scheme).
830:
831: The ELC model as applied to LMC X-1 has several parameters, many of
832: which can be set to reasonable values based on various observed
833: properties of the system. The important parameters are related to
834: either the geometrical properties of the system or to the radiative
835: properties of the system.
836:
837: The free parameters which control the basic system geometry include
838: the orbital period $P$, the orbital separation $a$ (which then gives
839: the total mass via Kepler's third law), the ratio of the masses
840: $Q=M/M_2$, the
841: inclination $i$, and the
842: Roche lobe filling factor $f_2$ of the
843: secondary star.
844: We find that the search of
845: parameter space is made more efficient if the $K$-velocity of the
846: secondary star $K_2$ and the mass of the secondary star $M_2$ are used
847: to set the system scale instead of $Q$ and $a$. When modeling actual
848: data, one must also specify the phase zero point $T_0$, which is
849: either the time of the inferior conjunction of the secondary star (in
850: the case of a circular orbit) or the time of periastron passage (in
851: the case of an eccentric orbit). We initially
852: assume the secondary star is
853: rotating synchronously with its orbit, and that the star's rotational
854: axis is perpendicular to the orbital plane.
855: We discuss in
856: \S\ref{systematic} how our results depend on the assumption of
857: synchronous rotation.
858:
859: The parameters which control the radiative properties include the
860: average temperature of the secondary star $T_2$, its gravity darkening
861: exponent $\beta$, and its bolometric albedo $A$. Following standard
862: practice, we set $\beta=0.25$ and $A=1$, which are values appropriate
863: for stars with a radiative envelope. ELC uses specific intensities
864: derived from model atmosphere calculations, so no parameterized limb
865: darkening law is needed.
866: In modeling the LMC
867: X-1 data, we use model atmospheres computed for the LMC metalicity
868: from the OSTAR2002 grid (Lanz \& Hubeny 2003), supplemented by models
869: from the BSTAR2006 grid at lower temperatures (Lanz \& Hubeny 2007).
870:
871:
872:
873: ELC can include light from a flared accretion disk, and geometrical
874: effects due to the disk (e.g.\ eclipses of the secondary star). If the
875: disk contributes a significant amount of the light from the binary the
876: observed amplitude of the ellipsoidal light curve will decrease. Thus
877: it is critical to understand how luminous the accretion disk is relative
878: to the star. We show in the appendix that the accretion disk
879: contributes at most $\approx 10\%$ of the light in the optical and
880: most likely contributes less than
881: a few percent of the light. We
882: therefore omit the disk in the models. We discuss in \S\ref{systematic}
883: how our results depend on the assumption of no disk light.
884:
885: The effects of X-ray heating are computed
886: using a scheme adapted from Wilson (1990). The source of the
887: X-rays is assumed to be a thin disk with a radius that is
888: very small compared to the orbital separation.
889: The extent to
890: which X-ray heating changes the light (and velocity) curves depends
891: primarily on how large the X-ray luminosity is compared to the
892: bolometric luminosity of the secondary star. From our ongoing work on
893: modeling 53 {\it RXTE} PCA X-ray spectra of LMC X-1, we find that the
894: isotropic bolometric X-ray luminosity of LMC X-1 is quite steady over
895: a time span of several years and is $\approx 2.3 \times
896: 10^{38}$\ergsec. For comparison, the bolometric luminosity of the
897: star is $L\approx 1\times 10^{39}$ erg s$^{-1}$.
898: Because the star intercepts only a modest fraction of the total
899: X-ray flux, we thus
900: expect that
901: the effects of X-ray heating on the light and velocity curves will be
902: minor.
903: We show in \S\ref{systematic} that our main results are
904: insensitive to reasonable changes in $L_x$, and that systematic errors
905: in the inclination and the derived masses due to improper treatment of
906: X-ray heating are very small.
907: In the following analysis we adopt $L_x=2.3\times 10^{38}$ erg s$^{-1}$.
908:
909:
910:
911:
912: To summarize, we have seven free parameters: $i$, $K_2$, $M_2$, $P$,
913: $T_0$, $T_2$, and $f_2$. The data we model include 58
914: $B$-band and 57 $V$-band measurements from SMARTS, and 37 radial
915: velocity measurements. The $J$-band light curve from SMARTS proved to
916: be too noisy to model. The uncertainties on the individual
917: observations were scaled to give $\chi^2\approx N$ for each set
918: separately. The mean of the error bars after the scaling are 0.012
919: mag for $B$, 0.015 mag for $V$, and 2.62 km s$^{-1}$ for the radial
920: velocities. Finally, we have three additional observational
921: constraints that we include in the model ``fitness'' (see Orosz et
922: al.\ 2002): the projected rotational velocity of the star ($V_{\rm
923: rot}\sin i=129.9\pm 2.2$ km s$^{-1}$), the radius of the secondary star
924: ($R_2=17.0 \pm 0.8\,R_{\odot}$), and the fact that the X-ray source
925: is not eclipsed.
926:
927: Parameter space was thoroughly explored by running ELC's genetic
928: optimizer several times with the order of the parameters changed (this
929: gives different initial populations while leaving the volume of the
930: explored parameter space unchanged). The value of $T_2$ was confined to
931: the $2\sigma$ range $32,200\le T_2\le 34,200$~K as determined from the
932: spectra, while the other parameters were given generous ranges. The
933: final values of the parameters were refined using the grid search
934: optimizer. The uncertainties on the fitted and derived parameters were
935: found using the technique described in Orosz et al.\ (2002).
936: Table \ref{parm} gives the fitted parameters, some derived parameters,
937: and their uncertainties for our adopted circular orbit model and for an
938: eccentric model discussed in more detail below.
939: Figure \ref{plotfitted} shows curves of $\chi^2$ vs.\
940: various fitting and derived parameters of interest for the circular
941: orbit model, from which one may read off the best-fitting value of a
942: given parameter and
943: the $1\sigma$, $2\sigma$, and $3\sigma$ confidence ranges. Each
944: parameter is well constrained by the ellipsoidal model, with the
945: exception of the mean temperature of the secondary star $T_2$.
946: The fact
947: that $T_2$ is not constrained by the ellipsoidal model is not
948: surprising in this case (e.g.\ a single optical star in a non-eclipsing
949: system) and is not a cause for concern since the optical spectra
950: constrain the temperature quite nicely. We discuss in
951: \S\ref{systematic} how our results depend on the adopted value of $T_2$
952: for a wide range of values.
953: Figure \ref{lcfig1} shows the folded light curves with the
954: corresponding model curves,
955: while the folded radial velocity curve and the
956: corresponding model curves are shown in Figure \ref{rvfig1}.
957:
958:
959:
960: \subsubsection{Possible Systematic Errors in the Light Curve
961: Models}\label{systematic}
962:
963:
964: Our adopted parameters depend, among other things,
965: an assumption that the orbit is circular and the star rotates
966: synchronously with the orbit, and
967: on two measured
968: quantities that are not directly constrained (or only weakly
969: constrained) by the light or velocity curves: the mean temperature of
970: the secondary star $T_2$ and its radius $R_2$. The radius of the
971: star, when used as an observed quantity, helps constrain the scale of
972: the binary and hence the component masses. The temperature is needed
973: to find the radius independently.
974: We did numerical experiments to
975: see how our results change when nonsynchronous rotation
976: and eccentric orbits are allowed, when
977: the input temperature of the star is
978: changed, and when the radius (used as an extra constraint) is changed.
979: We also explored how our results depend on our assumption that
980: there is no disk light contamination in the optical and how our results
981: change with different assumed values of the X-ray luminosity.
982: But first, we briefly discuss the slight difference between the $K$-velocity
983: derived from simple sinusoid fits to the radial velocities and
984: the $K$-velocity found from the joint radial velocity and light curve fits.
985:
986: \paragraph{Systematic Errors in $K_2$.}
987:
988: The estimated $K$-velocity for LMC X-1 depends slightly on which lines
989: are used and how the spectra are normalized (\S\ref{radvel} and
990: Fig.\
991: \ref{Kfig}).
992: A four-parameter sinusoid fit to the thirty seven
993: radial velocities measured using all of the suitable lines
994: yields an orbital period of $P=3.909139\pm 0.000054$ d,
995: and $K_2=69.79\pm 0.65$ km s$^{-1}$, with $\chi^2=36.50$.
996: This $K$-velocity is
997: slightly smaller than the $K$-velocity found from the
998: detailed modeling of the light and velocity
999: curves ($K_2=71.61 \pm 0.67$ km s$^{-1}$).
1000: [For comparison Hutchings et al.\ (1983) derived $K_2=65\pm 7$
1001: km s$^{-1}$ for a period of $P=3.909$ d.]
1002: There are two reasons for the difference. The main
1003: reason is that the orbital phases of the radial velocities are tightly
1004: constrained when they are modeled together with the light curves. In
1005: this case, the time of the inferior conjunction of the secondary star
1006: found from a simple sinusoid fit to the velocities is slightly
1007: different from the same time found from the combined light and
1008: velocity curve modeling (see below).
1009: The second reason for the slight difference
1010: in the $K$-velocities is that ELC computes corrections to the radial
1011: velocities due to the distorted nature of the star, so the model curve
1012: is not precisely a sinusoid, even for a circular orbit.
1013:
1014: Fitting the radial velocities along with the light curves provides a
1015: physical scale for the model. If one knew the $K$-velocity ahead of
1016: time, then one in principle could fit the light curves by themselves.
1017: However, since the $K$-velocity is not known exactly, it is better to
1018: use an additional $\chi^2$ constraint for $K_2$ rather than a fixed input
1019: value. We ran fits to the light curves by themselves using several
1020: values of $K_2$ as an extra constraint, assuming an uncertainty of 0.65
1021: km s$^{-1}$. The period was fixed at the value found from the
1022: four-parameter sinusoid fit, leaving six free parameters. The results
1023: are shown in Table \ref{fixK}. Not surprisingly, the masses scale as
1024: $K_2$. Since the computation of the light curves is not completely
1025: independent of the scale of the system (the model atmosphere specific
1026: intensities are tabulated in terms of the temperature and gravity in cgs
1027: units), there is a little bit of variation in the derived values of $i$,
1028: $T_0$, and $\chi^2$.
1029: The most important result contained in Table \ref{fixK} is the fact that
1030: the
1031: derived values of the time of inferior conjunction of the secondary
1032: determined from the light curves is $\approx 2\sigma$ different than
1033: that derived from the four-parameter sinusoid fit to the radial
1034: velocities. Since the photometrically determined time of conjunction
1035: should agree with the spectroscopically determined one, it is clearly
1036: better to fit the radial velocities together with the light curves. The
1037: $K$-velocity derived from the simultaneous fit will be closer to the
1038: actual value, but at the expense of a larger $\chi^2$ value contribution
1039: from the radial velocities.
1040:
1041: \paragraph{Possible Nonzero Eccentricity and Nonsynchronous Rotation.}
1042:
1043: Although the timescales for a close binary to evolve to a circular orbit
1044: and synchronous rotation are relatively short, we should nevertheless
1045: entertain the possibility that the orbit of LMC X-1 could be slightly
1046: eccentric and/or the rotation could be nonsynchronous because the age of
1047: the massive secondary star in LMC X-1 might be less than the timescales
1048: for synchronization or circularization.
1049: Also, some examples of high mass
1050: X-ray binaries with eccentric orbits are known and include Vela X-1
1051: (Rappaport, Joss, \& McClintock 1976; Bildsten et al.\ 1997) and M33 X-7
1052: (Orosz et al.\ 2007). Having an eccentric orbit adds the eccentricity
1053: $e$ and the argument of periastron $\omega$ as free parameters (see Avni
1054: 1976 and Wilson 1979 for a discussion of generalized Roche potentials
1055: for eccentric orbits). As before, parameter space was thoroughly
1056: explored by using the genetic optimizer, and the results are shown in
1057: Table \ref{parm}.
1058:
1059: Avni (1976) also discussed the generalization of the Roche potential for
1060: nonsynchronous rotation. The figure of the star depends somewhat on how
1061: fast it is rotating, and a convenient parameterization of the rotation
1062: rate is $\Omega$, which is the ratio of the rotation frequency of the
1063: star to the orbital frequency.
1064: For a circular orbit and synchronous rotation, we have $\Omega=1$.
1065: For an eccentric orbit, the idea of synchronous rotation is somewhat
1066: harder to define because the angular orbital speed of the star changes
1067: over its orbit.
1068: One can compute $\Omega$ so that the rotation frequency
1069: of the star matches its orbital frequency at periastron (Hut 1981).
1070: We computed models for a wide range of $\Omega$ for both circular
1071: and eccentric orbits, and the results are given in Table
1072: \ref{nonsync}.
1073:
1074:
1075:
1076: The eccentricity is formally nonzero at the $\approx 3\sigma$ level
1077: since the total $\chi^2$ of the fit decreased by about 14 with
1078: the addition of two more free parameters.
1079: However, there are several reasons for believing that the nonzero
1080: eccentricity is spurious:
1081: (i) The sampling of the radial velocities is such that a best-fit sinusoid
1082: is shifted in phase relative to the light curves. In an eccentric
1083: orbit, the phases of the minimum or the maximum velocity are shifted
1084: relative to a sine curve (see Fig.\ \ref{rvfig1}), so that in the
1085: combined light+velocity curve modeling the total $\chi^2$ can be lower.
1086: Indeed,
1087: the $\chi^2$ value of the fit to the radial velocities for the eccentric
1088: model (Table \ref{parm}) in the combined light+velocity curve analysis
1089: is
1090: comparable
1091: to the $\chi^2$ of the
1092: four-parameter best-fit sinusoid fit discussed above.
1093: In other words,
1094: an eccentric orbit fitted to the radial velocities
1095: {\em alone} is not significantly better
1096: than a simple sine curve fit.
1097: (ii) The argument of periastron $\omega$ is
1098: consistent with 270$^{\circ}$. This gives one pause because tidal
1099: distortions of the secondary can give rise to distorted line profiles,
1100: which in turn give rise to systematic errors in the measured radial
1101: velocities. These velocity errors can result in fits with spurious
1102: eccentricities and with the argument of periastron at a quadrature phase
1103: (e.g.\ Wilson \& Sofia 1976; Eaton 2008).
1104: ELC does compute corrections to the radial velocities in
1105: the manner of Wilson \& Sofia (1976), but these corrections may not be
1106: completely accurate. (iii) When nonsynchronous models are considered
1107: (Table \ref{nonsync}) we find that the eccentricity is correlated with
1108: the
1109: parameter $\Omega$. As the value of $\Omega$ gets smaller (i.e.\ as the
1110: star rotates slower and slower), the eccentricity and the inclination
1111: get larger. In fact, values of $\Omega$ smaller than about 0.65 are not
1112: possible since the best-fitting inclination would produce an X-ray
1113: eclipse, which is not observed at a high level of confidence.
1114: Hence, we would have for the best-fitting model in Table 5 a star
1115: rotating at $\approx 65\%$ of
1116: the synchronous value in a moderately eccentric orbit where the X-ray
1117: source passes just below the limb of the star as seen in the plane of
1118: the sky.
1119: If we ignored the fact that there is no X-ray eclipse, our best-fitting
1120: model would have $i=90^{\circ}$.
1121: (iv) The differences between the model light curves for the
1122: eccentric orbit model and the circular orbit model are quite subtle.
1123: (v) Detailed model
1124: computations specific to LMC X-1 give timescales for synchronization and
1125: circularization that are both a few orders of magnitude shorter than the
1126: lifetime of the secondary star for most values of the initial rotation
1127: rate (Valsecchi, Willems, \& Kalogera, private communication 2008).
1128:
1129: Given these arguments which cast doubt on the reality of a nonzero
1130: eccentricity, we
1131: adopt the parameters derived from the circular orbit model. To be
1132: conservative with the uncertainties,
1133: the
1134: small differences between the parameters derived from the circular
1135: orbit model and the eccentric orbit model are taken to be a measure of
1136: the systematic errors. These systematic errors are added in
1137: quadrature to the uncertainties found for the circular orbit model to
1138: produce the final adopted uncertainties (fourth column in Table
1139: \ref{parm}).
1140:
1141:
1142: Regarding the models for nonsynchronous rotation, some trends
1143: are evident in Table \ref{nonsync}. First, as $\Omega$ goes from high
1144: values to lower values, the mass of the secondary star and the orbital
1145: increase, and the mass of the black hole decreases (i.e.\ the mass ratio
1146: $Q$ decreases).
1147: The reason for the
1148: mass ratio change is easy to understand. One can show that the
1149: rotational velocity of a star that fills its Roche lobe and is in
1150: synchronous rotation depends on the mass ratio (Wade \& Horne 1988). The
1151: same holds true for a star that underfills its Roche lobe, although the
1152: function relating the velocity to the mass ratio becomes more
1153: complicated. When the rotation is nonsynchronous, the relation between
1154: the velocity and the mass ratio changes further. The reason for the
1155: change in the inclination is also easy to understand.
1156: As the value of $\Omega$ decreases (i.e.\ as the star rotates more
1157: slowly), the star becomes less distorted. As a result,
1158: higher inclinations are needed to get roughly the same amplitudes
1159: for the light curves.
1160:
1161: Finally, as the value of $\Omega$ goes down, the value of $\chi^2$ for
1162: the fit goes down as well. However, the change in $\chi^2$ over the
1163: entire range of $\Omega$ for the circular orbit is modest
1164: ($\Delta\chi^2\approx 3.5$), whereas for the eccentric orbit the change
1165: is much larger ($\Delta\chi^2\approx 14$).
1166: We have already discussed why we believe that the non-zero eccentricity
1167: may be spurious.
1168: For the circular
1169: orbit model, we have no strong observational constraint on the value of
1170: $\Omega$. Based on other considerations, we give two reasons why we
1171: believe $\Omega=1$.
1172: First, our formal best solution corresponds to a near-grazing X-ray
1173: eclipse, which although
1174: possible, seems unlikely.
1175: Second, as discussed above, the timescales
1176: for both synchronization and circularization are much shorter than the
1177: age of the secondary star for most initial conditions. We conclude that
1178: $\Omega$ is most likely to have a value very close to unity.
1179:
1180:
1181:
1182: \paragraph{Fixed Temperature.}
1183:
1184:
1185:
1186: In our next experiment, we imagine that the temperature of the
1187: secondary is known exactly. In this case, the bolometric correction
1188: is known exactly (to the extent that one believes the model atmosphere
1189: computations). Once the bolometric correction is known, the radius of
1190: the star can be found using the apparent $K$ magnitude, the distance,
1191: and the extinction $A_V$. We ran eight simulations where the
1192: temperature was fixed at values between 30,000~K and 37,000~K in steps
1193: of 1000~K. We assumed a distance modulus of $18.41\pm 0.10$ mag and
1194: an extinction of $A_V=2.28\pm 0.06$ mag.
1195: Table \ref{fixT} gives the derived radii and luminosities for
1196: these eight cases. For each case, we fit the data using the genetic
1197: code, assuming the orbit is circular. Table \ref{fixT} also
1198: gives the resulting values of the inclination $i$, secondary star mass
1199: $M_2$, the black hole mass $M$, and the $\chi^2$ of the fit. A few
1200: trends are obvious from Table \ref{fixT}. First, as the temperature
1201: increases, the derived radius is roughly constant
1202: until $T=34,000$, after which it decreases.
1203: This is due in part to the way the bolometric corrections
1204: change with temperature. The luminosity also increases with increasing
1205: temperature.
1206: Second, as the temperature increases, the
1207: $\chi^2$ of the fit generally goes down.
1208: On the other hand, the derived component masses do not change that
1209: much over the range of temperatures. The extreme values of the black
1210: hole mass are $8.79\pm 0.75\,M_{\odot}$ and
1211: $11.21\pm 1.35\,M_{\odot}$.
1212:
1213:
1214:
1215: \paragraph{Fixed Radius.}
1216: Next, we can imagine that the radius of the
1217: secondary star is known exactly. In this case, we do not need to know
1218: either the distance to the source or the extinction. Using a
1219: temperature of $T_2=33,200\pm 500$~K, we can compute
1220: the luminosity of the star and its uncertainty.
1221: We ran eight simulations. In each we fit the data using the genetic
1222: code (again using a circular orbit) with the secondary star radius
1223: fixed at one of the eight equally-spaced values between
1224: $15.0\,R_{\odot}$ and $19.0\,R_{\odot}$.
1225: The results of this exercise (see Table \ref{fixR})
1226: can be used to judge what would happen
1227: to our results if our adopted distance to the LMC is in error, or if
1228: our adopted value of the extinction $A_V$ were in error. For example,
1229: if the LMC were actually closer than what we assume, the radius that
1230: we would derive would be smaller (with all other things being equal).
1231: As
1232: before, a few trends are evident when inspecting Table \ref{fixR}.
1233: First, it is obvious that the luminosity of the star will go up as its
1234: radius goes up. As the radius increases, the scale of the binary
1235: increases leading to larger masses. As the star's radius goes
1236: up, it fills more of its Roche lobe, resulting in lower inclinations.
1237: In fact,
1238: good solutions with
1239: radii larger than $\approx 19.0\,M_{\odot}$ are not possible since the
1240: star would then
1241: overfill its Roche lobe. Unlike the case when the temperature
1242: was fixed at various values, the $\chi^2$ of the fit changes very
1243: little when the radius is fixed at different values, except when
1244: $R_2\gtrsim 18\,R_{\odot}$.
1245:
1246:
1247: \paragraph{Variations in X-ray Heating.}
1248: As noted above, ELC assumes the source of X-rays is a thin disk
1249: with a small diameter compared to the orbital separation.
1250: Our value of $L_x$ ($\approx 2.3\times 10^{38}$ erg
1251: s$^{-1}$,
1252: which includes the geometrical factor to account for the inclination
1253: angle of the disk as seen from Earth) may need two correction factors
1254: when
1255: used as an input to the simple way in which ELC computes the effects
1256: of X-ray heating. First, if there is extinction local to the source
1257: (for example due to the outer edge of the accretion disk),
1258: some of the X-rays that would have reached the secondary star could be
1259: absorbed. If a flared accretion disk is present in the model, ELC
1260: can test which grid elements on the star are shielded from the X-ray
1261: source by the disk rim. However, it is hard to know what parameters
1262: to choose for the disk without more observational constraints.
1263: In any case, the first correction factor would be less than unity.
1264: The
1265: second correction factor comes from the fact that the X-ray source
1266: may not be a perfectly flat and thin disk
1267: that is in the orbital plane (i.e.\ the disk may be warped or there
1268: may be some kind of spherical corona that emits some of the X-rays).
1269: When there is a perfectly thin and flat disk,
1270: an observer on the star would see the X-ray
1271: emitting thin disk at an angle, and the intensity of X-rays observed
1272: at that point would be lower than it would be for the simple point source
1273: case owing to foreshortening of the disk (the foreshortening of the
1274: surface element itself is accounted for in the code).
1275: A warp in the disk or the presence of a quasi-spherical corona could
1276: increase the effective emitting area, leading to more X-rays hitting the
1277: secondary star. The second correction factor would be greater than
1278: unity.
1279:
1280: To see how much the derived parameters depend on our
1281: adopted value of $L_x$, we ran nine numerical experiments with $\log
1282: L_x$ fixed at values between 38.0 and 39.0 in steps of 0.25 dex. We
1283: also set $L_x=0$ (e.g.\ the case for no X-ray heating). We assume a
1284: circular orbit, and that the free parameters and the ranges thereof were
1285: the same as in our main work. Table \ref{fixX} gives the results.
1286: As
1287: the value of $L_x$ goes up, the $\chi^2$ of the fit remains virtually
1288: the same.
1289: In addition, there is little variation in the inclination and in
1290: the derived masses, and we conclude that our overall results are
1291: rather insensitive to our treatment of the X-ray heating.
1292:
1293: \paragraph{Variations in the Disk Contribution.}
1294: We show in the Appendix that the accretion disk contributes a few percent
1295: or less of the flux in the optical bands, and at most $\approx 10\%$
1296: when extreme ranges of parameters are considered. In the analysis above
1297: we assumed there was zero disk contribution.
1298: We ran models which included an accretion disk to see how our
1299: results depend on this assumption of zero disk contribution
1300: in the optical. In the ELC code one must specify the inner and outer
1301: disk radii, the disk opening angle, the temperature at the inner edge,
1302: and the exponent on the power-law temperature profile.
1303: Since LMC X-1 is not eclipsing, there is little
1304: to constrain the parameters of the accretion disk. Consequently we
1305: fixed the inner radius, the opening angle, and the exponent
1306: on the power-law temperature profile. The genetic code was run
1307: four times, and
1308: models where the disk fraction was within small threshold of a specified
1309: value were selected by means of an additional term in the
1310: total $\chi^2$. The various runs had the disk fraction in $V$ set at
1311: 0.05, 0.10, 0.15, and 0.20.
1312: If there is optical light from the accretion disk, then the O-star
1313: by itself would be less luminous since the apparent magnitude of the
1314: disk+star combination is fixed. As a result its computed radius would
1315: be smaller for
1316: a fixed temperature.
1317: Therefore
1318: in each of the four cases
1319: the assumed radius of the O-star was adjusted downward
1320: accordingly.
1321: The results are shown in Table \ref{diskfrac}.
1322: In general, as the disk fraction increases, the inclination increases
1323: slightly
1324: and the masses go down slightly. For a disk fraction of 10\%, the mass
1325: of the black hole decreases by $\approx 1\sigma$. We therefore conclude
1326: that our
1327: results are relatively insensitive to our assumption of no disk
1328: contribution in the optical.
1329:
1330:
1331: \section{Discussion}\label{discuss}
1332:
1333: \subsection{The Mass of the Black Hole}
1334:
1335: The precision of our measurement of the black hole mass ($10.91\pm
1336: 1.55\,M_{\odot}$) represents a great improvement over the pioneering
1337: work of Hutchings et al.\ (1987) who give a mass ``near
1338: $6\,M_{\odot}$'' for the black hole.
1339: Prior to this work, relatively
1340: little progress was made on the dynamics of LMC X-1; for example,
1341: in their recent compilation Charles \& Coe
1342: (2006) give a range of $8-20\,M_{\odot}$ for the mass of the black
1343: hole in LMC X-1.
1344: Most of the binaries that contain dynamically confirmed black holes
1345: have low mass secondary stars
1346: (where $M_2\lesssim 2.5\,M_{\odot}$), and the sample size of
1347: black hole binaries with high mass secondaries
1348: (where $M_2\gtrsim 10\,M_{\odot}$),
1349: is rather limited at
1350: the moment.
1351: There are three systems with high mass secondaries where dynamical
1352: studies unequivocably show that the compact object must be a black
1353: hole: LMC X-1, M33 X-7, and Cyg X-1.
1354: The component masses of M33 X-7 are well constrained ($M=15.65\pm
1355: 1.45\,M_{\odot}$ and $M_2=70.0\pm 6.9\,M_{\odot}$, Orosz et al.\
1356: 2007), whereas there is some disagreement on the parameters of Cyg X-1
1357: (Herrero et al.\ 1995 give masses of $\approx 10\,M_{\odot}$ and
1358: $18\,M_{\odot}$ for the black hole and secondary star, respectively
1359: while Charles \& Coe give $>4.8\,M_{\odot}$ for the black hole mass).
1360: There are two other eclipsing high mass X-ray binaries to mention
1361: in this context.
1362: 4U 1700-377 has a compact object with a mass of
1363: around $2.4\,M_{\odot}$, but this object may be a massive neutron
1364: star (Clark et al.\ 2002). IC 10 X-1
1365: has a large optical mass function that was measured using the He II
1366: $\lambda4686$ emission line ($7.64\pm 1.26\,M_{\odot}$,
1367: Silverman \& Filippenko 2008; Prestwich et al.\ 2007). Assuming the
1368: He II line properly tracks the motion of
1369: the Wold-Rayet (WR) secondary star, and assuming a mass for
1370: the WR star based on evolutionary models, the compact object in
1371: IC 10 X-1 has a mass of $\approx 30\,M_{\odot}$ or more.
1372: Fortunately, the future for studies of black hole binaries with high
1373: mass secondaries appears promising.
1374: Deep X-ray surveys of nearby galaxies are finding more and more
1375: high mass X-ray binaries that can be followed up in ground-based
1376: studies with the new generation of large telescopes and high-performance
1377: focal-plane instrumentation.
1378:
1379: \subsection{Evolutionary Status of the Secondary Star}\label{evol}
1380:
1381: Using our measured temperature and luminosity of the secondary star
1382: star ($33,200\pm 500$~K, and $\log L/L_{\odot}=5.50 \pm
1383: 0.05$) we can place the star on a temperature-luminosity
1384: diagram (Figure \ref{plothr}). The position of the star in the
1385: diagram can be compared with theoretical evolutionary tracks for
1386: single stars (Meynet et al.\ 1994) and with theoretical isochrones
1387: (Lejeune \& Schaerer et al.\ 2001).
1388: Although we recognize that
1389: the evolutionary models for massive stars are uncertain
1390: owing to assumptions regarding stellar winds and rotation,
1391: for the purposes of this discussion
1392: we will take the evolutionary tracks and the isochrones at face value.
1393: The position of the LMC X-1
1394: secondary is very close to the 5 Myr isochrone. Meynet et al.\
1395: (1994) give evolutionary tracks for ZAMS masses of 20, 25, 40, and
1396: $60\, M_{\odot}$. To get a crude idea of where the evolutionary track
1397: of a star with an initial mass of $35\,M_{\odot}$ would be, we plotted
1398: the initial luminosity vs.\ the initial mass and the initial
1399: temperature vs.\ the initial mass for the tabulated models and used
1400: quadratic interpolation to determine the initial temperature and
1401: luminosity of a star with a ZAMS mass of $35\,M_{\odot}$. The
1402: evolutionary track for the star with a ZAMS mass of $40\,M_{\odot}$
1403: was then shifted down to the interpolated position of the
1404: $35\,M_{\odot}$ star and plotted. The interpolated track with a ZAMS
1405: mass of $35\,M_{\odot}$ passes very near the position occupied by the
1406: LMC X-1 secondary. The star with a ZAMS mass of $40\,M_{\odot}$ has
1407: lost about $2.6\,M_{\odot}$ by the time its temperature has fallen to
1408: $\approx 33,000$~K. If a star with a ZAMS mass of $35\,M_{\odot}$
1409: behaves in a similar manner, then its mass would be $\approx
1410: 32.4\,M_{\odot}$ when its temperature has dropped to 33,000~K, which
1411: is consistent with what we measure ($M_2=31.79\pm 3.48\,M_{\odot}$,
1412: Table \ref{parm}).
1413:
1414: As one can see from the increasing space between the isochrones, as
1415: the age increases these massive stars move quickly across the
1416: temperature-luminosity diagram. Since the secondary star in LMC X-1
1417: nearly fills its Roche lobe (the fraction filled is about 90\%), it
1418: will soon enter into an interesting stage of binary star evolution.
1419: In round numbers, the star's current radius is $17\,R_{\odot}$ and
1420: the radius of its Roche lobe is $19\,R_{\odot}$. From interpolation
1421: of the evolutionary track, the star with a ZAMS mass of
1422: $40\,M_{\odot}$ has a radius of $17\,R_{\odot}$ at an age of 4.05
1423: Myr and a radius of $19\,R_{\odot}$ at an age of 4.22 Myr. If the
1424: evolution of a star with a ZAMS mass of $35\,M_{\odot}$ proceeds at a
1425: similar pace, the secondary star in LMC X-1 will encounter its Roche
1426: lobe in only a few hundred thousand years. Since the secondary star
1427: is substantially more massive than the black hole, the resulting mass
1428: transfer will be rapid and
1429: may be unstable (Podsiadlowski, Rappaport, \& Han 2003).
1430: If the mass transfer is stable and the binary survives the initial phase
1431: of rapid mass transfer, the present-day secondary may become detached again
1432: in a binary with a longer orbital period. On the other hand,
1433: if the rapid mass transfer is unstable,
1434: the orbit will rapidly shrink
1435: and the binary will enter a
1436: ``common envelope'' (CE) phase.
1437: The outcome of the CE phase
1438: depends on how tightly bound the envelope of the star is. If the
1439: envelope is tightly bound, then the energy liberated as the two stars
1440: move towards each other will not be enough to expel the envelope, and
1441: the two stars will merge. On the other hand, if the envelope is
1442: loosely bound, then the envelope of the star can be expelled before
1443: the cores merge, leaving behind a tight binary consisting of the
1444: present-day black hole and the core of the present-day secondary star.
1445: Although detailed computations are required to assess these scenarios,
1446: it seems that the most
1447: likely outcome of the CE phase in LMC X-1 would be a merger
1448: since the envelopes of massive stars are tightly bound, and the
1449: density gradient in a main sequence star is not nearly as steep as it
1450: is in a giant (e.g.\ Podsiadlowski et al.\ 2003). What the merger
1451: product would look like is an open question.
1452:
1453: \subsection{The O-star Wind and the ASM X-ray Light Curves}\label{wind}
1454:
1455: The X-ray minimum/maximum in the ASM X-ray light curves occurs at
1456: inferior/superior conjunction of the secondary star (Fig.\
1457: \ref{ASMfold}), as expected if the modulation is caused by scattering
1458: or absorption in a stellar wind. A simple cosine wave gives a
1459: reasonable fit to the light curves; the respective full amplitudes for
1460: the A, B and C channels are $0.072 \pm 0.010$, $0.077 \pm 0.11$, and
1461: $0.038 \pm 0.029$ (Fig.\ \ref{ASMfold} and Table \ref{ASMfits}). Thus,
1462: the modulation is largely independent of energy, which indicates that
1463: electron scattering is the main source of opacity. We now show that a
1464: simple model of Thomson scattering in a spherically symmetric stellar
1465: wind is not only reasonably consistent with the observed X-ray
1466: modulation but also yields plausible values of the wind mass flux,
1467: accretion rate, and accretion efficiency. For this section we use the
1468: values of component masses and other system parameters in the
1469: right-hand column of Table~\ref{parm}.
1470:
1471:
1472: We assume that the O star has a radiatively driven wind like those
1473: from other hot stars and that the wind velocity has a radial
1474: dependence that follows a ``beta law'', i.e., $v(r) = v_{\infty}(1 -
1475: b/r)^\beta$ where $b$ is approximately equal to the O-star
1476: photospheric radius $R_2$ \citep[e.g.,][and references
1477: therein]{dess05,kudpuls00}. The terminal velocity $v_{\infty}$ can be
1478: estimated from the escape velocity at the stellar surface and the
1479: photospheric temperature \citep{kudpuls00}. The escape velocity
1480: $v_{\rm esc}$ can be estimated in turn from the mass and radius
1481: results in Table~\ref{parm}. We obtain $v_{\rm esc} \approx 590 \pm
1482: 40$ km s$^{-1}$. In the calculations below, we take the values of
1483: $v_{\infty}/v_{\rm esc}$ and of $\beta$ to be like those in the wind
1484: model presented for M33 X-7 by Orosz et al.\ (2007), and so we adopt
1485: the values $v_{\infty} = 1400$ km s$^{-1} \sim 2.4 v_{\rm esc}$ and
1486: $\beta = 1$. This value of $v_{\infty}/v_{\rm esc}$ is appropriate
1487: for an O-star with photospheric temperature $T_2 \approx 33,000$ K
1488: \citep{kudpuls00}. The wind velocity at the radius corresponding to
1489: the location of the black hole ($r = a = 36.5$ \rsun) is then $v(a)
1490: \sim 740$ km s$^{-1}$. Adding in quadrature the transverse velocity
1491: of the black hole through the wind, we find that the gas flows by the
1492: black hole at a velocity $V \sim 880$ km s$^{-1}$. Given the
1493: substantial uncertainties in both the form of the velocity law and the
1494: values of the parameters, these estimates of $v(a)$ and $V$ must each
1495: be uncertain by as much as 50\%.
1496:
1497: In any spherically-symmetric wind model, the scattering optical depth
1498: along the line of sight to the X-ray source at superior conjunction
1499: only differs from that at inferior conjunction by the optical depth
1500: along an additional path of length $d = 2a \sin i = 43.3R_{\odot}$ at
1501: superior conjunction. Let $N_{\rm e}$ be the electron column density
1502: along this part of the superior conjunction line of sight. Then there
1503: must be a scattering optical depth along this part of the line of
1504: sight of ${N_{\rm e}}{\sigma_{\rm T}} \approx 0.07$, where
1505: $\sigma_{\rm T}$ is the Thomson cross section. This, in turn, implies
1506: $N_{\rm e} \sim 1.1 \times 10^{23}$ cm$^{-2}$. The mean electron
1507: number density along the path of length d is then $n_{\rm e} \sim 3.5
1508: \times 10^{10}$ e$^-$ cm$^{-3}$. In our adopted wind model, this mean
1509: density is $\sim 33$\% higher than the density at the black hole
1510: orbital radius (and it is also $\sim 33$\% lower than at the point of
1511: highest density along this section of the line of sight). The gas
1512: density at the black hole orbital radius is then $\rho(a) \sim 5.3
1513: \times 10^{-14}$ g cm$^{-3}$. The corresponding mass loss rate in the
1514: O-star wind is $\dot M_{W} = 4\pi{a^2}{\rho(a)}v(a) \sim 5 \times
1515: 10^{-6}$ \msun/yr$^{-1}$.
1516:
1517: If the wind of the O star in LMC X-1 is only partially ionized like
1518: the winds of ordinary O-stars, then photoelectric absorption should be
1519: manifest via a larger degree of modulation in the lower energy band
1520: ASM light curves. Is it a reasonable assumption in the case of LMC X-1
1521: that iron and the other metals in the wind have been fully stripped of
1522: their K shell electrons? Some support for this idea is provided by the
1523: remarkable photoionized He III nebula that surrounds LMC X-1 (Pakull
1524: \& Angebault 1986). More directly, at every point on the path
1525: described above the ionization parameter (Tarter et al.\ 1969) is
1526: large, $L/nR^2 > 1000$. Furthermore, the spectrum of LMC X-1 is soft
1527: (disk blackbody temperature $kT \approx 0.9$ keV), which makes it an
1528: effective ionizing agent. Under these conditions, studies of similar
1529: high-mass X-ray binaries support our conclusion that the wind is
1530: close to fully ionized in the region in question (McCray et al. 1984;
1531: Vrtilek et al. 2008).
1532:
1533:
1534:
1535: A crude estimate of the accretion rate is given by the rate at which
1536: the matter in the wind enters a cylinder of radius $r_{\rm c}
1537: \approx 2GM/V^2$ centered on the black hole, where $G$ is the
1538: gravitational constant, $M$ is the black-hole mass, and $V$ is the
1539: (above estimated) velocity of the wind relative to the black hole
1540: \citep[see, e.g.,][]{st83}. The accretion rate is then given by $\dot
1541: M_{\rm B-H} \sim {\pi}r_{\rm cap}^2{\rho(a)}{V}$. Using the value of
1542: $M$ in Table \ref{parm} and the values of $\rho(a)$ and $V$ estimated
1543: above, we obtain $r_{\rm c} \sim 5$ \rsun\ and $\dot M_{\rm B-H}
1544: \sim 3 \times 10^{-8}$ \msun\, yr$^{-1}$ with uncertainties in each
1545: value of roughly a factor of two.
1546:
1547:
1548:
1549: As noted above, we find that the isotropic bolometric X-ray luminosity
1550: of LMC X-1 is quite steady over a time span of several years and is
1551: $\approx 2.3 \times 10^{38}$\,\ergsec. With $M = 10.91$ \msun \ and the
1552: above estimated accretion rate, the implied accretion efficiency is
1553: $\eta \sim 0.1$. This is comparable to the 0.06 efficiency of a
1554: Schwarzschild black hole or the canonical value of 0.1 that is
1555: commonly used, but, considering that this efficiency estimate is
1556: uncertain by at least a factor of two, this may be fortuitous.
1557:
1558:
1559: \section{Summary}\label{summary}
1560:
1561: We present a new dynamical model of the high mass X-ray binary LMC X-1
1562: based on high quality echelle spectra, extensive optical light curves,
1563: and infrared magnitudes and colors. From our optical data we find an
1564: orbital period of $P=3.90917 \pm 0.00005$ days. We present a refined
1565: analysis of the All Sky Monitor data from {\em RXTE} and find a period
1566: of $P=3.9094 \pm 0.0008$ days, which is consistent with the optical
1567: period. A simple model of Thomson scattering in the stellar wind
1568: accounts for the 7\%
1569: modulation seen in the X-ray light curves. We find
1570: that the $V$-band extinction to the source is much higher than
1571: previously assumed. The $V-K$ color of the star ($1.17\pm 0.05$)
1572: implies $A_V=2.28\pm 0.06$, whereas several determinations of the
1573: column density from X-ray observations give $A_V>2.57\pm 0.12$. The
1574: color excess of $E(B-V)=0.44$ together with $A_V=2.27$ gives
1575: $R_V=5.18$, which is much higher than the nominal mean value in the
1576: LMC of $R_V=3.1$. For the secondary star we measure a radius of
1577: $R_2=17.0 \pm 0.8 \,R_{\odot}$ and a projected rotational velocity
1578: of $V_{\rm rot}\sin i= 129.9 \pm 2.2$ km s$^{-1}$. Using these measured
1579: properties of the companion star to constrain the dynamical model of
1580: the light and velocity curves, we find
1581: an inclination of $i=36.38 \pm 1.92^{\circ}$, a
1582: secondary star mass of $M_2=31.79\pm 3.48\,M_{\odot}$, and a black
1583: hole mass of $10.91\pm 1.41\,M_{\odot}$. The present location
1584: of the secondary star in a temperature-luminosity diagram is
1585: consistent with that of a star with an initial mass of $35\,M_{\odot}$
1586: that is 5 Myr past the zero-age main sequence. The star nearly fills
1587: its Roche lobe ($\approx 90\%$)
1588: and, owing to the rapid
1589: change in radius with time in its present evolutionary state, it will
1590: encounter its Roche lobe and begin rapid and
1591: possibly unstable mass transfer in only a
1592: few hundred thousand years.
1593:
1594:
1595: \acknowledgments
1596:
1597: This publication makes use of data products from the Two Micron All
1598: Sky Survey, which is a joint project of the University of
1599: Massachusetts and the Infrared Processing and Analysis
1600: Center/California Institute of Technology, funded by the National
1601: Aeronautics and Space Administration and the National Science
1602: Foundation. CBD, MMD, and MB gratefully acknowledge support from the
1603: National Science Foundation grant NSF-AST 0707627. The work of JEM
1604: was supported in part by NASA grant NNX08AJ55G. MB gratefully
1605: acknowledges support from the National Science Foundation through
1606: grant AST 04-53609 for the California State University Undergraduate
1607: Research Experiences (CSUURE) program at San Diego State University.
1608: DS acknowledges a STFC Advanced Fellowship as well as support
1609: through the NASA Guest Observer Program.
1610: We thank Thierry Lanz for providing the OSTAR2002 bolometric
1611: corrections computed for the $K$ band. We acknowledge helpful
1612: discussions with Bram Boroson and Tim Kallman on photoionized winds.
1613: We thank Phil Massey for helpful discussion on O-stars.
1614:
1615:
1616: \appendix
1617:
1618: \section{Accretion Disk Contamination in the Optical and Infrared Bands}
1619:
1620: The optical and infrared bands of interest to us
1621: for the ellipsoidal modeling ($B$, $V$,
1622: and $J$) are in
1623: the Rayleigh-Jeans part of the spectrum for the emission from both the
1624: secondary star and the disk (see below). Therefore, the flux we
1625: receive in any particular band is proportional to the product of the
1626: projected area of the radiating surface $A$ and the temperature $T$.
1627:
1628: The radius of the secondary star is $17.0\,R_\odot$ and its
1629: photospheric temperature is $33,200\,{\rm K}$. Therefore, for the
1630: radiation from the star, we have
1631: \begin{equation}
1632: \langle AT\rangle_{\rm sec} = 1.5\times10^{29} ~{\rm cm^2\,K}.
1633: \label{ATsec}
1634: \end{equation}
1635: In the case of the disk, we note that the circularization radius of
1636: the captured wind material is likely to be quite small. Nevertheless,
1637: the steady state disk will probably have a large outer radius since
1638: the gas needs to get rid of its angular momentum. An extreme
1639: possibility is that the angular momentum is removed solely by tidal
1640: torques from the companion star acting on the outer regions of the
1641: disk. From the discussion in Frank, King \& Raine (2002; eq.\ 5.122),
1642: the outer radius of the disk is then approximately equal to 0.9 times
1643: the Roche lobe radius. Using a standard approximation for the Roche
1644: lobe radius (Frank et al. 2002), we find for the outer radius
1645: \begin{equation}
1646: R_{\rm out} = 0.9\times0.462 \left({M_{\rm BH}\over
1647: M_{\rm BH}+M_{\rm sec}}\right)^{1/3} a =6.6\times10^{11} ~{\rm cm}.
1648: \end{equation}
1649:
1650: We further note that most of the disk will be dominated, not by local
1651: viscous dissipation, but by reprocessing of radiation emitted by the
1652: inner region of the disk. Assuming that $10\xi_{10}\%$ of the
1653: bolometric luminosity of the disk is reprocessed (the exact fraction
1654: is uncertain), the luminosity of the reprocessed radiation from the
1655: disk is $2.3\times10^{37}\xi_{10} ~{\rm erg\,s^{-1}}$. For a typical
1656: reprocessing geometry, the effective temperature of the reprocessed
1657: radiation will vary as $R^{-1/2}$ (as compared to the steeper
1658: $R^{-3/4}$ variation one expects for viscously generated flux). Let
1659: us write the disk temperature as
1660: \begin{equation}
1661: T(R) = T_{10}\,(R/10^{10}\,{\rm cm})^{-1/2},
1662: \end{equation}
1663: and let us assume that the reprocessing region of the disk extends
1664: from $R_{\rm in} \sim100R_S = 3.1\times 10^8 ~{\rm cm}$ to $R_{\rm
1665: out}$ (the precise value of $R_{\rm in}$ is not important for what
1666: follows). Then we have the condition (in cgs units)
1667: \begin{equation}
1668: 2.3\times10^{37}\xi_{10} = \int_{3.1\times10^8}^{6.6\times10^{11}}
1669: 4\pi R\sigma T_{10}^4 (R/10^{10})^{-2}dR.
1670: \end{equation}
1671: Solving, we obtain
1672: \begin{equation}
1673: T_{10} = 8.1\times10^4\xi_{10}^{1/4} \,{\rm K}.
1674: \end{equation}
1675: The disk temperature thus varies from $4.6\times
1676: 10^5\xi_{10}^{1/4}\,$K at $R_{\rm in}$ to $1.0
1677: \times10^4\xi_{10}^{1/4}\,$K at $R_{\rm out}$. Even the outermost
1678: region of the disk is hot enough that we may treat its optical/IR
1679: emission in the Rayleigh-Jeans limit.
1680:
1681: We can now estimate the value of $\langle AT\rangle$ for the disk
1682: emission:
1683: \begin{equation}
1684: \langle AT\rangle_{\rm disk} = \int_{3.1\times10^8}^{6.6\times10^{11}}
1685: 2\pi R(\cos i) T_{10}\,(R/10^{10})^{-1/2} dR
1686: =1.5\times10^{28}\xi_{10}^{1/4} ~{\rm cm^2\,K}.
1687: \label{ATdisk}
1688: \end{equation}
1689: Comparing this estimate with equation (\ref{ATsec}), we see that the
1690: disk will contribute about 10\% of the flux in the optical and
1691: infrared. Note that the uncertain parameter $\xi_{10}$ appears only
1692: with a 1/4 power, so we are not sensitive to its value. However, the
1693: result is very sensitive to the assumed outer radius of the disk. If
1694: the disk is smaller than we have assumed, the disk flux will be
1695: reduced substantially, roughly as $R_{\rm out}^{3/2}$.
1696:
1697: For a wind-fed system like LMC X-1, the disk will probably get rid of
1698: much of its angular momentum by interacting directly with the incoming
1699: material. The efficiency of this process is very difficult to
1700: estimate from first principles, so we will use the eclipsing system
1701: M33 X-7 as a guide to what the outer radius
1702: of the accretion disk
1703: might be. Orosz et al.\ (2007) determined an outer radius
1704: of $0.45\pm
1705: 0.03$ times the Roche radius for M33 X-7. If the accretion disk in
1706: LMC X-1 fills 45\% of its Roche lobe then the disk contribution will
1707: be $\langle AT \rangle_{\rm disk} = 5\times 10^{27}\xi_{10}^{1/4}$,
1708: which is about 3.5\% of the optical/infrared flux. Even this may be
1709: an overestimate. It is quite conceivable that the disk is much
1710: smaller than 45\% of the Roche lobe radius, which would make the disk
1711: contamination completely insignificant.
1712:
1713: We have ignored reprocessing of radiation from the secondary star, but
1714: one can show that it is again not important.
1715:
1716:
1717:
1718: \begin{thebibliography}{}
1719:
1720: \bibitem[Avni(1976)]{avi76}
1721: Anvi, Y. 1976, \apj, 209, 574
1722:
1723: \bibitem[Bernstein et al.(2002)]{ber02}
1724: Bernstein, R., Shectman, S., Gunnels, S.,
1725: Mochnacki, S., \& Athey, A. 2002, Proc.\ SPIE 4841, 1694
1726:
1727: \bibitem[Bianchi \& Pakull(1985)]{bia85}
1728: Bianchi, L., \& Pakull, M. 1985, \aap, 146, 242
1729:
1730: \bibitem[Bildsten et al.(1997)]{bil97}
1731: Bildsten, L., Chakrabarty, D., Chiu, J., et al.\
1732: 1997, \apjs, 113, 367
1733:
1734: \bibitem[Boroson et al.(2008)]{bor08}
1735: Vrtilek, S. D., Boroson, B. S., Hunacek, A., Gies, D., \& Bolton,
1736: C. T. 2008, ApJ, 678, 1248
1737:
1738:
1739:
1740: \bibitem[Cardelli, Clayton, \& Mathis(1989)]{car1989}
1741: Cardelli, J. A., Clayton, G. C., \& Mathis, J. S. 1989, \apj, 345, 245
1742:
1743: \bibitem[Charles and Coe(2006)]{cha06}
1744: Charles, P. A., \& Coe, M. J. 2006, In
1745: Compact stellar X-ray sources. Eds.\
1746: Walter Lewin \& Michiel van der Klis.
1747: Cambridge Astrophysics Series, No. 39.
1748: (Cambridge, UK: Cambridge University Press)
1749:
1750: \bibitem[Clark et al.(2002)]{cla02}
1751: Clark, J. S.,
1752: Goodwin, S. P.,
1753: Crowther, P. A.,
1754: Kaper, L.,
1755: Fairbairn, M.,
1756: Langer, N., \& Brocksopp, C. 2002, \aap, 392, 909
1757:
1758:
1759: \bibitem[Cowley, Crampton, \& Hutchings(1978)]{cow78}
1760: Cowley, A. P., Crampton, D., \& Hutchings, A. P. 1978,
1761: \aj, 83, 1619
1762:
1763: \bibitem[Cowley et al.(1995)]{cow95}
1764: Cowley, A. P., Schmidtke, P. C., Anderson, A. L., \& McGrath, T. K.
1765: 1995, \pasp, 107, 145
1766:
1767: \bibitem[Cui et al.(2002)]{cui02}
1768: Cui, W., Feng, Y. X., Shang, S. N., Bautz, M. W., Garmire, G. P.,
1769: \& Schulz, N. S. 2002, \apj, 576, 357
1770:
1771: \bibitem[Dessart \& Owocki(2005)]{dess05}
1772: Dessart, L., \& Owocki, S.~P.\ 2005, \aap, 432, 281
1773:
1774:
1775: \bibitem[Eaton(2008)]{eat08}
1776: Eaton, J. 2008, \apj, 681, 562
1777:
1778: \bibitem[Frank, King, \& Raine(2002)]{fra02}
1779: Frank, J., King, A., \& Raine, D. J. 2002, Accretion Power in Astrophysics,
1780: Third Edition (Cambridge: Cambridge University Press)
1781:
1782: \bibitem[Gierli\'nski, Macio\l ek-Nied\'zwiecki, \& Ebisawa(2001)]{gie01}
1783: Gierli\'nski, M., Macio\l ek-Nied\'zwiecki, A., \& Ebisawa, K. 2001,
1784: \mnras, 325, 1253
1785:
1786: \bibitem[Haardt et al.(2001)]{haa01}
1787: Haardt, F.,
1788: Galli, M. R.,
1789: Treves, A.,
1790: Chiappetti, L.,
1791: Dal Fiume, D.,
1792: Corongiu, A.,
1793: Belloni, T.,
1794: Frontera, F.,
1795: Kuulkers, E., \&
1796: Stella, L. 2001, \apjs, 133, 187
1797:
1798:
1799:
1800: \bibitem[Heap, Lanz, \& Hubeny(2006)]{hea06}
1801: Heap, S. R., Lanz, T., \& Hubeny, I. 2006, \apj, 638, 409
1802:
1803: \bibitem[Herrero et al.(1995)]{her95}
1804: Herrero, A., Kudritzki, R. P., Gabler, R., Vilchez, J. M.,
1805: \& Gabler, A. 1995, \aap, 297, 556
1806:
1807: \bibitem[Hut(1981)]{hut81}
1808: Hut, P. 1981, \aap, 99, 126
1809:
1810: \bibitem[Hutchings et al.(1983)]{hut83}
1811: Hutchings, J. B., Crampton, D., \& Cowley, A. P. 1983, \apj, 275, L43
1812:
1813: \bibitem[Hutchings et al.(1987)]{hut1987}
1814: Hutchings, J. B.,
1815: Crampton, D.,
1816: Cowley, A. P.,
1817: Bianchi, L., \&
1818: Thompson, I. B. 1987, \aj, 94, 340
1819:
1820: \bibitem[Kelson(2003)]{kel03}
1821: Kelson, D. D. 2003, \pasp, 115, 688
1822:
1823: \bibitem[Kudritzki \& Puls(2000)]{kudpuls00}
1824: Kudritzki, R.-P., \& Puls, J.\ 2000, \araa, 38, 613
1825:
1826: \bibitem[Landolt(1992)]{lan1992}
1827: Landolt, A. U. 1992, \aj, 104, 340
1828:
1829: \bibitem[Lanz \& Hubeny(2003)]{lan02}
1830: Lanz, T., \& Hubeny, I. 2003, \apjs, 146, 417
1831:
1832: \bibitem[Lanz \& Hubeny(2007)]{lan07}
1833: Lanz, T., \& Hubeny, I. 2007, \apjs, 169, 83
1834:
1835: \bibitem[Lejeune \& Schaerer(2001)]{lej01}
1836: Lejeune, T., \& Schaerer, D. 2001, \aap, 366, 538
1837:
1838: \bibitem[Levine et al.(1996)]{asm96} Levine, A.~M., Bradt,
1839: H., Cui, W., Jernigan, J.~G., Morgan, E.~H., Remillard, R., Shirey,
1840: R.~E.,
1841: \& Smith, D.~A.\ 1996, \apjl, 469, L33
1842:
1843: \bibitem[Levine \& Corbet(2006)]{lev06}
1844: Levine, A., \& Corbet, R. 2006, ATel \#940
1845:
1846: \bibitem[Leong et al.(1971)]{leo71}
1847: Leong, C., Kellogg, E., Gursky, H., Tannenbaum, H., \& Giacconi, R.
1848: 1971, \apj, 170, L67
1849:
1850: \bibitem[Mark et al.(1969)]{mar69}
1851: Mark, H., Price, R., Rodrigues, R., Seward, F. D., \& Swift,
1852: C. D. 1969, \apj, 155, L143
1853:
1854: \bibitem[Martini et al.(2004)]{mar04}
1855: Martini, P., Persson, S.E., Murphy, D.C., Birk, C., Shectman, S.A.,
1856: Gunnels, S.M., and Koch, E. 2004, Proc.\ SPIE, 5492, 1653
1857:
1858: \bibitem[Martins \& Plez(2006)]{mar06}
1859: Martins, F., \& Plez, B. 2006, \aap, 457, 637
1860:
1861: \bibitem[McCray et al.(1984)]{mcc84}
1862: McCray, R., Kallman, T. R., Castor, J. I., \& Olson, G. L. 1984, \apj,
1863: 282, 245
1864:
1865: \bibitem[Meynet et al.(1994)]{mey94}
1866: Meynet, G.,
1867: Maeder, A.,
1868: Schaller, G.,
1869: Schaerer, D., \& Charbonnel, C. 1994, \aaps, 103, 97
1870:
1871: \bibitem[Negueruela \& Coe(2002)]{neg02}
1872: Negueruela, I., \& Coe, M. J. 2002, \aap, 385, 517
1873:
1874: \bibitem[Nowak et al.(2001)]{now01}
1875: Nowak, M. A.,
1876: Wilms, J.,
1877: Heindl, W. A.,
1878: Pottschmidt, K.,
1879: Dove, J. B., \&
1880: Begelman, M. C. 2001, \mnras, 320, 316
1881:
1882: \bibitem[Orosz \& Hauschildt(2000)]{oro2000}
1883: Orosz, J. A., \& Hauschildt, P. H. 2000, \aap, 364, 265
1884:
1885: \bibitem[Orosz et al.(2002)]{oro02}
1886: Orosz, J. A., Groot, P. J., van der Klis, M., McClintock, J. E.,
1887: Garcia, M. R., Zhao, P., Jain, R. K., Bailyn, C. D., \& Remillard,
1888: R. A. 2002, \apj, 568, 845
1889:
1890: \bibitem[Orosz et al.(2007)]{oro07}
1891: Orosz, J. A.,
1892: McClintock, J. E.,
1893: Narayan, R.,
1894: Bailyn, C. D.,
1895: Hartman, J. D.,
1896: Macri, L.,
1897: Liu, J.,
1898: Pietsch, W.,
1899: Remillard, R. A.,
1900: Shporer, A., \& Mazeh, T. 2007, \nat, 449, 872
1901:
1902: \bibitem[Pakull \& Angebault(1986)]{pak86}
1903: Pakull, M. W., \& Angebault, L. P. 1986, Nature, 322, 511
1904:
1905: \bibitem[Podsiadlowski et al.(2003)]{pod03}
1906: Podsiadlowski, Ph.,
1907: Rappaport, S., \& Han, Z. 2--3, \mnras, 341, 385
1908:
1909: \bibitem[Predehl, P.; Schmitt(1995)]{pre1995}
1910: Predehl, P., \& Schmitt, J. H. M. M. 1995, \aap, 293, 889
1911:
1912: \bibitem[Prestwich et al.(2007)]{pre07}
1913: Prestwich, A. H.,
1914: Kilgard, R.,
1915: Crowther, P. A.,
1916: Carpano, S.,
1917: Pollock, A. M. T.,
1918: Zezas, A.,
1919: Saar, S. H.,
1920: Roberts, T. P., \&
1921: Ward, M. J. 2007, \apj, 669, L21
1922:
1923: \bibitem[Rappaport, Joss, \& McClintock(1976)]{rap76}
1924: Rappaport, S., Joss, P. C., \& McClintock, J. E. 1976,
1925: \apj, 206, L103
1926:
1927: \bibitem[Schlegel et al.(1994)]{sch94}
1928: Schlegel, E. M.,
1929: Marshall, F. E.,
1930: Mushotzky, R. F.,
1931: Smale, A. P.,
1932: Weaver, K. A.,
1933: Serlemitsos, P. J.,
1934: Petre, R., \& Jahoda, K. M.
1935: 1994, \apj, 422, 243
1936:
1937: \bibitem[Shapiro \& Teukolsky(1983)]{st83}
1938: Shapiro, S.~L., \& Teukolsky, S.~A.\ 1983, {\it Black holes,
1939: white dwarfs, and neutron stars: The physics of compact objects},
1940: (Wiley-Interscience, New York)
1941:
1942: \bibitem[Silverman \& Filippenko(2008)]{sil08}
1943: Silverman, J. M., \& Filippenko, A. V. 2008, \apj, 678, L17
1944:
1945: \bibitem[Skrutskie et al.(2006)]{skr06}
1946: Skrutskie, M. F.,
1947: Cutri, R. M.,
1948: Stiening, R.,
1949: Weinberg, M. D.,
1950: Schneider, S.,
1951: Carpenter, J. M.,
1952: Beichman, C.,
1953: Capps, R.,
1954: Chester, T.,
1955: Elias, J.,
1956: Huchra, J.,
1957: Liebert, J.,
1958: Lonsdale, C.,
1959: Monet, D. G,
1960: Price, S.,
1961: Seitzer, P.,
1962: Jarrett, T.,
1963: Kirkpatrick, J. D.,
1964: Gizis, J.,
1965: Howard, E.,
1966: Evans, T.,
1967: Fowler, J.,
1968: Fullmer, L.,
1969: Hurt, R.,
1970: Light, R.,
1971: Kopan, E. L.,
1972: Marsh, K. A.,
1973: McCallon, H. L.,
1974: Tam, R.,
1975: Van Dyk, S., \&
1976: Wheelock, S. 2006, \aj, 131, 1163
1977:
1978:
1979: \bibitem[Stetson(1987)]{ste87}
1980: Stetson, P. B., 1987, PASP, 99, 191
1981:
1982: \bibitem[Stetson(1990)]{ste90}
1983: Stetson, P. B., 1990, PASP, 102, 932
1984:
1985: \bibitem[Stetson, Davis, \& Crabtree(1991)]{ste91}
1986: Stetson, P. B., Davis, L. E., Crabtree, D. R., 1991, in
1987: ``CCDs in Astronomy,'' ed.\ G. Jacoby, ASP Conference Series, Volume
1988: 8, page 282
1989:
1990: \bibitem[Stetson(1992a)]{ste92a}
1991: Stetson P. B., 1992a, in ``Astronomical Data Analysis Software and
1992: Systems I,'' eds.\ D. M. Worrall, C. Biemesderfer, \& J. Barnes, ASP
1993: Conference Series, Volume 25, page 297
1994:
1995: \bibitem[Stetson(1992b)]{ste92b}
1996: Stetson, P. B., 1992b, in ``Stellar Photometry--Current Techniques and
1997: Future Developments,'' IAU Coll.\ 136, eds.\ C. J. Butler, \& I. Elliot,
1998: Cambridge University Press, Cambridge, England,
1999: page 291
2000:
2001:
2002: \bibitem[Tonry \& Davis(1979)]{ton1979}
2003: Tonry, J., \& Davis, M. 1979, \aj, 84, 1511
2004:
2005: \bibitem[Tarter et al.(1969)]{tar69}
2006: Tarter, C. B., Tucker, W. H., \& Salpeter, E. E. 1969, ApJ, 156, 493
2007:
2008: \bibitem[van der Meer(2007)]{mee07}
2009: van der Meer, A., Kaper, L., van Kerkwijk, M. H.,
2010: Heemskerk, M. H. M., \& van den Heuvel, E. P. J. 2007, \aap, 473, 523
2011:
2012: \bibitem[Vrtilek et al.(2008)]{vrt}
2013: Vrtilek, S. D., Boroson, B. S., Hunacek, A., Gies, D., \& Bolton,
2014: C. T. 2008, \apj, 678, 1248
2015:
2016: \bibitem[Wade \& Horne(1988)]{wad88}
2017: Wade, R. A., \& Horne, K. 1988, \apj, 324, 411
2018:
2019: \bibitem[Walborn(1972)]{wal72}
2020: Walborn, N. 1972, \aj, 77, 312
2021:
2022: \bibitem[Walborn(1973)]{wal73}
2023: Walborn, N. 1973, \aj, 78, 1067
2024:
2025: \bibitem[Walborn \& Fitzpatrick(1990)]{wal90}
2026: Walborn, N. R., \& Fitzpatrick, E. L. 1990, \pasp, 102, 379
2027:
2028: \bibitem[Wilson(1979)]{wil79}
2029: Wilson, R. E. 1979, \apj, 234, 1054
2030:
2031: \bibitem[Wilson(1990)]{wil90}
2032: Wilson, R. E. 1990, \apj, 356, 613
2033:
2034: \bibitem[Wilson \& Sofia(1976)]{wil76}
2035: Wilson, R. E., \& Sofia, S. 1976, \apj, 203, 182
2036:
2037: \bibitem[Xiang, Zhang, \& Yao(2005)]{xia05}
2038: Xiang, J., Zhang, S. N., Yao, Y. 2005, \apj, 628, 769
2039:
2040: \end{thebibliography}
2041:
2042: \clearpage
2043:
2044:
2045:
2046: \begin{deluxetable}{cccccc}
2047: \tablecaption{Fitting Parameters for ASM Light
2048: Curves\tablenotemark{a}\label{ASMfits}}
2049: \tablewidth{0pt}
2050: \tablehead{
2051: \colhead{Channel} &
2052: \colhead{$a_0$} &
2053: \colhead{$a_1$} &
2054: \colhead{$a_2$} &
2055: \colhead{$\chi^2_{\nu}$} &
2056: \colhead{$A$\tablenotemark{b}}
2057: }
2058: \startdata
2059: A & $ 0.7291 \pm 0.0025$ & $-0.0263 \pm 0.0036$ & $-0.0031 \pm 0.0035$
2060: & 1.942 & $0.072 \pm 0.010$ \cr
2061: B & $ 0.5275 \pm 0.0020$ & $-0.0204 \pm 0.0028$ & $ 0.0037 \pm 0.0028$
2062: & 1.151 & $0.077 \pm 0.011$ \cr
2063: C & $ 0.2602 \pm 0.0026$ & $-0.0050 \pm 0.0038$ & $ 0.0005 \pm 0.0037$
2064: & 0.681 & $0.038 \pm 0.029$ \cr
2065: SUM & $ 1.5201 \pm 0.0042$ & $-0.0491 \pm 0.0060$ & $-0.0004 \pm 0.0059$
2066: & 1.236 & $0.065 \pm 0.008$ \cr
2067: \enddata
2068: \tablenotetext{a}{$f(\phi) = a_0 + a_1\cos(2\pi\phi) + a_2\sin(2\pi\phi)$}
2069: \tablenotetext{b}{fractional amplitude $A=(2|a_1|/a_0)$}
2070: \end{deluxetable}
2071:
2072:
2073: \begin{deluxetable}{cccccc}
2074: \tablecaption{LMC X-1 Column Density Measurements\label{xray}}
2075: \tablewidth{0pt}
2076: \tablehead{
2077: \colhead{Model Number} &
2078: \colhead{Model} &
2079: \colhead{Mission} &
2080: \colhead{$N_H$ ($10^{21}$cm$^{-2}$)} &
2081: \colhead{$A_V$ (mag)} &
2082: \colhead{Reference}
2083: }
2084: \startdata
2085: 1 & DBB & {\em Chandra} HETG & $4.6\pm 0.2$ & $2.57\pm 0.12$
2086: & Cui et al.\ 2002 \cr
2087: 2 & COMPTT & {\em Chandra} HETG & $5.9\pm 0.3$ & $3.30\pm 0.18$
2088: & Cui et al.\ 2002 \cr
2089: 3 & DBB & {\em BeppoSAX} & $8.1\pm 0.5$ & $4.53\pm 0.29$
2090: & Haardt et al.\ 2001 \cr
2091: 4 & GRdisk & {\em ASCA} & $5.3\pm 0.2$ & $2.96\pm 0.12$
2092: & Gierlinski et al.\ 2001 \cr
2093: 5 & DBB & {\em ASCA} & $6.3\pm 0.1$ & $3.52\pm 0.08$
2094: & Nowak et al.\ 2001 \cr
2095: 6 & DBB & {\em BBXRT} & $5.8\pm 0.9$ & $3.24\pm 0.51$
2096: & Schlegel et al.\ 1994 \cr
2097: 7 & X-ray halo & {\em Chandra} & $6.5\pm 0.1$ & $3.65\pm 0.10$
2098: & Xiang et al.\ 2005 \cr
2099: \enddata
2100: \end{deluxetable}
2101:
2102:
2103: \clearpage
2104:
2105: \begin{deluxetable}{rccc}
2106: \tablecaption{LMC X-1 Adopted Parameters\label{parm}}
2107: \tablewidth{0pt}
2108: \tablehead{
2109: \colhead{parameter} &
2110: \colhead{value} &
2111: \colhead{value} &
2112: \colhead{Adopted} \\
2113: \colhead{ } &
2114: \colhead{(circular orbit)} &
2115: \colhead{(eccentric orbit)} &
2116: \colhead{value\tablenotemark{a}}
2117: }
2118: \startdata
2119: $P$ (days)
2120: & $3.90914\pm 0.00005$
2121: & $3.90917\pm 0.00005$
2122: & $3.90917\pm 0.00005$ \\
2123: $T_0$ (HJD 2,453,300+)\tablenotemark{b}
2124: & $91.3436\pm 0.0078$
2125: & $91.3419\pm 0.0087$
2126: & $91.3436\pm 0.0080$ \\
2127: $K_2$ (km s$^{-1}$)
2128: & $71.61\pm 0.67$
2129: & $70.74\pm 0.75$
2130: & $71.61\pm 1.10$ \\
2131: $i$ (deg)
2132: & $36.38\pm 1.92$
2133: & $37.00\pm 1.78$
2134: & $36.38\pm 2.02$ \\
2135: $f_2$
2136: & $0.886 \pm 0.036$
2137: & $0.894 \pm 0.040$
2138: & $0.886 \pm 0.037$ \\
2139: $M_2$ ($M_{\odot}$)
2140: & $31.79\pm 3.48$
2141: & $30.62\pm 3.17$
2142: & $31.79\pm 3.67$ \\
2143: $e$
2144: & \nodata
2145: & $0.0256 \pm 0.0066$
2146: & \nodata \\
2147: $\omega$ (deg)
2148: & \nodata
2149: & $260.5 \pm 16.8$
2150: & \nodata \\
2151: $T_0$ (HJD 2,453,300+)\tablenotemark{c}
2152: & \nodata
2153: & $91.3072\pm 0.2073$
2154: & \nodata \\
2155: $R_2$ ($R_{\odot}$)
2156: & $16.89\pm 0.87$
2157: & $16.41\pm 0.71$
2158: & $17.00\pm 0.80$\tablenotemark{d} \\
2159: $\log g$ (cgs)
2160: & $3.485\pm 0.014$
2161: & $3.497\pm 0.011$
2162: & $3.485\pm 0.018$ \\
2163: $a$ ($R_{\odot}$)
2164: & $36.49\pm 1.42$
2165: & $35.97\pm 1.28$
2166: & $36.49\pm 1.51$ \\
2167: $M$ ($M_{\odot}$)
2168: & $10.91\pm 1.41$
2169: & $10.30\pm 1.18$
2170: & $10.91\pm 1.54$ \\
2171: $\chi^2$
2172: ($B$, $V$ bands)
2173: & 59.18, 60.44
2174: & 59.00, 59.70
2175: & \nodata \\
2176: $\chi^2$
2177: (radial velocities)
2178: & 49.63
2179: & 36.20
2180: & \nodata \\
2181: $\chi^2$
2182: (total)
2183: & 169.27
2184: & 155.39
2185: & \nodata
2186: \enddata
2187: \tablecomments{The effective temperature of the secondary has
2188: been constrained to the range $32,200\le T_{\rm eff}\le 34,200$~K.}
2189: \tablenotetext{a}{The uncertainties include systematic errors}
2190: \tablenotetext{b}{Time of inferior conjunction of secondary}
2191: \tablenotetext{c}{Time of periastron passage of secondary}
2192: \tablenotetext{d}{Determined from the temperature, $K$ magnitude,
2193: distance, and extinction}
2194: \end{deluxetable}
2195:
2196:
2197:
2198: \begin{deluxetable}{cccccc}
2199: \tablecaption{LMC X-1 Parameters as a Function
2200: of $K_2$\label{fixK}}
2201: \tablewidth{0pt}
2202: \tablehead{
2203: \colhead{Assumed $K_2$} &
2204: \colhead{$i$} &
2205: \colhead{$\Delta\phi$\tablenotemark{a}} &
2206: \colhead{$M_2$} &
2207: \colhead{$M$} &
2208: \colhead{$\chi^2_{\rm min}$} \\
2209: \colhead{(km s$^{-1}$)} &
2210: \colhead{(deg)} &
2211: \colhead{} &
2212: \colhead{($M_{\odot}$)} &
2213: \colhead{($M_{\odot}$)} &
2214: \colhead{}
2215: }
2216: \startdata
2217: $68.50\pm 0.65$ & $36.20\pm 2.02$ & $0.0111\pm 0.0055$ & $31.46\pm 3.51$
2218: & $10.34\pm 1.31$ & 115.68 \\
2219: $69.50\pm 0.65$ & $36.17\pm 2.02$ & $0.0111\pm 0.0057$ & $31.72\pm 3.53$
2220: & $10.58\pm 1.33$ & 115.66 \\
2221: $69.80\pm 0.65$ & $36.17\pm 2.03$ & $0.0111\pm 0.0055$ & $31.79\pm 3.48$
2222: & $10.65\pm 1.33$ & 115.65 \\
2223: $70.50\pm 0.65$ & $36.16\pm 2.03$ & $0.0111\pm 0.0055$ & $32.00\pm 3.49$
2224: & $10.82\pm 1.36$ & 115.63 \\
2225: $71.50\pm 0.65$ & $36.23\pm 2.01$ & $0.0112\pm 0.0055$ & $32.10\pm 3.51$
2226: & $11.00\pm 1.38$ & 115.60 \\
2227: $72.50\pm 0.65$ & $36.19\pm 2.02$ & $0.0112\pm 0.0055$ & $32.37\pm 3.51$
2228: & $11.26\pm 1.41$ & 115.57 \\
2229: \enddata
2230: \tablenotetext{a}{$\Delta\phi$ is the phase shift of the
2231: photometric $T_0$ relative to the spectroscopically determined
2232: value.}
2233: \tablecomments{The assumed distance is $18.41\pm 0.10$ mag,
2234: the assumed temperature is in the range
2235: $32,200\le T_{\rm eff}\le 34,200$ K, and the assumed
2236: extinction is $A_V=2.28\pm 0.06$ mag.}
2237: \end{deluxetable}
2238:
2239:
2240:
2241: \begin{deluxetable}{ccccccc}
2242: \tablecaption{LMC X-1 Parameters for Nonsynchronous
2243: Rotation\label{nonsync}}
2244: \tablewidth{0pt}
2245: \tablehead{
2246: \colhead{$\Omega$\tablenotemark{a}} &
2247: \colhead{$i$} &
2248: \colhead{$e$} &
2249: \colhead{$\omega$} &
2250: \colhead{$M_2$} &
2251: \colhead{$M$} &
2252: \colhead{$\chi^2_{\rm min}$} \\
2253: \colhead{ } &
2254: \colhead{(deg)} &
2255: \colhead{} &
2256: \colhead{(deg)} &
2257: \colhead{($M_{\odot}$)} &
2258: \colhead{($M_{\odot}$)} &
2259: \colhead{}
2260: }
2261: \startdata
2262: 1.15 & $32.0\pm 1.0$ & 0 (fixed) & \nodata & $30.7\pm 2.3$
2263: & $12.5\pm 1.0$ & 171.53 \\
2264: 1.10 & $33.0\pm 1.4$ & 0 (fixed) & \nodata & $31.4\pm 2.4$
2265: & $12.1\pm 1.2$ & 170.45 \\
2266: 1.05 & $34.9\pm 1.5$ & 0 (fixed) & \nodata & $31.0\pm 1.5$
2267: & $11.3\pm 1.4$ & 169.91 \\
2268: 1.00 & $36.4\pm 1.9$ & 0 (fixed) & \nodata & $31.2\pm 3.5$
2269: & $10.9\pm 1.4$ & 169.38 \\
2270: 0.95 & $38.5\pm 2.2$ & 0 (fixed) & \nodata & $32.2\pm 3.1$
2271: & $10.3\pm 1.0$ & 169.11 \\
2272: 0.90 & $42.8\pm 2.5$ & 0 (fixed) & \nodata & $31.6\pm 2.9$
2273: & ~$9.3\pm 1.1$ & 168.73 \\
2274: 0.85 & $44.5\pm 2.6$ & 0 (fixed) & \nodata & $32.3\pm 2.9$
2275: & ~$8.4\pm 0.9$ & 168.51 \\
2276: 0.80 & $48.1\pm 2.6$ & 0 (fixed) & \nodata & $33.0\pm 2.2$
2277: & ~$8.4\pm 1.1$ & 168.18 \\
2278: 0.75 & $52.1\pm 2.5$ & 0 (fixed) & \nodata & $33.8\pm 3.1$
2279: & ~$7.9\pm 0.9$ & 167.84 \\
2280: 0.70 & $56.3\pm 2.7$ & 0 (fixed) & \nodata & $35.6\pm 2.9$
2281: & ~$7.7\pm 0.7$ & 168.17 \\
2282: 0.65 & $60.9\pm 1.8$ & 0 (fixed) & \nodata
2283: & $37.4\pm 2.0$ & ~$7.5\pm 0.4$ & 168.06 \\
2284: \hline
2285: 1.15 & $33.8\pm 1.5$ & $0.0215\pm 0.0038$ & $267.0\pm 17.2$
2286: & $28.2\pm 2.9$ & $11.0\pm 1.2$ & 162.45 \\
2287: 1.10 & $34.6\pm 1.7$ & $0.0210\pm 0.0063$ & $265.5\pm 16.4$
2288: & $29.4\pm 3.2$ & $11.0\pm 1.3$ & 160.46 \\
2289: 1.05 & $36.0\pm 1.8$ & $0.0245\pm 0.0036$ & $261.3\pm 16.8$
2290: & $29.6\pm 3.3$ & $10.7\pm 1.4$ & 157.03 \\
2291: 1.00 & $37.4\pm 2.1$ & $0.0269\pm 0.0038$ & $266.3\pm 16.9$
2292: & $30.3\pm 3.4$ & $10.3\pm 1.4$ & 154.93 \\
2293: 0.95 & $39.4\pm 2.0$ & $0.0287\pm 0.0048$ & $266.3\pm 17.2$
2294: & $31.4\pm 3.2$ & ~$9.8\pm 1.1$ & 153.19 \\
2295: 0.90 & $42.1\pm 1.7$ & $0.0293\pm 0.0042$ & $264.5\pm 16.6$
2296: & $32.2\pm 2.4$ & ~$9.2\pm 0.9$ & 152.35 \\
2297: 0.85 & $44.2\pm 2.2$ & $0.0311\pm 0.0046$ & $265.2\pm 17.0$
2298: & $33.5\pm 2.8$ & ~$9.0\pm 1.2$ & 150.76 \\
2299: 0.80 & $47.6\pm 2.6$ & $0.0336\pm 0.0045$ & $266.2\pm 17.5$
2300: & $34.5\pm 3.6$ & ~$8.6\pm 1.1$ & 149.94 \\
2301: 0.75 & $50.3\pm 3.5$ & $0.0360\pm 0.0038$ & $268.3\pm 17.1$
2302: & $36.4\pm 3.5$ & ~$8.4\pm 1.0$ & 149.48 \\
2303: 0.70 & $56.8\pm 4.0$ & $0.0372\pm 0.0035$ & $268.2\pm 16.5$
2304: & $36.1\pm 3.9$ & ~$7.5\pm 0.9$ & 148.32 \\
2305: 0.65 & $61.7\pm 2.6$ & $0.0372\pm 0.0037$ & $259.9\pm 17.0$
2306: & $38.0\pm 2.4$ & ~$7.4\pm 0.4$ & 148.46 \\
2307: \enddata
2308: \tablenotetext{a}{$\Omega$ is the ratio of the rotation
2309: frequency of the star to the orbital frequency.}
2310: \tablecomments{The temperature is in the range
2311: $32,200 \le T_{\rm eff} \le 34,200$~K,
2312: the assumed distance is $18.41\pm 0.10$ mag, and the assumed
2313: extinction is $A_V=2.28\pm 0.06$ mag.}
2314: \end{deluxetable}
2315:
2316:
2317:
2318:
2319: \begin{deluxetable}{ccccccc}
2320: \tablecaption{LMC X-1 Parameters as a Function
2321: of Temperature\label{fixT}}
2322: \tablewidth{0pt}
2323: \tablehead{
2324: \colhead{temperature} &
2325: \colhead{derived radius} &
2326: \colhead{$\log L$} &
2327: \colhead{$i$} &
2328: \colhead{$M_2$} &
2329: \colhead{$M$} &
2330: \colhead{$\chi^2_{\rm min}$} \\
2331: \colhead{(K)} &
2332: \colhead{($R_{\odot}$)} &
2333: \colhead{($L_{\odot}$)} &
2334: \colhead{(deg)} &
2335: \colhead{($M_{\odot}$)} &
2336: \colhead{($M_{\odot}$)} &
2337: \colhead{}
2338: }
2339: \startdata
2340: 30,000 & $17.03\pm 0.83$ & $5.33\pm 0.04$ & $36.29\pm 2.10$ & $32.34\pm 3.48$
2341: & $11.09\pm 1.34$ & 174.14 \\
2342: 31,000 & $16.89\pm 0.82$ & $5.38\pm 0.04$ & $36.47\pm 2.11$ & $31.93\pm 1.6$
2343: & $10.95\pm 1.34$ & 174.66 \\
2344: 32,000 & $17.07\pm 0.83$ & $5.44\pm 0.04$ & $36.12\pm 2.03$ & $32.73\pm 3.56$
2345: & $11.21\pm 1.35$ & 173.94 \\
2346: 33,000 & $17.07\pm 0.82$ & $5.49\pm 0.04$ & $36.30\pm 2.02$ & $32.45\pm 3.73$
2347: & $11.07\pm 1.27$ & 171.30 \\
2348: 34,000 & $16.72\pm 0.81$ & $5.53\pm 0.04$ & $37.11\pm 2.17$ & $30.67\pm 3.61$
2349: & $10.44\pm 1.32$ & 169.24 \\
2350: 35,000 & $16.23\pm 0.78$ & $5.55\pm 0.04$ & $38.42\pm 2.33$ & $28.66\pm 3.21$
2351: & ~$9.66\pm 1.25$ & 169.55 \\
2352: 36,000 & $15.79\pm 0.76$ & $5.58\pm 0.04$ & $39.92\pm 2.13$ & $27.12\pm 3.00$
2353: & ~$8.96\pm 1.13$ & 170.07 \\
2354: 37,000 & $15.42\pm 0.74$ & $5.61\pm 0.04$ & $40.16\pm 1.71$ & $26.81\pm 1.91$
2355: & $~8.79\pm 0.75$ & 165.82 \\
2356: \enddata
2357: \tablecomments{The assumed distance is $18.41\pm 0.10$ mag and the assumed
2358: extinction is $A_V=2.28\pm 0.06$ mag.}
2359: \end{deluxetable}
2360:
2361:
2362:
2363: \begin{deluxetable}{cccccc}
2364: \tablecaption{LMC X-1 Parameters as a Function of
2365: Radius\label{fixR}}
2366: \tablewidth{0pt}
2367: \tablehead{
2368: \colhead{radius} &
2369: \colhead{derived luminosity} &
2370: \colhead{$i$} &
2371: \colhead{$M_2$} &
2372: \colhead{$M$} &
2373: \colhead{$\chi^2_{\rm min}$} \\
2374: \colhead{($R_{\odot}$)} &
2375: \colhead{($\log L/L_{\odot}$)} &
2376: \colhead{(deg)} &
2377: \colhead{($M_{\odot}$)} &
2378: \colhead{($M_{\odot}$)} &
2379: \colhead{}
2380: }
2381: \startdata
2382: 15.0 & $5.39\pm 0.03$ & $41.24\pm 0.87$ & $24.56\pm 0.95$
2383: & ~$8.15\pm 0.25$ & 168.52 \\
2384: 15.5 & $5.42\pm 0.03$ & $39.64\pm 0.80$ & $26.34\pm 0.90$
2385: & ~$8.86\pm 0.28$ & 168.71 \\
2386: 16.0 & $5.45\pm 0.03$ & $38.17\pm 0.81$ & $28.17\pm 0.97$
2387: & ~$9.62\pm 0.29$ & 168.99 \\
2388: 16.5 & $5.47\pm 0.03$ & $36.82\pm 0.71$ & $30.16\pm 0.89$
2389: & $10.43\pm 0.28$ & 169.19 \\
2390: 17.0 & $5.50\pm 0.03$ & $35.62\pm 0.71$ & $32.28\pm 0.97$
2391: & $11.29\pm 0.31$ & 169.44 \\
2392: 17.5 & $5.53\pm 0.03$ & $34.42\pm 0.70$ & $34.50\pm 1.24$
2393: & $12.22\pm 0.30$ & 169.77 \\
2394: 18.0 & $5.55\pm 0.03$ & $33.37\pm 0.65$ & $36.86\pm 0.87$
2395: & $13.20\pm 0.37$ & 170.42 \\
2396: 18.5 & $5.57\pm 0.03$ & $32.39\pm 0.56$ & $39.29\pm 0.59$
2397: & $14.22\pm 0.34$ & 170.36 \\
2398: 19.0 & $5.56\pm 0.03$ & $32.48\pm 0.59$ & $42.60\pm 0.56$
2399: & $15.44\pm 0.37$ & 171.27 \\
2400: \enddata
2401: \tablecomments{The assumed temperature is in the range
2402: $32,200 \le T_{\rm eff} \le 34,200$~K,
2403: the assumed distance is $18.41\pm 0.10$ mag, and the assumed
2404: extinction is $A_V=2.28\pm 0.06$ mag.}
2405: \end{deluxetable}
2406:
2407: \clearpage
2408:
2409: \begin{deluxetable}{ccccc}
2410: \tablecaption{LMC X-1 Parameters as a Function of
2411: X-ray Heating\label{fixX}}
2412: \tablewidth{0pt}
2413: \tablehead{
2414: \colhead{$\log L_x$ (erg s$^{-1}$)} &
2415: \colhead{$i$} &
2416: \colhead{$M_2$} &
2417: \colhead{$M$} &
2418: \colhead{$\chi^2_{\rm min}$} \\
2419: \colhead{(isotropic equivalent)} &
2420: \colhead{(deg)} &
2421: \colhead{($M_{\odot}$)} &
2422: \colhead{($M_{\odot}$)} &
2423: \colhead{}
2424: }
2425: \startdata
2426: no heating & $36.02\pm 1.78$ & $31.60\pm 3.11$
2427: & $10.98\pm 1.28$ & 169.57 \\
2428: 38.00 & $36.02\pm 1.82$ & $31.61\pm 3.03$
2429: & $10.99\pm 1.40$ & 169.57 \\
2430: 38.25 & $36.01\pm 1.28$ & $31.54\pm 2.27$
2431: & $10.99\pm 0.99$ & 169.39 \\
2432: 38.50 & $36.06\pm 1.83$ & $31.41\pm 3.28$
2433: & $10.95\pm 1.40$ & 169.35 \\
2434: 38.75 & $36.02\pm 1.39$ & $31.40\pm 2.50$
2435: & $10.97\pm 0.77$ & 169.48 \\
2436: 39.00 & $35.91\pm 1.77$ & $31.49\pm 3.17$
2437: & $11.05\pm 1.16$ & 170.30 \\
2438: \enddata
2439: \tablecomments{The assumed temperature is in the range
2440: $32,200 \le T_{\rm eff} \le 34,200$~K,
2441: the assumed distance is $18.41\pm 0.10$ mag, and the assumed
2442: extinction is $A_V=2.28\pm 0.06$ mag.}
2443: \end{deluxetable}
2444:
2445:
2446: \begin{deluxetable}{cccccc}
2447: \tablecaption{LMC X-1 Parameters as a Function of
2448: the Disk Fraction\label{diskfrac}}
2449: \tablewidth{0pt}
2450: \tablehead{
2451: \colhead{Disk fraction} &
2452: \colhead{derived radius}&
2453: \colhead{$i$} &
2454: \colhead{$M_2$} &
2455: \colhead{$M$} &
2456: \colhead{$\chi^2_{\rm min}$} \\
2457: \colhead{($V$-band)} &
2458: \colhead{($R_{\odot}$)} &
2459: \colhead{(deg)} &
2460: \colhead{($M_{\odot}$)} &
2461: \colhead{($M_{\odot}$)} &
2462: \colhead{}
2463: }
2464: \startdata
2465: 0.05 & $16.58\pm 0.80$ & $37.31\pm 1.84$ & $28.88\pm 2.56$
2466: & $10.03\pm 1.10$ & 169.78 \\
2467: 0.10 & $16.15\pm 0.80$ & $37.99\pm 2.28$ & $27.25\pm 2.65$
2468: & ~$9.52\pm 1.12$ & 170.16 \\
2469: 0.15 & $15.69\pm 0.80$ & $38.44\pm 1.43$ & $25.96\pm 1.89$
2470: & ~$9.15\pm 0.73$ & 171.04 \\
2471: 0.20 & $15.22\pm 0.80$ & $39.10\pm 1.36$ & $24.43\pm 1.70$
2472: & ~$8.69\pm 0.60$ & 175.73 \\
2473: \enddata
2474: \tablecomments{The assumed temperature is in the range
2475: $32,200 \le T_{\rm eff} \le 34,200$~K,
2476: the assumed distance is $18.41\pm 0.10$ mag, and the assumed
2477: extinction is $A_V=2.28\pm 0.06$ mag.}
2478: \end{deluxetable}
2479:
2480:
2481: \clearpage
2482:
2483:
2484:
2485:
2486: \begin{figure}
2487: \includegraphics[scale=.85,angle=-90]{compressf01.ps}
2488: %%\plotone{compressf01.ps}
2489: %\includegraphics[scale=.85,angle=-90]{f01.ps}
2490: %\plotone{f01.ps}
2491: \caption{$1\times 1$ arcminute finding
2492: charts for the field surrounding LMC X-1 obtained with the
2493: Magellan-Baade telescope. The left panel is a $V$-band
2494: image obtained with the
2495: IMACS imaging spectrograph, sampled at
2496: 0\farcs 22 pixel$^{-1}$ under seeing of 0\farcs 8. The
2497: right panel is the same field observed in the
2498: $K$-band using the PANIC camera
2499: with
2500: 0\farcs 125 pixel$^{-1}$
2501: sampling and a seeing of 0\farcs 55. The correct counterpart
2502: together with the nearby B5 supergiant R148 are marked.}
2503: \label{fc}
2504: \end{figure}
2505:
2506:
2507: \begin{figure}
2508: \epsscale{0.95}
2509: \plotone{f02.eps}
2510: \caption{The phased light curves of LMC X-1. Phase zero corresponds
2511: to the time of the inferior conjunction of the secondary star. Shown
2512: from the top are the $B$, $V$, $J$ light curves with the best-fitting
2513: ellipsoidal models assuming a circular orbit (solid lines) and
2514: assuming an eccentric orbit (dashed lines). Owing to the large scatter,
2515: the $J$ band data were not used in the modeling.}
2516: \label{lcfig1}
2517: \end{figure}
2518:
2519:
2520:
2521: \begin{figure}
2522: \epsscale{0.95}
2523: \plotone{compressf03.ps}
2524: %\plotone{f03.eps}
2525: \caption{The average spectrum of LMC X-1 in the restframe of the
2526: secondary star (black dots) obtained with the MIKE echelle.
2527: Most of the strong features are labeled, and interstellar
2528: lines are marked with `IS'.
2529: The best-fitting model (red line) has
2530: $T_{\rm eff}=33,225$~K, $\log g=3.56$, and $V_{\rm rot}\sin i
2531: = 128.0$ km s$^{-1}$. The noise near 5060~\AA\
2532: is caused by the transition from the blue arm to the red arm.
2533: }
2534: \label{newspecfig}
2535: \end{figure}
2536:
2537:
2538:
2539:
2540: \begin{figure}
2541: \epsscale{1.05}
2542: \plotone{f04.eps}
2543: \caption{Summary of fitted values of the velocity semiamplitude $K$
2544: (top) and the time of maximum velocity $T_{\rm max}$ (bottom) for
2545: different combinations of bandpasses, template spectra and modes of
2546: filtering. The lines or line group are indicated at the bottom of the
2547: figure (see text). The open and filled circles are for template HD
2548: 93843 and Legendre filtering. Triangles: HD 101205, Legendre filtering.
2549: Squares: Synthetic template, Legendre filtering. Crosses: Synthetic
2550: template, Fourier filtering. The filled circles and dashed lines
2551: correspond to the values of $K$ and $T_{\rm max}$ given in the text,
2552: which were computed for the mean velocities. The uncertainties shown
2553: are purely statistical, correspond to $\chi_{\nu}^{2}=1$, and are at the
2554: $1\sigma$ level of confidence. No results are shown for the He I
2555: $\lambda4471$ line for HD 101205 because the template spectrum was
2556: corrupted in this band.}
2557: \label{Kfig}
2558: \end{figure}
2559:
2560: \begin{figure}
2561: \epsscale{0.85}
2562: \plotone{f05.eps}
2563: \caption{The equivalent widths of H10 (top), He I $\lambda 4471$
2564: (center), and He II $\lambda 4200$ (bottom) as a function of
2565: orbital phase, where phase zero corresponds
2566: to the time of the inferior conjunction of the secondary star.
2567: }
2568: \label{ewfig}
2569: \end{figure}
2570:
2571:
2572:
2573:
2574: \begin{figure}
2575: \epsscale{0.85}
2576: \plotone{f06.eps}
2577: \caption{Top:
2578: The phased velocity curve of LMC X-1. Phase zero corresponds
2579: to the time of the inferior conjunction of the secondary star.
2580: The MIKE radial velocities are shown with the filled circles and the
2581: MagE radial velocities are shown with the filled triangles.
2582: The
2583: model curve for the best-fitting orbital model assuming a circular
2584: orbit is shown with the solid line, and the model curve for the
2585: best-fitting model assuming an eccentric orbit is shown with the dashed
2586: lines.
2587: Center: The residuals with respect to the circular orbit model.
2588: Bottom: The residuals with respect to the eccentric orbit model.
2589: }
2590: \label{rvfig1}
2591: \end{figure}
2592:
2593:
2594:
2595:
2596: \begin{figure}
2597: \includegraphics[scale=.7]{f07.eps}
2598: \caption{Top: $\chi^2$ computed using a three-parameter
2599: sinusoid vs.\ the trial period for the Magellan radial velocities.
2600: Center: The periodogram (computed in a similar manner as
2601: above) for the radial velocities from Hutchings et
2602: al.\ (1983, 1987) combined with the Magellan
2603: radial velocities.
2604: Bottom: The periodogram derived from the SMARTS photometry and
2605: the Magellan radial velocities (see text).
2606: The best-fitting period is found to
2607: be $3.90917 \pm 0.00005$, and all other possible alias periods are
2608: ruled out at high confidence.
2609: In each of the three panels
2610: the X-ray
2611: period given in Levine \& Corbet (2006) is indicated with an open
2612: circle, and the refined X-ray period is denoted by the filled circle.
2613: }
2614: \label{plotperiodogram}
2615: \end{figure}
2616:
2617: \begin{figure}
2618: \epsscale{0.85}
2619: \plotone{f08.eps}
2620: \caption{Power density spectrum (PDS) of a 1.5-12 keV ASM light curve
2621: of LMC X-1 that has been modified to remove variability on time scales
2622: longer than $\approx 30$ days (see text). The original FFT was
2623: oversampled by a factor of four and had approximately 62,000
2624: frequencies from 0 cycles day$^{-1}$ to the Nyquist frequency of 3.3
2625: cycles day$^{-1}$. Top: The low frequency part of a rebinned PDS in
2626: which the number of frequency bins was reduced by a factor of 10 by
2627: using the maximum power in each contiguous set of 10 frequency bins in
2628: the original PDS as the power of the corresponding bin in the rebinned
2629: PDS. Bottom: A portion of the original PDS. The horizontal bar above
2630: the peak shows our estimate of the possible values of the centroid
2631: frequency, i.e., $0.25580 \pm 0.00005 $ d$^{-1}$. In both panels, the
2632: power is normalized relative to the PDS-wide average.}
2633: \label{ASM_PDS}
2634: \end{figure}
2635:
2636: \begin{figure}
2637: \epsscale{0.85}
2638: \plotone{f09.eps}
2639: \caption{The {\it RXTE} ASM light curves of LMC X-1 folded using a
2640: period of 3.9094 days and an epoch of phase zero of MJD 53390.75174.
2641: The four panels from bottom to top show the data from the A (1.5-3
2642: keV), B (3-5 keV), and C (5-12 keV) bands, and the sum of the three
2643: bands (nominally comprising photon energies 1.5-12 keV). The error
2644: bars indicate $\pm 1 \sigma$ statistical uncertainties. The smooth
2645: curves are the best-fit functions of the form $f(\phi) = a_0 +
2646: a_1\cos(2\pi\phi) + a_2\sin(2\pi\phi)$ (see Table
2647: \protect\ref{ASMfits} for the best-fitting parameters).}
2648: \label{ASMfold}
2649: \end{figure}
2650:
2651:
2652:
2653: \begin{figure}
2654: \includegraphics[scale=0.75,angle=-90]{f10.eps}
2655: \caption{The average profiles of the four helium lines used to
2656: determine the rotational velocity (``histograms''), and
2657: the best-fitting model profiles for each (smooth curves).
2658: The line identification and the
2659: derived value of $V_{\rm rot}\sin i$ in km s$^{-1}$ are given
2660: in each panel.
2661: }
2662: \label{vrotfig}
2663: \end{figure}
2664:
2665:
2666: \begin{figure}
2667: \epsscale{0.85}
2668: \plotone{f11.eps}
2669: \caption{The radius of the companion as a function of the extinction
2670: $A_V$ for the $V$-band (triangles) and the $K$-band (circles). The
2671: radius is the same for both bands when the extinction is
2672: $\approx A_V=2.2$
2673: mag.}
2674: \label{radfig}
2675: \end{figure}
2676:
2677: \begin{figure}
2678: \includegraphics[scale=.7,angle=-90]{f12.eps}
2679: \caption{The average IUE spectrum of LMC X-1, dereddened using
2680: $A_V=2.27$ and $R_V=5.16$, is shown with a model spectrum from the
2681: OSTAR2002 grid having $T_{\rm eff}=32,500$~K and $\log g=3.5$, scaled
2682: using a distance modulus of 18.41 and a stellar radius of
2683: $17.08\,R_{\odot}$.}
2684: \label{plotiue}
2685: \end{figure}
2686:
2687:
2688: \begin{figure}
2689: \includegraphics[scale=.7,angle=-90]{f14.eps}
2690: \caption{Curves of $\chi^2$ vs.\ various fitting and derived
2691: parameters of interest for the circular orbit model. The horizontal
2692: dashed lines denote the 1, 2, and $3\sigma$ confidence limits.}
2693: \label{plotfitted}
2694: \end{figure}
2695:
2696:
2697: \begin{figure}
2698: \epsscale{0.85}
2699: \plotone{f15.eps}
2700: \caption{The solid lines show evolutionary tracks for an LMC
2701: metalicity on a temperature-luminosity diagram for stars with ZAMS
2702: masses of $60\,M_{\odot}$, $40\,M_{\odot}$, $25\,M_{\odot}$,
2703: $20\,M_{\odot}$ taken from Meynet et al.\ 1994). The track for a
2704: ZAMS mass of $35\,M_{\odot}$ was crudely interpolated. The dotted
2705: lines show isochrones (also for an LMC metalicity, Lejeune \& Schaerer
2706: 2001) from left to right of 0, 1, 2, 3, 4, 5, and 6 Myr. The location
2707: of the LMC X-1 secondary is very close to the interpolated track for
2708: the $35\,M_{\odot}$ star and the 5 Myr isochrone.}
2709: \label{plothr}
2710: \end{figure}
2711:
2712:
2713: \clearpage
2714:
2715:
2716:
2717: \end{document}
2718: