1: % Mass and Radius Constraints on XTE J1814-338 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: % One and only one of the following should be commented out:%%%%%%%%%%%%
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5:
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %%%%%%%%%%%%%%%%%%%%%% Use these for ApJ formatted style (preprint)%%%%%
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: \documentclass{emulateapj} %
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: % \documentclass{article}
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: %%%%%%%%%%%%%%%%%%%%%% Use this for ApJ electronic submission %%%%
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: %\documentclass[12pt,preprint]{aastex} %
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17:
18:
19:
20: %\usepackage{epsfig}
21:
22: \newcommand{\be}{\begin{equation}}
23: \newcommand{\ee}{\end{equation}}
24: \newcommand{\ba}{\begin{eqnarray}}
25: \newcommand{\ea}{\end{eqnarray}}
26:
27: \newcommand{\xte}{\mbox{XTE~J1814 }}
28: \newcommand{\saxj}{\mbox{SAX~J1808.4-3658 }}
29: \newcommand{\herx}{\mbox{Her~X-1 }}
30:
31:
32: \newcommand{\textdegree}{\ensuremath{^\circ}}
33: \newcommand{\bigo}{\mathcal{O}}
34:
35: \newcommand{\km}{\hbox{km}}
36:
37:
38: %\slugcomment{Submitted to ApJ: June 1, 2008}
39: \slugcomment{Accepted by ApJ: September 26, 2008}
40: \shorttitle{The Properties of XTE J1814-338}
41:
42: \begin{document}
43:
44: \title{Constraints on the Properties of the Neutron Star XTE~J1814-338 from Pulse Shape Models}
45: \author {Denis A. Leahy\altaffilmark{1}, Sharon M. Morsink\altaffilmark{2},
46: Yi-Ying Chung\altaffilmark{3}, \& Yi Chou\altaffilmark{3}}
47:
48:
49: \altaffiltext{1}{Department of Physics and Astronomy, University of Calgary,
50: Calgary AB, T2N~1N4, Canada; leahy@ucalgary.ca}
51: \altaffiltext{2}{Department of Physics,
52: University of Alberta, Edmonton, AB, T6G~2G7, Canada; morsink@phys.ualberta.ca}
53: \altaffiltext{3}{Graduate Institute of Astronomy, National Central University, Jhongli 32001,
54: Taiwan; yichou@astro.ncu.edu.tw}
55:
56: \begin{abstract}
57:
58: The accretion-powered (non-X-ray burst) pulsations of
59: XTE J1814-338 are modeled to determine neutron star parameters and their uncertainties.
60: The model is a rotating circular hot spot and includes: (1) an isotropic
61: blackbody spectral component; (2) an anisotropic Comptonized spectral component;
62: (3) relativistic time-delays and light-bending; and (4) the oblate shape of the star
63: due to rotation. This model is the simplest possible
64: model that is consistent with the data. The resulting best-fit parameters of the model
65: favor stiff equations of state, as can be seen from the 3-$\sigma$
66: allowed regions in the mass-radius diagram. We analyzed all data combined from a 23 day period
67: of the 2003 outburst, and separately analyzed data from 2 days of the outburst.
68: The allowed mass-radius regions for both cases only allow equations of state (EOS)
69: that are stiffer than EOS APR \citep{APR}, consistent with
70: the large mass that has been inferred for the pulsar NGC 6440B \citep{Fre08}.
71: The stiff EOS inferred by this analysis is not compatible with the soft EOS
72: inferred from a similar analysis of SAX J1808.
73: \end{abstract}
74:
75:
76: \keywords{stars: neutron --- stars: rotation --- X-rays: binaries --- relativity
77: --- pulsars: individual: XTE J1814-338}
78:
79:
80: \section{Introduction}
81: \label{s:intro}
82:
83: The accretion-powered millisecond-period X-ray pulsars are promising targets for
84: constraining the neutron star equation of state (EOS) through the modeling of
85: emission from hot spots on the pulsar's surface.
86: The first pulsar discovered in this class, \saxj \citep{Wij98}, has a spectrum
87: consistent \citep{Gie02} with emission from a hot spot on the star's surface.
88: Pulse shape modeling of rapidly rotating neutron stars relies on two relativistic
89: effects: the gravitational bending of light rays reduces the modulation of
90: the pulsed emission and depends on the mass to radius ratio $M/R$; and the
91: Doppler boosting due to the star's rotation creates an asymmetry in the
92: pulse shape and depends on the star's radius $R$. These features, combined with
93: reasonable models of the emission properties at the neutron star's surface can
94: be used to constrain the neutron star's mass and radius and hence the EOS
95: of supra-nuclear density matter.
96:
97: XTE J1814-338 (hereafter XTE~J1814) was discovered during outburst in June 2003 \citep{MS03}, and
98: is an accretion powered millisecond pulsar with spin frequency 314.36 Hz
99: and orbital period of 4.3 hr \citep{Metal03}. A detailed timing analysis
100: for \xte was performed by \citet{Pap07} to obtain accurate values for
101: orbital period, projected semi-major axis, pulse spin frequency and spin down rate.
102: A similar analysis of the pulse arrival times was carried out by \citet{Wat06} and \citet{Chu07},
103: which both included an analysis of phase lags. Soft lags were found in the 2-10 keV energy
104: band, similar to those for SAX J1808-3658 and consistent with an origin in Doppler boosting
105: of a Comptonized pulse component
106: with a much broader emission pattern than the blackbody component.
107:
108: \citet{Str03} found the same frequency in the X-ray bursts as was found in
109: the persistent emission, but with a lower second harmonic content. \citet{Wat08}
110: showed that the X-ray burst oscillations are tightly phase-locked with the
111: non-burst pulsations.
112: \citet{Bhat05} modeled the oscillations during X-ray bursts with a hot spot model
113: for a spherical star and for 2 equations of state. Using a large grid of models they
114: found an upper limit on compactness $R_{S}/R<0.48$, with $R_{S}$, the Schwarzschild radius.
115:
116: There are pulse shape models for a few other X-ray pulsars. The 1.2 s period X-ray pulsar
117: \herx was modeled by \citet{Leahy04} using a model that includes accretion columns.
118: The model for \herx constrains the neutron star EOS to a fairly moderate stiffness \citep{Leahy04}.
119: %Old Text
120: %\citet{Bog07} have modeled the X-ray emission from the 5.8 ms period radio pulsar PSR~J0437-4715
121: %and shown that its radius must be larger than 6.7 km if the mass is $1.4 M_\odot$.
122: %New Text
123: \citet{Zav98} and \citet{Bog07} have modeled the X-ray emission from the 5.8 ms period radio pulsar PSR~J0437-4715
124: using a Hydrogen atmosphere model. In the case of PSR~J0437-4715,
125: \citet{Bog07} found that a simple isotropic blackbody model is inconsistent with the data.
126: In their models, \citet{Bog07} showed that the radius of PSR~J0437-4715
127: must be larger than 6.7 km if the mass is $1.4 M_\odot$. Unfortunately
128: the mass of this pulsar is not well-constrained. \citet{Bog08} have shown that constraints on
129: radius for a number of other ms radio pulsars are also possible.
130:
131: Constraints on \saxj (with a spin period of 2.5 ms) were made by \citet{PG03}
132: using data from the 1998 outburst. The modeling
133: done by \citet{PG03} included blackbody emission from a hot spot that is Compton scattered by
134: electrons above the hot spot. Their model makes use of a spherical model for the star's surface
135: and does not include the effects of relative time-delays caused by the different time of flights
136: for photons emitted from different parts of the star's surface. More recently \citet{CLM05}
137: and \citet{CMLC07} have shown that time-delays and the star's oblate shape are important factors
138: that can affect the outcome of pulse-shape modeling for rapidly rotating pulsars such as \saxj.
139: The 1998 outburst data for \saxj was revisited using a models that included time-delays
140: and oblateness \citep{LMC08} with the result that the EOS for \saxj is constrained to be very soft.
141:
142: In this paper we model the accretion-powered pulsations of \xte using a hot-spot model.
143: The hot-spot model allows for one or two circular hot spots with a two-component spectrum. The
144: spectral model includes isotropic blackbody emission and an anisotropic Compton-scattered
145: component described by a powerlaw. The photons are propagated to the observer using the
146: oblate Schwarzschild approximation \citep{MLCB07} which allows the photon initial conditions to
147: be placed on an oblate-shaped initial surface determined by an empirical formula. The Schwarzschild
148: metric is used to compute the photon bending angles and time delays since it has been
149: shown \citep{CMLC07} that the
150: corrections induced by the Kerr black hole metric or a numerical metric for a rotating star
151: are insignificant
152: compared to the corrections induced by the oblate shape. In order to do the pulse-shape
153: modeling, we
154: construct light curves in two narrow energy bands, 2-3 keV and 7-9 keV. We first analyse
155: a composite pulse-shape
156: constructed from 23 days of data and then consider pulse-shapes constructed from single
157: days of data in order
158: to determine whether variations of the pulse-shape with time are significant.
159:
160: The outline of this paper is as follows. In Section \ref{s:method} the
161: method used to construct the light curves
162: and analyse them is outlined. The results of the best-fit models are presented in Section \ref{s:results}.
163: A discussion of the results is presented in Section \ref{s:discuss}.
164:
165: \section{Method}
166: \label{s:method}
167:
168: \subsection{Construction of Light Curves}
169:
170: Pulse shapes for the accretion-powered pulsations are constructed
171: using the ephemeris given by \citet{Chu07}. Data is limited to
172: the first 23 days of the 2003 outburst, June 5 to June 27 2003 (MJD 2452795-2452817), in order to
173: avoid the later period of the outburst when the flux and pulse shape
174: became more erratic (see for example, \citet{Wat05}).
175: X-ray bursts were cut out of the data during the interval between 100 s
176: before and after the start of each burst.
177:
178: Although the RXTE observations include data in the range of 2 - 50 keV,
179: we have chosen to concentrate on the lower energy range from 2 - 10 keV for
180: two reasons. First, the data is noisier at energies above 10 keV. Second,
181: the Chandra observations by \citet{Kra05} constrain the spectrum in the
182: 2 - 10 keV range.
183: It is also useful to separate the data into narrow energy bands
184: in order to separate the different spectral components. We have chosen
185: two narrow bands, the 2-3 keV band and the 7-9 keV band based on the Chandra
186: spectrum. The narrow-band pulse shapes constructed using data
187: from the full observation period (June 5 - 27) are shown in
188: Figure~\ref{fig:all-days}.
189:
190:
191: %%
192: \begin{figure}
193: \plotone{f1.eps}
194: \caption{Pulse profiles for \xte constructed with data from
195: all days between June 5 - 27 (excluding X-ray bursts). Light curves
196: for two energy bands, 2-3 keV and 7-9 keV are shown.
197: }
198: \label{fig:all-days}
199: \end{figure}
200:
201:
202: We have also investigated the variability of the pulse shape with time.
203: In order to do this, the data was separated into one-day segments and
204: separate light curves constructed for each day. It is not computationally
205: feasible to model all days simultaneously, so we instead focus on two
206: separate days. The days were chosen by comparing light curves in the 2 - 10 keV range
207: for different days using a $\chi^2$ test and selecting two days which differ the most from each other.
208: This also has the effect of selecting days with intrinsically smaller error bars.
209: The days resulting from this selection process correspond to June 20 and 27. Light curves for these
210: two days, in the two narrow energy bands are shown in Figure \ref{fig:2-days}.
211:
212:
213: %%
214: \begin{figure}
215: \plotone{f2.eps}
216: \caption{One-day pulse profiles for \xte constructed with data from
217: days June 20 and June 27 (MJD 2452810 and 2452817). Two energy bands are shown for each day.
218: }
219: \label{fig:2-days}
220: \end{figure}
221:
222:
223: \subsection{Analysis of Light Curves}
224: \label{s:spectrum}
225:
226: \citet{Kra05} observed \xte with Chandra on June 20, 2003. They modeled the spectrum in the 0.5 - 10 keV
227: range and found that the
228: best fit solution corresponds to a combination of a 0.95 keV blackbody and powerlaw emission with a photon
229: spectral index of $\Gamma = 1.4$. The ratio of flux from the blackbody to the powerlaw in their model
230: is about 10\%. We use the \citet{Kra05} spectral model and assume that it holds for the other days covered
231: by the RXTE data. This assumption is motivated by the fact that the relative normalization of different
232: energy bands is approximately constant from day to day, although the overall flux at all wavelengths changes
233: with time.
234:
235: The spectral model of \citet{Kra05} motivates the use of two narrow bands in our pulse shape models.
236: A low energy band is necessary in order to capture the blackbody component, so we choose the lowest possible
237: XTE energy band at 2 - 3 keV. The spectrum in this band is dominated by the powerlaw component, but the
238: blackbody contribution is still important. We also choose the 7 - 9 keV band as the highest energy band
239: covered by the Chandra observation. In this high energy band the blackbody flux is negligible.
240:
241: Our method for modeling the observed emission is very similar to the method presented in
242: \citet{LMC08}. The spectral model has three components: (1) Comptonized flux in the high energy band
243: (7-9 keV); (2) Comptonized flux in the low energy band (2-3 keV); and (3) blackbody flux in the
244: low energy band. The observed flux for the $i$th component, $F_i$, integrated
245: over the appropriate observed energy band is given by
246: \be
247: F_i(E) = I_i \eta^{3+\Gamma_i} (1 - a_i \mu).
248: \label{eq:flux}
249: \ee
250: In equation (\ref{eq:flux}), $I_i$ is a constant amplitude, $\eta$ is the Doppler boost factor,
251: $\Gamma_i$ is the photon spectral index in the star's rest frame, $\mu$ is the cosine of the
252: angle between the normal to the star's surface and the initial photon direction, and the
253: constant $a_i$ describes the anisotropy of the emitted light. For a definition of $\eta$
254: as well as a more complete description of the modeling method, please see \citet{LMC08}.
255:
256: In our modeling, the amplitudes $I_1$ and $I_2$ are free parameters
257: while the third amplitude
258: $I_3$ is defined through the constant $b = \bar{F_3}/\bar{F_2}$, the ratio of the
259: phase-averaged blackbody to Comptonized flux in the low-energy band. In the spectral
260: model by \citet{Kra05}, $b=0.1$, but we include this parameter as a fitting parameter
261: with 1$\sigma$ limits from their spectral model.
262: The photon spectral indices for the Comptonized
263: components are fixed at $\Gamma_1=\Gamma_2 = 1.4$ as given by their model. In the
264: narrow range of the low-energy band the blackbody component of $0.95$ keV can be
265: modeled by a powerlaw with photon spectral index $\Gamma_3=0.85$. The
266: anisotropy parameters for the Comptonized components $a_1 = a_2 = a$ are assumed to
267: be equal, and the parameter $a$ is kept as a free parameter.
268:
269: %% old
270: %The anisotropy
271: %parameter for the blackbody component is set to zero ($a_3=0$). In past modeling
272: %\citep{LMC08} we have found that adding a limb-darkening function such as
273: %$e^{-\tau/\mu}$ to the blackbody component does not change the final outcome of the
274: %best-fit models, so we do not include this type of factor. The final set of free
275: %parameters describing the spectrum are $I_1$, $I_2$, $b$ and $a$.
276: %% old
277:
278: %%New paragraph
279: In the modelling of the non-accreting ms pulsars, it
280: was found \citep{Zav98,Bog07,Bog08} that a limb-darkened
281: Hydrogen atmosphere spectral model is required by the data. It is
282: reasonable to expect that that the blackbody component of the
283: spectrum should also be limb-darkened. We tested this hypothesis
284: by multiplying the blackbody flux by a limb-darkening function of the form $e^{-\tau/\mu}$.
285: We then computed the bestfit neutron star models for two type of models: (1) models with
286: non-zero optical depth $\tau$ and (2) models with zero optical depth. The bestfit
287: models for these two cases are almost identical: the mass and radius of the bestfit model
288: changes by less than 0.5\% when a nonzero optical depth is added, and the value of $\delta \chi^2 = 0.1$
289: when the limb-darkening is added. Since the change in $\chi^2$ and the physical parameters are negligible
290: we conclude that adding an extra parameter to model limb-darkening is not warranted by the data. The
291: reason for this is due to the Chandra model which restricts the blackbody contribution in the 2-3 keV
292: band to only 10\% of the Comptonized contribution, and effectivly sets the blackbody component to zero
293: in the high energy band. Since the Comptonized flux is dominant and has fan-beaming included, small changes
294: to the anisotropy of the blackbody component don't affect the final models. For this reason we have
295: set the anisotropy parameter for the blackbody component to zero ($a_3=0$). This is consistent with the
296: results found for \saxj \citep{LMC08} which also did not require any limb-darkening. The final set of free
297: parameters describing the spectrum are $I_1$, $I_2$, $b$ and $a$.
298: %%
299: %% End of new paragraph
300:
301: In order to fit a set of light curves we also need to introduce a set of parameters
302: describing the star and the emission geometry. These parameters are the mass $M$
303: and equatorial radius $R$ of the star, the co-latitude of the spot $\theta$,
304: the inclination angle $i$ as well as a free phase $\phi$. The radius of the
305: circular spot (in the star's rest frame) is kept fixed at 1.5 km, as
306: given by the Chandra spectral model \citep{Kra05}.
307:
308: Our models make use of light curves for two different days' data, which requires a
309: separate set of parameters for each day. However, on the two different days the
310: parameters $M$, $R$ and $i$ do not change. In order to simplify the analysis,
311: we also assume that the photon spectral indices and the parameters $a$ and $b$
312: are also fixed. The full set of free parameters are:
313: $\{I_1, I_2, \theta, \phi\}$ for each day plus $M$, $R$, $i$, $a$ and $b$,
314: for a total of 13 free parameters. However, for each of our fits, the ratio
315: $M/R$ is kept fixed, so for any one value of $M/R$, there are only 12 free
316: parameters.
317:
318: We use the oblate Schwarzschild approximation \citep{MLCB07} to connect
319: photons emitted at the star's surface with those detected by the observer.
320: In previous studies \citep{CLM05,CMLC07} we have shown that, to the
321: accuracy required for extracting the parameters of a rapidly rotating
322: neutron star, it is sufficient to use the Schwarzschild metric to
323: compute the bending of light rays and the relative time delays of
324: photons emitted at different locations on the star. The extra time
325: delays and light bending caused by frame-dragging or higher order
326: rotational corrections in the metric are negligible. However, the
327: rotation of the star causes a deformation of the star into an oblate shape,
328: which changes (relative to a sphere) the directions that photons can
329: be emitted into. We have developed a simple approximation \citep{MLCB07}
330: that allows an empirical fit to the oblate shape of a rotating star to
331: be embedded in the Schwarzschild metric and make use of it in this
332: analysis.
333:
334:
335: \section{Results}
336: \label{s:results}
337:
338: \subsection{Evidence for a Second Spot}
339: \label{s:bump}
340:
341: The pulse profiles in Figure \ref{fig:all-days} show a feature
342: in the phase interval between 0.24 and 0.4. This feature
343: is seen in all of the other energy bands as well. In order to
344: investigate the nature of this feature, we restrict the analysis here to
345: just the 7-9 keV light curve. Since the blackbody contribution in this
346: energy band is negligible, we use a simplified model which only includes
347: the Comptonized component of the radiation.
348:
349: The simplest model for the emission is a single spot. We fitted
350: the 7-9 keV light curve shown in Figure \ref{fig:bump} with a single
351: spot model by first fixing $2M/R=0.4$ and varying the parameters $M,i,
352: \theta, a, I$ and $\phi$. (Similar results are obtained for other
353: values of $2M/R$.) Since there are 32 data points this corresponds to
354: 25 degrees of freedom. The best fit solution for a single spot model has
355: $\chi^2 = 50.7$, which is not a very good fit. This best-fit solution is
356: shown as a solid curve in Figure \ref{fig:bump}.
357:
358: We now turn to a two-spot model, where the second spot is allowed to have
359: an arbitrary location relative to the first spot, but the spectrum
360: is assumed to be the same. The introduction of a second spot
361: introduces three new parameters to the model: an intensity and two
362: angles. The resulting fit for $2M/R=0.4$ has $\chi^2 = 28.3$ for 23 degrees of
363: freedom, which is a significant improvement. The light curve for this
364: model is shown with a dotted curve in Figure \ref{fig:bump}.
365: The mass for this model is 2.04 $M_\odot$.
366: In this model, the
367: second spot's location is situated so that the second spot is
368: almost never seen and its light only contributes during the
369: phase interval between 0.24 and 0.4. Outside of this
370: interval the light received only originates from the primary
371: spot. This suggests a simpler one-spot model where bins
372: in the phase interval 0.24 to 0.4 (marked with
373: a circle in Figure \ref{fig:bump}) are removed from the data
374: set. This reduces the degrees of freedom to 19 (32-6 data
375: points and 5 parameters for a one-spot model). The resulting
376: best-fit model for $2M/R=0.4$ has $\chi^2 = 19.2$, which is also
377: a significant improvement from the one-spot model that uses
378: all of the data. The mass for this model is $2.08 M_\odot$
379: and the light curve for this model is shown as a dashed line
380: in Figure \ref{fig:bump}.
381:
382: Comparing the two-spot model with the one-spot model with
383: the circle-bins excluded, we see that the difference in
384: $\chi^2$ is not significant, and there is little change
385: between the best-fit values of mass and radius. This
386: leads us to the conclusion that there is good evidence
387: for a second spot (or a feature that mimics a second spot),
388: but that the amount of data encoded in those bins affected by the
389: second spot is not sufficient to allow us to model the details of
390: the second spot with any confidence. Since the
391: inclusion or exclusion of the second spot does not
392: change the best-fit values of the main physical
393: parameters of the star ($M, R, i$) it is more
394: appropriate to choose the simpler one-spot model.
395: The qualitative results do not change when we
396: look at different energy bands. As a result,
397: for the remaining modeling reported in this paper
398: we only use one-spot models where data in the
399: 0.24 - 0.4 phase range is removed from the analysis.
400:
401: \begin{figure}
402: \plotone{f3.eps}
403: \caption{Models used to fit the data in the 7-9 keV
404: band (data combined from all 23 days). A one-spot model
405: that uses all data points (both squares and circles) results
406: in the best fit solid curve with $\chi^2/\hbox{dof} = 50.7/25$.
407: A two-spot model that uses all data points results in the best fit
408: dotted curve with $\chi^2/\hbox{dof} = 28.3/23$.
409: A one-spot model that omits the circle bins results in
410: the best fit dashed curve with $\chi^2/\hbox{dof} = 19.2/19$.
411: }
412: \label{fig:bump}
413: \end{figure}
414:
415:
416: \subsection{Best-fit Models using Data from All Days}
417:
418: Our procedure for modeling the two-energy band data shown in
419: Figure \ref{fig:all-days} is to assume a one-spot model
420: that includes both blackbody and Comptonized emission.
421: Data in the phase period between 0.24 and 0.4 is
422: omitted, as described in section \ref{s:bump}. We do a
423: number of fits, each with a fixed value of $2M/R$. (Fixing
424: the ratio of $M/R$ simplifies the fitting procedure since
425: the light-bending and time delays depend on this ratio.)
426: Once $2M/R$ is fixed, all other parameters are allowed to vary
427: and the minimum value of $\chi^2$ is found. In addition to the
428: parameters described in Section \ref{s:spectrum}, we also added
429: two parameters corresponding to DC offsets for the two energy
430: bands. This allows for small errors in the background subtraction.
431: Once $2M/R$ is fixed, we have a total of 10 free parameters:
432: $M$, $\theta$, $i$, $a$, $b$, 2 amplitudes, 2 DC offsets
433: and one overall phase. (The parameter $b$ is restricted to have
434: a value that is within 1 $\sigma$ of the value found by
435: \citet{Kra05}.) Since each energy band has 32 points,
436: but we exclude 6 of these points we have $64-12-10 = 42$
437: degrees of freedom.
438:
439: For a fixed value of $2M/R$, we find that there are two
440: local minima. These two minima are shown in Table~
441: \ref{tab:minima} and we label these two best-fit solutions
442: as the high and low mass solutions. The lowest value of $\chi^2$
443: corresponds to the high mass solution and the lower mass solution
444: has a higher value of $\chi^2$. Although the high mass solution
445: is a better fit, we exclude this solution on physical grounds.
446: First of all, it requires a neutron star radius of 28 km, which is
447: not allowed by any known equation of state. Secondly, once the
448: neutron star mass and the inclination angle are known, the
449: companion's mass can be calculated (shown in the column labeled
450: $M_c$ in Table~\ref{tab:minima}). In the high mass case
451: the companion's mass is $1.7 M_\odot$. Due to the dim nature
452: of the companion, \citet{Kra05} have shown that the companion
453: (if a main sequence star) would have to have a mass that is
454: no bigger than $0.5 M_\odot$. Clearly this excludes the
455: high mass solution but allows the lower mass solution.
456: For these reasons, we exclude the high mass solutions.
457: %for all other values of $M/R$.
458:
459:
460: \begin{deluxetable}{lrrrrrrrr}
461: \tablecaption{Comparison of the Two Minima.\label{tab:minima}}
462: \tablewidth{0pt}
463: \tablehead{
464: \colhead{Model} &
465: \colhead{$2M/R$}&\colhead{$M$}&\colhead{$R$}&\colhead{$\theta$}&
466: \colhead{$i$}&\colhead{$a$}&\colhead{$M_c$}&\colhead{$\chi^2/$dof}\\
467: \colhead{}&\colhead{}&\colhead{$M_\odot$}&\colhead{km}&\colhead{deg.}&
468: \colhead{deg.}&\colhead{}&\colhead{$M_\odot$}&\colhead{}
469: }
470: \startdata
471: High Mass & 0.3 & 2.86 & 28.7 & 66.8 & 11.8 & 0.81 & 1.70 & 55.9/42 \\
472: Low Mass & 0.3 & 1.95 & 19.8 & 42.4 & 25.0 & 0.59 & 0.55 & 61.0/42
473: \enddata
474: \end{deluxetable}
475:
476:
477: In Table \ref{tab:all} the best-fit solution for each value of $2M/R$
478: is shown. In each case only the lower mass solution is shown. The
479: best-fit solution shown as a solid curve in Figure \ref{fig:all-days}
480: corresponds to the $2M/R=0.4$ solution shown in the Table \ref{tab:all}.
481: Although we call this these solutions ``lower mass'', clearly they
482: still correspond to high mass neutron stars. In Table \ref{tab:all},
483: only solutions for $2M/R = 0.3$ to $0.6$ are shown. In the case of less
484: compact stars, such as $2M/R = 0.2$, the only solution corresponded to the
485: ``high mass'' branch of unphysical solutions. We did not test solutions that
486: were more compact than $2M/R=0.6$ since these solutions would allow spots to
487: have multiple images, and our program is unable to handle multiple images.
488: Technically, the solutions with $2M/R=0.6$ could have spots with multiple
489: images, but we have checked that the relative values of $\theta$ and $i$
490: do not lead to this problem for our solutions for the most compact stars.
491:
492:
493: \begin{deluxetable}{rrrrrrrr}
494: \tablecaption{Best-fit solutions using data from all days.\label{tab:all}}
495: \tablewidth{0pt}
496: \tablehead{
497: \colhead{$2M/R$}&\colhead{$M$}&\colhead{$R$}&\colhead{$\theta$}&
498: \colhead{$i$}&\colhead{$a$}&\colhead{$M_c$}&\colhead{$\chi^2/$dof}\\
499: \colhead{}&\colhead{$M_\odot$}&\colhead{km}&\colhead{deg.}&
500: \colhead{deg.}&\colhead{}&\colhead{$M_\odot$}&\colhead{}
501: }
502: \startdata
503: 0.3 & 1.95 & 19.8 & 42.4 & 25.0 & 0.59 & 0.55 & 61.0/42 \\
504: 0.4 & 2.45 & 18.4 & 47.0 & 24.2 & 0.61 & 0.65 & 61.2/42 \\
505: 0.5 & 2.38 & 14.2 & 36.7 & 39.4 & 0.59 & 0.39 & 61.9/42 \\
506: 0.6 & 2.42 & 11.9 & 40.9 & 42.8 & 0.59 & 0.37 & 62.6/42 \\
507: \enddata
508: \end{deluxetable}
509:
510:
511: In Figure \ref{fig:b1} we show a number of mass-radius curves
512: for stars spinning at 314 Hz as well as the 2- and 3-$\sigma$ confidence
513: regions for the ``low mass'' solutions. (In this figure, the radius $R$
514: refers to the equatorial radius.) These regions are found by
515: fixing the value of $2M/R$, and varying all other parameters and finding
516: the solutions that have $\chi^2$ larger than the global
517: minimum of $\chi^2_{min}=61.0$ by
518: $\delta \chi^2 = 4$ (2 $\sigma$) or 9 (3 $\sigma$).
519: The region allowed with 99.7\% confidence (3 $\sigma$) only includes
520: large stars with high mass. Of the equations of state displayed in
521: Figure \ref{fig:b1}, the only one lying in the 3 $\sigma$ allowed
522: region is L, corresponding to pure neutron matter computed in
523: a mean field approximation \citep{PPS76}. A pure neutron core
524: is unlikely to be the correct description of supra-nuclear
525: density matter. However it is possible for an EOS that includes
526: some softening due to the presence of other species to be allowed
527: by this data.
528:
529: \begin{figure}
530: \plotone{f4.eps}
531: \caption{Best-fit mass and radius values using the combined data from all days,
532: separated into two narrow energy bands. Contours shown are for 2- and 3-$\sigma$
533: confidence levels. Mass-Radius curves for stars spinning at 314 Hz are shown
534: as solid curves. The EOS shown are: APR \citep{APR}, BBB1 \citep{BBB},
535: ABPR2-3 \citep{ABPR}, H3-7 \cite{LNO06} and L (mean-field theory, pure
536: neutrons \citep{PPS76}).
537: }
538: \label{fig:b1}
539: \end{figure}
540:
541: \subsection{Best-fit Models Using Data From 2 Days}
542:
543: It is possible for variability of the data to affect the fit results,
544: as appears to be the case for SAX J1808 \citep{LMC08}. For this reason
545: we have rebinned the data into one-day segments in order to see if there
546: is any significant change in the pulse shape on a day to day basis.
547: We performed a $\chi^2$ test to see how closely each day's data matched
548: the other days' data. Comparisons of one day with an adjacent day
549: gave values of $\chi^2/\hbox{dof}$ ranging from 0.8 to 1.6, indicating
550: day-to-day changes are small. The largest change is between
551: day 810 (June 20) and day 817 (June 27) with $\chi^2/\hbox{dof}$ = 4.8.
552: Light curves in the two narrow energy bands for these two days are shown
553: in Figure \ref{fig:2-days}.
554:
555: The data corresponding to these two days is binned into 32 bins per period.
556: Since we continue to remove the 6 bins corresponding to the second ``spot'',
557: this corresponds to a total of 104 data points.
558: We fit the data for these two days by assuming that the parameters
559: $M, R, i, a$ are the same for both days. The spot's latitude is allowed to
560: vary, as are the amplitudes of the energy bands and the value of phase.
561: Since the DC contributions found in our previous fits are very small, we
562: do not include terms for DC offsets. In addition, we keep the value of $b$
563: fixed at the \citet{Kra05} value in order to simplify the fits. This corresponds
564: to 12 parameters, and a total of 92 degrees of freedom.
565:
566: The best-fit solutions for this 2-day joint fit are shown in Table \ref{tab:2days}.
567: For each fixed value of $2M/R$ we only found one minimum, unlike the case with
568: all days included. The angular locations of the spot on the two days are
569: are labeled $\theta_1$ and $\theta_2$.
570: The change in angular location of the spot between the
571: two days is less than $2^\circ$ in all cases. The solutions continue to have
572: large masses and radii, as in the case of fits using all of the data. In the
573: case of $2M/R= 0.2$ a low mass (1.2 $M_\odot$) solution is allowed, but it has a very large radius.
574: The 2- and 3-$\sigma$ confidence regions for the two-day joint fits are shown
575: in Figure \ref{fig:a}. The 3-$\sigma$ confidence region is somewhat larger than
576: the same region computed using all of the data, but the two methods have a significant
577: overlap. The 2-day joint fit also only allows the stiffest EOS L.
578:
579:
580: \begin{deluxetable}{rrrrrrrrr}
581: \tablecaption{Best-fit solutions using data only from Days 810 \& 817.\label{tab:2days}}
582: \tablewidth{0pt}
583: \tablehead{
584: \colhead{$2M/R$}&\colhead{$M$}&\colhead{$R$}&\colhead{$\theta_1$}& \colhead{$\theta_2$} &
585: \colhead{$i$}&\colhead{$a$}&\colhead{$M_c$}&\colhead{$\chi^2/$dof}\\
586: \colhead{}&\colhead{$M_\odot$}&\colhead{km}&\colhead{deg.}& \colhead{deg.} &
587: \colhead{deg.}&\colhead{}&\colhead{$M_\odot$}&\colhead{}
588: }
589: \startdata
590: 0.2 & 1.18 & 18.4 & 33.6 & 34.7 & 33.8 & 0.55 & 0.29 & 123.9/92 \\
591: 0.3 & 1.71 & 17.3 & 34.1 & 35.3 & 32.0 & 0.55 & 0.39 & 124.6/92 \\
592: 0.4 & 2.13 & 16.1 & 35.1 & 36.6 & 33.6 & 0.55 & 0.43 & 125.4/92 \\
593: 0.5 & 2.41 & 14.4 & 37.1 & 38.7 & 36.3 & 0.55 & 0.43 & 127.5/92 \\
594: 0.6 & 2.55 & 12.6 & 39.5 & 41.4 & 39.9 & 0.56 & 0.41 & 132.0/92 \\
595: \enddata
596: \end{deluxetable}
597:
598:
599: \begin{figure}
600: \plotone{f5.eps}
601: \caption{Best-fit mass and radius values using data only from day 810 and 817,
602: separated into two narrow energy bands. Contours shown are for 2- and 3-$\sigma$
603: confidence levels. EOS labels are the same as in Figure \ref{fig:b1}.
604: }
605: \label{fig:a}
606: \end{figure}
607:
608:
609: \subsection{Dependence of Models on Assumed Parameters}
610:
611: The models in this paper depend on the results of the spectral models of
612: \citet{Kra05}. We now consider the effect of allowing the parameters in the
613: spectral model to vary within the error bars. As mentioned in Section~\ref{s:spectrum} we
614: already allow the ratio of the blackbody to powerlaw components ($b$) to vary within the
615: 1$\sigma$ limits given by the \citet{Kra05} spectral model.
616: Another spectral parameter that could %potentially
617: affect the fits is the photon spectral index $\Gamma$ for the powerlaw
618: component of the spectrum. \citet{Kra05} found a value of $\Gamma = 1.41 \pm 0.06$
619: and in all of our models presented in the previous section we kept the photon spectral
620: index fixed at a value of $\Gamma = 1.40$. We would expect that a larger value of $\Gamma$
621: would allow for smaller stars. This is because the flux in Equation (\ref{eq:flux}) is
622: proportional to the Doppler boost factor $\eta$ raised to the power $\Gamma + 3$.
623: The Doppler boost factor is mainly responsible for introducing higher harmonics into the
624: signal, so a larger value of $\Gamma$ creates a more asymmetric pulse shape. In order to
625: compensate, the best-fit solution will require a smaller value of stellar radius $R$
626: in order to decrease the value of $\eta$. In order to test the dependence of the best-fit
627: values of the parameters on $\Gamma$, we chose a value of $\Gamma=1.50$ which is somewhat
628: larger than the range allowed by \citet{Kra05} and fit the data using the same method
629: described earlier in this paper. The results of the best-fit parameters for the two
630: values of $\Gamma$ for the case of $2M/R=0.4$ are shown in Table~\ref{tab:test}.
631: As expected, increasing $\Gamma$ allowed for a smaller star, but the decrease is
632: only by 3\%. Similar results occur for other values of $M/R$. Clearly the
633: dependence on the photon spectral index is not sensitive enough to affect the
634: resulting large size of the best-fit stars.
635:
636: In our models we keep the
637: spot size (as measured on the star's surface) fixed at a diameter of 3 km. In previous
638: modeling \citep{LMC08} of \saxj we found that the final values of the best-fit
639: parameters were not sensitive to changes in the spot size, assuming that the spot is
640: small compared to the size of the star. For this reason we have kept the spot size
641: fixed at 3 km for all models in this analysis. We have also assumed the hot spot
642: is circular, although recent MHD models \citep{KR05} more complicated spot shapes.
643: In this paper we have only attempted to use the simplest possible model that is
644: still consistent with the data.
645: Adding extra parameters to our models in order to describe more complicated spot
646: shapes is not yet warranted by the quality of the data.
647:
648:
649:
650:
651: For all models computed so far, we have made use of an empirical formula for the
652: oblate shape of the star, and have included the relative time-delays for photons
653: emitted from different parts of the star. In Table \ref{tab:test} the
654: effects of oblateness and time-delays on the fits are shown. For the model labeled ``sphere'',
655: a spherical initial surface was assumed, but relative time-delays were included
656: in the computation. The resulting best-fit solution is about 10\% larger than
657: the corresponding oblate model (labeled ``oblate'' and $\Gamma=1.4$ in Table
658: \ref{tab:test}). This shrinkage of the star's radius when oblateness is
659: included has been observed in the modeling of \saxj \citep{LMC08}.
660: For the model labeled ``no td'' time-delays were omitted from the calculation
661: and a spherical surface was used.
662: Comparison of the two spherical models
663: in Table \ref{tab:test} shows that the model that includes time-delays
664: is about 3\% smaller than the model that omits time-delays.
665:
666: \subsection{Comparison with X-ray Burst Data}
667:
668: In our analysis of \xte we have only included data from the accretion-powered
669: pulsations and have omitted any data corresponding to an X-ray burst. \citet{Bhat05}
670: analyzed the light curves constructed from the X-ray bursts for this neutron star.
671: In their analysis they assumed a spherical surface for the star and traced the
672: paths of the X-rays using the Kerr metric. They also made use of a limb-darkened
673: blackbody emission (2 keV) spectral model appropriate for X-ray bursts. Due to
674: the method that they adopted, it was necessary for them to assume one of two
675: different equations of state. The stiffer EOS used by \citet{Bhat05} is
676: the same as the EOS that we label ``APR'' and is the A18+$\delta v$+UIX
677: computed by \citet{APR}. The analysis of the X-ray bursts by \citet{Bhat05} allows the APR
678: EOS, while our analysis of the accretion-powered pulsations only allows stiffer EOS.
679: From their analysis it is difficult to determine whether or not their analysis of
680: the X-ray burst data is consistent with a very stiff EOS, as indicated by our analysis.
681:
682: \citet{Wat05} provide a detailed analysis of many aspects of both the
683: X-ray bursts and the non-burst emission. One of the quantities that they
684: measured was the fractional amplitude of the pulsations at the fundamental
685: frequency and the first harmonic for both the burst and non-burst emission.
686: They found that the non-burst
687: emission (modeled in this paper) has a larger harmonic content than the
688: burst emission studied by \citet{Bhat05}. Since Doppler boosting is partially
689: responsible for the harmonic content, it is perhaps not surprising that our
690: analysis of the non-burst pulsations implies a larger Doppler factor for the
691: star, which in turn implies a larger radius than a similar analysis for
692: the X-ray burst oscillations.
693:
694:
695: \begin{deluxetable}{lrrrrrrrrrr}
696: \tablecaption{Dependence of Models on Parameters. Joint fits for two energy bands
697: for two separate days (810 and 817).\label{tab:test}}
698: \tablewidth{0pt}
699: \tablehead{
700: \colhead{Model}& \colhead{$\Gamma$} &
701: \colhead{$\frac{2M}{R}$}&\colhead{$M$}&\colhead{$R$}&\colhead{$\theta_1$}& \colhead{$\theta_2$} &
702: \colhead{$i$}&\colhead{$a$}&\colhead{$M_c$}&\colhead{$\chi^2/$dof}\\
703: \colhead{}& \colhead{} &
704: \colhead{}&\colhead{$M_\odot$}&\colhead{km}&\colhead{deg.}& \colhead{deg.} &
705: \colhead{deg.}&\colhead{}&\colhead{$M_\odot$}&\colhead{}
706: }
707: \startdata
708: oblate & 1.4 &0.4 & 2.13 & 16.1 & 35.1 & 36.6 & 33.6 & 0.55 & 0.43 & 125.4/92 \\
709: oblate & 1.5 &0.4 & 2.07 & 15.6 & 34.5 & 35.9 & 34.3 & 0.55 & 0.41 & 126.6/92 \\
710: sphere & 1.4 &0.4 & 2.38 & 17.6 & 32.8 & 34.3 & 31.6 & 0.54 & 0.49 & 124.7/92 \\
711: no td & 1.4 & 0.4 & 2.45 & 18.5 & 49.0 & 50.9 & 21.1 & 0.60 & 0.76 & 126.8/92
712: \enddata
713: \end{deluxetable}
714:
715:
716: \section{Discussion}
717: \label{s:discuss}
718:
719: We have analyzed the accretion-powered (non-X-ray burst) pulsations of
720: \xte using a hot spot model. Our modeling includes (1) an isotropic
721: blackbody spectral component; (2) an anisotropic Comptonized component;
722: (3) relativistic time-delays; (4) the oblate shape of the star due to
723: rotation. The model presented in this paper is the simplest possible
724: model that is consistent with the data. The resulting best-fit models
725: favor stiff equations of state, as can be seen from the 3-$\sigma$
726: allowed regions in Figures \ref{fig:b1} and \ref{fig:a}. In Figure
727: \ref{fig:b1} all data from a 23 day period of the 2003 outburst were
728: included, while for Figure \ref{fig:a} data from only 2 days were
729: included. The allowed regions for the two data sets differ slightly,
730: but both only allow equations of state that are stiffer than
731: EOS APR \citep{APR}.
732:
733: It is interesting that a large mass has been inferred for the pulsar
734: NGC 6440B \citep{Fre08} through measurements of periastron precession.
735: Assuming that the observed periastron precession is purely from
736: relativistic effects, the pulsar's mass is $M = 2.74 \pm 0.21 M_\odot$
737: (1-$\sigma$ error bars) \citep{Fre08}. If the mass really is this high,
738: it would be consistent with the stiff equations of state allowed by
739: our analysis of \xte. However, it is still possible that the
740: large periastron precession observed for NGC 6440B
741: could be caused by a very rapidly rotating companion \cite{Fre08},
742: in which case the pulsar's mass would be smaller and compatible with
743: more moderate equations of state. A high mass for \saxj is also
744: inferred by observations during its quiescent state \citep{Hei07}.
745: Modelling of the neutron star X7 in 47 Tuc \citep{Hei06} allows
746: for a high mass neutron star, although for X7 a low mass neutron star
747: is also allowed.
748:
749: Similar hot spot models of \saxj imply a soft equation of state
750: and a column model for Her X-1 also implies a soft EOS (see \cite{LMC08}
751: and \cite{Leahy04} for details).
752: The best-fit pulse-shape models found for \xte have mass and radius incompatible
753: with the 3-$\sigma$ allowed regions of mass and radius
754: found for \saxj or Her X-1. Some possible
755: interpretations could be (1) time-variations in the pulse profile
756: of \saxj led to an underestimate of the star's radius; (2) the equation of state
757: for dense matter has a two-phase nature allowing both large and
758: small compact stars; or (3) the simple hot-spot model doesn't
759: describe one or more of these pulsars.
760:
761:
762: The first reason is a factor for \saxj.
763: It was demonstrated in \cite{LMC08} that the analysis of the pulse profile averaged
764: over a long observation gave significantly different results than the pulse
765: profile obtained from a much shorter observation.
766: Supporting this conclusion is a recent analysis
767: by \cite{Har08} of data from the
768: 1998, 2002 and 2005 outbursts of \saxj showing a great deal of
769: variation in the pulse shape over time (see Figure 3 of \citet{Har08}).
770: The pulse-shape analysis by \citet{LMC08} only made use of the data
771: from the 1998 outburst. The 1998 data is very sinusoidal in nature
772: and has very little harmonic content. The results of \citet{Har08}
773: show that the later outbursts have a stronger harmonic content.
774: Since a larger radius star can produce a stronger harmonic content,
775: it is possible that the addition of data from 2002 and 2005 will
776: alter the conclusions of \citet{LMC08} about SAX J1808.
777: But the effect of pulse shape variability is not
778: a factor for Her X-1 where the pulse shape has high stability.
779:
780: The second reason above, i.e. a bimodal equation of state, is a possible, but
781: speculative, solution to the greatly different allowed regions for M and R.
782: In this scenario, there is still only one baryonic equation of state and one quark
783: matter EOS, but above a certain critical density, $\rho_{crit}$, the whole star
784: makes a transition from baryonic matter to quark matter.
785: Then for stars with central density $\rho_c$ having $\rho_c<\rho_{crit}$,
786: the M vs. R relation follows a stiff baryonic EOS, somewhat like EOS L, whereas
787: for $\rho_c>\rho_{crit}$ the star has converted to a quark star and lies on a
788: quark matter EOS curve in the mass-radius diagram. In the case of EOS L,
789: the $3 \sigma$ region with $M<2.7 M_\odot$ would require a value of
790: $\rho_{crit} \sim 10^{15} g/cm^3$.
791: Since quark matter EOS curves have a lower maximum mass than baryonic EOS curves,
792: any baryonic star that makes
793: the transition and has mass above the quark star maximum mass must lose mass to
794: end up as a stable quark star. The mass loss depends on the physics of the transition
795: process and is likely vary from star to star. In this scenario, we interpret
796: \xte to lie on the baryonic branch of the mass vs. radius diagram and
797: \saxj and Her X-1 to lie on the quark matter branch.
798:
799: The third listed reason for the discrepancy, that the emission region models are
800: too simple to represent the actual emission regions on the stars, is a definite
801: possibility. For instance, if the emission is coming from the magnetosphere, then
802: our models are incorrect. Alternatively, the emission may arise from surface spots, but
803: the region's shape might be more complicated than a circle.
804: This can only be tested by constructing more complex models and applying
805: them to the observed pulse shapes. However with a more complex model with more parameters
806: describing the model, better pulse shape data is required to constrain the model parameters.
807: Future work is planned to explore more complex emission models, and to test whether
808: these resolve the apparent discrepancy in mass and radius for different pulsars.
809:
810:
811: \acknowledgments
812: This research was supported by grants from NSERC to D. Leahy and S. Morsink.
813: Y. Chung and Y. Chou acknowledge partial support from Taiwan National
814: Science Council grant NSC 95-2112-M-008-026-MY2.
815: We also thank the Theoretical Physics
816: Institute at the University of Alberta for supporting D. Leahy's visits
817: to the University of Alberta.
818:
819: \begin{thebibliography}{}
820: %
821: \bibitem[Akmal et al.(1998)]{APR}
822: Akmal, A., Pandharipande, V. R., \& Ravenhall, D. G. 1998,
823: Phys. Rev. C, 58, 1804
824: %
825: \bibitem[Alford et al.(2005)]{ABPR}
826: Alford, M., Braby, M., Paris, M., \& Reddy, S. 2005,
827: ApJ 629, 969
828: %
829: \bibitem[Baldo et al.(1997)]{BBB}
830: Baldo, M., Bombaci, I., \& Burgio, G. F., 1997, \aap, 328, 274
831: %
832: \bibitem[Bhattacharyya et al.(2005)]{Bhat05} Bhattacharyya,
833: S., Strohmayer, T.~E., Miller, M.~C., \& Markwardt, C.~B.\ 2005, \apj, 619,
834: 483
835: %
836: \bibitem[Bogdanov et al.(2007)]{Bog07} Bogdanov, S., Rybicki,
837: G.~B., \& Grindlay, J.~E.\ 2007, \apj, 670, 668
838: %
839: \bibitem[Bogdanov et al.(2008)]{Bog08} Bogdanov, S.,
840: Grindlay, J.~E.,
841: \& Rybicki, G.~B.\ 2008, ArXiv e-prints, 801, arXiv:0801.4030
842: %
843: \bibitem[Cadeau et al.(2005)]{CLM05}
844: Cadeau, C., Leahy, D.~A., \& Morsink, S.~M. 2005, \apj, 618, 451
845: %
846: \bibitem[Cadeau et al.(2007)]{CMLC07}
847: Cadeau, C., Morsink, S. M., Leahy, D. A., \& Campbell, S. S. 2007,
848: \apj, 654, 458
849: %
850: \bibitem[Chung(2007)]{Chu07}
851: Chung, Y. 2007, Study of Orbital and Pulsation Properties of Accretion-Powered Millisecond
852: Pulsar XTE J1814-338,
853: MSc. Thesis, National Central University, Taiwan.
854: %
855: \bibitem[Freire et al.(2008)]{Fre08}
856: Freire, P. C. C., et al. 2008, \apj, 675, 670
857: %
858: \bibitem[Gierli\'{n}ski et al.(2002)]{Gie02}
859: Gierli\'{n}ski, M., Done, C., \& Barret, D. 2002,
860: \mnras, 331, 141
861: %
862: \bibitem[Hartman et al.(2008)]{Har08}
863: Hartman, J. M., et al. 2008,
864: \apj, 675, 1468
865: %
866: \bibitem[Heinke et al.(2006)]{Hei06}
867: Heinke, C.~O., Rybicki, G.~B., Narayan, R., \& Grindlay, J.~E.\ 2006, \apj, 644, 1090
868: %
869: \bibitem[Heinke et al.(2007)]{Hei07} Heinke, C.~O., Jonker,
870: P.~G., Wijnands, R., \& Taam, R.~E.\ 2007, \apj, 660, 1424
871: %
872: \bibitem[Krauss et al.(2005)]{Kra05}
873: Krauss, M.~I., et al.\ 2005, \apj, 627, 910
874: %
875: \bibitem[Kulkarni \& Romanova(2005)]{KR05}
876: Kulkarni, A.~K. \& Romanova, M.~M. 2005, \apj, 633, 349
877: %
878: \bibitem[Lackey, Nayyar, \& Owen(2006)]{LNO06}
879: Lackey, B.~D., Nayyar, M., \& Owen, B.~J.\ 2006, \prd, 73, 024021
880: %
881: \bibitem[Leahy(2004)]{Leahy04}
882: Leahy, D.~A.\ 2004, \apj, 613, 517
883: %
884: \bibitem[Leahy et al.(2008)]{LMC08}
885: Leahy, D. A., Morsink, S. M., \& Cadeau, C. 2008,
886: \apj, 672, 1119
887: %
888: \bibitem[Markwardt \& Swank (2003)]{MS03}
889: Markwardt, C. B., \& Swank, J. H. 2003, IAU Circ., 8144, 1
890:
891: \bibitem[Markwardt et al.(2003)]{Metal03}
892: Markwardt, C. B., Strohmayer, T. E., \& Swank, J. H. 2003, Astron. Telegram,
893: 164, 1
894: %
895: \bibitem[Morsink et al.(2007)]{MLCB07}
896: Morsink, S. M., Leahy, D. A., Cadeau, C. \& Braga, J. 2007,
897: \apj, 663, 1244
898: %
899: \bibitem[Pandharipande et al.(1976)]{PPS76}
900: Pandharipande, V. R., Pines, D., \& Smith, R. A. 1976,
901: \apj, 208, 550
902: %
903: \bibitem[Papitto et al.(2007)]{Pap07} Papitto, A., di Salvo,
904: T., Burderi, L., Menna, M.~T., Lavagetto, G., \& Riggio, A.\ 2007, \mnras,
905: 375, 971
906: %
907: \bibitem[Pechenick et al.(1983)]{PFC83}
908: Pechenick, K. R., Ftaclas, C., \& Cohen, J. M. 1983, \apj, 274, 846
909: %
910: \bibitem[Poutanen \& Gierli\'{n}ski(2003)]{PG03}
911: Poutanen, J. \& Gierli\'{n}ski, M. 2003, \mnras, 343, 1301
912: %
913: \bibitem[Strohmayer et al.(2003)]{Str03}
914: Strohmayer, T. E., Markwardt, C. B., Swank, J. H., \& in't Zand, J. 2003, ApJ,
915: 596, L67
916: %
917: \bibitem[Sunyaev \& Titarchuk(1985)]{Sun85} Sunyaev, R.~A.,
918: \& Titarchuk, L.~G.\ 1985, \aap, 143, 374
919: %
920: \bibitem[Watts et al.(2005)]{Wat05} Watts, A.~L., Strohmayer,
921: T.~E., \& Markwardt, C.~B.\ 2005, \apj, 634, 547
922: %
923: \bibitem[Watts \& Strohmayer(2006)]{Wat06}
924: Watts, A.~L., \& Strohmayer, T.~E. \ 2006, \mnras, 373, 769
925: %
926: \bibitem[Watts et al.(2008)]{Wat08} Watts, A.~L., Patruno,
927: A., \& van der Klis, M.\ 2008, ArXiv e-prints, 805, arXiv:0805.4610
928: %
929: \bibitem[Wijnands \& van der Klis(1998)]{Wij98}
930: Wijnands, R., \& van der Klis, M. 1998, \nat, 394, 344
931: %
932: \bibitem[Zavlin \& Pavlov(1998)]{Zav98}
933: Zavlin, V. E., \& Pavlov, G. G. 1998, \aap, 329, 583
934:
935: \end{thebibliography}
936:
937: \end{document}
938:
939:
940:
941:
942:
943:
944:
945:
946:
947:
948:
949: