0810.1758/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: \usepackage{amsmath}
3: \bibliographystyle{apj}
4: \usepackage{graphicx}
5: \clubpenalty=9999 % no orphans
6: 
7: \newcommand{\Mdot}{\dot{M}}
8: \newcommand{\rin}{R_{\rm in}}
9: \newcommand{\risco}{R_{\rm ISCO}}
10: \newcommand{\mdot}{\dot{m}}
11: \newcommand{\msun}{\rm M_{\sun}}
12: \newcommand{\rchi}{\chi^{2}_\nu}
13: \newcommand{\rchinu}{\chi^{2}/\nu}
14: \newcommand{\fsc}{f_{\rm SC}}
15: \newcommand{\nh}{N_{\rm H}}
16: \newcommand{\NH}{N_{\rm H}}
17: \newcommand{\kpc}{\rm kpc}
18: \newcommand{\kev}{\rm keV}
19: \newcommand{\keV}{\rm keV}
20: \newcommand{\erg}{\rm erg}
21: \newcommand{\cm}{\rm cm}
22: \newcommand{\km}{\rm km}
23: \newcommand{\nin}{n_{\rm in}}
24: \newcommand{\nout}{n_{\rm out}}
25: 
26: \newcommand{\new}{***}
27: 
28: 
29: \shorttitle{A Simple Comptonization Model}
30: \shortauthors{Steiner et al.}
31: 
32: 
33: \begin{document}
34: 
35: \title{A Simple Comptonization Model}
36: \author{James F. Steiner, Ramesh Narayan, Jeffrey E. McClintock}
37: \affil{Harvard-Smithsonian Center for Astrophysics, Cambridge MA 02138}
38: \email{jsteiner@cfa.harvard.edu}
39: \and
40: \author{Ken Ebisawa}
41: \affil{Institute of Space and Astronautical Science/JAXA, 3-1-1
42: Yoshinodai, Sagamihara, Kanagawa 229-8510, Japan}
43: 
44: 
45: \begin{abstract}
46: 
47: We present an empirical model of Comptonization for fitting the
48: spectra of X-ray binaries.  This model, {\sc simpl}, has been
49: developed as a package implemented in XSPEC.  With only two free
50: parameters, {\sc simpl} is competitive as the simplest empirical model of
51: Compton scattering.  Unlike other empirical models, such as the
52: standard power-law model, {\sc simpl} incorporates the basic physics
53: of Compton scattering of soft photons by energetic coronal electrons.
54: Using a simulated spectrum, we demonstrate that {\sc simpl} closely
55: matches the behavior of physical Comptonization models which consider
56: the effects of optical depth, coronal electron temperature, and
57: geometry.  We present fits to {\it RXTE} spectra of the black-hole
58: transient H1743--322 and a {\it BeppoSAX} spectrum of LMC X--3 using
59: both {\sc simpl} and the standard power-law model.  A comparison of
60: the results shows that {\sc simpl} gives equally good fits and a
61: comparable spectral index, while eliminating the troublesome
62: divergence of the standard power-law model at low energies.
63: Importantly, {\sc simpl} is completely flexible and can be used self-consistently with
64: any seed spectrum of photons. 
65: We show that {\sc simpl} -- unlike the standard power law -- teamed up
66: with {\sc diskbb} (the standard model of disk accretion) gives results for the
67: inner-disk radius that are unaffected by strong Comptonization, a result
68: of great importance for the determination of black hole spin via the
69: continuum-fitting method.
70: 
71: 
72: \end{abstract}
73: 
74: \keywords{Astrophysical Data:  Data Analysis and Techniques}
75: 
76: \section{Introduction}
77: 
78: Spectra of X-ray binaries typically consist of a soft (often blackbody
79: or bremsstrahlung) component and a higher-energy tail component of
80: emission, which we refer to generically as a ``power law'' throughout
81: this work.  The origin of the power-law component in both neutron-star
82: and black-hole systems is widely attributed to Compton up-scattering
83: of soft photons by coronal electrons \citep[][hereafter
84: RM06]{White_1995,RR_JEM_review_2006}.  This component is present in
85: the spectra of essentially all X-ray binaries, and it occurs for a
86: wide range of physical conditions.
87: 
88: The tail emission is generally modeled by adding a simple power-law
89: component to the spectrum, e.g., via the model {\sc powerlaw} in the
90: widely used fitting package XSPEC \citep{XSPEC}.  A few of the many
91: applications where   power-law models are employed include: modeling
92: the thermal continuum \citep{Shafee_spin} or the
93: relativistically-broadened Fe K line \citep{Miller_GX339} in order to
94: obtain estimates of black-hole spin; modeling the surrounding
95: environment of compact X-ray sources, such as a tenuous accretion-disk
96: corona \citep{White_1982} or a substantial corona that scatters
97: photons up to MeV energies \citep{Gierlinski_1999}; and classifying
98: patterns of distinct X-ray states, e.g., in black-hole binaries
99: (RM06).
100: 
101: Because of the importance of the power-law component, several physical
102: models have been developed to infer the conditions of the hot plasma
103: that causes the Comptonization.  Models of this variety that are
104: available in XSPEC are {\sc compTT} \citep{COMPTT}, {\sc eqpair} 
105: \citep{EQPAIR}, {\sc compTB}
106: \citep{COMPTB}, {\sc bmc} \citep{BMC}, {\sc compbb} \citep{COMPBB},
107: {\sc thcomp} \citep{THCOMP}, {\sc compls} \citep{COMPLS}, and {\sc compps}
108: \citep{COMPPS}.  It is
109: essential to use such physical models when one is focused on
110: understanding the physical conditions and structure of a scattering
111: corona or other Comptonizing plasma.
112: 
113: Often, however, the physical conditions of the Comptonizing medium are
114: poorly understood or are not of interest, and one is satisfied with an
115: empirical model that seeks to match the data with no pretense that the
116: model describes the physical system.  The model {\sc powerlaw} is one
117: such empirical model which has been extraordinarily widely used in
118: modeling black-hole and neutron-star binaries (see text \& references in
119: \citealt{White_1995,Tanaka_1995,Brenneman_2006}; RM06) and AGN
120: \citep[e.g.,][]{Zdziarski_2002,Brenneman_2006}.  However, {\sc
121: powerlaw} introduces a serious flaw: at low energies it rises without limit.
122: The divergence at low energies is unphysical, and it often
123: significantly corrupts the parameters returned by the model component
124: with which it is teamed (e.g., the widely used disk blackbody
125: component {\sc diskbb}; \S3).
126: 
127: An excellent alternative to the standard power-law model for describing
128: Compton scattering is provided by a convolution model that is based on a
129: Green's function that was formulated decades ago \citep{Shapiro_1976,Rybicki_Lightman,Sunyaev_1980,
130: COMPTT}.  In this approach the power-law is generated self-consistently
131: via Compton up-scattering of a seed photon distribution; consequently,
132: the power-law naturally truncates itself as the seed distribution falls
133: off at low energies.  
134: 
135: In this paper, we present our implementation of a flexible convolution
136: model named {\sc simpl} that can be used with any spectrum of seed
137: photons.  For a Planck distribution we show that {\sc simpl} gives 
138: identical results to {\sc bmc}, as expected since the two models are
139: functionally equivalent (\S2.3).  Although {\sc simpl} has only two
140: free parameters, the same number as the standard {\sc powerlaw}, this
141: empirical model is nevertheless able to very successfully fit data
142: simulated using {\sc compTT}, a prevalent physical model of
143: Comptonization (\S2.2).
144: 
145: We analyze data for two black hole binaries and illustrate the
146: flexibility of {\sc simpl} by convolving {\sc simpl} with {\sc diskbb},
147: the workhorse accretion disk model that has been used for decades
148: \citep{DISKBB}.  Our principal result is that {\sc simpl} in tandem with
149: {\sc diskbb} enables one to obtain fitted values for the inner-disk
150: radius $R_{\rm in}$ for strongly-Comptonized data that are
151: consistent with those obtained for weakly-Comptonized data (see \S3.2).
152: The standard power law, on the other hand, delivers very inconsistent
153: values of $R_{\rm in}$.  As we show in \S4.2, this result is
154: consequential for the measurement of black hole spin via the
155: continuum-fitting method: It implies that using {\sc simpl} in place of
156: the standard power law one can obtain reliable measurements of spin for
157: a far wider body of data than previously thought possible, and for more
158: sources (e.g., Cyg X-1).
159: 
160: In \S2 we outline the model and in \S3 we present a case study with
161: several examples.  We discuss the implications and intended applications
162: of the model in \S4 and conclude with a summary in \S5.
163: 
164: %*********************************************************************
165: 
166: \section{The Model: {\sc simpl}}
167: 
168: The model {\sc simpl} (SIMple Power Law) functions as a convolution
169: that converts a fraction of input seed photons into a power law (see
170: eq. [\ref{eqn:convolution}]).  The model is currently available in
171: XSPEC\footnote{see
172: http://heasarc.nasa.gov/xanadu/xspec/manual/XSmodelSimpl.html}.  In
173: addition to {\sc simpl-{\small 2}}, which is our implementation of the classical
174: model described by \citet{Shapiro_1976} and \citet{Sunyaev_1980}, which 
175: corresponds to both up- and down-scattering of photons, we offer an 
176: alternative ``bare-bones'' implementation in which photons are only 
177: up-scattered in energy.  The physical motivations behind the two versions
178:  of the model are described in \S2.1, and the corresponding scattering 
179: kernels --- the Green's functions --- are given in equation (2) and 
180: equation (3), respectively.
181: 
182: The parameters of {\sc simpl} and the standard {\sc powerlaw} model
183: are similar.  Their principal parameter, the photon index $\Gamma$, is
184: identical.  However, in the case of {\sc simpl} the normalization
185: factor is the scattered fraction $\fsc$, rather than the photon flux.
186: The goal of {\sc simpl} is to characterize the effects of
187: Comptonization as simply and generally as possible.  In this spirit,
188: all details of the Comptonizing medium, such as its geometry (slab
189: vs. sphere) or physical characteristics (optical depth, temperature
190: ,thermal vs. non-thermal electrons
191: ), which would require additional parameters for their description, 
192: are omitted.
193: 
194: It is appropriate to employ {\sc simpl} when the physical conditions of
195: the Comptonizing medium are poorly understood or are not of interest.
196: When the details of the Comptonizing medium are known, or are the main
197: object of study, one should obviously use other models (e.g., {\sc
198: compTT}, {\sc compps}, {\sc thcomp}, etc.), which are designed
199: specifically for such work.  {\sc simpl}, on the other hand, is meant
200: for those situations in which a Compton power-law component is present
201: in the spectral data and needs to be included in the model but is not
202: the primary focus of interest.  {\sc simpl} should thus be viewed as a
203: broad-brush model with the same utility as {\sc powerlaw} but designed
204: specifically for situations involving Comptonization.
205: 
206: By virtue of being a convolution model, {\sc simpl} mimics physical
207: reprocessing by tying the power-law component directly to the energy
208: distribution of the input photons.  The most important feature of the
209: model is that it produces a power-law tail at energies larger than the
210: characteristic energy of the input photons, and that the power law does
211: not extend to lower energies.  This is precisely what one expects any
212: Compton-scattering model to do and is a general feature of all the
213: physical Comptonization models mentioned above.  In contrast, the model
214: {\sc powerlaw} simply adds to the spectrum a pure power-law component
215: that reaches all the way downward %from high energy 
216: to arbitrarily low
217: energies.  The difference between {\sc simpl} and {\sc powerlaw} is thus
218: most obvious at soft X-ray bands where {\sc simpl} cuts off in a
219: physically natural way whereas {\sc powerlaw} continues to rise without
220: limit \citep[e.g., see][]{Yao_2005}.
221: 
222: Two assumptions underlie {\sc simpl}.  The first is that all soft
223: photons have the same probability of being scattered (e.g., the
224: Comptonizing electrons are distributed spatially uniformly).  This is
225: a reasonable assumption when one considers that, even in the best of
226: circumstances, almost nothing is known about the basic geometry of the
227: corona.  For example, usually the corona is variously and crudely
228: depicted as a sphere, a slab, or a lamp post.  The second assumption
229: is that the scattering itself is energy independent.  This is again
230: reasonable given the soft thermal spectra of the seed photons that are
231: observed for black-hole and neutron-star accretion disks with typical
232: temperatures of $\sim 1$ keV and a few keV, respectively.  For
233: example, in the extreme case of a $180^\circ$ back-scatter off a
234: stationary electron, a 3 keV seed photon suffers only a 1\% loss of
235: energy, and even a 10 keV photon loses only 4\% of its initial energy.
236: 
237: Figure~1 shows sample outputs from {\sc simpl} when the input soft
238: photons are modeled by the multi-temperature disk blackbody model {\sc
239: diskbb} \citep{DISKBB}.  Results are shown for both {\sc simpl-{\small 2}} and
240: {\sc simpl-{\small 1}}, our alternative version of {\sc simpl} that
241: includes only up-scattering of photons; the spectra are shown for
242: $\Gamma=2.5$ and a range of values of $\fsc$.  Note the power-law
243: tails in the model spectra at energies above the peak of the soft
244: thermal input and the absence of an equivalent power-law component at
245: lower energies.  This is the primary distinction between {\sc simpl}
246: and {\sc powerlaw}.  {\sc simpl-{\small 2}} and {\sc simpl-{\small 1}} give
247: similar spectra, but the spectrum from {\sc simpl-{\small 1}} has a
248: somewhat stronger power-law tail for the same value of $\fsc$.  This
249: is because {\sc simpl-{\small 1}} transfers all the scattered photons
250: to the high energy tail, whereas {\sc simpl-{\small 2}} has double-sided
251: scattering.  Therefore, for the same value of $\fsc$, fewer photons
252: are scattered into the high-energy tail with {\sc simpl-{\small 2}}.
253: Correspondingly, when fitting the same data, {\sc simpl-{\small 2}} returns a
254: larger value of $\fsc$ compared to {\sc simpl-{\small 1}} (for
255: examples, see \S3 and Table~2).
256: 
257: \subsection{Green's Functions}
258: 
259: Given an input distribution of photons $\nin (E_0)dE_0$ as a function of
260: photon energy $E_0$, {\sc simpl} computes the output distribution $\nout
261: (E)dE$ via the integral transform:
262: \begin{equation}
263: \nout(E)dE = (1-\fsc)\nin(E)dE+\fsc
264: \left[\int_{E_{\rm min}}^{E_{\rm max}} \nin(E_0) G(E;E_0)dE_0\right] dE.
265: \label{eqn:convolution}
266: \end{equation}
267: A fraction $(1-\fsc)$ of the input photons remains unscattered (the
268: first term on the right), and a fraction $\fsc$ is scattered (the
269: second term).  Here, $E_{\rm min}$ and $E_{\rm max}$ are the minimum
270: and maximum photon energies present in the input distribution, and
271: $G(E;E_0)$ is the energy distribution of scattered photons for a
272: $\delta$-function input at energy $E_0$, i.e., $G(E;E_0)$ is the
273: Green's function describing the scattering.
274: 
275: We now describe the specific prescriptions we use for {\sc simpl-{\small 2}} and
276: {\sc simpl-{\small 1}}.  We also discuss the physical motivations behind
277: these prescriptions, drawing heavily on the theory of Comptonization as
278: described by \citet[][hereafter RL79]{Rybicki_Lightman}.
279: 
280: \subsubsection{{\sc simpl-{\small 2}}}
281: 
282: In sec.~7.7, RL79 discuss the case of unsaturated repeated scattering
283: by nonrelativistic thermal electrons.  Following \citet*{Shapiro_1976},
284: they solve the Kompaneets equation and show that
285: Comptonization produces a power-law distribution of photon energies
286: (eq. 7.76d in RL79).  There are two solutions for the photon index
287: $\Gamma$:
288: \begin{eqnarray*}
289: \Gamma_1 &=& -{1\over2} + \sqrt{{9\over 4}+{4\over y}}, \\
290: \Gamma_2 &=& -{1\over2} - \sqrt{{9\over 4}+{4\over y}},
291: \end{eqnarray*}
292: where the Compton $y$ parameter is given by $y=(4kT_e /m_e c^2){\rm
293: Max}(\tau_{\rm es}, \tau_{\rm es}^2)$.  Up-scattered photons have a
294: power-law energy distribution with photon index $\Gamma_1$ and
295: down-scattered photons have a different power-law distribution with
296: photon index $\Gamma_2$.
297: 
298: We model this case of nonrelativistic electrons with the following
299: Green's function \citep{Sunyaev_1980, COMPTT,Ebisawa_99}, 
300:  which corresponds to the model {\sc simpl-{\small 2}}:
301: \begin{eqnarray}\label{eqn:simpl2}
302: G(E;E_0)dE=\frac{(\Gamma-1)(\Gamma+2)}{(1+2\Gamma)} \begin{cases}
303:    (E/E_0)^{-\Gamma} dE/E_0, \;\qquad E\geq E_{0} \\
304:    (E/E_0)^{\Gamma+1} dE/E_0,\qquad E < E_0. 
305: \end{cases}
306: \end{eqnarray}
307: The function is continuous at $E=E_0$, is normalized such that it
308: conserves photons, and holds for all $\Gamma>1$.  Substituting
309: (\ref{eqn:simpl2}) in (\ref{eqn:convolution}) we see that {\sc simpl-{\small 2}} has
310: two parameters: $\fsc$ and $\Gamma$.  Although the model makes use of
311: two power laws, their slopes are not independent.
312: 
313: As in the case of the standard power law, {\sc simpl} includes no high
314: energy cutoff.  Technically, for any complete model of Comptonization,
315: the up-scattered power-law distribution is cut off for photon energies
316: larger than $kT_e$.  To avoid increasing the complexity of our model,
317: we have ignored this detail; extra parameters could easily be added to
318: account for high energy attenuation if desired.  By keeping the model
319: very basic, {\sc simpl} is a direct two-parameter replacement for the
320: standard power law while bridging the divide between the latter model
321: and physical Comptonization models.
322: 
323: \subsubsection{{\sc simpl-{\small 1}}}
324: 
325: The Green's function (2) is obtained by solving the Kompaneets
326: equation, which assumes that the change in energy of a photon in a
327: single scattering is small.  This assumption is not valid when the
328: Comptonizing electrons are relativistic.
329: 
330: In sec.~7.3 of their text, RL79 discuss Compton scattering by
331: relativistic electrons with a power-law distribution of energy:
332: $n_e(E_e)dE_e \propto E_e ^{-p}dE_e$.  In the limit when the optical
333: depth is low enough that we only need to consider single scattering,
334: they show that the Comptonized spectral energy distribution (SED) is a
335: power law of the form $P(E)dE \propto E^{-(p-1)/2}$.  Equivalently,
336: the photon energy distribution takes the form $n(E)dE \propto
337: E^{-\Gamma}$, with a photon index $\Gamma = (p+1)/2$.  Hardly any
338: photons are down-scattered in energy.
339: 
340: In sec.~7.5, RL79 show that repeated scatterings produce a power-law SED
341: even when the relativistic electrons have a non-power-law distribution
342: \citep[see also][]{Titarchuk_95}.  In terms of the mean amplification of
343: photon energy per scattering $A$ and the optical depth to electron
344: scattering $\tau_{\rm es}$, the Comptonized photon energy distribution
345: takes the form $n(E)dE \propto E^{-\Gamma}$ with a photon index $\Gamma
346: = 1-\ln\tau_{\rm es}/\ln A$.  For the specific case of a thermal
347: distribution of electrons with a relativistic temperature $kT_e \gg
348: m_ec^2$, the amplification factor is given by $A=16(kT_e /m_e c^2)^2$.
349: Once again, hardly any photons are down-scattered.
350: 
351: For both cases discussed above, Comptonization is dominated by
352: up-scattering and produces a nearly one-sided power-law distribution
353: of photon energies.  This motivates the following Green's function,
354: valid for $\Gamma>1$, which we refer to as the model {\sc
355: simpl-{\small 1}}:
356: \begin{eqnarray}\label{eqn:simpl1}
357: G(E;E_0)dE = \begin {cases}
358:   (\Gamma-1)({E/E_0})^{-\Gamma}dE/E_0, ~~E \geq E_0 \\
359:   0, \qquad\qquad\qquad\qquad\quad  E < E_0. 
360: \end{cases}
361: \end{eqnarray}
362: The normalization factor $(\Gamma-1)$ ensures that we conserve
363: photons.  
364: 
365: Although {\sc simpl-{\small 1}} is most relevant for relativistic
366: Comptonization, it can also be used as a stripped-down version of
367: {\sc simpl-{\small 2}} for non-relativistic coronae.  The reason is that the
368: low-energy power-law $(E/E_0)^{\Gamma+1}$ in equation (2) almost never
369: has an important role.  There is not much power in this component, and
370: what little contribution it makes is indistinguishable from the input
371: soft spectrum.  Therefore, even for the case of nonrelativistic
372: thermal Comptonization, for which the Green's function (2) is
373: designed, there would be little difference if one were to use {\sc
374: simpl-{\small 1}} instead of {\sc simpl-{\small 2}}.
375: 
376: \subsection{Comparison to {\sc compTT}}
377: 
378: To illustrate the performance of {\sc simpl} relative to other
379: Comptonization models, we have simulated a $2 \times 10^6$-count {\it BeppoSAX}
380: \citep{BEPPOSAX} observation using the {\sc compTT} model in XSPEC
381: v12.4.0x.
382: 
383: For our source spectrum, we adopt disk geometry, a Wien distribution of
384: seed photons at $kT_0 = 1 \; \keV$, and a hydrogen column density of
385: $\nh = 10^{21} \; \cm ^{-2}$.  We set the optical depth and
386: temperature of the Comptonizing medium to $\tau_c = 2$ and $kT_e = 40
387: \; \keV$.  Our simulation uses the LECS, MECS, and PDS detectors on {\it
388: BeppoSAX}, which span a wide energy range $\sim 0.1-200$ keV (for
389: details on the instruments, see \S3). The total
390: number of counts in the simulated spectra ($\sim 2 \times 10^6$) corresponds to a
391: 3 ks observation of a 1 Crab source.
392: 
393: We analyze the simulated data with a model consisting of a blackbody
394: ({\sc bb}) coupled with {\sc simpl}.  We refer to this model
395: as {\sc simpl$\otimes$bb} (the {\sc$\otimes$} is to emphasize that
396: {\sc simpl} represents a convolution).  The best fits achieved have
397: reduced chi-squared values of $\chi_\nu^2=1.00$ ({\sc simpl-{\small 1}}) 
398: and $\chi_\nu^2=1.06$ ({\sc simpl-{\small 2}}).  The
399: fitted {\sc bb} temperatures are respectively $1.14 \pm 0.02 $ keV and
400: $1.29 \pm 0.01$ keV compared to 1 keV in the original {\sc compTT}
401: model.   Figure~2 shows the fit using {\sc simpl-{\small 1}} 
402: and Table~1 lists the best-fit parameters for both models. 
403: 
404: 
405: In comparison, {\sc compbb}, an alternative model of Compton
406: scattering that assumes slab geometry, fits our simulated spectrum 
407: comparably well as 
408:  {\sc simpl}, with $\rchi = 1.05$ (Table~1).  {\sc compbb}
409: returns the same temperature as {\sc simpl-{\small 2}}, $kT_{\rm
410: bb}=1.29 \pm 0.01$ keV.  Compared to the {\sc compTT} progenitor, {\sc
411: compbb} gives similar estimates of the coronal temperature
412: $kT_e$ and optical depth $\tau_c$ (Table~1).  Even though {\sc
413: compbb} is a physically more realistic model of coronal scattering
414: than {\sc simpl}, it does not outperform {\sc simpl} in terms of
415: fitting the {\sc compTT}-generated data.  Meanwhile, the model 
416: {\sc bb}+{\sc powerlaw} performs quite poorly, yielding $\rchi > 2$.  
417: Parameters for this fit
418: are given in Table~1.  Note that the derived $\nh$ using {\sc
419: powerlaw} is much higher than either the original value or those from
420: fits with {\sc simpl}.
421: 
422: Though {\sc simpl} is a purely empirical model, we see that it can
423: deliver a remarkably successfully fit to data simulated using the
424: physical model {\sc compTT}.  Even for a very cool corona with electron
425: temperatures as low as $kT_e = 20$ keV, which causes {\sc compTT} to
426: produce noticeable curvature in the high-energy spectrum, we find that
427: {\sc simpl-{\small 2}} and {\sc simpl-{\small 1}} achieve reasonable fits with
428: $\rchi < 1.2$.
429: 
430: A significant virtue of {\sc simpl} relative to the physical
431: Comptonization models in XSPEC is that {\sc simpl} can be employed in
432: conjunction with any source of seed photons.  The physical models, on
433: the other hand, are typically restricted to treating only one or two
434: predefined photon distributions.  One standard choice of continuum
435: model that is widely used in fitting Comptonized accretion disks is
436: {\sc diskbb}+{\sc compTT}.  With {\sc simpl}, one would instead employ
437: the model {\sc simpl$\otimes$diskbb}.  The latter not only generates
438: the power law self-consistently via up-scattering of the seed photons,
439: but it also has two fewer parameters.
440: 
441: \subsection{Bulk Motion Comptonization}
442: 
443: The model {\sc bmc} describes the Comptonization of blackbody seed
444: photons by a converging flow of isothermal gas that is freely falling
445: toward a compact object, i.e., bulk motion Comptonization \citep[see,
446: e.g., ][]{Shrader_1998,BMC}.  {\sc bmc} is an alternative to coronal
447: Comptonization models and is structured {\it identically} to {\sc
448: simpl-{\small 2}$\otimes$bb}; both models are specified with just four parameters.
449: As a direct demonstration in XSPEC that {\sc simpl-{\small 2}$\otimes$bb} and {\sc
450: bmc} are identical, we analyzed our simulated {\it BeppoSAX} spectrum
451: described above using both models.  We found that the returned values of
452: the column density $N_{\rm H}$, the blackbody temperature $kT$, and the
453: photon index $\Gamma$ agreed in each case to four or more significant
454: figures.
455: 
456: {\sc bmc} has been variously used to support claims that Compton
457: scattering off in-falling gas within several gravitational radii gives
458: rise to the observed high energy power law in several black-hole
459: binaries \citep[e.g.,][]{Shrader_1998,Shrader_1999,Borozdin_1999}.
460: However, this is only one interpretation of the model; {\sc
461: simpl-{\small 2}}$\otimes${\sc bb}, and therefore {\sc bmc}, can equally
462: be used to support a more standard model of coronal scattering
463: (operating with uniform efficiency at all energies, see \S2 and
464: \S2.1.2).  Thus, although {\sc bmc} is designed specifically to model
465: relativistic accretion inflows, its function is actually quite
466: general.
467: 
468: One virtue of {\sc simpl} relative to {\sc bmc} is that {\sc simpl}
469: does depend upon discerning 
470: the nature of the Comptonizing region, be
471: it corona, relativistic in-falling gas, or other.  Another virtue of
472: {\sc simpl} is that it fully incorporates the utility of {\sc bmc}
473: while allowing complete flexibility in the choice of the spectrum of
474: seed photons, e.g., {\sc simpl$\otimes$diskbb} is more appropriate for
475: modeling Comptonization in accretion disks than {\sc bmc}, which is
476: hardwired to a Planck function.
477: 
478: The theory of bulk motion Comptonization is developed further and
479: rigorously in \citet{BMC}.  This paper describes a Green's function
480: that is more appropriate than the one used in {\sc bmc}.  A complete
481: version of this Green's function is incorporated into the more
482: sophisticated model {\sc compTB}.  However, this model is again
483: limited to treating scattering from a predefined set of
484: (blackbody-like) seed photon distributions and includes additional
485: free parameters.  We find that the fitting results obtained using this
486: Green's function are intermediate between those given by {\sc simpl-{\small 1}}
487: and {\sc simpl-{\small 2}} so long as the temperature of the
488: in-flowing electrons, $T_e$, is above the observed energy range.
489: 
490: %**************************************************************
491: 
492: \section{Data Analysis}
493: 
494: In this section, we apply {\sc simpl} to a sample of observations to
495: illustrate how {\sc simpl} compares with {\sc powerlaw}.  To this end,
496: we have selected two black-hole binaries, H1743--322 and LMC~X--3.
497: H1743--322 (hereafter H1743) is an especially pristine black-hole
498: transient \citep[see ][]{RR_H1743_2006} since, for much of its 2003
499: outburst, its spectrum can be satisfactorily modeled with just absorbed
500: ($\nh \approx 2.2\times10^{22} \cm^{-2}$) thermal-disk and power-law
501: components \citep[][hereafter M07]{JEM_H1743_2007}.  In particular, the
502: 122 days of contiguous spectral data on which we focus do not require
503: any additional components to accommodate the reflection or absorption
504: features that are often present in the spectra of black hole binaries.
505: 
506: The spectra of H1743 were acquired by the {\it Rossi X-ray Timing
507: Explorer} ({\it RXTE}) PCU-2 module \citep{RXTE}, {\it RXTE}'s
508: best-calibrated PCU detector, and were taken in ``standard 2'' format.
509: All spectra have been background subtracted and have typical exposure
510: times $\sim3000\;$s.  The customary systematic error of 1\% has been
511: added to all energy channels.  
512: The resultant pulse-height spectra
513: are analyzed from $2.8 - 25$ keV using XSPEC v12.4.0x (see M07 for
514: further details).
515: 
516: While {\it RXTE} provides good spectral coverage in hard X-rays
517: ($\gtrsim 10 \;\kev$), which is most important for constraining the
518: power-law component, it is not sensitive at low energies ($<2.5
519: \;\kev$).  Therefore, {\it RXTE} data are generally insensitive to
520: $\nh$.  To complement the {\it RXTE} observations presented here, we
521: have selected a {\it BeppoSAX} observation of LMC X--3, a persistent and
522: predominantly thermal black-hole source with a very low hydrogen column
523: \citep[$\nh \approx 4\times10^{20} \cm^{-2}$;][]{Page_2003, Yao_2005}.
524: 
525: The {\it BeppoSAX} narrow-field instruments provide sensitive
526: measurements spanning a wide range in energy, from tenths to hundreds of
527: keV.  The low-energy concentrator system (LECS) and the medium-energy
528: concentrator system (MECS) probe soft fluxes, from $\sim0.1-4$ keV and
529: $\sim1.5-10$ keV, respectively.  The phoswich detector system (PDS) is
530: sensitive to hard X-rays from $\sim15-200$ keV, and the high-pressure
531: gas scintillation counter (HPGSPC) covers $\sim4-100$ keV.  In this
532: analysis, we consider only the LECS, MECS, and PDS because the
533: statistical quality of the HPGSPC data is relatively poor.
534: 
535: In reducing {\it BeppoSAX} data, we have followed the protocols given in
536: the Cookbook for {\it BeppoSAX} NFI Spectral Analysis
537: \citep{BepposaxABC}.  We use pipeline products and extract spectra from
538: 8$\arcmin$ apertures centered on LMC X--3 for both the LECS and
539: (combined) MECS detectors.  For the PDS, which is a simple collimated
540: phoswich detector, we selected the fixed rise-time spectrum.  In our
541: analysis, we have used standard response matrices and included
542: blank-field background spectra with the appropriate scalings.  No
543: pile-up correction is necessary.
544: 
545: \subsection{Steep Power Law State}
546: 
547: About a third of the way through its nine-month outburst cycle, H1743
548: repeatedly displayed spectra in the steep power-law (SPL) state that were 
549: devoid of  absorption features.  A salient feature of the SPL state is the 
550: presence of a strong power-law component of emission.  (For a review of
551: black-hole spectral states and a precise definition of the SPL state,
552: see Table~2 and text in RM06.) Twenty-eight such featureless spectra
553: were consecutively observed over a period of about three weeks (spectra
554: \#58--85; M07).  We focus here on one representative spectrum, \#77.  In
555: Figure~3 we show our fits and the associated unabsorbed models obtained
556: using {\sc diskbb}+{\sc powerlaw} and {\sc simpl$\otimes$diskbb}.
557: Fitted spectral parameters are presented in Table~2.
558: 
559: The quality of fit (as measured by $\rchi$) using either model is
560: comparable.  Nevertheless, there are distinct differences between the
561: models.  The fits with {\sc simpl} have a $\sim 50$\% larger disk
562: normalization compared to {\sc powerlaw} and a $\sim 40$\% lower $\nh$
563: (Table~2).  The fit using {\sc powerlaw} diverges at low energies, as
564: revealed by removing photoabsorption from the fitted models (panels on
565: the right in Fig. 3).  The effect is quite severe and has no obvious
566: physical explanation.  In contrast, the fit using {\sc simpl} is well
567: behaved and the unabsorbed model is not divergent.
568: 
569: \subsection{Thermal Dominant State}
570: 
571: The key feature of the thermal dominant (TD) state is the presence of a totally dominant
572: and soft ($kT \sim 1$ keV) blackbody-like component of emission that
573: arises in the innermost region of the accretion disk.  The TD state is
574: defined by three criteria, the most relevant of which here is that the
575: fraction of the total 2--20 keV unabsorbed flux in the thermal component
576: is $\ge 75$\%.  For the full definition of this state, see Table~2 in
577: RM06.
578: 
579: Here we have chosen H1743 spectrum \#91 which belongs to a sequence of
580: $\sim$50 featureless spectra (\#86--136; M07) in the TD state.  
581: This spectrum has $\Gamma \sim 2$, which is somewhat harder
582: than usual, but is otherwise typical of H1743's TD state.  Spectral fit
583: results are shown in Figure~4.  In addition, in order to further
584: illustrate for the TD state the differences between {\sc simpl} and {\sc
585: powerlaw} at energies below the $\approx 2.5$ keV response cutoff of
586: {\it RXTE}, we use a {\it BeppoSAX} observation of LMC X--3; our results
587: are illustrated in Figure~5.  This observation was carried out on 1996
588: November 28 with exposure times of 1.8, 4.5, and 2 ks respectively for
589: the LECS, MECS and PDS.
590: 
591: As in \S3.1, we fit these data using {\sc diskbb+powerlaw}
592: and {\sc simpl$\otimes$diskbb}.  The best-fit spectral parameters are
593: listed in Table~2.  Due to a calibration offset between
594: the various {\it BeppoSAX} instruments, we follow standard procedure and
595: fit for the normalization of the LECS and PDS relative to the MECS, the
596: best-calibrated of the three.  We adopt the canonical limits of $0.7-1$
597: for LECS/MECS and $0.77-0.93$ for PDS/MECS.  These normalizations are
598: included in the tabulated results.
599: 
600: A comparison of the results obtained with {\sc powerlaw} and {\sc simpl}
601: confirms the trends highlighted in \S3.1, namely the
602: differences in normalization and $\nh$.  However, they are more modest
603: here because the Compton component is weaker in the TD state.
604: 
605: \subsection{Comparison of {\sc simpl} and {\sc powerlaw}}
606: 
607: An examination of Table~2 reveals the following systematic
608: differences in the derived spectral parameters returned when fitting
609: with {\sc simpl} vs. {\sc powerlaw}: {\sc simpl} yields (i) a stronger
610: and softer thermal disk component, i.e., a larger normalization and
611: lower $kT_*$; (ii) a generally steeper power law component (larger
612: $\Gamma $); and (iii) a systematically lower $\nh$.  As we now show, all
613: of these effects can be simply understood.
614: 
615: Because {\sc powerlaw} produces higher fluxes than {\sc simpl} at low
616: energies, it tends to suppress the flux available to the (soft)
617: thermal component, namely {\sc diskbb} in the examples given here.
618: This explains why {\sc powerlaw} tends to harden the {\sc diskbb}
619: component and to steal flux from it (i.e., reduce its normalization
620: constant).  Meanwhile, at low energies the {\sc powerlaw} component
621: predicts artificially high fluxes that, in order to conform to the
622: observed spectrum, depress the value of $\Gamma$.  These differences
623: between {\sc simpl} and {\sc powerlaw} are most pronounced when the
624: power law is relatively steep, i.e., typically when $\Gamma \gtrsim
625: 3$.
626: 
627: Modest and reasonable values of $\nh$ are returned in fits using {\sc
628: simpl}, as well as {\sc compTT} and other Comptonization models, because
629: the Compton tail is produced by the up-scattering of seed photons and
630: there is no power-law component at low energies.  In contrast, {\sc
631: powerlaw} continues to rise at low energies, which forces $\nh$ to
632: increase in order to allow the model to fit the observed spectrum.  This
633: systematic difference is apparent in our fit results for the H1743
634: spectra and is especially prominent in the case of the LMC X-3 spectrum
635: for which $\nh$ differs by a factor of two.  For H1743, the discrepancy
636: in $\nh$ is much less for the TD spectrum than for the SPL spectrum
637: because the SPL state has both a steeper and relatively stronger
638: power-law component.
639: 
640: We turn now to consider the {\sc diskbb} normalization constant, which
641: is proportional to $R_{\rm in}^2$, the square of the inner disk radius
642: (see footnotes to Table~2).  For the pair of H1743 spectra, we see that
643: the disk normalization obtained with {\sc powerlaw} is $\approx$35\%
644: smaller in the SPL state than in the TD state (Table~2), indicating that
645: $R_{\rm in}$ is smaller for the SPL state.  With {\sc simpl}, on the
646: other hand, there is no significant change in the normalization, and
647: hence both the SPL and TD states can be modeled with a disk that has the
648: same inner radius.  The radius is constant because {\sc simpl} recovers
649: the original (unscattered) flux emitted by the disk, which {\sc
650: powerlaw} cannot do.  
651: 
652: 
653: This crucial ability to unify the inner regions of the accretion disk 
654: in thermally active states exhibiting both high and low levels of Comptonization 
655: (i.e., TD and SPL states) paves the way for a full general relativistic analysis which 
656: can formally link $\rin$ to black-hole spin (see discussion in \S\ref{subsec:spin}). 
657: \citet{Kubota_2004} similarly identified a constant radius for the black hole binary
658: XTE~J1550--564 between the TD and SPL states in an analysis using the
659: model {\sc diskbb + thcomp}.  Because {\sc thcomp} is implemented as an 
660: additive (i.e., non-convolution) model, Kubota \& Makishima had to employ an awkward and
661: ad hoc procedure to obtain their result (see their Appendix).  Their work improved upon a similar result obtained for black-hole GRO~J1655--40 \citep{Kubota_2001}.  
662: With {\sc
663: simpl}, the modeling is significantly easier. 
664: 
665: 
666: %********************************************************************
667: 
668: \section{Discussion}
669: 
670: \subsection{Black Hole X-ray States}
671: 
672: A standard method of classifying X-ray states in black hole binaries
673: involves spectral decomposition into two primary components -- a
674: multi-temperature blackbody disk, {\sc diskbb}, and a Compton power law,
675: {\sc powerlaw} (RM06).  This method is compromised by the use of the
676: standard power law when the photon index is large ($\Gamma \gtrsim 3$).
677: In this case, at low energies the flux from the power law can rival or
678: exceed the thermal component and thereby pollute it.  As discussed in
679: \S2, intrusion of the power-law component at low
680: energies is fundamentally inconsistent with Compton scattering.
681: 
682: This difficulty in classifying states, which is caused by the use of
683: {\sc powerlaw}, is remedied by the use of {\sc simpl} because the latter
684: model naturally truncates the power-law component at low energies.  It
685: is useful to consider the intrinsic differences between the two models
686: and how they influence the classification of black-hole X-ray states.
687: Using {\sc powerlaw}, the thermal disk and tandem power-law emission are
688: modeled independently.  On the other hand, under {\sc simpl} all
689: photons originate in the accretion disk.  Some of these disk photons
690: scatter into a power law en~route from the disk to the observer.   As
691: described in \S3.3, fits employing {\sc simpl} imply stronger disk
692: emission and weaker Compton emission than those using {\sc powerlaw}.
693: As a result, state selection criteria would need to be modified for
694: classification using {\sc simpl}. This topic is beyond the scope of the
695: present paper.
696: 
697: 
698: 
699: \subsection{Application to the Measurement of Black Hole Spin}\label{subsec:spin}
700: 
701: During the past few decades, the masses of 22 stellar black holes have
702: been measured, 17 of which are found in transient black hole binaries
703: (RM06).  Recently, we have measured the spins of four of these stellar
704: black holes \citep{Shafee_spin, spin_1915, spin_m33} by fitting their
705: continuum spectra to our fully relativistic model of an accretion disk
706: {\sc kerrbb2} \citep{KERRBB, spin_1915} plus the standard power-law
707: component {\sc powerlaw}.  
708: In the continuum-fitting method, spin is measured
709: by estimating the inner radius of the accretion disk $R_{\rm in}$ 
710: \citep{Zhang_1997}.  We
711: identify $R_{\rm in}$ with the radius of the innermost stable circular
712: orbit $R_{\rm ISCO}$, which is predicted by general relativity.  
713: 
714: To date, we have conservatively selected only
715: TD data for analysis (\S3.2). 
716: %, i.e., X-ray spectra that are dominated by a soft blackbody-like component (\S3.2; RM06).  
717: Meanwhile, the transient black
718: hole binaries spend only a modest fraction of their outburst cycle in
719: the TD state and are often found in the SPL or some intermediate state
720: (see Figs. 4--9 in RM06).  Since for each source we seek to obtain as
721: many independent measurements of the spin parameter as possible, our
722: sole reliance on TD data has been a significant limitation.  
723: Using {\sc simpl}, the highly Comptonized SPL state is now able to provide
724: estimates of spin, a matter to be discussed more fully in 
725: (J. Steiner et al.\ 2009, in preparation). 
726: Not only will this allow us to substantially
727: increase the size of our data sample for many sources, it also will
728: likely allow us to obtain spin measurements for sources such as Cygnus
729: X-1 that never enter the TD state \citep{MR06}.
730: 
731: %Black hole spin is commonly expressed in terms of the dimensionless
732: %quantity $a_* \equiv cJ/GM^2$ ($|a_*| \le 1$), where $M$ and $J$ are
733: %respectively the black hole mass and angular momentum
734: %\citep{Shapiro_1983}.  
735: %In the continuum-fitting method, spin is measured
736: %by estimating the inner radius of the accretion disk $R_{\rm in}$.  We
737: %identify $R_{\rm in}$ with the radius of the innermost stable circular
738: %orbit $R_{\rm ISCO}$, which is predicted by general relativity.  
739: %Strong
740: %support for linking $R_{\rm in}$ to $R_{\rm ISCO}$ is provided by
741: %decades of empirical evidence that $R_{\rm in}$ is constant in
742: %disk-dominated states of BH binaries \citep[e.g.,][]{Gierlinski_2004}
743: %and by recent MHD simulations of thin accretion disks
744: %\citep{Reynolds_2008, Shafee_2008}.  $R_{\rm ISCO}/M$ is a monotonic
745: %function of $a_*$, decreasing from $6M$ to $M$ (c~=~G~=~1) as spin
746: %increases from $a_*=0$ to $a_*=1$.  This relationship between $a_*$ and
747: %$R_{\rm ISCO}$ is the foundation of the continuum-fitting method: Given
748: %an independent dynamical measurement of $M$, a measurement of $R_{\rm
749: %ISCO}$ is equivalent to a measurement of $a_*$.
750: 
751: Our reason for developing {\sc simpl} was to improve our methods for
752: analyzing TD-state data in order to determine more reliable values of
753: black hole spin.  That {\sc simpl} now allows us to determine spins
754: for SPL-state data was a serendipitous discovery and a major
755: bonus.  We were motivated to develop {\sc simpl} because we have been
756: hampered by the use of {\sc powerlaw} in two ways.  First, in all of
757: our work we have exclusively used TD spectral data
758: \citep[e.g.,][]{Shafee_spin}, which is maximally free of the uncertain
759: effects of Comptonization.  The selection of these data is problematic
760: because, as indicated in \S4.1 above, it can be affected in unknown
761: ways when the spectral index of the Compton component is large.
762: 
763: Secondly, of greater concern is the potential adulteration in the TD
764: state of the thermal component by the power-law component, which can
765: have an uncertain and sizable effect on the fitted value of the
766: black-hole spin parameter.  In the case of a number of
767: spectra with steep power-law components, we found that the fitted values
768: of the spin parameter were affected by the contribution of the power law
769: flux at energies below $\sim 5$ keV; e.g., see \S~4.2 in
770: \cite{spin_1915}.  In order to mitigate this problem, we applied in turn
771: two alternative models: {\sc compTT}, and a standard power-law component
772: curtailed by an exponential low-energy cutoff, {\sc
773: expabs$\times$powerlaw}.  The former model was unsatisfactory because we
774: were unable to fit for reasonable values of both the coronal temperature
775: $kT_e$ and the optical depth $\tau_c$.  The latter model was likewise
776: unsatisfactory because it requires the use of an arbitrary cutoff energy
777: $E_c$.
778: 
779: We developed {\sc simpl} in order to sidestep these difficulties and
780: uncertainties.  At low energies, the model truncates the power law in
781: the same physical manner as {\sc compTT} and other sophisticated
782: Comptonization models.  {\sc simpl} self-consistently ties the
783: emergent power-law flux to the seed photons in order to deliver the
784: power-law component via coronal reprocessing.  An application of {\sc
785: simpl} to the measurement of the spin of the black-hole primary in LMC
786: X-1 is described in L.\ Gou et al. (2009).
787: 
788: 
789: 
790: %******************************************************************
791: 
792: \section{Summary}
793: 
794: We present a new prescription for treating Comptonization in X-ray
795: binaries.  While no new physics has been introduced by this model, its
796: virtues lie in its simplicity and natural application to a wide range of
797: neutron-star and black-hole X-ray spectra.  {\sc simpl} offers a generic
798: and empirical approach to fitting Comptonized spectra using the minimum
799: number of parameters possible (a normalization and a slope), and it is
800: valid for a broad range of geometric configurations (e.g.,
801: uniform slab and spherical geometries).  The scattering of a seed
802: spectrum occurs via convolution, which self-consistently mimics physical reprocessing of
803: photons from, e.g., an accretion disk.  In addition to this physically
804: motivated underpinning, {\sc simpl} remains as unassuming as {\sc
805: powerlaw} but without its troublesome divergence at low energies.
806: 
807: Our model is valid for all $\Gamma>1$.  We have shown that {\sc simpl}
808: is able to provide a good fit to a demanding simulated data set, which
809: was generated with the widely-used Comptonization model {\sc compTT}.
810: Furthermore, we have demonstrated that {\sc simpl} and {\sc powerlaw}
811: give very comparable results when fitting spectral data in terms of
812: quality-of-fit and spectral index (see Table \ref{tab:H1743}).  This
813: quality of performance holds true not only for spectra with weak
814: Compton tails (TD state) but also for spectra requiring a large
815: Compton component (SPL state).  In the latter case, the model based on
816: {\sc simpl} gives physically more reasonable results for the soft end
817: of the spectrum (e.g., see \S3.3).
818: 
819: Using {\sc simpl$\otimes$diskbb} it will be important to revisit the
820: classification of black hole states (RM06) for two reasons.  First, the
821: selection of TD data will no longer be adversely affected by the
822: presence of a steep power-law component.  Secondly, this model will
823: allow some degree of unification of the TD state and SPL state, the
824: latter being a more strongly Comptonized version of the former.  In
825: determining black hole spin via the continuum-fitting method using {\sc
826: kerrbb2}, {\sc simpl} is a significant advance on three fronts: It will
827: (1) enable the selection of data with a dominant thermal component that
828: is not mucked up by the effects of a divergent power-law component; (2)
829: allow reliable spin measurements to be obtained using
830: strongly-Comptonized SPL data, thereby substantially increasing the data
831: sample for a given source; and (3) likely make possible the
832: determination of the spins of some black holes that do not enter the TD
833: state (e.g., Cygnus X-1).
834: 
835: %***************************************************************
836: 
837: \acknowledgements The authors would like to thank George Rybicki for
838: discussions on the physics of Comptonization as well as Jifeng Liu,
839: Lijun Gou, Rebecca Shafee, and Ron Remillard for their input on {\sc
840: simpl}.  JFS thanks Joey Neilsen for enthusiastic discussions as well as
841: comments on the manuscript, Ryan Hickox for suggestions which improved
842: this paper, and Keith Arnaud for helping implement {\sc simpl} in XSPEC.
843: The authors thank Tim Oosterbroek for his indefatigable assistance with
844: the {\it BeppoSAX} reduction software.  JFS was supported by the
845: Smithsonian Institution Endowment Funds and RN acknowledges support from
846: NASA grant NNX08AH32G and NSF grant AST-0805832.  JEM acknowledges
847: support from NASA grant NNX08AJ55G.
848: 
849: %****************************************************************
850: 
851: 
852: 
853: \appendix
854: \section{XSPEC Implementation}
855: 
856: {\sc simpl} is presently implemented in XSPEC.  This version includes
857: three parameters (two that can be fitted), the power-law photon index
858: ($\Gamma$), the scattered fraction ($\fsc$), and a switch to set 
859: up-scattering only ({\sc simpl-{\small 1}}: switch~$>0$) and
860: double-sided scattering ({\sc simpl-{\small 2}}: switch~$\leq 0$). 
861: Since {\sc simpl} redistributes input photons to higher (and lower) energies, 
862: for detectors with limited response matrices (at high or low energies), or
863: poor resolution, the sampled energies should be extended or resampled
864: within XSPEC to adequately cover the relevant range.  For example,
865: when treating the {\it RXTE} data in \S3, which has no response
866: defined below $1.5$ keV, the command ``{energies 0.05 50 1000 log}''
867: was used to explicitly extend and compute the model over 1000
868: logarithmically spaced energy bins from $0.05-50$ keV.
869: 
870: Using {\sc simpl} can be problematic when $\Gamma$ is large, especially
871: if the power-law component is faint or the detector response extends
872: only to $\sim 10$ keV (e.g., {\it Chandra}, {\it XMM} or {\it ASCA}).
873: When the photon index becomes sufficiently large, a runaway process can
874: occur in which $\Gamma$ steepens and the scattered fraction becomes
875: abnormally high (typically $\gtrsim 50\%$, inconsistent with a weak
876: power law).  This occurs because scattering redirects photons from
877: essentially a $\delta$-function into a new function with characteristic
878: width set by $\Gamma$.  If $\Gamma$ reaches large values ($\gtrsim 5$),
879: the scattering kernel will also act like a $\delta$-function, and the
880: convolved spectrum will be nearly identical to the seed spectrum.
881: 
882: In such circumstances, we recommend bracketing $\Gamma$.  In practice,
883: the power-law spectral indices of black-hole systems are found to lie in
884: the range $1.4 \lesssim \Gamma \lesssim 4$ \citep{RR_JEM_review_2006}.
885: An upper limit of $\Gamma \sim4-4.5$ is typically sufficient to prevent
886: this runaway effect, and this constraint should be applied if it is
887: deemed appropriate for the source in question.
888: 
889: We advise against applying {\sc simpl} to sharp components such as
890: spectral lines or reflection components.  {\sc simpl} broadens spectral
891: lines, and for prominent higher energy emission features (such as Fe
892: K$\alpha$), {\sc simpl} can saturate the power-law flux with scattering
893: from the line itself.  To prevent this from occurring, such components should be invoked
894: outside the scope of {\sc simpl}.  For example, the XSPEC model
895: declaration ``{model phabs$\times$(smedge$\times$simpl(kerrbb)+laor)}''
896: would satisfy this recommendation.
897: 
898: 
899: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
900: 
901: \newcounter{BIBcounter} % Make a counter to reference
902: \refstepcounter{BIBcounter} \bibliography{ms} \mbox{~}
903: 
904: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
905: 
906: \input{tab1.tex}
907: 
908: \input{tab2.tex}
909: 
910: 
911: \clearpage
912: 
913: \begin{figure}
914: \plotone{f1.eps}
915: \caption{Spectral energy density vs.\ photon energy for a sample
916: spectrum calculated with {\sc simpl-{\small 1}} (solid lines) and {\sc
917: simpl-{\small 2}} (dashed lines).  The models conserve photons and
918: Comptonize a seed spectrum, which in the case shown is {\sc diskbb}
919: with $kT_*=1 \; \keV$ (black line).  Ascending colored lines show
920: increasing levels of scattering, from $\fsc = 1-100$\%.}
921: \end{figure}
922: 
923: \begin{figure}
924: \plotone{f2.eps} 
925: \caption{The data correspond to a simulated {\it BeppoSAX} observation
926: with a total of 2.1$\times 10^6$ counts; the spectrum was generated
927: using {\sc compTT}.  The histogram shows the fit achieved using {\sc
928: simpl-{\small 1}}.  This fit is performed over the recommended energy
929: ranges of the narrow-field instruments (NFI), as given by the Cookbook
930: for {\it BeppoSAX} NFI Spectral Analysis, yielding $\chi^2_\nu=1.00$.
931: For details, see Table~1.  This example demonstrates the ability of
932: {\sc simpl} to match a representative spectrum generated by a physical
933: model of Comptonization.}
934: \end{figure}
935: 
936: \begin{figure}
937: \plotone{f3.eps}
938: \caption{{\it left:} Unfolded spectral fits to an {\it RXTE}
939: observation of H1743 in the SPL state and {\it right:} the corresponding
940: unabsorbed models.  Data are fitted using {\it (a,b):} {\sc
941: phabs$\times$(diskbb+powerlaw)}, {\it (c,d):} {\sc
942: phabs$\times$(simpl-{\small 1}$\otimes$diskbb)}, {\it (e,f):} {\sc
943: phabs$\times$(simpl-{\small 2}$\otimes$diskbb)}.  The composite model is
944: represented by a solid black line and the emergent disk and Compton
945: components are shown as red and blue dashed lines respectively.  The
946: seed spectrum for {\sc simpl} is shown (dashed) in green.  Contrasting
947: behaviors between {\sc simpl} and {\sc powerlaw} are most clearly
948: revealed in the unabsorbed models at low energies.  Spectral parameters
949: are given in Table~2.}
950: \end{figure}
951: 
952: \begin{figure}
953: \plotone{f4.eps}
954: \caption{Same as Figure~3 except that the results shown here are for an
955: {\it RXTE} observation of H1743 in the TD state.  The systematic
956: differences between the {\sc simpl} and {\sc powerlaw} fits are greatly
957: reduced compared to the differences shown for the SPL example in
958: Figure~3.}
959: \end{figure}
960: 
961: \begin{figure}
962: \plotone{f5.eps}
963: \caption{Same as Figures 3 and 4 for a TD {\it BeppoSAX} spectrum of LMC
964: X--3.  The data have been rebinned for plotting purposes only and both
965: LECS and PDS counts have been rescaled by the fitted normalizations
966: given in Table~2.  At low energies (below $\sim$0.5 keV),
967: the unabsorbed model is strongly compromised for fits with {\sc
968: powerlaw}}.
969: \end{figure}
970: 
971: \end{document}
972: