1: \documentclass[12pt,preprint]{aastex}
2: %%%%\usepackage{color}
3: \usepackage{emulateapj5}
4: \newcommand{\et}{et al.}
5: \newcommand{\lum}{{\rm luminosity}}
6: \newcommand{\fv}{F_{\rm var}}
7: \newcommand{\fvar}{F_{\rm var}}
8: \newcommand{\Fvar}{F_{\rm var}}
9: \newcommand{\mbh}{M_{\rm BH}}
10: \newcommand{\Tb}{T_{\rm b}}
11: \newcommand{\fracfvar}{\frac{F_{var,soft}}{F_{var,hard}}}
12: \newcommand{\Dtsamp}{{\Delta}T_{{\rm samp}}}
13: \newcommand{\Dtsm}{{\Delta}T_{{\rm sim}}}
14: \newcommand{\rxte}{{\it RXTE}}
15: \newcommand{\xte}{{\it RXTE}}
16: \newcommand{\exosat}{{\it EXOSAT}}
17: \newcommand{\xmm}{{\it XMM-Newton}}
18: \newcommand{\nh}{N_{\rm H}}
19: \newcommand{\Msun}{\hbox{$\rm\thinspace M_{\odot}$}}
20: \def\lineindex#1{{\thinspace\small#1}}
21: \def\I{\lineindex I} \def\II{\lineindex{II}} \def\III{\lineindex{III}}
22: \slugcomment{}
23: \shorttitle{X-ray Spectrum of NGC 3227}
24: \shortauthors{Markowitz et al.}
25:
26: \begin{document}
27:
28: \title{A Comprehensive X-ray Spectral Analysis of the Seyfert 1.5 NGC 3227}
29:
30: \author{A. Markowitz\altaffilmark{1}, J.N. Reeves\altaffilmark{2}, I.M. George\altaffilmark{3,4}, V.\ Braito\altaffilmark{5}, R.\ Smith\altaffilmark{5}, S.\ Vaughan\altaffilmark{5}, P.\ Ar\'{e}valo\altaffilmark{6}, F.\ Tombesi\altaffilmark{4,7,8,9}
31: \altaffiltext{1}{Center for Astrophysics and Space Sciences, University of California, San Diego, M.C.\ 0424, La Jolla, CA, 92093-0424, USA}
32: \altaffiltext{2}{Astrophysics Group, School of Physical and Geographical Sciences, Keele University, Keele, Staffordshire, ST5 5BG, UK}
33: \altaffiltext{3}{Department of Physics, University of Maryland, Baltimore County, 1000 Hilltop Circle, Baltimore, MD, 21250, USA}
34: \altaffiltext{4}{X-ray Astrophysics Laboratory, Code 662, NASA Goddard Space Flight Center, Greenbelt, MD, 20771, USA}
35: \altaffiltext{5}{X-ray Astronomy Group, University of Leicester, Leicester LE1 7RH, UK}
36: \altaffiltext{6}{Department of Physics and Astronomy, University of Southampton, Highfield, Southampton, SO17 1BJ}
37: \altaffiltext{7}{Department of Physics and Astronomy, Johns Hopkins University, 3400 N.\ Charles Street, Baltimore, MD 21218, USA}
38: \altaffiltext{8}{INAF-IASF Bologna, via Gobetti 101, 40129, Bolognia, Italy}
39: \altaffiltext{9}{Departimento di Astronomia, Universit\`{a} degli Studi di Bologna, via Ranzini 1, 40127, Bologna, Italy}
40: }
41:
42: \begin{abstract}
43: We present results of a 100 ks {\it XMM-Newton} observation of the
44: Seyfert 1.5 AGN NGC 3227.
45: Our best-fit broadband model to the EPIC pn spectrum consists of a moderately
46: flat (photon index of 1.57) hard X-ray power-law absorbed by cold
47: gas with a column density of $3 \times 10^{21}$ cm$^{2}$, plus
48: a strong soft excess, modeled as a steep power law with a photon index of 3.35,
49: absorbed by cold gas with a column density of 9 $\times$ 10$^{20}$ cm$^{-2}$.
50: The soft excess increases in normalization by $\sim$20$\%$ in $\sim$20 ks,
51: independently of the hard X-ray emission component, and
52: the UV continuum, tracked via the OM,
53: also shows a strong increasing trend over the observation, consistent with reprocessing of soft X-ray emission.
54: Warm absorber signatures are evident in both the EPIC and RGS spectra;
55: we model two absorbing layers, with ionization parameters log$\xi$ = 1.2 and 2.9 erg cm s$^{-1}$, and
56: with similar column densities ($\sim$ 1--2 $\times 10^{21}$ cm$^{-2}$).
57: The outflow velocities relative to systemic of the high- and low-ionization absorbers are
58: estimated to be --(2060$^{+240}_{-170}$) km s$^{-1}$ and --(420$^{+430}_{-190}$) km s$^{-1}$, respectively.
59: The Fe K$\alpha$ line FWHM width is 7000 $\pm$ 1500 km s$^{-1}$;
60: its inferred distance from the black hole is consistent with the BLR and with the inner radius of the dust
61: reverberation-mapped by Suganuma et al. An emission feature near 6.0 keV is
62: modeled equally well as a narrow redshifted Fe K line, possibly associated with a
63: disk ``hot-spot,'' or as the red wing to a relativistically broadened Fe line profile.
64: {\it Swift}-BAT and archival {\it RXTE} data suggest at most a weak Compton reflection
65: hump ($R \lesssim 0.5$), and a high-energy cutoff near 100 keV. From {\it RXTE} monitoring,
66: we find tentative evidence for a significant fraction of the Fe line flux
67: to track variations in the continuum on time scales $<$ 700 days.
68:
69: \end{abstract}
70:
71:
72: \keywords{galaxies: active --- galaxies: Seyfert --- X-rays: galaxies --- galaxies: individual (NGC 3227) }
73:
74: \section{Introduction}
75:
76:
77:
78: A 1993 {\it Advanced Satellite for Cosmology and Astrophysics (ASCA)}
79: observation of the nucleus of the Seyfert 1.5
80: NGC 3227 first revealed evidence for
81: ionized absorbing gas with $N_{\rm H,WA}$ $\sim$ 1--4$\times$10$^{21}$ cm$^{-2}$
82: (Netzer \et\ 1994, Ptak \et\ 1994). Komossa \& Fink (1997b),
83: using 1993 {\it R\"{o}ntgen Satellite (ROSAT)}-PSPC data, and George \et\ (1998b),
84: using {\it ASCA} data taken in 1993 and 1995,
85: further modeled the warm absorber, finding it to be relatively
86: lowly-ionized relative to most other Seyferts' warm absorbers.
87: Given the low-resolution spectra,
88: the solution for further progress was to obtain a gratings spectrum.
89: Komossa (2002) reported on a {\it Chandra}-LETGS observation of NGC 3227
90: in October 2000, noting edge-like features near 0.7 keV. %%%% , associated with ionized absorption.
91: {\it XMM-Newton} first observed NGC 3227 in
92: November 2000, when the source was undergoing a 3-month long period
93: of very high levels of obscuration due to a compact cloud with
94: $N_{\rm H}$ near $3 \times 10^{23}$ cm$^{-2}$, covering 90$\%$ of the central source, and
95: likely associated with the BLR, traversing the line of sight (Lamer \et\ 2003).
96: The soft X-ray data from this observation indicated the
97: warm absorber to have ionization parameter log$\xi$\footnote{The ionization parameter $\xi \equiv L_{\rm ion} n_{\rm e}^{-1} r^{-2}$, where $L_{\rm ion}$ is the isotropic 1--1000 Ryd ionizing continuum luminosity, $n{\rm e}$ is the electron number density, and $r$ is the distance from the central continuum source to the absorbing gas.} = 1.7--2.0 erg cm s$^{-1}$ and
98: a column density $N_{\rm H,WA}$ = 2--9$\times$10$^{21}$ cm$^{-2}$ (Gondoin \et\ 2003).
99:
100:
101:
102: Optical spectra of the nucleus of NGC 3227 have indicated strong reddening
103: due to the presence of dust along the line of sight.
104: The narrow line H$\alpha$/H$\beta$ ratio, for instance, indicates reddening,
105: with measured ratios in NGC 3227 implying visual extinctions $A_{\rm V}$ = (H$\alpha$/H$\beta$)/3.1
106: near 1.2--1.7 (Cohen 1983; Gonzalez Delgado \& Perez 1997),
107: though $A_{\rm V}$ values of 4.5--4.9 have been reported
108: (Mundell \et\ 1995a, Rubin \& Ford 1968)\footnote{The
109: H$\alpha$/H$\beta$ ratio is intrinsically 3.1, assuming case B recombination and assuming that collisional excitation
110: is negligible, applicable for typical NLR conditions, but see warnings by Netzer 1990 regarding
111: uncertainties in the intrinsic Balmer decrement.}.
112: The broad H$\alpha$/H$\beta$ line ratio shows a similar level of reddening,
113: $A_{\rm V}$ = 1.4 (Cohen 1983, Winge \et\ 1995).
114: The steep drop in NGC 3227's continuum flux from the near-UV
115: to the far-UV also supports reddening: the spectral index from {\it International Ultraviolet Explorer (IUE)}
116: measurements is $\alpha_{\rm IUE} = -2.9$ for NGC 3227, whereas $\alpha_{\rm IUE} = -1.4$
117: is more typical for
118: Seyfert 1s (Komossa \& Fink 1997b; Courvoisier \& Paltani 1992; Kinney \et\ 1991).
119: Assuming a standard Galactic gas/dust ratio of
120: %%%%% $N_{\rm H}$ / $E_{\rm (B-V)}$ = 6$\times$10$^{21}$ cm$^{-2}$ magn$^{-1}$ or
121: $N_{\rm H}$/$A_{\rm V}$ = 2$\times$10$^{21}$ cm$^{-2}$ magn$^{-1}$
122: (e.g., Shull \& van Steenburg 1985), the amount of NLR and BLR
123: reddening implies that the dust is associated with
124: a column of gas with $N_{\rm H}$ $\sim$ 2--3 $\times$10$^{21}$ cm$^{-2}$.
125:
126:
127:
128: Komossa \& Fink (1997b) and George \et\ (1998b) also reported local absorption due to cold material
129: with a column density $N_{\rm H}$ = 3--6 $\times$10$^{20}$ cm$^{-2}$,
130: in addition to the Galactic column (2.1$\times$10$^{20}$ cm$^{-2}$,
131: Dickey \& Lockman 1990, Murphy \et\ 1996). The inferred local cold column density
132: agrees with that derived from 21cm VLA observations (Mundell \et\ 1995b);
133: however, it is not sufficient to produce the observed optical reddening,
134: assuming a standard Galactic gas/dust ratio.
135: The agreement between the inferred column density associated with the
136: dust and the column density of the ionized gas, however, prompted Komossa \& Fink (1997b) to
137: suggest that the dust was embedded in the warm absorber and not
138: associated with the cold gas. Brandt, Fabian, \& Pounds (1996) first suggested
139: this "dusty warm absorber" concept for the quasar IRAS 13349+2438.
140:
141: %%%%KF97b: dust-free and dust-contained: same fit results
142:
143: However, Kraemer \et\ (2000) suggested that it was
144: unlikely dust could survive within the high-ionization gas.
145: They proposed a configuration consisting of {\it two}
146: warm absorber zones, in addition to the cold column.
147: One warm absorber is dust-free and highly-ionized,
148: responsible for the oxygen edges; the other is a very lowly-ionized absorber with dust mixed in,
149: referred to as a "dusty lukewarm absorber." Those authors
150: modeled the latter absorber with a density of 20 cm$^{-3}$ and
151: a location encompassing or lying just outside the NLR, which has a
152: radius of $\sim$100 pc (based on [O III] imaging; Schmitt \& Kinney 1996).
153: Using {\it Hubble Space Telescope (HST)}-STIS, Crenshaw \et\ (2001) first observed
154: the features of the dusty lukewarm absorber, evident in the form of
155: optical/UV absorption lines due to intermediate species of C, N and Si;
156: Crenshaw \& Kraemer (2001)
157: suggested that the absorber may possibly be a relatively highly-ionized
158: component of the NLR seen in absorption.
159: Kraemer \et\ (2000) suggested that the dust may evaporate off the
160: putative molecular torus and be swept up in an outflowing wind
161: (e.g., Reynolds 1997).
162: Based on V-band-to-K-band reverberation mapping (the latter band dominated by
163: thermal dust emission) conducted by the MAGNUM collaboration,
164: Suganuma \et\ (2006) concluded that
165: the innermost extent the dust was at a radius of $\sim$5--20 light-days,
166: suggesting an alternate or additional site from which an outflowing wind
167: could sweep up dust.
168:
169:
170:
171: In 2006, we obtained a new, 100 ks {\it XMM-Newton} observation of NGC 3227 in an unobscured
172: state, to better characterize the broadband X-ray continuum and X-ray
173: warm absorber features, search for signatures of dust in the X-ray spectrum,
174: relate the X-ray and dusty optical/UV warm absorbers,
175: and study Fe K bandpass emission features.
176: $\S$2 describes the observations and data reduction.
177: $\S$3 and $\S$4 describe EPIC spectral fits to the time-averaged spectrum,
178: focusing on the Fe K bandpass and then extending to the
179: 0.2--10 keV bandpass, respectively.
180: $\S$5 describes the fits to the RGS spectra.
181: $\S$6 describes spectral analysis of archival {\it RXTE} data to constrain
182: the hard X-ray continuum.
183: $\S$7 describes time- and flux-resolved spectral fits to the EPIC data,
184: including $F_{\rm var}$ spectra, and time-resolved spectral fits
185: to the {\it RXTE} data to quantify the temporal behavior of the Fe line.
186: X-ray/UV correlations are presented in $\S$8.
187: The results are discussed in
188: $\S$9, and a brief summary is given in $\S$10.
189:
190: \section{Observations and Data Reduction}
191:
192: \subsection{{\it XMM-Newton} data reduction}
193:
194: NGC 3227 was observed by {\it XMM-Newton} during Revolution 1279, on
195: December 3--4, 2006, for a duration of 99 ks.
196: {\it XMM-Newton}'s European Photon Imaging Camera (EPIC) consists of one pn CCD
197: back-illuminated array sensitive to 0.15--15 keV photons (Str\"{u}der
198: \et\ 2001), and two MOS CCD front-illuminated arrays sensitive to
199: 0.15--12 keV photons (MOS 1 and MOS 2, Turner \et\ 2001).
200: Data from the pn were taken in Large Window mode; data from both MOSes
201: were taken in PrimePartialW2/small window mode.
202: The medium filter was used for all three EPIC instruments.
203: Spectra were extracted using {\it XMM-Newton}-SAS version 7.1.0 and using standard
204: extraction procedures. For all three cameras, source data were
205: extracted from a circular region of radius 40$\arcsec$; backgrounds were
206: extracted from circles of identical size, centered $\sim$3$\arcmin$ away
207: from the core. Hot, flickering, or bad pixels were excluded. Data were
208: selected using event patterns 0--4 for the pn and 0--12 for the MOSes.
209: Using the SAS task {\sc epatplot}, we verified that pile-up was negligible
210: for the pn. For the MOS, pile-up was minimal ($<$3$\%$) up to 8 keV,
211: and a bit higher (3--10$\%$) above 8 keV.
212: The effect of such pile-up is to artifically flatten the spectrum.
213: However, we also extracted counts from the MOS CCDs
214: using an annular region with inner radius 5$\arcsec$ and outer
215: radius 40$\arcsec$ to minimize pile-up; the resulting spectra
216: were virtually identical to those extracted using the circular region.
217: We elected to fit the spectra extracted using circular regions
218: in order to maximize the photon statistics associated with
219: emission features in the Fe K bandpass.
220:
221: We inspected the 10--12 keV pn background
222: light curve; there were no significant flares (the 10--12 keV rate
223: never exceeded 0.04 ct s$^{-1}$). The total good exposure time
224: was 90.7 ks for pn and 96.4 ks for each MOS.
225: The 0.2--10 keV pn light curve, binned to 600 s and normalized by its
226: mean count rate (11.74 ct s$^{-1}$), is displayed in the top
227: panel of Figure 1. Similarly displayed in Figure 1 are the 0.2--1 and 3--10 keV
228: light curves; in these energy ranges, the total continuum is dominated
229: by the soft excess and hard X-ray power-law components, respectively,
230: as will be demonstrated in $\S$4.
231:
232:
233: The Reflection Grating Spectrometer (RGS) data were extracted using standard extraction
234: procedures, including the SAS task {\sc rgsproc} and the most recent calibration files.
235: We used the first-order data only; spectra were binned every 10 channels ($\sim$0.1$\AA$, or roughly the RGS resolution).
236: Virtually all bins contained at least 20 counts, allowing use of the $\chi^2$ statistic.
237: Bad RGS channels were ignored. The good exposure time was 99.2 ks for each RGS.
238:
239:
240: The Optical Monitor (OM) was in Science User Defined
241: mode, with one ``image'' window $5.17\arcmin \times 5.17\arcmin$ centered on NGC 3227.
242: The UVW1 filter, whose effective area peaks near 260 nm, was used throughout.
243: Extraction proceeded in a manner similar to $\S$2--3 of Smith \& Vaughan (2007).
244: Source photons were collected using a 12-pixel radius circle centered on NGC 3227.
245: Emission from the host galaxy was not subtracted; the effect of including
246: such a constant component is that any variability we observe is thus
247: likely a lower limit to the intrinsic level of variability associated with the AGN.
248: The background was extracted from a circular region away from the host galaxy.
249: The resulting net light curve, extracted in 1400 s bins,
250: is displayed in Figure 1.
251:
252:
253: \subsection{{\it RXTE} PCA and HEXTE extraction}
254:
255: To investigate the form of the $>$10 keV continuum and estimate
256: the strength of any Compton reflection hump present, we
257: examined archival
258: Rossi X-ray Timing Explorer {\it RXTE} Proportional Counter Array (PCA;
259: Swank \et\ 1998)
260: and High-Energy X-Ray Timing Experiment (HEXTE; Rothschild \et\ 1998) monitoring data.
261: NGC 3227 was observed in November 1996 for $\sim$260 ks.
262: NGC 3227 was also monitored from 1999 January 2 until 2005 December 4,
263: with one visit every 2--7 days, along with a period of more intensive monitoring
264: approximately four times daily from 2000 April 2 to 2000 June 7
265: (see Uttley \& M$^{\rm c}$Hardy 2005 for further details regarding sampling patterns).
266: Each monitoring snapshot typically lasted 1 ks.
267: There were no observations simultaneous with the 2006 {\it XMM-Newton} observation, so we considered
268: all available data in the public archive in order to estimate average long-term properties.
269: However, data taken from late 2000 to early 2001 (modified Julian day [MJD] 51850--52050),
270: affected by the passage of the compact cloud across the line of sight, were ignored
271: during spectral fitting.
272:
273:
274: PCA data were extracted and screened using standard
275: methods and tools.
276: The 'L7-240' background models, appropriate for faint sources, were used.
277: See e.g., Markowitz, Edelson \& Vaughan (2003)
278: for details on PCA data extraction and background subtraction, the
279: dominant source of systematic uncertainty (e.g., in total broadband count rate)
280: in these data.
281: Counts were extracted from Proportional Counter Units (PCUs) 0--2, 0 and 2, and 2 only for data
282: observed before 1999 December 23, between 1999 December 23 and 2000 May 12, and after
283: 2000 May 12, respectively.
284: Response files were generated for gain epochs 3, 4 and 5 separately
285: (data observed before 1999 March 22, between 1999 March 22 and 2000 May 12, and
286: after 2000 May 12, respectively).
287: As the response of the PCA slowly hardens
288: over time due to the gradual leak of propane into the xenon layers,
289: data within each gain epoch were further split into roughly equal segments, each
290: spanning durations of roughly
291: a couple hundreds days, and separate response files were
292: generated for each segment. Spectra and responses were then added,
293: weighting by exposure times, using the FTOOLS {\sc sumpha} and {\sc addrmf}, respectively.
294: The total exposure time for all PCA data was 842.7 ks.
295:
296:
297: The HEXTE instrument consists of
298: two independent clusters (A and B), each containing four scintillation
299: counters which share a common 1$\degr$ FWHM
300: field of view. Source and background spectra were extracted from each
301: individual \xte\ visit using Science Event data and standard extraction
302: procedures. The same good time intervals used for the PCA data (e.g.,
303: including Earth elevation and SAA passage screening) were applied to the
304: HEXTE data. To measure real-time background measurements, the two HEXTE
305: clusters each undergo two-sided rocking to offset positions, in this
306: case, to 1.5$\arcdeg$ off-source, switching every 32 seconds
307: (16 seconds before 1998 Jan.). No strong
308: contaminating hard X-ray source within 2$\degr$ is evident
309: (e.g., from the {\it RXTE}-PCA or {\it Swift}-BAT all-sky slew survey data
310: available at the HEASARC's online SkyView service).
311: Cluster A data taken between
312: 2004 Dec 13 and 2005 Jan 14 were excluded, as the cluster did not rock
313: on/off source. Detector 2 aboard Cluster B lost spectral capabilities
314: in 1996; these data were excluded from spectral analysis. Cluster A and B
315: data were extracted separately and not combined, as their response matrices
316: differ slightly. Deadtime corrections were applied.
317: All data within each cluster were combined,
318: except for the 1996 data; %%%\footnote{REMOVE THIS FOOTNOTE BEFORE SUBMITTING:
319: %%%%%% As part of my duties as part of the HEXTE team, I'd spent some time
320: %%%%%% investigating HEXTE performance for faint objects (e.g., AGN).
321: %%%%%% I verified that we can go down to 1$\%$ of the background; the HEXTE
322: %%%%%% spectrum in this paper might be the faintest published yet,
323: %%%%%% esp.\ for a Seyfert (to my knowledge).
324: %%%%%% But there was a weird background subtraction error appearing when I
325: %%%%%% summed the 1996 spectrum (16 s rocking) and 1999-2005 spectrum (32 s rocking) together (each indiv.\ spectrum looks fine).
326: %%%%%% Rick R.\ thinks it has something to do
327: %%%%%% with the fact that between successive off-source pointings,
328: %%%%%% one assumes a linear background vs.\ time relation.
329: %%%%%% But the real background count rate decays exponentially.
330: %%%%%% This overestimate of the background is larger
331: %%%%%% for 32-sec rocking than for 16-sec rocking.
332: %%%%%% This may lead to the bkgd subtraction issue, but ONLY for very faint
333: %%%%%% sources ($\lesssim1-2\%$ of the background).}
334: that is, the 1996 (16 s rocking)
335: and 1999--2005 (32 s rocking) spectra were fit separately.
336: We note that the 1996 and 1999--2005 spectra agree well with each other for each cluster,
337: illustrating HEXTE's performance down to source fluxes which are
338: 1$\%$ of the background.
339: Good exposure times for the four HEXTE source spectra
340: were 237.1 ks (cluster A) and 234.2 ks (cluster B)
341: for the 1999--2005 data, and 41.5 (A) and 41.8 ks (B) for
342: the 1996 data. All data were grouped as follows:
343: channels 17--30, 31--39, 40--47, 48--67, 68--79 and $\geq80$
344: were grouped by factors of 2,3,4,5,6 and 10, respectively.
345: The LLAGN/LINER source NGC 3226 is located about 2$\arcmin$ away and is thus
346: in the field of view of both the PCA and HEXTE, but its flux
347: is a factor of $\sim$ 50 fainter than that of NGC 3227 in the 2-10 keV band
348: (Binder et al, in prep.; George \et\ 2001).
349:
350:
351:
352:
353:
354:
355:
356: \section{EPIC pn spectral fits to the Fe K Bandpass}
357:
358:
359: {\sc xspec} (Arnaud 1996) v.11.3.2ag was used for all spectral fitting.
360: All errors below correspond to $\Delta$$\chi^2$ = 2.71
361: (90$\%$ confidence level for one interesting parameter when the errors are symmetric) unless
362: otherwise noted.
363: The abundances of Lodders (2003) were used.
364: NGC 3227's redshift is 0.00386 (de Vaucouleurs et al.\ 1991);
365: a distance of 20.3 Mpc was used (Mould et al.\ 2000)
366: was used to calculate luminosities.
367: In all models, we included a
368: column of neutral absorption fixed at the Galactic column
369: of 2.1$\times$10$^{20}$ cm$^{-2}$ (Dickey \& Lockman 1990, Murphy \et\ 1996).
370:
371: We started by fitting the EPIC-pn data, 4--10 keV only.
372: Residuals to a simple power-law fit
373: ($\chi^2$/$dof$ = 1515/999) showed an obvious Fe K$\alpha$ emission
374: line, but also some interesting structure near 6.0 keV; see Figure 2a.
375: In our fits, we tested two competing models that explain the
376: emission features roughly equally well. As explained below,
377: in both models, narrow emission features are detected at
378: 6.40 keV and 7.37 keV (Fe K$\alpha$ and Ni K$\alpha$ respectively).
379: The models differ in how emission near 6.0 keV is modeled:
380: as a narrow Gaussian emission line, or as a relativistic diskline
381: (hereafter ``GA'' and ``DL'' models).
382:
383: We started with the former case. First, we included
384: a Gaussian component at 6.40 keV to model the
385: Fe K$\alpha$ line. In the best-fit ``GA'' model,
386: the Fe K$\alpha$ line best-fit energy, intensity and $EW$
387: were 6.403$^{+0.011}_{-0.010}$ keV, consistent with neutral Fe.
388: Its intensity and $EW$ were
389: 3.5$\pm$0.4 $\times 10^{-5}$ ph cm$^{-2}$ s$^{-1}$ and 91$\pm$10 eV,
390: respectively. Its width $\sigma$ was 65$\pm$14 eV, which corresponds to a
391: FWHM velocity of 7000$\pm$1500 km s$^{-1}$.
392:
393:
394: Figure 3 shows contour plots resulting from adding a
395: ``sliding Gaussian'' to the ``power-law + Fe K$\alpha$ line'' model
396: to trace the emission residuals after the narrow Fe K$\alpha$ line
397: was modeled. That is, we added a Gaussian with width $\sigma$ fixed at 10 eV
398: and searched over 4--9 keV in units of 0.1 keV, i.e., in
399: steps smaller than the instrumental resolution. The narrow feature
400: near 6.0 keV is clear; other residuals near 7.0 and 7.4 keV
401: also warrant further investigation.
402:
403:
404: To model Fe K$\beta$ emission,
405: we added a Gaussian to the model, keeping the
406: energy centroid fixed and width $\sigma$ tied to that of the Fe K$\alpha$ line.
407: However, the line was not significantly detected according to an $F$-test.
408: The upper limit to the line's $EW$ was 15 eV;
409: the upper limit on the K$\beta$/K$\alpha$ normalization ratio
410: was 0.19 (yielding no useful constraints on the ionization
411: state of the Fe-line emitting material).
412: We included this feature in the rest of the fits for completeness,
413: with intensity fixed to 0.13 times that of the Fe K$\alpha$ line.
414: Residuals to a model consisting of the power-law and
415: the K$\alpha$ and K$\beta$ lines (where $\chi^2$/$dof$ was 1028.3/996)
416: are plotted in Figure~2b.
417:
418:
419: We next added an Fe K edge at 7.11 keV (energy fixed)
420: likely associated with reflection; $\chi^2$/$dof$ dropped to 1017.6/995;
421: residuals to a model including the edge are plotted in Figure~2c.
422: In the best-fit model, the optical depth $\tau$ was 0.05$\pm$0.03.
423:
424:
425: Next, we used a Gaussian component
426: to model an emission line at 7.39$^{+0.08}_{-0.07}$ keV,
427: with width $\sigma$ tied to that of the Fe K$\alpha$ line,
428: consistent with emission from Ni K$\alpha$.
429: Its intensity and equivalent width were
430: $4^{+4}_{-3} \times 10^{-6}$ ph cm$^{-2}$ s$^{-1}$ and
431: $13^{+13}_{-10}$ eV, respectively. $\chi^2$/$dof$ dropped to 1010.2/993;
432: it was significant at 97.3$\%$
433: confidence according to an $F$-test to include this line.
434: Residuals to a model including this line are plotted in Figure~2d.
435:
436:
437: Next, we added a narrow Gaussian emission profile near 6.0 keV
438: to model those residuals;
439: $\Delta$$\chi^2$ was --24.94 for 3 less $dof$,
440: significant at $>$99.99$\%$ confidence according a standard $F$-test.
441: This line is not resolved
442: (width $\sigma$ $<$ 200 eV). Its intensity and
443: $EW$ are $9^{+8}_{-4} \times 10^{-6}$ ph cm $^{-2}$ s$^{-1}$ and
444: 21$^{+19}_{-9}$ eV, respectively.
445: A contour plot of intensity versus energy (confidence levels are for
446: two interesting parameters) is shown in Figure 4.
447: The best-fit line energy is 6.04$^{+0.18}_{-0.04}$ keV
448: inconsistent with 6.24 keV, the lowest photon energy
449: associated with a Compton shoulder scattering feature.
450: Using a Gaussian to model a Compton shoulder, we should expect
451: a centroid energy $>$ 6.24 keV; fixing the Gaussian line energy at 6.24 keV
452: failed to fully model away the 6.0 keV residuals.
453: We performed Monte Carlo simulations to assess the statistical
454: significance of this emission feature, as standard usage of
455: the $F$-test may overestimate the statistical significance of
456: lines at arbitrary energies such as this one (see e.g., warnings by Protassov \et\ 2002).
457: The simulations were conducted following $\S$4.3.3 of Markowitz, Reeves \& Braito (2006); see
458: also $\S$3.3 of Porquet \et\ (2004).
459: The simulations indicated that the likelihood that
460: the 6.0 keV emission line is photon noise is $<$0.1$\%$, i.e., the line is
461: significant at $>$99.9$\%$ confidence.
462:
463:
464: In this best-fit ``GA'' model, $\chi^2$/$dof$ was 983.2/990;
465: $\Gamma$ was 1.57$\pm$0.03. Other best-fit parameters are shown in Table~1.
466: Residuals are plotted in Figure~2e.
467:
468:
469: Next, we tried to model the 6.0 keV feature as the red wing
470: of a relativistic diskline profile.
471: We used a {\sc Laor} model profile (Laor 1991) to
472: model the diskline, with emissivity index
473: $\beta$\footnote{radial emissivity per unit area is quantified as a power law, $r^{-\beta}$}
474: fixed at 3. $\chi^2$/$dof$ was 988.7/989 ($\Delta$$\chi^2$ = --16.8;
475: significant at 99.8$\%$ confidence in an $F$-test)
476: in the best-fit ``DL'' model. The $EW$ of the diskline was 81$^{+42}_{-30}$ eV;
477: its inclination was $<$25$\degr$, and the inner radius was $<$22 $R_{\rm g}$\footnote{1 $R_{\rm g} \equiv GM_{\rm BH}/c^2$}. Visually, there still remained correlated, emission-like residuals from 5.9--6.1 keV,
478: as shown in Figure 2f, which might at first
479: suggest that the 6.0 keV emission feature may be better modeled
480: as a narrow feature than as a red wing of a diskline. However, given the similar values of
481: $\chi^2$/$dof$ and the instrument resolution, it would be difficult to demonstrate
482: that the difference in residuals between the two models is not consistent with
483: photon noise.
484:
485: %%%%%%The 3-Gaussian model seems preferred.
486:
487: \subsection{MOS 1+2 spectral fits to the Fe K bandpass}
488:
489: As a double-check on spectral modeling, we applied
490: our best-fit ``GA'' model to both time-averaged
491: MOS spectra (fit simultaneously, with all line
492: parameters tied between both MOS spectra). The power-law
493: indices and normalizations were allowed to differ between MOS 1
494: and MOS 2. Overall,
495: the MOS and PN yielded qualitatively similar results.
496: A simple power-law fit resulted in $\chi^2$/$dof$ = 771.9/558;
497: data/model residuals are plotted in the top panel of Figure 5.
498: Adding Gaussian components to represent Fe K$\alpha$ and K$\beta$ emission
499: in a manner similar to the pn fits caused $\chi^2$/$dof$ to drop to
500: 546.86/555. Again we performed a ``sliding narrow ($\sigma$ = 10 eV)
501: Gaussian'' test; the resulting contours are plotted in Figure 6 and again indicate
502: emission near 6.0-6.1 keV.
503:
504: Finally, adding a narrow Gaussian (width $\sigma$ fixed
505: at 10 eV) at $6.11^{+0.06}_{-0.13}$
506: keV further improved the fit; $\chi^2$/$dof$ fell to
507: 535.87/553 ($\Delta\chi^2 = -11.0$ for two less $dof$;
508: significant at 99.7$\%$ confidence
509: according to an $F$-test used in the standard way).
510: The best-fit values of the intensity and $EW$ were $6.4 \pm 3.4 \times 10^{-6}$ ph cm$^{-2}$ s$^{-1}$
511: and $13 \pm 7$ eV, respectively.
512: Data/model residuals for the best-fit model are plotted in the bottom panel of
513: Figure 5 (There remain $\sim$10--15$\%$
514: residuals in ratio space near 7.0 and 7.5 keV
515: in Figure 5, but Figure 6 indicates that those features are not detected
516: with high significance.).
517: A contour plot of the intensity of the emission line at 6.11 keV versus line energy
518: is plotted in Figure 7.
519: Again we performed Monte Carlo simulations
520: to gauge the detection significance, finding the line to be detected at 95$\%$ confidence
521: in the MOS 1+2 spectrum i.e., independent of the pn detection.
522: In fact, we can multiply the independent null hypothesis probabilities
523: for the pn and MOS ($<$ 0.001 and 0.05)
524: to yield a ``effective'' null hypothesis probability of $<$ $5\times10^{-5}$.
525:
526:
527: The detection of features at 6.0--6.1 keV in simultaneous pn and
528: MOS 1+2 spectra illustrates the importance of observing with
529: multiple X-ray instruments to attempt to
530: distinguish between narrow and/or weak features which may be intrinsic to the target and
531: features which are artifacts associated with photon noise
532: (even when Monte Carlo simulations are used to gauge the detection
533: significances in individual spectra). The likelihood that a strong emission feature could appear
534: at the same energy in both the pn and the MOS 1+2 spectra and be due to photon noise
535: is likely quite small.
536:
537: The photon indices for the MOS 1 and MOS 2
538: spectra were 1.37$\pm$0.04 and 1.46$\pm$0.04, respectively; the flatter
539: values of the photon index compared to the pn are
540: likely the result of pile-up (MOS spectra extracted using only pattern 0 events
541: has photon indices which were $\sim$0.3 steeper than the MOS pattern 0--12 spectra).
542: Finally, we added an Fe K edge (energy fixed at 7.11 keV);
543: $\chi^2$/$dof$ fell to 517.4/552 for $\tau = 0.09 \pm 0.04$.
544: Emission from Ni K$\alpha$ was not significantly
545: detected. All other parameters were consistent at
546: the 90$\%$ confidence level with those measured by the pn.
547:
548:
549: We also applied the best-fit ``DL'' model to the MOS 1+2 spectrum.
550: $\chi^2$/$dof$ was 504.9/550, with all diskline and
551: Fe K$\alpha$ line parameters consistent with those in the pn fits.
552:
553:
554: %%%%% comparing PN vs MOS hardband only:
555: %%%%% MOS
556: %%%%% HX Gamma = 1.41 (+/-0.04) PN:
557: %%%%% Tau = 0.16 +/- 0.05 PN: 0.05+/-0.03
558: %%%%% Int = 3.2e-5 (+0.4e-5,-0.5e-5) PN: 3.5+/-0.4 e-5
559: %%%%% EW = 70 eV (+9,-11) PN: 91+/- 10
560:
561:
562:
563: %%%%%%%%%%%%%%%%
564: \section{EPIC pn Broadband Spectral Modeling}
565:
566: We started with the best-fit ``GA'' model and
567: included data down to 0.2 keV.
568: In addition to the Galactic column, we included
569: a {\sc zwabs} component at the systemic redshift
570: to model any excess cold absorption
571: associated with e.g., the host galaxy or circumnuclear material,
572: $N_{\rm H,local}$.
573:
574: The residuals (Figure 8a) showed a large soft excess
575: and a large absorption trough at
576: $\sim$0.73 to $\sim$0.92 keV which we identify as
577: an Fe UTA feature. A narrow absorption feature near
578: 1.34 keV is apparent, likely due to Mg XI.
579: A narrow absorption feature near 1.85 keV could be due to Si XIII,
580: but could also be due to calibration uncertainties associated with
581: the instrumental Si K edge.
582:
583:
584: We first modeled the soft excess using a steep power-law with
585: photon index $\Gamma_{\rm SX}$ near 3. Our final, best-fit model (including
586: modeling the warm absorbers and \ion{O}{7} emission line; see below)
587: is henceforth referred to as the ``SXPL (soft X-ray power law)'' model.
588: In the best-fit SXPL model, $\Gamma_{\rm SX}$ was 3.35$^{+0.27}_{-0.10}$.
589: We also added another layer of cold
590: absorption at the systemic redshift, applying it only to the hard X-ray power
591: law; in the best-fit SXPL model, its column density $N_{\rm H,HX}$ was
592: 2.9$^{+0.3}_{-0.8}$$\times$10$^{21}$ cm$^{-2}$.
593: This improved the overall fit substantially
594: ($\chi^2$/$dof$ fell to 3070.3/1730), but the 0.7--0.9 keV trough remained; see
595: Figure 8b. In this model, $\Gamma_{\rm HX}$ was 1.64;
596: we note that forcing the photon indices of
597: the soft and hard power-law components $\Gamma_{\rm SX}$
598: and $\Gamma_{\rm HX}$ to be equal resulted in
599: a very poor fit, with $\chi$/$dof$ = 6057/1731.
600:
601:
602: Next, we used an {\sc xstar} grid, which
603: assumed a turbulent velocity width of 100 km s$^{-1}$,
604: to model a layer of warm absorption with ionization
605: parameter log$\xi$ and column density $N_{\rm H}$.
606: The grid assumed an underlying optical to X-ray spectral energy distribution described as follows:
607: Below 0.001 keV, a power-law component with spectral index $\alpha$ = 1.0 ($\Gamma=2.0$);
608: from 0.001 to 0.04 keV, a power law with $\alpha$ = 0.2 ($\Gamma=1.2$), following e.g.,
609: Elvis \et\ (1994) and Netzer (1996);
610: from 0.04 to 1.0 keV, a steep power law with $\alpha$ = 1.9 ($\Gamma=2.9$; the intrinsic EUV continuum is not well studied,
611: but we assume that the soft excess seen in the {\it XMM-Newton} spectrum extends down to 0.04 keV);
612: and above 1.0 keV, a power law with $\alpha$ = 0.5 ($\Gamma=1.5)$, with the
613: $>$0.04 keV power law indices based on the best-fitting SXPL model.
614: Initially, solar abundances and a zero velocity offset relative to systemic
615: were assumed. Applying one zone of warm absorption to the model,
616: $\chi^2$/$dof$ fell to 2184.1/1728 for log$\xi$ $\sim$ 2.2 and
617: $N_{\rm H}$ $\sim$ 2$\times$10$^{21}$ cm$^{-2}$. As shown in Figure 8c, however,
618: significant data/model residuals remained. A model incorporating
619: two zones of warm absorption successfuly modeled all the absorption-like residuals,
620: including the trough near 0.75 keV; $\chi^2$/$dof$ fell to 1942.1/1726.
621: In this model, the two X-ray absorbers had ionization parameters
622: log$\xi_{\rm lo}$ near 0.3 and log$\xi_{\rm hi}$ near 2.5.
623: We henceforth refer to the two X-ray absorbers as the
624: low-ionization X-ray absorber (though the ionization parameter
625: is still higher than that of the optical/UV lukewarm absorber; see $\S$9),
626: and high-ionization X-ray absorber, respectively.
627:
628: By now, the residuals (see Figure 8d) showed a narrow emission
629: feature near 0.56 keV, likely due to \ion{O}{7}.
630: Adding a narrow Gaussian component
631: at 0.58$\pm$0.01 keV with width $\sigma$ fixed at 1 eV caused $\chi^2$ to drop
632: by 75 for two less $dof$.
633: However, we note that the warm absorbers each predict
634: narrow absorption lines due to \ion{O}{7}, so there are likely large systematic
635: uncertainties associated with the measured intensity of the \ion{O}{7} emission line.
636: This model, with $\chi^2$/$dof$=1866.9/1724,
637: is our best-fit SXPL model. In this model,
638: the column densities and ionization parameters of the
639: low- and high-ionization absorbers are:
640: $N_{\rm H,lo} = 1.0^{+0.3}_{-0.1} \times 10^{21}$ cm$^{-2}$,
641: log$\xi_{\rm lo}$ = $1.45^{+0.16}_{-0.07}$,
642: $N_{\rm H,hi} = 1.8^{+1.2}_{-0.6} \times$10$^{21}$ cm$^{-2}$,
643: and log$\xi_{\rm hi}$ = $2.93^{+0.15}_{-0.09}$.
644: Other best-fit parameters are listed in Table 2.
645: The data/model residuals are shown in Figure 8e.
646: Residuals near 1.8 and 2.2 keV are instrumental
647: (Si K and Au M). Narrow emission-like features near 0.42 and 0.92
648: are tempting to identify as \ion{N}{6} and \ion{Ne}{9} emission, but they are
649: likely narrower than the instrument resolution.
650: There also appear to be some negative residuals near 6.8 keV, close to
651: the energies associated with \ion{Fe}{25} and \ion{Fe}{26}. However,
652: fitting this feature with an inverted Gaussian does not yield
653: significant improvement to the fit; it is likely a spurious feature
654: and will not be discussed further.
655:
656: Note that residuals near the expected O and Fe L3
657: edge energies due to dust (0.53 and 0.71 keV) are fine; we do not require
658: additional edges at those energies (though we revisit this issue with the
659: higher resolution RGS data below).
660: If dust is present in the warm absorber,
661: abundances of O and Fe (and also Si and C)
662: may be expected to be low due to depletion onto grains (see Snow \& Witt 1996).
663: However, the measured O and Fe abundances of the warm absorber
664: are consistent with solar, and left fixed at solar values for
665: the remainder of the paper.
666:
667:
668:
669: In the best-fit SXPL model, the hard X-ray power-law component has
670: a photon index $\Gamma_{\rm HX} = 1.57 \pm 0.02$,
671: a couple tenths lower than ``canonical'' values of 1.8--1.9 in many broad-line
672: Seyferts. We therefore explored the possibility that the observed
673: low value of the photon index is an artifact caused by a
674: yet-unmodeled partial-covering absorbing component
675: with column density $N_{\rm H,PC} \sim 10^{23}$ cm$^{-2}$
676: along the line of sight to the nucleus.
677: Such material is expected to exist within light-days
678: of the black hole, as demonstrated by the 2000-1 obscuration event.
679: Furthermore, such a column is required to explain the origin of the
680: Fe K$\alpha$ line (see $\S$9.4) in the absence of strong Compton
681: reflection (see $\S$6).
682: Modifying the hard X-ray power-law component
683: to be absorbed by a partial-covering neutral
684: absorber with $N_{\rm H,PC}$
685: set to 0.3, 1.0 or 3.0 $\times 10^{23}$ cm$^{-2}$, we find
686: only upper limits to the covering fraction in the range 5--15$\%$.
687: In the best fit models, $\Gamma_{\rm HX}$ was never higher than about 1.6.
688: Partial-covering models are not analyzed further.
689:
690:
691: As before, we applied our best-fit SXPL model to the time-averaged
692: MOS 1+2 spectra as a double-check.
693: The MOS fits, using this model and fitting over 0.4--10 keV,
694: and allowing power-law component parameters to differ between MOS 1 and MOS 2,
695: yielded $\chi^2$/$dof$ = 1406.33/1017, and
696: required slightly higher values of the column density and ionization parameter
697: of the high-ionization warm absorber
698: ($N_{\rm H,hi} = 1.2^{+0.3}_{-0.5} \times 10^{21}$ cm$^{-2}$ and
699: log$\xi_{\rm hi}$ = 3.22$^{+0.04}_{-0.11}$ erg cm s$^{-1}$, respectively)
700: compared to the pn.
701: %%%%% G1 = 1.43 G2 = 1.47 <---pat0-12
702: %%%%% G1 = 1.58 G2 = 1.64 <---pat0
703: $\Gamma_{\rm SX}$ for MOS1 and MOS2 were consistent with
704: that for the pn. $\Gamma_{\rm HX}$ for MOS1 was $1.52 \pm 0.01$, slightly flatter
705: than that for the pn. $\Gamma_{\rm HX}$ for MOS2 was $1.57 \pm 0.02$, consistent with the pn.
706: %%%% Due to pile-up, the measured photon indices for these pattern 0--12
707: %%%% spectra (1.43$^{+0.06}_{-0.02}$ and 1.48$^{+0.06}_{-0.02}$ for
708: %%%% MOS 1 and 2, respectively) were flatter by
709: %%%% 0.15--0.17 compared to spectra extracted from single pattern event data.
710: All other parameters were consistent at
711: the 90$\%$ confidence level with those measured by the pn.
712:
713: \subsection{Alternate parametrizations of the soft excess}
714:
715: We returned to the pn data to consider alternate methods of modeling
716: the soft excess. An alternate parametrization of the soft excess, a blackbody,
717: yields a nearly identical value of $\chi^2$/$dof$ (1870.4/1724).
718: Data/model residuals are plotted in Figure 8f.
719: The blackbody temperature $k_{\rm B}T$ = 83$^{+1}_{-4}$ eV
720: in the pn fit ($77 \pm 3$ eV for the MOS), similar
721: to values found for model fits
722: incorporating blackbody components for most other Seyferts,
723: e.g., $k_{\rm B}T$ $\sim$100--150 eV for many narrow-line Seyfert
724: 1s. It is generally accepted now that a blackbody is a physically unplausible
725: description of the soft excess in AGN.
726: Bechtold et al.\ (1987) were the first to point out
727: that the best-fit blackbody temperatures are generally too high to
728: be associated with accretion disks around supermassive black holes.
729: The consistency of blackbody temperatures across a wide range of
730: Seyfert 1 properties, including black hole mass,
731: furthers suggests that a blackbody is an unphysical
732: parametrization (e.g., Gierli\'{n}ski \& Done 2004).
733: However, the best-fit blackbody temperature in the current fit is
734: consistent with previous results for NGC 3227 (e.g., Komossa \& Fink 1997b).
735: We note that in this model (henceforth denoted simply as the
736: ``BB'' model), $N_{\rm H,HX}$ has dropped to
737: $\leq$ $N_{\rm H,local}$,
738: i.e., the value of $N_{\rm H,HX}$ is model dependent.
739:
740:
741: We next tried fitting the soft excess using
742: the Comptonization model {\sc CompST} (Sunyaev \&
743: Titarchuk 1980). Such components have been used
744: e.g., by Gierli\'{n}ski \& Done (2004) to
745: model the soft excess emission as Comptonization of
746: accretion disk seed photons in
747: a cool ($k_{\rm B}T \sim 0.3$ keV),
748: optically-thick corona, distinct from the optically-thin, hot
749: ($k_{\rm B}T \sim 100$ keV) corona
750: generally thought to be responsible
751: for the hard X-ray power-law component in Seyferts.
752: Physically, such a component could be potentially
753: associated with the transition region between
754: an optically-thick accretion disc and
755: an optically-thin inner corona
756: (Magdziarz et al.\ 1998), or in the hot, ionized surface of
757: the accretion disk (Hubeny \et\ 2001; Janiuk, Czerny \& Madejski 2001).
758: However, similar to the blackbody component, using such Comptonization
759: components to model soft excesses tends to yield a
760: narrow range of best-fit temperatures across a range
761: of Seyfert properties.
762: In our best-fit model (``COMPST'', where $\chi^2$/$dof$ = 1863.4/1723),
763: $k_{\rm B}T = 0.35^{+0.02}_{-0.03}$ keV,
764: and optical depth $\tau = 24^{+2}_{-4}$,
765: similar to previous fits to other AGN
766: (however, see e.g., Vaughan \et\ 2002 for warnings regarding
767: the covariance of optical depth and temperature in this model).
768: Best-fit parameters for the models using
769: the blackbody and {\sc CompST} components are listed in Table 2,
770: though the reader must bear in mind the difficulties
771: each of these models faces in terms of physical plausibility.
772: Data/model residuals are not plotted, as they are virtually identical
773: to those for the SXPL model (Figure 8e).
774:
775: Figure 9 shows unfolded model spectra for the SXPL, BB and COMPST
776: models, demonstrating how the steep soft excess is produced in each
777: (and demonstrating why the value of $N_{\rm H,HX}$ depends on how
778: the soft excess is modeled).
779:
780:
781: Many recent forays into modeling soft excesses have found success
782: using ionized reflection models
783: modified by relativistic smearing, i.e., a blurred
784: ionized reflector (e.g., Crummy \et\ 2006);
785: an origin in atomic features could plausibly explain the consistency
786: of soft excess features across a range of Seyfert black hole masses.
787: Starting with the best-fit SXPL model, we removed
788: the soft X-ray power-law and the 6.04 keV emission line,
789: and added a Ross \& Fabian (2005) reflection
790: component convolved with the kernel associated with
791: relativistic motion around a Kerr black hole
792: (a {\sc Laor} profile). The photon index of the illuminating
793: continuum was tied to that of the hard X-ray power-law.
794: The disk outer radius was fixed at 400 $R_{\rm g}$;
795: the emissivity index was fixed at 3.
796: The 6.4 keV Fe K$\alpha$ and Ni K$\alpha$ line parameters
797: were fixed at the values in the time-averaged spectrum.
798: The best-fit model had $\chi^2$/$dof$ = 2698.4/1721,
799: with large ($\pm$10$\%$) correlated residuals below $\sim$2 keV
800: (see Figure 8g).
801: The best-fit Fe abundance was $<$0.15,
802: and the best-fit value of the reflector ionization
803: parameter $\xi$ pegged at the lower limit of 30 erg cm s$^{-1}$,
804: likely indicating
805: that lower values of $\xi$ may be more appropriate.
806: The low level of Compton reflection $>$ 10 keV in NGC 3227,
807: $R$ $\lesssim$ 0.5,
808: as demonstrated in $\S$6, also suggests that
809: ionized reflection in NGC 3227, if it exists, is not strong.
810: We do not consider this model further.
811:
812:
813: Finally, we tried to model the soft excess
814: using smeared absorption to model
815: ionized absorbing material moving at relativistic velocity,
816: such as a wind launched from the inner accretion disk
817: {\sc swind1} (Gierli\'{n}ski \& Done 2006).
818: Assuming a single power-law to model the intrinsic
819: continuum emission, modified by smeared absorption,
820: yielded a poor fit ($\chi^2/dof$ = 8388/1728).
821: Modeling the intrinsic continuum using two power-laws,
822: as in the SXPL model, and adding smeared absorption
823: plus the low-ionization X-ray absorber,
824: yielded a good fit ($\chi^2/dof$ = 1892/1723), in
825: which a smeared ionized absorber with
826: log$\xi$ $\sim$2.7 erg cm s$^{-1}$ was able to
827: mimic the high-ionization X-ray absorber,
828: given the resolution of the pn.
829: However, analysis of the RGS spectrum ($\S$5)
830: supports the need for
831: both the low- and high-ionization X-ray absorbers, and so
832: we do not consider this model further.
833:
834:
835:
836:
837:
838: \section{RGS modeling of the ionized absorbers and emission lines}
839:
840: We first modeled the continuum of the RGS spectrum
841: with a model of the same form as the SXPL model from the pn fits:
842: a steep soft power-law with $\Gamma_{\rm SX}$ $\sim$ 3.2
843: absorbed by a cold column $N_{\rm H,local} = 1 \times 10^{21}$ cm$^{-2}$,
844: and a hard power-law absorbed by a column $N_{\rm H,HX} \sim 1 \times 10^{22}$ cm$^{-2}$,
845: and dominating the continuum only above $\sim$ 1.5 keV
846: (and with both $N_{\rm H,HX}$ and $\Gamma_{\rm HX}$ very poorly constrained,
847: given that the RGS only covers up to 2 keV).
848: %%%%% Given the RGS/EPIC cross calibration, we do not expect to measure identical photon indices.
849: The residuals to this model ($\chi^2$/$dof$ = 1156/433)
850: are shown in Figure 10.
851:
852: There is a broad absorption feature at $\sim$740--780 eV, which we identify as
853: an Fe M-shell UTA trough. Its energy range (assuming a velocity offset from systemic near zero)
854: indicates that log$\xi$ of the gas where this feature originates must be in
855: the range $\sim$ +0.3 to +1.4, according to an {\sc xstar} table; and absorption is mainly due to
856: species of Fe in the ionization range $\sim$ \ion{}{6}--\ion{}{13}, and likely not due to
857: Fe $\sim$ \ion{}{14}--\ion{}{16} (Gu et al.\ 2006)
858:
859:
860: %%%%%% high-xi UTA's energy indicates gas must have log xi in the range $\sim$ +1.7 to +2.9.
861:
862: %%%%% UTA -- the bottom isn't flat. Sub-dips at 755 and 780 eV
863: %%%%% RGS 2 has dip in eff area at 755 eV, but RGS 1 does not
864:
865: Many narrow absorption features are evident; the tentative identifications (and lab-frame energies) of
866: the most visually prominent ones are labeled in Figure 10 and include
867: \ion{C}{6} Ly$\alpha$ (368 eV), \ion{C}{6} Ly$\beta$ (436 eV),
868: \ion{N}{7} (500 eV), \ion{O}{8} Ly$\alpha$ (653 eV), \ion{O}{7} He$\beta$ (666 eV; perhaps blended with the \ion{N}{7} edge),
869: \ion{O}{7} He$\gamma$ (698 eV), \ion{Fe}{17} 3d--2p (812 eV),
870: \ion{Ne}{9} (922 eV, though this line may
871: be blended with Fe absorption lines at $\sim$ 918--934 eV), \ion{Ne}{10} (1022 eV) and \ion{Mg}{11} (1352 eV).
872:
873: %%%%%O VII $\beta$ (664 eV), Absn lines around 806, 816eV (Fe L 17 lines at 812,819eV(lab)?)
874: %%%%% 695 eV (instrumental fluorine edge)
875:
876: A strong edge at 739 eV due to \ion{O}{7} is not obvious, though it could be
877: blended with the Fe UTA feature. Similarly, an edge at 870 eV
878: due to \ion{O}{8} is not obvious. The lack of a very strong ($\gtrsim 20-30 \%$ drop)
879: \ion{O}{8} edge suggests that any absorbing material containing \ion{O}{8}
880: has a hydrogen column density $N_{\rm H}$ $<$ 10$^{22.5}$ or so.
881: There are also narrow emission features near the energies for
882: the \ion{O}{7} and \ion{N}{6} resonance lines (574 eV and 428 eV, lab-frame); these are discussed
883: further below.
884:
885: We next modeled the two ionized X-ray absorbers discussed in the pn fits above, using the same
886: {\sc xstar} table models, with their respective
887: velocity offsets relative to
888: systemic initially frozen at zero, and with abundances fixed to solar values.
889: $\chi^2$/$dof$ fell to 955.40/429 for values of log$\xi_{\rm lo}$ and log$\xi_{\rm hi}$ near
890: 1.2 and 2.8, respectively, and with column densities for each near 1 $\times$ 10$^{21}$ cm$^{-2}$.
891: However, many of the absorption lines seemed to be blueshifted slightly.
892:
893: We tried various combinations of allowing the absolute redshift of each
894: absorber, $z_{\rm lo}$ and $z_{\rm hi}$, to be thawed from the sytemic value.
895: Thawing $z_{\rm hi}$ only yielded $\chi^2$/$dof$ = 884.87/428 for a best-fit value of
896: $z_{\rm hi}$ near --0.0036 (--0.0075 relative to systemic).
897: Thawing both $z_{\rm hi}$ and $z_{\rm lo}$, but keeping their values tied, yielded
898: $\chi^2$/$dof$ = 887.46/428 for a best-fit value of $z_{\rm hi}$ = $z_{\rm lo}$ =
899: --0.0033 (--0.0072 relative to systemic).
900: Finally, thawing both $z_{\rm hi}$ and $z_{\rm lo}$ and allowing them to vary independently,
901: yielded $\chi^2$/$dof$ = 864.09/427 for best-fit values of
902: $z_{\rm hi}$ = --0.0034 (--0.0073 relative to systemic) and
903: $z_{\rm lo}$ = +0.0016 (--0.0023 relative to systemic).
904:
905:
906: Finally, to this last model (which had the lowest $\chi^2$ value),
907: we added five narrow (width $\sigma$=0.5 eV) Gaussian components to represent emission
908: at the systemic redshift from the \ion{O}{7} $(f)$, $(i)$ and $(r)$, \ion{N}{7},
909: and \ion{N}{6} $(r)$ (note that we model the sum of the two intercombination lines for \ion{O}{7}).
910: Their measured energies, intensities, and fit parameters are listed in Table 3.
911: \ion{N}{6} $(f)$ and \ion{N}{6} $(i)$ emission lines were detected only as upper limits,
912: with intensities $< 6.7 \times 10^{-5}$ and $< 3.0 \times 10^{-5}$ ph cm$^{-2}$ s$^{-1}$, respectively, and so
913: are not listed in Table 3.
914: However, as for the pn fits, we caution that since both warm absorbers require some \ion{O}{7}, \ion{N}{6} and \ion{N}{7} absorption,
915: the emission line intensities likely have large systematic uncertainties.
916: Final parameters for the ionized absorbers in our best-fit SXPL model (with $\chi^2$/$dof$ = 785.7/417) are
917: log$\xi_{\rm lo} = 1.21^{+0.18}_{-0.08}$,
918: $N_{\rm H,lo} = 1.1^{+0.1}_{-0.2} \times 10^{21}$ cm$^{-2}$,
919: log$\xi_{\rm hi} = 2.90^{+0.21}_{-0.26}$, and
920: $N_{\rm H,hi} = 2.4^{+2.0}_{-1.2} \times 10^{21}$ cm$^{-2}$; other model parameters
921: are listed in Table 4. We note that $\Gamma_{\rm SX} = 3.00 \pm 0.25$ and
922: $N_{\rm H,local} = 10.5^{+1.2}_{-1.9} \times 10^{20}$ cm$^{-2}$, both consistent with the best-fit SXPL model
923: for the EPIC data.
924:
925:
926: In this model, the best-fit value for the redshift of the high-ionization absorber is
927: $z_{\rm hi}$ (absolute) = $-0.00302^{+0.00057}_{-0.00080}$ or
928: $z_{\rm hi}$ (relative to systemic) = $-0.00688^{+0.00057}_{-0.00080}$, which corresponds to an outflow velocity
929: of --(2060$^{+240}_{-170}$) km s$^{-1}$ relative to systemic.
930: For the low-ionization absorber,
931: $z_{\rm lo}$ (absolute) = $+0.00246^{+0.00144}_{-0.00064}$ or
932: $z_{\rm lo}$ (relative to systemic) = $-0.00140^{+0.00144}_{-0.00064}$, corresponding to an outflow velocity
933: of --(420$^{+430}_{-190}$) km s$^{-1}$ relative to systemic.
934: Given the best-fit values of the two warm absorbers here and the relative
935: line strengths expected in those cases, the outflow velocity for the
936: high-ionization absorber is constrained primarily by the narrow absorption
937: lines due to \ion{Ne}{9}, \ion{Ne}{10}, and \ion{Mg}{11}; constraining the
938: respective energies of each of these lines separately yielded corresponding
939: outflow velocities relative to systemic of $1300^{+650}_{-300}$,
940: $600\pm600$, and $1800\pm600$ km s$^{-1}$, respectively. Isolating
941: the observed \ion{C}{6} Ly$\beta$, \ion{N}{7} and \ion{O}{8} Ly$\alpha$
942: absorption lines yielded similar velocity offsets ($2700\pm700$,
943: $1800\pm600$ and $1840\pm460$, respectively), though contributions to
944: these lines from both the low- and high-ionization absorber are
945: expected, given the best-fit ionization levels.
946: The only strong absorption features which are unique to only the low-ionization absorber
947: and unambiguously studied here are the (possibly blended) Fe UTA feature and the
948: \ion{O}{7} edge. Models wherein the outflow
949: velocity of the low-ionization absorber is either much closer to systemic,
950: equal to that of the dusty lukewarm UV absorber (tens of km s$^{-1}$; Crenshaw \et\ 2001),
951: or equal to that of the high-ionization absorber are not definitively ruled out at
952: high confidence. Setting the absolute value of $z_{\rm lo}$ equal to systemic,
953: +0.00353 (--0.00033 relative to systemic, corresponding to an outflow
954: velocity of 100 km s$^{-1}$), or $z_{\rm hi}$
955: or +0.003859 and refitting yielded increases in $\chi^2$ of only 9, 6, and 16 respectively.
956: In each case, the outflow velocity for the high-ionization absorber
957: remained virtually unchanged from the best-fitting SXPL model.
958:
959: As an additional confirmation on the {\sc xstar} table results,
960: we measured the $EW$s of selected individual absorption features.
961: The measured $EW$ for the Fe UTA (--(13$^{+4}_{-8}$)) eV is consistent with the prediction from
962: an {\sc xstar} table with $N_{\rm H,lo} = 1.1 \times 10^{21}$ cm$^{-2}$ and log$\xi_{\rm lo}$=1.21.
963: The measured $EW$s for \ion{Ne}{9} ($-5.5\pm2.0$ eV), \ion{Ne}{10} ($-4.4\pm1.6$ eV) and \ion{Mg}{11} ($-4.1\pm2.0$ eV) are
964: in good or reasonable agreement (given the data quality)
965: with the predicted $EW$s from an {\sc xstar} table with
966: $N_{\rm H,hi} = 2.4 \times 10^{21}$ cm$^{-2}$ and log$\xi_{\rm hi}$=2.90 (these lines are not expected
967: to be generated with significantly noticeable $EW$ in the low-ionization absorber).
968: The measured $EW$s for \ion{C}{6} Ly $\beta$ ($-1.0\pm0.9$ eV), \ion{N}{7} (--(1.1$^{+1.3}_{-0.2}$) eV) and \ion{O}{8} ($-3.3\pm0.7$ eV)
969: are in good or reasonable agreement (given the data quality)
970: with the sums of the predicted $EW$s from the two {\sc xstar} tables.
971:
972: Most of the remaining data/model residuals are consistent with RGS calibration uncertainties or
973: are instrumental in nature (Pollack \et\ 2007; observed energies listed): a small dip near 394 eV;
974: large residuals near the O instrumental edge, 524--530 eV;
975: a small dip at 573--577 eV (which further adds to the systematic uncertainties associated
976: with \ion{O}{7} emission lines intensity); %%%%%a small residual near 584 eV;
977: residuals at $\sim$680--695 eV, associated with an instrumental fluorine feature;
978: an apparent edge at 850 eV; and
979: a dip at 937--943 eV.
980: Many of these instrumental features are also seen in other RGS spectra of Seyferts (e.g., Braito \et\ 2007).
981:
982: We rely on the RGS primarily for determination of
983: warm absorber and soft X-ray emission line parameters;
984: all hard X-ray parameters are best constrained by the EPIC data.
985: Still, for the SXPL model, most best-fit parameters (including those for the warm
986: absorbers) were consistent at the $\Delta\chi^2 = 2.71$ significance level between the RGS and the pn;
987: $N_{\rm H,HX}$ and $\Gamma_{\rm HX}$ were the exceptions,
988: and clearly better constrained by the EPIC.
989: Simultaneous RGS/pn fitting using the SXPL model ($\chi^2$/$dof$ = 3220/2161)
990: yielded best-fit values for $\Gamma_{\rm SX}$, $N_{\rm H,local}$
991: and warm absorber column densities and ionization parameters
992: consistent with those measured by the RGS and EPIC-pn separately.
993: Similarly, modeling the continuum in the RGS data
994: using a blackbody or a {\sc CompST} component yielded good fits
995: ($\chi^2$/$dof$ = 805.7/417 and 806.4/416, respectively), with
996: a best-fit blackbody temperature and {\sc CompST} parameters consistent with
997: those measured using only EPIC data, and
998: with no significant changes to the warm absorber or emission line parameters
999: compared to using a soft power-law component to model the soft excess.
1000:
1001: To this point, we have been assuming that the ionized absorbers along the
1002: line of sight have been full-covering. Using the RGS data, we
1003: tested a model wherein only a fraction of the continuum radiation
1004: is absorbed by the ionized absorbers. We found the upper limit to
1005: the fraction of continuum radiation that remains unattenuated by
1006: ionized absorption to be 30$\%$, i.e., a covering fraction of $>70\%$
1007: for the RGS data. For the pn data, the covering fraction is inferred
1008: to be $>93\%$.
1009:
1010: We also tested for the presence of radiative recombination continuum
1011: (RRC) features in the RGS data associated with
1012: H- or He-like ions, expected if the emission lines are associated with gas that is
1013: photo-ionized. Using a {\sc redge} component in {\sc xspec} with
1014: energy fixed and with $k_{\rm B}T$ fixed at
1015: 0.5 keV (arbitrary value), we found upper limits in the range 5--8 eV for
1016: \ion{C}{5}, \ion{C}{6}, \ion{N}{6} and \ion{O}{7} RRCs, upper limits $\leq$ 20 eV for \ion{O}{8} and
1017: \ion{Ne}{9}, and an upper limit near 50 eV for an \ion{Ne}{10}
1018: RRC. A \ion{N}{7} RRC at 667 eV significantly improved the fit, but
1019: that was due to fitting the instrumental fluorine feature,
1020: and therefore is likely not real.
1021: Finally, we note no significant emission from the \ion{Fe}{17} L line (3d--2p)
1022: at 826 eV (upper limit of 1 eV), limiting the possibility that the observed
1023: emission lines are generated via collisional ionization processes.
1024:
1025:
1026: We now discuss the search for X-ray spectral features in the RGS
1027: associated with dust embedded in the warm absorbers.
1028: The expected effects of embedded dust on the X-ray spectrum
1029: are: 1) Edges due to neutral Fe, O, C, Si and Mg
1030: (depending on dust composition) in excess from the absorption expected from gas.
1031: The O edge would be at 531 eV. The three Fe edges are the L3, L2, and L1 edges,
1032: at 707, 721 and 846 eV (assuming oxidized Fe; edges due to pure Fe would be $\sim$3 eV higher),
1033: roughly in a 10:5:1 ratio (e.g., Kortright \& Kim 2000; Bearden \& Burr 1967; Schulz \et\ 2002).
1034: 2) There may be weak, narrow absorption lines due to Fe oxide species, in
1035: particular at 702 eV (e.g., Crocombette \et\ 1995).
1036: 3) Fe, C and O abundances in the warm absorber would be
1037: low due to depletion onto dust grains (Snow \& Witt 1996; Komossa \& Fink 1997a).
1038:
1039: %%%%%% Komossa \& Fink 1997a (327,483) ).
1040:
1041: Near 707 eV and 721 eV (the Fe L3 and Fe L2 edge energies)
1042: the residuals in the RGS spectrum of NGC 3227
1043: do not appear obviously edge-like.
1044: Inserting an edge at any of neutral Fe or O energies does not improve the
1045: fit; upper limits to the optical depths of O, Fe L3 and Fe L2 in excess of the neutral absorption
1046: associated with the gas components modeled above (both local and Galactic)
1047: are 0.17, 0.05 and 0.21, respectively.
1048: In contrast, the Fe L3 edges seen in Cyg X-1 (Schulz \et\ 2002) and MCG--6-30-15 (Lee \et\ 2001)
1049: with the {\it Chandra} HETGS are both $\gtrsim$50$\%$ drop across the edge ($\tau \gtrsim 0.7$).
1050: From the Fe L3 edge in NGC 3227, the implied upper limit on the \ion{Fe}{1} column density is
1051: $\sim10^{17}$ cm$^{-2}$.
1052: Assuming the abundances of Lodders (2003), the corresponding
1053: equivalent hydrogen column density is constrained to be less than
1054: $\sim10^{21.5}$ cm$^{-2}$. This limit is consistent with the
1055: equivalent hydrogen column density implied by optical line absorption (see $\S$1).
1056: The limit on the $EW$ (relative to locally-determined and unabsorbed continuum) of
1057: a narrow (width $\sigma$ = 0.5 eV) absorption line at 702 eV is --0.5 eV.
1058: Finally, thawing the abundances of Fe, O, and Mg in the warm absorbers
1059: (abundances in both absorbers tied) does not yield significant evidence for
1060: deviations from solar abundances.
1061:
1062:
1063: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1064:
1065:
1066:
1067:
1068: \section{Constraining the $>$10 keV continuum with RXTE and Swift-BAT}
1069:
1070: We fit archival {\it RXTE} data and {\it Swift} Burst Alert Telescope (BAT) 9-month survey
1071: data to constrain
1072: the amount of the Compton reflection and constrain any high-energy power-law cutoff.
1073: We did not do simultaneous fitting with the EPIC spectrum,
1074: since the {\it XMM-Newton} and latest {\it RXTE} and BAT data obtained
1075: were separated by a year and since
1076: the only two broadband model components in the range of overlap
1077: are the variable hard X-ray power-law and the hard X-ray absorber,
1078: whose column $N_{\rm H,HX}$ $\lesssim 10^{21}$ cm$^{-2}$ is not well constrained by the PCA.
1079: However, the observed 2--10 keV flux during the 2006 {\it XMM-Newton} observation,
1080: $F_{2-10}$ = 3.5$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$, fell in the approximate range
1081: probed by the {\it RXTE} monitoring; only during MJD $\sim$51850--52050
1082: was $F_{2-10}$ observed to be consistently below 2$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$.
1083: This suggests that {\it XMM-Newton} caught the source in a ``typical'' flux state.
1084:
1085:
1086: We fit the PCA data over the 3.3--30 keV range; the source is
1087: not well detected at higher energies.
1088: HEXTE data below 24 keV were ignored as the responses of
1089: clusters A and B diverge somewhat for faint sources (N.\ Shaposhnikov,
1090: private communication, 2005); there is good agreement above this energy.
1091: HEXTE data above 100 keV were also ignored.
1092: 14--195 keV BAT data were included; four-channel spectra and response matrices
1093: were taken from the BAT 9-month AGN catalog
1094: website\footnote{http://swift.gsfc.nasa.gov/docs/swift/results/bs9mon/}.
1095: A constant coefficient was included in all spectral models
1096: to account for minor normalization offsets between the PCA and HEXTE; its value for
1097: the HEXTE spectra relative to the PCA spectrum was typically 0.8--0.9.
1098: No constant was needed for the BAT spectrum relative to the PCA spectrum.
1099:
1100:
1101: We applied a model consisting of a power-law (initially with a high-energy cutoff
1102: set arbitrarily at $E_{\rm c}$ 500 keV),
1103: a Gaussian emission line at 6 keV to model Fe K$\alpha$ emission,
1104: and a component to model Compton reflection from neutral material, using
1105: {\sc pexrav} (Magdziarz \& Zdziarski 1995), all modified by
1106: a cold column $N_{\rm H,HX}$ constrained to be less than $6 \times 10^{21}$ cm$^{-2}$.
1107: The {\sc pexrav} component's inclination was fixed at 45$\degr$, solar Fe abundances were assumed,
1108: and the high energy cutoff was tied to that of the power-law.
1109: The best-fit model had $\chi^2$/$dof$ = 297/111, the
1110: photon index $\Gamma_{HX}$ was 1.66$\pm$0.01, and the
1111: value of the Compton reflection strength
1112: $R$\footnote{$R \equiv \Omega/2\pi$, where $\Omega$ is the solid angle subtended by the reflector}
1113: was $0.16^{+0.08}_{-0.02}$. Allowing the high-energy cutoff to be free and re-fitting,
1114: $\chi^2$/$dof$ fell to 222/110 (significant at 5.6$\sigma$ confidence according to an $F$-test).
1115: $\Gamma_{HX}$ was 1.63$\pm$0.02, $R$ was 0.40$\pm$0.07, and $E_{\rm c}$ = 90$\pm$20 keV.
1116: The cold column $N_{H,HX}$ was $6.0^{+0}_{-0.8} \times 10^{21}$ cm$^{-2}$
1117: (uncertainty pegged at upper limit).
1118: The residuals to the best-fit models with and without a high-energy cutoff
1119: are shown in Figure 11\footnote{The presence of the high-energy cutoff is also inferred from fitting PCA+HEXTE
1120: only: excluding the BAT data, and refitting, we find that including a high-energy cutoff
1121: still significantly improves the fit, with best-fit values consistent with the
1122: PCA + HEXTE + BAT fit, though the statistical uncertainty on $E_{\rm c}$ is 1.8 times larger.
1123: Ignoring the HEXTE data, and fitting PCA + BAT only,
1124: the best-fit values of $\Gamma_{HX}$, $R$, and $E_{\rm c}$ are 1.62$^{+0.02}_{-0.04}$,
1125: 0.53$^{+0.19}_{-0.08}$ and $65^{+10}_{-25}$ keV, respectively.}.
1126:
1127:
1128: Because the {\it RXTE} and {\it XMM-Newton}
1129: observations were not simultaneous, it is not clear whether the
1130: slightly flatter value of $\Gamma_{\rm HX}$ for the pn in the SXPL model (1.57$\pm$0.02)
1131: compared to that for the time-averaged {\it RXTE} spectrum
1132: is due to source variability or associated with systematic differences in
1133: modeling spectra taken with two instruments with very different energy resolution.
1134: For example,
1135: the average Fe K$\alpha$ line flux as measured by the PCA was $9 \pm 1 \times 10^{-5}$ ph cm$^{-2}$ s$^{-1}$,
1136: but this is not automatically
1137: an indication of line variability between the {\it RXTE} and {\it XMM-Newton}
1138: observations given the differences in PCA and EPIC resolution, the
1139: systematic uncertainties associated with fitting such a complex model to the low-resolution
1140: PCA data, and possible uncertainties associated with PCA/pn cross-calibration.
1141: %%%%% \footnote{See warnings by Yaqoob \et\ 2003 and in an {\it ASCA} Guest Observatory Facility memo at http://heasarc.gsfc.nasa.gov/docs/asca/calibration/3c273\_results.html.}.
1142: All uncertainties listed here are
1143: one-dimensional statistical errors only and do not account
1144: for systematic uncertainties associated with correlations between
1145: parameters or for systematic instrumental uncertainties associated
1146: with observing faint objects.
1147: However, we can safely conclude that the value of $R$ must be relatively low:
1148: Figure 12 shows contour plots for $E_{\rm c}$ and $R$ as a
1149: function of $\Gamma_{HX}$ from the joint {\it RXTE}+{\it Swift}-BAT fit,
1150: demonstrating that values of $R$ $>$ 0.6 are rejected at high confidence.
1151:
1152:
1153:
1154: \section{Time- and Flux-Resolved Spectral Fits}
1155:
1156: We investigated the temporal behavior of the X-ray spectrum of NGC 3227 on
1157: a range of time scales. Any rapid variation in ionization parameter
1158: would rule out models where the warm absorber is spatially extended
1159: and/or located very far from the black hole;
1160: tracking ionization responses to continuum flux variations
1161: can yield constraints on the location of the warm absorbers
1162: (e.g., Krongold \et\ 2007).
1163: Similarly, any rapid variation in the strength of an Fe K emission feature
1164: (e.g., Tombesi \et\ 2007) would suggest an origin close to the X-ray illuminating source.
1165: We used the {\it XMM-Newton} spectrum to study variability on time scales $\lesssim$ 1 day
1166: (hereafter ``short'' time scales);
1167: we used the {\it RXTE} intensive monitoring from 2000 April to June to investigate
1168: variability on time scales from $\sim$2 weeks to 2 months (hereafter ``medium'' time scales),
1169: and the 1999--2005 monitoring to investigate variability on time scales from
1170: months to several years (``long'' time scales'').
1171:
1172: \subsection{{\it XMM-Newton} time-resolved spectral fits}
1173:
1174: We investigated the $<$1-day time-resolved spectral behavior
1175: of NGC 3227 during the {\it XMM-Newton} observation
1176: by first splitting the EPIC-pn data into five segments
1177: of equal duration (good exposure time 18 ks); shorter durations would not have yielded
1178: adequate photon statistics in narrow features.
1179: In the Fe K bandpass, spectral fitting revealed several
1180: ``candidate'' narrow transient emission features between 4.5 and 5.2 keV,
1181: each with $\Delta$$\chi^2$ in the range --11 to --14.
1182: Monte Carlo simulations (see $\S$3) showed that each of these emission
1183: features were significant at 92--97$\%$ confidence.
1184: However, when one takes into consideration the number of segments searched over
1185: (Vaughan \& Uttley 2008),
1186: the significance levels drop to 60--85$\%$ confidence; these features are
1187: not discussed further. Considering the broadband spectrum, all
1188: parameters for the neutral and ionized absorbers and the \ion{O}{7}
1189: emission line were consistent with those found in the
1190: time-averaged spectrum.
1191: Similarly, flux-resolved spectral analysis,
1192: performed on high- and low-flux pn spectra (extracted above and
1193: below a 0.2--12 keV pn count rate of 11.0 ct s$^{-1}$, respectively),
1194: yielded no significant evidence for variability of the neutral or
1195: ionized absorbers, the 6.4 keV Fe K$\alpha$ line, or the \ion{O}{7} line.
1196: The 6.04 keV line was not significantly detected with these
1197: short exposures.
1198:
1199: However, we noticed in these time-resolved fits that
1200: the normalization of the soft excess increased by $\sim$25$\%$
1201: from the first 20 ks segment to the second segment.
1202: Similar results were obtained when modeling the soft excess as
1203: a steep power-law, a blackbody, or a {\sc CompST} component.
1204:
1205: The rapid (ks and shorter) variations seen simultaneously in both the
1206: hard and soft X-ray light curves (Figure 1) are similar to those routinely
1207: detected in other Seyferts; the associated variability
1208: mechanisms will be explored in depth in a future paper (Ar\'{e}valo
1209: et al., in prep) and are not discussed further here.
1210: Here, we are concentrating on the soft X-ray band's
1211: relatively slower and distinct trend.
1212: Since this trend occurs over tens of ks, we will refer to it
1213: as "rapid" for the remainder of this paper.
1214:
1215:
1216: We repeated the time-resolved analysis, dividing the data into
1217: ten segments of equal duration (9 ks good exposure time), fitting
1218: the SXPL model, and
1219: keeping all warm and cold absorber, \ion{O}{7} line
1220: and Fe emission line parameters frozen at their time-averaged values.
1221:
1222: The resulting light curves of $\Gamma_{\rm HX}$, $\Gamma_{\rm SX}$
1223: and the normalizations of the SXPL and HXPL components
1224: are plotted in Figure 13. The SXPL normalization
1225: seems to track the light curve of 0.2--1 keV flux in Figure 1.
1226: To further illustrate the rapid variability of the soft excess, the
1227: data for segments 1 (the lowest soft excess flux bin), 2, and 6
1228: (the highest soft excess flux bin) are plotted in Figure 9.
1229:
1230: Finally, fractional variability amplitudes ($F_{\rm var}$
1231: see Vaughan \et\ 2003 for a definition) were calculated for
1232: 16 energy bands. The resulting $F_{\rm var}$ for the entire duration, for
1233: the first 20 ks, and for the final 80 ks are plotted in Figure 14,
1234: further illustrating the rapid variability in the soft excess.
1235: Further detailed variability analysis of the variable X-ray continuum,
1236: including coherence, intra-X-ray time lags and power spectral
1237: density function measurement, will be presented in a future
1238: paper (Ar\'{e}valo et al., in prep).
1239:
1240:
1241:
1242: \subsection{{\it RXTE} time-resolved spectral fits}
1243:
1244:
1245: Time-resolved spectral fitting of the {\it RXTE} data
1246: closely followed Markowitz, Edelson \& Vaughan (2003).
1247: Bin sizes were chosen to optimize the
1248: trade off between minimizing the uncertainties on
1249: Fe line flux and maximizing the number of bins.
1250: This yielded 5 bins, each of duration 13 days, for the medium
1251: time scale. On the long time scale, start/stop times were
1252: chosen to avoid 60-days gaps where {\it RXTE} did not observe the
1253: source due to sun-angle constraints, with most
1254: durations roughly 100 days before loss of PCU0 on 2000 May 12,
1255: and 150 days afterward, yielded 16 bins.
1256: Response matrices were generated separately for each segment.
1257:
1258: We used a model consisting of a power law component,
1259: a Gaussian component to model Fe K emission, and
1260: a {\sc pexrav} component to model Compton reflection,
1261: all absorbed by cold material (with a column density $N_{\rm H, HX}$
1262: constrained to be $ < 6 \times 10^{21}$ cm$^{-2}$,
1263: unless the bins included data taken during the 2000-1 obscuring event).
1264: In the best fit models, the best-fit values of $R$, Fe line rest-frame
1265: energy centroid, and Fe line width $\sigma$ were
1266: $0.4^{+0.2}_{-0.1}$ ($0.2 \pm 0.1$),
1267: $6.26^{+0.10}_{-0.11}$ ($6.27^{+0.05}_{-0.04}$) keV, and
1268: $0.50^{+0.32}_{-0.14}$ ($0.33^{+0.09}_{-0.08}$) keV
1269: in the medium (long) time scale fits, respectively.
1270: To determine which parameters to leave free during each
1271: time-resolved fit, we first fit all segments simultaneously with
1272: all parameters tied, and thawed one parameter at a time,
1273: testing for significant improvement in the fit according to an $F$-test.
1274: We found it was significant to thaw the power law normalization, $\Gamma_{\rm HX}$,
1275: Fe line intensity $I_{\rm FeK\alpha}$, and $N_{\rm H,HX}$ in the individual,
1276: time-resolved fits.
1277:
1278: %%%% The PCA has crappy energy resolution, but XMM has shown that the narrow line dominates,
1279:
1280: %%%%% MED:
1281: %%%%% R = 0.4 (+0.2,-0.1)
1282: %%%%% Ener = 6.26 (+0.10,-0.11) (rest-frame)
1283: %%%%% sigma = 0.50 (+0.32,-0.14) keV
1284:
1285: %%%%% LONG
1286: %%%%% R = 0.2 (+/-0.1)
1287: %%%%% Ener = 6.27 (+0.05,-0.04) (rest-frame) par12
1288: %%%%% sigma = 0.33 (+0.09,-0.08) keV par13
1289:
1290:
1291:
1292: Errors for $I_{\rm FeK\alpha}$ and $\Gamma_{\rm HX}$
1293: were derived using the point-to-point
1294: variance\footnote{See $\S$3.3 of Markowitz, Edelson \& Vaughan
1295: (2003) for further details and the definition of the point-to-point variance}.
1296: Errors on values of $F_{2-10}$ within a time bin
1297: were determined from the 2--10~keV continuum light curve,
1298: using the mean flux error on the $\sim$1~ks exposures in that time bin.
1299: Figure 15 shows the resulting light curves for $F_{2-10}$,
1300: $I_{\rm FeK\alpha}$, and $\Gamma_{\rm HX}$.
1301:
1302:
1303: Fractional variability amplitudes $F_{\rm var}$ were calculated.
1304: For the medium time scale, $F_{\rm var}$ for
1305: the $F_{2-10}$ and $I_{\rm FeK\alpha}$ light curves were
1306: 38.4$\pm$0.3$\%$ and 22.7$\pm$10.3$\%$, respectively,
1307: although we caution that with only 5 data points, these measured
1308: values of $F_{\rm var}$ are likely not highly reliable.
1309: For the long time scale,
1310: $F_{\rm var}$ for $F_{2-10}$ and $I_{\rm FeK\alpha}$ were
1311: 32.7$\pm$0.7$\%$ and 20.0$\pm$4.3$\%$.
1312: Qualitatively, these
1313: results are consistent with what Markowitz, Edelson
1314: \& Vaughan (2003) found for a small sample of Seyferts:
1315: the Fe line does not vary as strongly as the continuum flux.
1316:
1317: %%%%%%%%%%%%%%%%%%% new paragraph.
1318:
1319: Inspection of the $F_{2-10}$ and $I_{\rm FeK\alpha}$
1320: light curves would seem to indicate, to the human eye at least,
1321: similar variability trends. However, with so few points
1322: (especially for the medium time scale),
1323: such a statement is not highly significant.
1324: On the medium time scale, the best-fit linear relation is
1325: $I_{\rm FeK\alpha} = 2.05 \pm 1.31 \times F_{2-10} + 4.1 \pm 3.1$,
1326: with $I_{\rm FeK\alpha}$ in units of 10$^{-5}$ ph cm$^{-2}$ s$^{-1}$ and
1327: $F_{2-10}$ in units of $10^{-11}$ erg cm$^{-2}$ s$^{-1}$.
1328: Figure 16 shows the zero-lag correlation plot of
1329: $I_{\rm FeK\alpha}$ as a function of $F_{2-10}$,
1330: with the best-fit linear relations plotted as dashed lines.
1331: The zero-lag Spearman rank correlation coefficient\footnote{calculated at http://www.wessa.net/rankcorr.wasp} is 0.700; the corresponding null hypothesis
1332: probability (the probability that the
1333: correlation could arise from randomly chosen data points)
1334: is 16$\%$; i.e., the correlation is significant only at 84$\%$
1335: confidence.
1336:
1337: For the long time scale data, the best-fit linear relation is
1338: $I_{\rm FeK\alpha} = 1.44 \pm 0.30 \times F_{2-10} + 3.5 \pm 1.2 $.
1339: The zero-lag Spearman rank correlation coefficient is
1340: r=0.739, significant at 99.6$\%$ confidence.
1341: We searched for lags using an Interpolated Cross Correlation Function
1342: (ICF; Gaskell \& Peterson 1987, White \& Peterson 1994), with errors determined
1343: using the bootstrap method of Peterson et al.\ (1998).
1344: The ICF peak correlation coefficient $r_{max}$=0.854
1345: was reached at a delay of 75 $\pm$ 690 days ($F_{2-10}$
1346: leading $I_{\rm FeK\alpha}$), i.e., consistent with
1347: zero lag.
1348:
1349: %%%%%%%%%%%%%%%%%%%%%%%
1350: With such few data points and low correlation
1351: coefficients on both medium and long time scales,
1352: any claim of a correlation must be deemed tentative at best.
1353: Cross correlation analysis is further complicated by the red-noise nature
1354: of the continuum (and likely line) light curves. Specifically,
1355: cross-correlation between two unrelated red-noise
1356: light curves can randomly yield spurious correlations
1357: with higher than expected values of the
1358: correlation strength $r_{\rm max}$ (e.g., Welsh 1999).
1359: Lags where $r_{\rm max}$ is not very close to 1.0
1360: should thus be treated with skepticism.
1361: Additional monitoring, spanning much longer durations and
1362: wider ranges in both $F_{2-10}$ and $I_{\rm FeK\alpha}$ and encompassing
1363: additional upward/downward trends in the light curves
1364: of both $F_{2-10}$ and $I_{\rm FeK\alpha}$ are required to
1365: critically test for any significant correlation.
1366: We therefore conclude that there is, at best, tentative
1367: evidence from the {\it RXTE} time-resolved spectral fits for a significant
1368: fraction of the the Fe line intensity to respond to continuum
1369: variations on time scales shorter than 700 days.
1370: The lack of strong hard X-ray variability during the
1371: {\it XMM-Newton} observation ($F_{\rm var} = 8.5\pm0.2\%$ for the
1372: 3--10 keV light curve, binned to 600 s)
1373: means we cannot draw any conclusions about the response
1374: of the line on $\lesssim1$ day time scales.
1375:
1376: %%%%%%%%%%%%%%%%%%%
1377:
1378:
1379:
1380: \section{X-ray/UV continuum light curve correlations}
1381:
1382:
1383: %%%%%Fvar for optical is 2.7 +/- 1 $\%$
1384:
1385: The OM light curve of NGC 3227 displayed in Figure 1 shows a
1386: $\sim$10$\%$ increase across the observation, with an RMS of
1387: 2.8$\%$. Rapid optical/UV variability in Seyferts on time scales
1388: of $\lesssim$ 1 day is relatively rare. Of the sample of 8
1389: Seyferts examined by Smith \& Vaughan (2007), NGC 3783 displayed
1390: the strongest variability on $\lesssim$ 1 day time scales, with
1391: an RMS of 2.9$\%$ in the UVW2 filter; there were only
1392: 3 additional observations using either the UVW2 or U filter
1393: where the RMS was $> 1.8\%$.
1394: The current {\it XMM-Newton} data would thus seem to indicate that
1395: NGC 3227 is towards the top of the list of Seyferts which display
1396: strong UV continuum variability on $\lesssim$ 1 day time scales.
1397: In NGC 3227, the X-rays are much more strongly variable on these
1398: time scales (RMS for the 0.2--1 and 3--10 keV light curves
1399: were 15.8$\%$ and 8.9$\%$, respectively), a result similar to
1400: what Smith \& Vaughan (2007) found for their sample.
1401:
1402: Visually, it is tempting to connect the gradual brightening
1403: observed in the OM light curve with that observed simultaneously
1404: in the soft X-ray band during the {\it XMM-Newton} observation.
1405: However, the rise in soft X-ray flux ($\sim$40$\%$) is much greater
1406: than that observed in the UV band ($\sim$10$\%$).
1407: We calculated ICFs and Discrete Correlation Functions (DCF; Edelson \& Krolik 1988) between the UV and soft X-ray light curves (the latter rebinned to 1400 s); the results are
1408: plotted in Figure 17, along with 90 and 95$\%$ confidence limits
1409: from the Bartlett method. We found no significant correlation:
1410: $r_{\rm max}$ peaks at only $\sim$0.5; there are no ``bends'' or
1411: multiple trends in the OM light curve to drive a correlation.
1412: The warning regarding cross-correlation analysis performed on two
1413: unrelated red-noise light curves bear repeating:
1414: beware of spuriously high values of $r_{\rm max}$ (e.g., Welsh 1999)
1415: and treat values of $r_{\rm max}$ which are not very close to 1.0
1416: with skepticism. The suggestion of a correlation between the
1417: soft X-ray band and the UV band, while visually tempting, is
1418: thus speculative at best.
1419:
1420:
1421: \section{Discussion}
1422:
1423: We have analyzed a $\sim$100 ks {\it XMM-Newton}
1424: long-look of NGC 3227, observed in December 2006.
1425: In both the EPIC-pn and RGS spectra, we have modeled the ionized
1426: X-ray absorber using two components, with very similar column
1427: densities, and
1428: with ionization parameters log$\xi$ near 1.2 and 2.9 (in the RGS spectrum),
1429: With the RGS, we have constrained
1430: the outflow velocity of the high-ionization absorber to be
1431: --(2060$^{+240}_{-170}$) km s$^{-1}$ relative to systemic.
1432: The best estimate for the outflow velocity relative to systemic for the low-ionization
1433: absorber is --(420$^{+430}_{-190}$) km s$^{-1}$,
1434: though a wider range of possible outflow velocities
1435: cannot be ruled out at very high confidence.
1436: In $\S$9.1, we discuss further details of these ionized outflows,
1437: including exploring connections to the low-velocity, very low ionization,
1438: dusty lukewarm absorber observed in the UV.
1439:
1440: The steep soft excess, which does not seem be to affected by the same neutral material
1441: obscuring the hard X-ray continuum, is shown with the EPIC pn to be rapidly variable
1442: compared to other Seyferts on $\lesssim$ 1 day time scales; its
1443: normalization increases by $\sim$25$\%$ in $\sim$20 ks.
1444: The UV continuum light curve during the {\it XMM-Newton}
1445: observation also shows a relatively strong
1446: increase compared to most other Seyferts on $\lesssim$ 1 day time scales.
1447: Possible origins for these behaviors are discussed in $\S$9.2.
1448:
1449:
1450: The hard X-ray power-law continuum in both the {\it XMM-Newton} pn and MOS
1451: fits and in the PCA + HEXTE + BAT fits is somewhat low, $\Gamma_{\rm HX} \sim 1.5-1.6$.
1452: Using accumulated {\it RXTE} PCA and HEXTE archival data from
1453: over 6 years of monitoring, plus {\it Swift}-BAT 9-month survey data,
1454: we constrain the strength of the Compton reflection component $R$
1455: to be $\lesssim$ 0.5. We also find the first evidence
1456: for a high-energy continuum cutoff in this source, at 90$\pm$20 keV.
1457: The high-energy continuum is discussed further in $\S$9.3.
1458:
1459: The narrow Fe K$\alpha$ emission line at 6.4 keV is resolved in the pn
1460: spectrum. From time-resolved spectral fits to the {\it RXTE}-PCA
1461: data, we find tentative evidence for a significant fraction
1462: of the Fe K line flux to track variations in
1463: the continuum flux $F_{2-10}$ on time scales of $<$700 days. The 6.4 keV
1464: Fe K$\alpha$ line properties are discussed further in $\S$9.4.
1465:
1466:
1467: In addition, we find significant evidence for emission near 6.0 keV
1468: in both the pn and MOS 1+2 spectra.
1469: It is modeled approximately equally well as a narrow
1470: emission line or as the red wing to a relativistically-broadened Fe K emission line,
1471: as discussed in $\S$9.5.
1472:
1473:
1474:
1475:
1476: \subsection{Overview of the X-ray absorbers}
1477:
1478: In the {\it XMM-Newton} RGS spectrum, we model two zones of
1479: absorption, with log$\xi_{\rm lo} = 1.21^{+0.18}_{-0.08}$ and log$\xi_{\rm hi} = 2.90^{+0.21}_{-0.26}$, and with very
1480: similar column densities,
1481: $N_{\rm H,lo} = 1.1^{+0.1}_{-0.2} \times 10^{21}$ cm$^{-2}$ and
1482: $N_{\rm H,hi} = 2.4^{+2.0}_{-1.2} \times 10^{21}$ cm$^{-2}$.
1483: Other AGN have been reported to host multiple
1484: warm absorber components with different
1485: ionization parameters but similar hydrogen column densities
1486: (see e.g., Blustin \et\ 2005).
1487: The blueshift relative to systemic of the high-ionization absorber was significantly constrained, yielding
1488: an outflow velocity relative to systemic of --(2060$^{+240}_{-170}$) km s$^{-1}$.
1489: The best-fit blueshift relative to systemic for the low-ionization absorber corresponded to
1490: an outflow velocity relative to systemic of --(420$^{+430}_{-190}$) km s$^{-1}$;
1491: however, as this is primarily constrained by only two (possibly blended) absorption features,
1492: models with blueshifts ranging from zero to identical to that for
1493: the high-ionization absorber were not clearly rejected.
1494: It is therefore not clear if the two zones are physically and kinematically distinct or
1495: if there exists a single ionized absorber spanning a broad
1496: range of ionization levels, a possibility to consider if
1497: the ionized X-ray absorbing gas is spatially extended.
1498: For instance, Gon\c{c}alves \et\ (2006) model the warm absorbing layers in NGC 3783
1499: as a single constant-density gas component in pressure equilibrium.
1500:
1501:
1502: To estimate the distance $r$ between the central black hole
1503: and the outflowing gas in NGC 3227, we can use $\xi = L_{\rm ion}/(nr^2)$,
1504: where $n$ is the number density, and $L_{\rm ion}$ is the 1--1000 Ryd
1505: illuminating continuum luminosity. We estimate the maximum possible
1506: distance to the material by assuming that the thickness
1507: $\Delta$$r$ must be less than the distance $r_{\rm max}$.
1508: The column density $N_{\rm H}$ = $n$$\Delta$$r$,
1509: yielding the upper limit $r_{\rm max}$ $<$ $L_{\rm ion}$/($N_{\rm H}$$\xi$).
1510: We estimate the unabsorbed 1--1000 Ryd flux to be
1511: $\sim 3 \times 10^{-10}$ erg cm$^{-2}$ s$^{-1}$, which
1512: corresponds to $L_{\rm ion} \sim 1.5 \times 10^{43}$ erg s$^{-1}$
1513: (using a distance of 20.3 Mpc, and assuming
1514: $H_{\rm o}$ = 70 km s$^{-1}$ Mpc$^{-1}$ and $\Lambda_{\rm o}$ = 0.73).
1515: Assuming two physically distinct warm absorbers,
1516: for the high-ionization X-ray absorber, this yields
1517: $r_{\rm max}$ = 3.6 pc. For the low-ionization X-ray absorber,
1518: the constraints are even weaker: $r_{\rm max}$ = 150 pc.
1519: These constraints are much poorer compared to that derived by Gondoin \et\
1520: (2003) for a one-zone absorber, $r_{\rm max} = 0.45$ pc, as those authors found a higher value
1521: for the column density and a lower value for the ionizing flux.
1522:
1523: However, we can also derive a minimum radial distance from the black hole $r_{\rm min}$ for the winds
1524: via the requirement for the outflow velocity $v$ to be
1525: greater than the escape velocity: For the high-ionization X-ray absorber, $r_{\rm min} = (c^2/v^2)R_{\rm g} =
1526: 17000 R_{\rm g} = 40$ light-days, placing it outside
1527: both the BLR and the inner radius of the dust as
1528: reverberation-mapped by Suganuma \et\ (2006).
1529:
1530:
1531: %%%%%%%%%%%%%%%%%%%%%%%%%
1532:
1533: Under the assumption that the gas is in equilibrium and that
1534: the outflow velocity is a constant, the mass outflow rate
1535: \.{M}$_{\rm out}$ of the X-ray absorbers can be derived via
1536: conservation of mass:
1537: \.{M}$_{\rm out}$ = $\Omega$$n$$r^2$$v$$m_{\rm p}$,
1538: where $v$ is the outflow velocity, $m_{\rm p}$ is the proton mass,
1539: and $\Omega$ is the covering fraction.
1540: We then substitute $n$$r^2$ = $L_{\rm ion}$/$\xi$.
1541: Assuming $\Omega$ = 0.3 (arbitrary), we find, for the high-ionization absorber,
1542: \.{M}$_{\rm out}$ $\sim 2 \times 10^{24}$ gm s$^{-1}$ $\sim$ 0.03 $\Msun$ yr$^{-1}$.
1543: For the low-ionization absorber, assuming $v = 420$ km s$^{-1}$,
1544: \.{M}$_{\rm out}$ $\sim 2 \times 10^{25}$ gm s$^{-1}$ $\sim$ 0.3 $\Msun$ yr$^{-1}$.
1545: We note that the actual outflow rate should be lower if there is an
1546: extreme degree of collimation along the line of sight.
1547: We can compare the outflow rate
1548: to the inflow accretion rate \.{M}$_{\rm acc}$ using
1549: $L_{\rm bol}$ = $\eta$\.{M}$_{\rm acc}$$c^2$,
1550: where $\eta$ is the accretion efficiency parameter, typically 0.1.
1551: Woo \& Urry (2002) estimate $L_{\rm bol}$ for NGC 3227
1552: to be $7.2 \times 10^{43}$ erg s$^{-1}$, which means an accretion rate
1553: relative to Eddington, $L_{\rm bol}/L_{\rm Edd}$, of 1.4$\%$.
1554: For $\eta=0.1$, \.{M}$_{\rm acc} = 0.01 \Msun$ yr$^{-1}$.
1555: The kinetic power associated with the outflow component,
1556: estimated as \.{M}$_{\rm out}$$v^2$, is thus in the approximate range
1557: $10^{40-41}$ erg s$^{-1}$.
1558:
1559:
1560: The outflow mass rate and kinetic energy are rough estimates only,
1561: but it does appear likely that the outflows represent at least a large fraction
1562: of the AGN's accretion rate. If the X-ray ionized outflows
1563: are long-lived, then a sustained feeding of the black hole would be difficult.
1564:
1565:
1566:
1567:
1568:
1569: %%%%%%%%%%%%%%%%%%%%%%%%%
1570:
1571: \subsubsection{Helium-like emission line diagnostics}
1572:
1573: It is not obvious whether the emission lines detected in the
1574: RGS spectrum are due to collisional ionization or
1575: photo-ionization. We do not detect a strong \ion{Fe}{17} L
1576: 3d--2p $^1$P$_1$ line at 826 eV, an indicator of collisional ionization.
1577: However, RRC lines, indicators of photo-excitation, are not obvious, either.
1578:
1579:
1580: We can attempt to
1581: use diagnostics associated with the helium-like O and N emission triplets.
1582: In a collisionally- (photo-) ionized plasma, the resonance (forbidden)
1583: line dominates. However, in NGC 3227, there is not
1584: significant dominance of one of these lines in either emission triplet.
1585: The presence of helium-like absorption lines may be a factor:
1586: the emission and absorption lines are
1587: not completely separated in energy given the modeled blueshift of
1588: the absorption lines, the RGS resolution and the signal/noise ratio
1589: of the RGS data. This fact also complicates our use of the density
1590: indicator $R \equiv f/i$ and the temperature indicator $G \equiv (f+i)/r$
1591: (e.g., Porquet \& Dubau 2000)
1592: where $f$, $i$ and $r$ are the intensities of the
1593: forbidden, intercombination, and resonance lines, respectively
1594: ($i$ is the summed intensity of both intercombination lines).
1595: As $G$ depends on $r$, it can indicate whether photo-ionization
1596: dominates or whether a photo-/collisional-ionization hybrid is
1597: applicable (Porquet \& Dubau 2000).
1598: Porter \& Ferland (2007) warn that although $G$ indicates temperature in
1599: a collisionally-ionized plasma, it should not be used as a temperature
1600: indicator in a photo-ionized plasma (in addition, see Porter \& Ferland 2007 for
1601: warnings regarding usage of $R$ as density indicator in photo-ionized plasmas).
1602:
1603:
1604: For the \ion{O}{7} emission triplet in NGC 3227, we measure
1605: $G = 4.7^{+9.3}_{-4.0}$; $G < 2.0$ for the \ion{N}{6} triplet.
1606: Uncertainties here are statistical only and do not include
1607: systematic effects associated with the presence of absorption lines.
1608: These values thus do not yield any useful constraints on whether
1609: collisional- or photo-ionization dominates.
1610: For the \ion{O}{7} emission triplet, $R = 1.4^{+1.0}_{-0.9}$.
1611: Assuming the plasma is either purely ionized or
1612: a hybrid of collisionally- and photo-ionized material,
1613: and assuming a temperature near $10^6$ K, electron densities
1614: of roughly $10^{10.5-11.5}$ cm$^{-3}$ are implied
1615: (Porquet \& Dubau 2000). For either collisional or photo-ionization,
1616: densities above $10^{12}$ cm$^{-3}$ are ruled out by the presence of the
1617: strong \ion{O}{7} $(f)$ line.
1618:
1619:
1620:
1621:
1622:
1623: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1624:
1625:
1626:
1627: \subsubsection{Long-term variability in the cold and ionized absorbers}
1628:
1629: Variations in the column density of neutral absorbing gas along the line of sight
1630: has been reported to occur on a wide range of time scales (days to years)
1631: in both Seyfert 1 AGN (e.g., I Zw 1, Gallo et al.\ 2007; NGC 4151, Puccetti et al.\ 2007;
1632: NGC 3516, Turner \et\ 2008, Markowitz \et\ 2008)
1633: and Seyfert 2 AGN (Risaliti et al.\ 2002, 2005).
1634: NGC 3227 is no exception, given the 2000-1 obscuring event by a lowly-ionized
1635: (log $\xi \sim 0$) dense cloud of column density $2.6 \times 10^{23}$ cm$^{-2}$
1636: and inferred to be located in the BLR (Lamer \et\ 2003).
1637: That cloud is likely physically distinct from the local cold
1638: gas and the ionized X-ray absorbers, as
1639: $N_{\rm H,local}$, $N_{\rm H,WA1}$ and $N_{\rm H,WA2}$
1640: modeled from the 2006 {\it XMM-Newton} observation
1641: are much lower than that for the 2000-1 obscuring cloud.
1642: Nonetheless, there does seem to be evidence for variations in
1643: $N_{\rm H,local}$ on time scales of years in NGC 3227:
1644: George \et\ (1998b) noted $N_{\rm H,local}$
1645: to increase from $\sim 3 \times 10^{20}$ cm$^{-2}$ in 1993 to $\sim 3 \times 10^{21}$ cm$^{-2}$ by 1995.
1646: Gondoin \et\ (2003) reported $N_{\rm H,local} = 7 \times 10^{20}$ cm$^{-2}$ from the 2000 {\it XMM-Newton}
1647: observation; that value is the same as the column here in the 2006 EPIC-pn spectrum.
1648: Assuming that discrepancies in values of $N_{\rm H,local}$ inferred from different missions
1649: are intrinsic to the source and not due to cross-instrumental calibration uncertainties,
1650: the range in measured $N_{\rm H,local}$ values would suggest the
1651: presence of cold gas along the line of sight at $<<$pc radii from the black hole.
1652:
1653:
1654: %%%%% The evidence is mounting that the level of observed absorption may not modeled
1655: %%%%% using a simple torus of gas; the absorbing gas is implied to be
1656: %%%%% inhomogeneous in either Seyfert type, and exist over a wide range in length scales.
1657:
1658:
1659:
1660: George \et\ (1998b) noted an increase in the ionized absorber column density, from
1661: $\sim 3 \times 10^{21}$ cm$^{-2}$ in 1993 to $\sim 3 \times 10^{22}$ cm$^{-2}$ 1995,
1662: ruling out a location outside the NLR. The value of $N_{\rm H,WA}$ obtained by
1663: Gondoin \et\ (2003) using {\sc xstar} modeling is $2.7 \pm 0.7 \times 10^{21}$ cm$^{-2}$.
1664: Variability in $N_{\rm H,WA}$ between 2000 and 2006 is thus implied to be significant only
1665: at the 2.3$\sigma$ confidence level. Furthermore, Gondoin \et\ (2003) did not include a component
1666: to model a soft X-ray excess; if the soft excess was intrinsically present during the 2000
1667: observation, then their estimate of $N_{\rm H,WA}$ would be too high.
1668: Assuming the variations in $N_{\rm H,WA}$
1669: since 1993 to be real, then at least some fraction of the ionized X-ray absorbing gas is likely present
1670: at $<<$pc radii.
1671:
1672:
1673:
1674:
1675: \subsubsection{Possible connections between the X-ray and UV ionized absorbers}
1676:
1677: The extended dusty lukewarm absorber (hereafter DLWA) studied by Crenshaw \et\ (2001)
1678: may be a physically and/or kinematically distinct component from the ionized X-ray absorbers.
1679: Crenshaw \et\ (2001) report the ionization parameter $U$ of the DLWA
1680: to be 0.13 (see George \et\ 1998a for definitions of the optical/UV ionization parameter
1681: $U$ and the X-ray ionization parameter $U_{\rm x}$).
1682: Assuming an optical-to-X-ray spectral index $\alpha_{\rm ox}$ (estimated from
1683: photometric measurements taken from NED), and using the conversions from
1684: Figure 1 of George \et\ (1998a), $U/U_{\rm x} \sim 250$ and
1685: $\xi/U_{\rm x} \sim 5000$, we can translate our best-fit ionization parameters
1686: for the low- and high-ionization X-ray absorbers,
1687: log$\xi$=1.21 and log$\xi$=2.90, respectively,
1688: into corresponding values of $U = 1.0$ and 40.
1689: The values of the ionization parameters for the DLWA and the low-ionization X-ray absorber
1690: are not too dissimilar, tentatively suggesting a possible physical connection.
1691: However, it is not possible to definitively link the DLWA and the
1692: low-ionization X-ray absorber in velocity space, as,
1693: other than the Fe UTA and \ion{O}{7} edge, there are no strong X-ray absorption features $>$0.35 keV
1694: endemic to only the low-ionization, and not the high-ionization,
1695: X-ray absorber at the best-fit ionization parameter values.
1696:
1697: %%%% The DLWA could not generate H- or He-like absorpion lines seen in the X-ray spectra
1698: %%%% but could contain Fe species associated with the Fe UTA near 0.7 keV. CONFIRM THIS
1699:
1700:
1701:
1702: %%%%%% U -- Ux -- xi -- logXi
1703: %%%% logXi=0.9===Xi=8 Ux = 0.002 U = 0.5
1704: %%%% logXi=2.3===Xi=200 Ux = 0.04 U = 13
1705: %%%% G98: Ux = 0.01 (xi=50; logXi = 1.7; U = )
1706:
1707: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Lukewarm NH U=0.13 NH=2e21
1708:
1709: Crenshaw \et\ (2001) note that the column density of the DLWA is $2 \times 10^{21}$ cm$^{-2}$,
1710: very similar to that measured for both
1711: X-ray absorbers here. The outflow velocities relative to systemic
1712: of DLWA lines (tens of km s$^{-1}$) are much lower than that of at
1713: least the high-ionization X-ray absorber ($\sim 1500-2000$ km s$^{-1}$).
1714: Furthermore, it is likely that the DLWA has a high covering fraction (Crenshaw \et\ 2001).
1715: We cannot rule out the possibility that the low-ionization X-ray absorber may be the inner
1716: edge of the extended DLWA.
1717: Moreover, it is conceivable that the outflowing, ionized X-ray absorbers
1718: may supply the extended, 100-pc scale DLWA gas (at least along the line of sight),
1719: with the velocity
1720: slowing and ionization parameter decreasing as the distance from the black hole increases.
1721: If the X-ray ionized absorbing gas is continuously feeding
1722: the UV absorber, we might expect the existence of an ``intermediate''
1723: zone of absorbing material with outflow velocity
1724: between those of the observed X-ray and UV ionized absorbers.
1725: Confirmation of such a zone would strengthen the physical connection
1726: between the X-ray and UV ionized absorbers.
1727: However, clarification of whether or not the X-ray absorbing outflow is
1728: sustained or intermittent would be needed to determine the
1729: exact physical and kinematic connection between the UV and X-ray absorbers.
1730:
1731:
1732: The properties we have derived for the ionized X-ray absorbers are
1733: consistent with the notion that dust may be swept up by an outflowing wind at radii
1734: $\gtrsim$ a few light-days (outside the BLR, where dust cannot survive).
1735: However, the current data cannot distinguish between a dusty and a dust-free
1736: X-ray ionized absorber; as discussed in $\S$5, the upper limit on the neutral Fe L3
1737: dust edge in the RGS spectrum corresponds to a
1738: hydrogren column density consistent with that
1739: implied by optical line absorption.
1740: %%%% {\it Chandra} HETGS spectra of MCG--6-30-15 (Lee \et\ 2001)
1741: %%%% and Cyg X-1 (Schulz \et\ 2002)
1742: %%%% reveal the presence of dust via Fe L3 edges at 707 eV; in each case
1743: %%%% the drop in continuum across the edge is $\sim2$ ($\tau \sim 0.7$),
1744: %%%% an order of magnitude larger than the upper limits inferred from the RGS spectrum of NGC 3227.
1745:
1746:
1747:
1748: %%%%%%%%%%%%%%%%%%%%%%%
1749: %%%%%%%%%%%%%%%%%%%%%%%% ---------------------- Section 9.2
1750:
1751: \subsection{The variable soft excess in NGC 3227}
1752:
1753: The origin of the soft excess emission in NGC 3227 is not immediately obvious.
1754: The soft excess does not undergo the same obscuration by cold material as that
1755: suffered by the hard X-ray continuum.
1756: However, the large discrepancy between $\Gamma_{\rm SX}$ and $\Gamma_{\rm HX}$
1757: argues against both partial covering scenarios and against the
1758: soft X-rays being nuclear power-law emission scattered in an extended region.
1759: We have modeled the soft excess in NGC 3227 using a blackbody component to represent
1760: direct thermal emission from accretion disk, but such a model seems to be unphysical
1761: given the apparent consistency of blackbody parameters across a wide range of Seyfert
1762: properties, including black hole mass (e.g., Gierli\'{n}ski \& Done 2004).
1763: We have also modeled the soft excess as inverse Comptonization of seed photons, likely
1764: thermal optical and UV photons from the accretion disk, by thermal electrons
1765: (Sunyaev \& Titarchuk 1980). Such a mechanism could conceivably operate in
1766: the hot, ionized surface of the accretion disk (Hubeny \et\ 2001; Janiuk,
1767: Czerny \& Madejski 2001).
1768:
1769:
1770:
1771:
1772: %%%%%%%%%%%% Var.
1773: %%%% {\bf In blazars, a rapidly variable soft excess has been observed in PKS 1510--089
1774: %%%% (Gambill \et\ 2003), a candidate target for the bulk-Compton emission. Is there
1775: %%%% any reason to discuss that mechanism here?}
1776:
1777:
1778:
1779: In Seyferts, soft excesses have been seen to vary on time scales of weeks, e.g.,
1780: as seen in the narrow line Seyfert 1s Ark 564 and Ton S180 (Edelson \et\ 2002).
1781: The soft excess in the quasar 3C~273 has also been known to vary on
1782: time scales of $\sim$ a week (Kim 2001 and references therein).
1783: Vaughan \et\ (2002) found the soft excess of the NLSy1 Ton S180
1784: to vary rapidly, though in concert with the hard X-ray power law
1785: component, leading to measurents of $F_{\rm var}$ roughly
1786: independent of energy band across the EPIC bandpass.
1787: However, most other Seyferts' $F_{\rm var}$ spectra peak near 1--2 keV (e.g., Ar\'{e}valo \et\ 2008;
1788: Vaughan \& Fabian 2004). This is commonly explained with a variable
1789: power-law component superimposed over a constant or relatively less variable hard component
1790: above $\sim$5 keV (likely the Compton reflection hump) and a constant or less variable
1791: soft component below $\sim$1 keV (the soft excess). In NGC 3227, however, $F_{\rm var}$ increases
1792: with decreasing energy below 1 keV (Figure 14) due to a soft excess that is strongly
1793: variable in normalization.
1794:
1795:
1796: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1797:
1798: The simultaneous increases in UV and soft X-ray continuum flux are not statistically significantly
1799: correlated, so a direct UV--soft X-ray connection is tenuous at best.
1800: However, if such a connection were real, and if both the UV and soft X-ray originate in the same
1801: location on the disk, a rapid change in illumination of the accretion disk could explain
1802: the simultaneous increases.
1803:
1804:
1805: We can investigate if the observed UV trend can be attributed to
1806: thermal reprocessing of the increasing soft X-ray flux trend
1807: by the accretion disk.
1808: Given the black hole mass, accretion rate, and assuming a standard thin
1809: disk with an inner radius of 6 $R_{\rm g}$,
1810: 90\% (50\%, 10\%) of the 260 nm continuum
1811: emission originates from within 60 (25, 10) $R_{\rm g}$.
1812: The light crossing time across the diameter of the disk at
1813: a 60 $R_{\rm g}$ radius is roughly 20 ks.
1814: In a thermal reprocessing scenario, the UV flux should track
1815: the soft X-ray flux and be smeared on a time scale
1816: $<$ 20 ks, assuming that the soft X-ray continuum emitting region is
1817: located near the central disk plane.
1818: If the soft X-ray emitting region is located well off the plane,
1819: the light travel time to the UV-emitting part of the disk
1820: is increased (e.g., 30 ks travel time for a height of 100 $R_{\rm g}$.)
1821: The light curves displayed in Figure 1 are consistent
1822: with this scenario, as any soft X-ray to UV lag is less than
1823: several tens of ks.
1824:
1825: Another possibility is that the soft X-ray continuum flux originates
1826: in inverse Comptonization of UV seed photons from the disk; e.g.,
1827: the scenario associated with the COMPST model.
1828: However, the fact that the observed increase in the soft X-ray flux (40$\%$)
1829: is stronger than the observed increase in the UV flux (10$\%$)
1830: would argue against all of the soft X-ray photons being passively-reprocessed
1831: UV photons, unless only the soft X-ray photons were anisotropically beamed
1832: along the line of sight, or unless the soft X-ray flux is responding to
1833: a comparative increase in the UV continuum flux that
1834: occurred less than a few tens of ks before the start of the
1835: {\it XMM-Newton} observation.
1836:
1837: In the inverse Comptonization scenario, and in the absence of thermal reprocessing,
1838: the UV variability would have
1839: to be intrinsic to the disk. We can consider inwardly-propagating
1840: variations in the local mass accretion rate (e.g., Lyubarski 1997,
1841: Ar\'{e}valo \& Uttley 2006), specifically,
1842: the case where the fluctuations travel on the viscous time scale.
1843: To produce the 10$\%$ observed variation, such
1844: a propagating fluctuation would have to modify
1845: 20\% of the flux contained within a radius of 25 $R_{\rm g}$,
1846: or 100\% of the flux contained within a radius of 10 $R_{\rm g}$.
1847: Let us consider the former case: a
1848: fluctuation that propagates with an inward velocity that takes it from
1849: a radius of 25 $R_{\rm g}$ to the innermost stable disk radius of
1850: 6 $R_{\rm g}$ over 100 ks, while modulating
1851: 20\% of the flux within 25 $R_{\rm g}$, could
1852: produce the observed UV variability. However, the required velocity,
1853: $\sim$1/5 of the orbital velocity, is very high for a standard disk
1854: (Shakura \& Sunyaev 1973):
1855: to be the viscous velocity for a standard disk, one would need
1856: $(H/R)^2\alpha = 0.2$ ($H$, $R$, and $\alpha$ are the height, radius
1857: and viscosity parameters, respectively), i.e., a very
1858: thick disk with a large viscosity parameter, which is probably
1859: not adequate for the optical-emitting portion of the disk.
1860: The UV variability is thus likely not associated with
1861: accretion rate fluctuations traveling on viscous time scales.
1862: The sound speed near 25 $R_{\rm g}$ (assuming $\alpha = 0.1$) is
1863: roughly two orders of magnitude slower than the orbital speed,
1864: and sound waves are thus likely
1865: inadequate to produce the required flux modification in 100 ks.
1866:
1867: %%%%% $F_{2-10}$ rapidly varies by 50$\%$ (maximum to minimum flux ratio)
1868: %%%%% during the {\it XMM-Newton} observation, but it is not clear what variability ``trend''
1869: %%%%% simultaneously observed in the hard X-ray band during the {\it XMM-Newton} observation
1870: %%%%% is driving the increasing trend on tens of ks in the soft excess. It is thus difficult
1871: %%%%% to attribute the soft excess to inverse Comptonization of disk seed photons
1872: %%%%% in the accretion disk skin unless some
1873: %%%%% disk skin property directly related to soft X-ray emission,
1874: %%%%% and not directly related to the hard X-ray coronal emission, were changing,
1875: %%%%% or unless the light travel time delay between the hard X-ray emission origin and
1876: %%%%% the soft X-ray emission origin was significant.
1877: %%%%% Perhaps the soft excess during the {\it XMM-Newton} observation is responding to
1878: %%%%% strong variations in $F_{2-10}$ which occurred before the start of the observation,
1879: %%%%% with a light travel time delay of at least tens of ks.
1880:
1881: %%%%%%%%%%%
1882:
1883:
1884: The rapidly variable soft excess is qualitatively similar to that
1885: observed in 3C 120 with {\it Suzaku} by Kataoka \et\ (2006), who model
1886: the variable soft excess in that object as the high-energy part of a
1887: synchrotron jet component. The rapid variability observed
1888: in the soft X-ray and UV bands in NGC 3227, and the fact that the
1889: soft excess undergoes less absorption by cold material
1890: compared to the hard X-ray continuum, may also be
1891: consistent with existence of such a jet component,
1892: in addition to and independent of the accretion disk corona
1893: (hard X-ray) emission.
1894: Unlike 3C 120, NGC 3227 is of course a radio-quiet AGN.
1895: A $\sim$40 pc scale jet was detected in NGC 3227 by the VLA in
1896: a 1991 observation (Kukula \et\ 1995), though 10-100 pc scale jets are
1897: common in Seyferts (e.g., Gallimore \et\ 2006).
1898: Broad band spectral energy distribution (SED) modeling
1899: of an accretion disk + jet is beyond the scope of this paper;
1900: constraining such an SED component would be complicated by the presence of absorbing dust in
1901: the poorly-studied EUV band. However, such a component is plausible
1902: if its peak in $\nu$$L_{\nu}$ space were near $10^{14-16}$ Hz
1903: (explained if the jet were young, e.g.,
1904: $10^{\sim4-5}$ yrs), and the total observed synchrotron power were
1905: no more than $10^{41-42}$ erg s$^{-1}$. In this scenario, the observed
1906: UV continuum emission would be a mixture of slowly-varying "big blue bump"
1907: emission from the accretion disk plus rapidly variable synchrotron emission
1908: (which would contribute less than $\sim$50$\%$ to the total observed UV
1909: continuum flux).
1910:
1911:
1912: %%%%%%%%%%%%%%%%%%%%%%%%% section 9.3 the high-energy continuum
1913:
1914: \subsection{The high energy emission components}
1915:
1916: The photon index of the power-law component used to model the
1917: hard X-ray continuum emission in both the
1918: {\it XMM-Newton} and {\it RXTE} data
1919: is rather low ($\Gamma_{\rm HX} \sim 1.5-1.6$) compared to
1920: traditional or "canonical" values of $\sim 1.8-1.9$ for most broad-line
1921: Seyfert 1 AGN. However, similarly low values of
1922: $\Gamma_{\rm HX}$ have been reported previously for this object (e.g.,
1923: George et al.\ 1998b). The presence of a cutoff near 100 keV, as implied by both
1924: the {\it Swift}-BAT and the {\it RXTE}-HEXTE data,
1925: would suggest thermal Comptonization, but better
1926: data in the 50--200 keV band and above are needed to confirm the
1927: cutoff and better constrain its energy.
1928: The low-energy rollover to the hard X-ray continuum component was modeled above
1929: as being due to cold absorption, but could also potentially be a signature of
1930: Comptonization, as it is possible for
1931: Comptonized continuua to have a low-energy rollover e.g., below 1 keV
1932: depending on parameters such as the distribution of input photon field
1933: (e.g., Titarchuk 1994).
1934:
1935:
1936: The 7.1 keV edge is likely due to reflection, as the absorbing
1937: components modeled here predict an edge depth much too small
1938: to detect here. However, the strength of the Compton
1939: reflection component $R$ is measured to be low ($\lesssim 0.5$),
1940: suggesting that Compton-thick material exists in only a small
1941: fraction of the sky as seen from the continuum source.
1942: For example, the putative Compton-thick
1943: molecular torus invoked in standard
1944: Seyfert 1/2 unification schemes (Urry \& Padovani 1995) could be
1945: weak or small. Another possibility is that the
1946: optically-thick accretion disk could be truncated,
1947: or the inner accretion disk could transition into
1948: an optically-thin, radiatively inefficient flow, e.g.,
1949: an ADAF or RIAF (Narayan \& Yi 1994, 1995; Blandford \& Begelman 1999; Narayan \et\ 2000).
1950: Such flows are frequently
1951: invoked when describing low accretion rate accreting black
1952: hole systems (e.g., low luminosity AGN).
1953: In addition, Liu \et\ (2007) forward a model for
1954: low accretion rate flows in which optically-thin coronal
1955: matter condenses into a cool, optically-thick inner disk,
1956: for $L_{\rm Bol}/L_{\rm Edd} = 0.1-2\%$ (assuming
1957: values for the viscosity parameter $\alpha \sim 0.1 - 0.4$).
1958: Condensation radii of a few tens of $R_{\rm g}$ are plausible.
1959:
1960: Wu \& Gu (2008) noted that for both black hole X-ray binary systems and AGN
1961: accreting above a critical "transition" value
1962: of $L_{\rm Bol}/L_{\rm Edd}$, $\Gamma_{\rm HX}$ correlates
1963: with $L_{\rm Bol}/L_{\rm Edd}$. Below this transition,
1964: $\Gamma_{\rm HX}$ and $L_{\rm Bol}/L_{\rm Edd}$ anti-correlate.
1965: Wu \& Gu (2008) suggest that the transition between
1966: a thin disk and an ADAF-type flow occurs near
1967: $L_{\rm Bol}/L_{\rm Edd} = 1\%$ for black hole X-ray binary
1968: systems and near 0.3$\%$ for AGN.
1969: The corresponding value of $\Gamma_{\rm HX}$
1970: at this transition point is 1.5, similar to that observed
1971: in NGC 3227 ($L_{\rm Bol}/L_{\rm Edd} = 1\%$). This suggests that the accretion flow in
1972: NGC 3227 may be consistent with either a thin disk, a ADAF-type flow,
1973: or a transition to both.
1974:
1975: %%%% ADAF flows are also suspected of generating continuum emission via thermal bremmstrahlung.
1976: %%% The soft excess in the EPIC-pn spectrum of NGC 3227
1977: %%% is also well modeled with such an emission component, with $\chi^2$/$dof$ = 1890.3/1724 and
1978: %%% residuals virtually identical to those in the BB and COMPST models for an electron
1979: %%% temperature $k_{\rm B}T_{\rm e} = 0.18^{+0.03}_{-0.04}$ keV.
1980: %%% However, ADAFs are generally expected to be associated with higher electron temperatures, e.g., a few keV
1981:
1982:
1983:
1984: %%%%%%%%%%%%%%%%%%%%%%%%% ------------- 8.4 FE K line emission features
1985:
1986: \subsection{The Narrow Fe K$\alpha$ line at 6.4 keV}
1987:
1988:
1989: The measured energy of the Fe K$\alpha$ emission line is consistent
1990: with an origin in neutral material. The line is resolved in the pn spectrum;
1991: the measured width $\sigma$ of
1992: 65$\pm$14 eV translates into a FWHM velocity $v_{\rm FWHM}$
1993: of 7000$\pm$1500 km s$^{-1}$.
1994: This velocity width is similar to that for the
1995: H$\beta$ line, 5500$\pm$500 km s$^{-1}$ (Wandel \et\ 1999), suggesting
1996: an origin for the Fe K line in material commensurate with the BLR.
1997:
1998: Assuming Keplerian motion, and assuming the velocity dispersion
1999: is related to the FWHM velocity as $<$$v^2$$> = \frac{3}{4} v_{\rm FWHM}^2$
2000: (Netzer \et\ 1990), assuming a black hole mass $M_{\rm BH}$ of
2001: $4.22 \pm 2.14 \times 10^7 \Msun$ (Peterson \et\ 2004), we can use
2002: $GM_{\rm BH} = r$$<$$v^2$$>$ to estimate the radius $r$ of the line-emitting
2003: material. We find $r = 7.2^{+12.7}_{-4.9}$ light-days, equivalent to
2004: $3000^{+5300}_{-2100}$ $R_{\rm g}$.
2005: This value is consistent with the radius of $<$ 700 light-days
2006: inferred from the very crude reverberation mapping discussed in $\S$6.2.
2007: It is also consistent with the BLR (H$\beta$) region radius of
2008: $9^{+6}_{-8}$ light-days (Peterson \et\ 2004),
2009: as well as with the $\sim$5--20 light-day
2010: inner radius of the dust as reverberation-mapped by Suganuma \et\ (2006).
2011:
2012: The Fe line-emitting material may be spatially extended in NGC 3227;
2013: only the innermost regions have had time to respond to
2014: continuum variations, which may get smoothed out by
2015: the outer regions anyway. Alternatively, we note in Figure 16 that extrapolation of the
2016: $F_{2-10}$--$I_{\rm FeK\alpha}$ relation to zero continuum flux results in a non-zero offset.
2017: That is, there may exist Fe-line emitting material which does not respond to the
2018: continuum flux, and only some fraction of the total observed
2019: Fe line flux responds to the continuum variations on $<$ 700 light-days.
2020: Either of these two ideas may explain why
2021: $F_{\rm var}$ for the 2--10 keV continuum flux is
2022: larger than that for the Fe line.
2023:
2024:
2025: The observed line $EW$ is 91$\pm$10 eV.
2026: The column density of the gas required to produce such an emission
2027: line must be greater than $10^{\sim 22}$ cm$^{-2}$,
2028: otherwise the optical depth would be insufficient to produce
2029: such a prominent line. However, the lack of an obvious
2030: Compton shoulder (the emission near 6.0 keV
2031: was not identified as such) would argue against the bulk of the line
2032: photons originating in Compton-thick material.
2033:
2034: Let us first assume an origin in optically-thin gas that completely surrounds
2035: a single, isotropically-emitting continuum source (covering fraction
2036: $f_{\rm c}$ = 1)
2037: and is uniform in column density.
2038: We can relate $EW$ to $N_{\rm H}$ using the following equation,
2039: taken from e.g., Markowitz \et\ (2007):
2040: \begin{equation}
2041: EW_{\rm calc} = f_{\rm c} \omega f_{\rm K\alpha} A \frac{\int^{\infty}_{E_{\rm K
2042: edge}}P(E) \sigma_{\rm ph}(E) N_{\rm H} dE}{P(E_{\rm line})}
2043: \end{equation}
2044: $\omega$ is the fluorescent yield: the value for Fe, 0.34, was taken from
2045: Kallman \et\ (2004). $f_{\rm K\alpha}$ is the fraction of
2046: photons that go into the K$\alpha$ line
2047: as opposed to the K$\beta$ line; this is 0.89 for \ion{Fe}{1}.
2048: $A$ is the number abundance relative to hydrogen;
2049: solar abundances, using Lodders (2003), were used.
2050: $P(E)$ is the spectrum of the illuminating continuum at energy $E$;
2051: $E_{\rm line}$ is the K$\alpha$ emission line energy.
2052: $\sigma_{\rm ph}(E)$ is the photo-ionization cross section
2053: assuming absorption by K-shell electrons only; all cross sections were taken
2054: from Veigele (1973\footnote{http://www.pa.uky.edu/$\sim$verner/photo.html}).
2055: The observed Fe K$\alpha$ $EW$ can thus be produced from material with
2056: $N_{\rm H} = 1.4 \pm 0.2 \times 10^{23}$ cm$^{-2}$.
2057: Similarly, the observed Ni K$\alpha$ line $EW$ of 13$^{+13}_{-10}$ eV
2058: indicates $N_{\rm H} = 3.9^{+3.9}_{-2.9} \times 10^{23}$ cm$^{-2}$.
2059:
2060: If the absorbers are instead clumpy and lie out of the line of sight, we
2061: can use Eqn.\ 1 of Wozniak \et\ (1998), which gives the expected Fe K$\alpha$
2062: line intensity for a cloud with a column $\gtrsim 10^{23}$ cm$^{-2}$
2063: lying off the line of sight and subtending a fraction $\Omega/4\pi$
2064: of the sky as seen from a single, isotropically-emitting continuum
2065: source. Assuming solar abundances and given the observed
2066: Fe K$\alpha$ line intensity, for values of $\Omega/4\pi$ =
2067: 0.1 (0.3), a column density of $1.5 \times 10^{23}$
2068: ($5 \times 10^{22}$) cm$^{-2}$.
2069:
2070:
2071: In either case, the derived column density is higher than
2072: that of the hard X-ray absorbing material in the SXPL
2073: and COMPST model fits to the {\it XMM-Newton} EPIC data.
2074: It is, however, closer to the column density
2075: of the obscuring cloud during the 2000-1 obscuring event (Lamer \et\ 2003).
2076: Based on the duration of the obscuring event and the implied
2077: number density, that cloud was inferred to be located in the BLR.
2078: The similarities in column density and inferred location
2079: suggest a connection between the Fe line-emitting material
2080: and the obscuring cloud of 2000-1.
2081:
2082:
2083: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2084:
2085: %%%%% 9.3.2-
2086: \subsection{Emission at 6.0 keV}
2087:
2088: The Fe K emission profiles in both the pn and MOS 1+2 spectra
2089: suggest emission redward of the Fe K$\alpha$ core. However, the present
2090: data cannot distinguish between a model wherein the emission
2091: near 6.0 keV is due to a narrow emission line or is instead the
2092: red wing of a relativistically broadened Fe K$\alpha$ line with
2093: an $EW$ near 80 eV.
2094:
2095: \subsubsection{A narrow emission feature?}
2096:
2097: We first discuss the emission in the context of modeling it as a narrow
2098: feature. In the pn spectrum, the line width $\sigma$ was $<$200 eV; the
2099: $EW$ was $21^{+19}_{-9}$ eV. High-resolution spectroscopy has yielded
2100: evidence for similar narrow emission features in
2101: roughly 16 Seyferts to date (see Vaughan \& Uttley 2008 for a review).
2102: These features are commonly interpreted as red- or blue-shifted Fe K features
2103: associated with "hot-spots," localized areas of the accretion disk
2104: illuminated by a localized flare just above the disc (possibly due
2105: to magnetic reconnection), rather than a central
2106: illuminator or an extended corona (e.g., Dov\u{c}iak et al.\ 2004,
2107: Goosmann et al.\ 2007).
2108:
2109: Vaughan \& Uttley (2008) point out that these narrow emission
2110: features are rarely reported at more than 3.0$\sigma$ confidence.
2111: Consideration of selection bias and publication bias
2112: brings skepticism to the validity of these features and
2113: raises the possibility that a non-trivial fraction of
2114: these published lines may be consistent with being due to photon noise.
2115: In the case of the pn spectrum of NGC 3227,
2116: the Monte Carlo simulations we performed suggested that
2117: the line was detected at $>$99.9$\%$ confidence (i.e., $<$0.1$\%$
2118: likelihood that it is due to photon noise). However,
2119: we caution that the Monte Carlo procedure yields an estimate of the
2120: detection significance independently of the existence of
2121: other high-resolution spectra of NGC 3227 and other Seyferts.
2122: On the other hand, the marginal detection of the emission feature near 6.1 keV
2123: in the MOS 1+2 spectrum, simultaneously to the pn data, reduces the
2124: likelihood that the feature is due to photon noise. In addition,
2125: the value of $\vert\Delta\chi^2\vert$ associated with modeling the
2126: line in the pn spectrum is among the largest reported for
2127: narrow emission lines (See Table 1 of Vaughan \& Uttley 2008).
2128:
2129:
2130:
2131:
2132: If the feature near 6.0 keV is indeed emission from
2133: a "hot spot" on the accretion disk, we can estimate its radial distance from the black hole.
2134: For instance, we can assume that the observed redshifting is due solely
2135: to Doppler shifting associated with Keplerian motion of the receding side
2136: of the disk. We use the 3$\sigma$ energy range of 5.8 to 6.3 keV,
2137: corresponding to radial velocities along the line of sight
2138: $v_{\rm r}$ of 0.02--0.09$c$ (assuming an origin in neutral Fe).
2139: Assuming the disk is inclined by $i$ = 30$\degr$ (45$\degr$) with respect to
2140: the plane of the sky, using $M_{\rm BH} = 4.2 \times 10^7 \Msun$
2141: (Peterson \et\ 2004), and ignoring gravitational redshifting, this
2142: corresponds to Keplerian motion at radii of roughly
2143: 12--430 (25--800) $R_{\rm Sch}$.
2144:
2145: If the flare is co-rotating with the disk, then a minimum radius
2146: independent of $i$ is defined by the fact that the line energy is
2147: consistent with being constant and redward of 6.4 keV for the 100 ks
2148: duration of the observation. Any orbital period $<$100 ks is excluded;
2149: otherwise the number of photons $>$6.4 keV would be greater than
2150: the number of photons $<$6.4 keV. Radii $<$ 9 $R_{\rm Sch}$ are
2151: thus excluded.
2152:
2153:
2154: An alternate possibility is that the line-emitting material is
2155: in free-fall, with no velocity component transverse to
2156: our line of sight. Yet another possibility is that the observed redshifting
2157: is purely due to energy losses of photons escaping from
2158: the vicinity of the black hole;
2159: assuming for simplicity emission from neutral Fe originating in gas lying
2160: along the line of sight to the black hole, with all
2161: velocity transverse to line of sight, radii of
2162: $\sim 6-30 R_{\rm Sch}$ are plausible.
2163:
2164: Another speculative possibility is that the line-emitting material could be
2165: associated with the base of the jet, forming above the accretion disk,
2166: as opposed to being a part of the accretion disk.
2167: Close to the rotation axis in AGN,
2168: magnetic fields are likely twisted by the differential
2169: rotation of the disk, serving to launch and collimate jets
2170: (e.g., Blandford \& Payne 1982).
2171: For instance, Marscher \et\ (2008) presented observational evidence for
2172: the jet in the radio-loud AGN BL Lac to follow a spiral flow as it is
2173: accelerated through a zone containing a helical magnetic
2174: field. The 6.0 keV line in NGC 3227 may be emission from Fe in a blob
2175: of material, with a column density $N_{\rm H}$ of 10$^{\sim 21-22}$ cm$^{-2}$,
2176: which has been caught up in the launching and formation
2177: of the weak jet. If the material is close to the axis,
2178: above the disk, and moving with a velocity with a large azimuthal
2179: component, then an observed redshift is possible unless the
2180: rotation axis lies very close to the line of sight.
2181: However, the weakness of Seyferts' jets may possibly be attributed to
2182: weaker acceleration and/or collimation mechanisms compared to
2183: those in blazars, reducing the likelihood that
2184: such azimuthal velocities can exist in Seyferts.
2185:
2186:
2187: \subsubsection{A relativistically broadened diskline component?}
2188:
2189: Evidence for relativistically broadened Fe line
2190: "diskline" profiles in a significant fraction of Seyferts
2191: has been accumulating since the days of {\it ASCA}
2192: (Tanaka \et\ 1995, Fabian \et\ 2000).
2193: In recent years, {\it XMM-Newton} and {\it Chandra}-HETGS
2194: observations have demonstrated that a narrow core at 6.4 keV
2195: is ubiquitous; accurately modeling that component, as well as
2196: absorption due to ionized gas in the line of sight, is critical
2197: for accurate measurement of the flux and profile of
2198: any broad Fe line present.
2199:
2200: In NGC 3227, our best-fit "DL"
2201: model incorporates a narrow Fe K$\alpha$ core,
2202: and suggests a relatively weak broad line, with an $EW$ of
2203: 81$^{+42}_{-30}$ eV. We constrained the inner radius to be $< 22 R_{\rm g}$ and the inclination to be
2204: $<$25$\degr$ (though these values were derived with the radial
2205: emissivity parameter fixed at --3).
2206:
2207: Assuming solar abundances, the abundances of Lodders (2003),
2208: and given the measured strength of the Compton reflection component
2209: $R = 0.40 \pm 0.07$, the predicted Fe line $EW$
2210: (George \& Fabian 1991)\footnote{George \& Fabian
2211: 1991 predict $R$ = $EW$/150 eV, using
2212: the abundances of Morrison \& McCammon (1983), which assumed
2213: $3.3 \times 10^{-5}$ Fe atoms per H atom. Lodders (2003) sets
2214: the Fe abundance to $2.95 \times 10^{-5}$ Fe atoms per H atom.}
2215: is 54$\pm$9 eV, consistent with the modeled $EW$ of the observed
2216: broad Fe line. The low value of the broad line $EW$ may also suggest
2217: that the optically-thick disk is not spatially
2218: extended, again consistent with the notion that
2219: the nucleus of NGC 3227 could harbor an ADAF or RIAF,
2220: and the upper limit on the inner radius of 22 $R_{\rm g}$ is
2221: potentially consistent with the model of Liu \et\ (2007) containing
2222: a small, optically-thin disk.
2223:
2224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2225:
2226: \section{Conclusions}
2227:
2228: We have observed the nucleus of the Seyfert 1.5 AGN NGC 3227
2229: with {\it XMM-Newton} in December 2006 for almost 100 ks.
2230: We have used EPIC and RGS spectra to study the Fe K bandpass,
2231: ionized absorbers, and 0.2--10 keV continuum in detail. We also present
2232: X-ray continuum light curves and a UV continuum light curve obtained with the
2233: Optical Monitor. We have combined these data with archival {\it RXTE}-PCA and HEXTE
2234: monitoring, plus {\it Swift}-BAT monitoring data, to constrain the level
2235: Compton reflection, study the high energy continuum up to 200 keV,
2236: and track the flux behavior of the Fe K line on time scales of weeks to years.
2237: Our main results are summarized as follows:
2238:
2239: The EPIC pn spectrum shows the prominent narrow Fe K$\alpha$ emission line
2240: to be consistent with an origin in neutral Fe (Gaussian energy centroid
2241: 6.403 $\pm$ 0.009 keV). Its intensity and
2242: $EW$ are $3.5 \pm 0.4 \times 10^{-5}$ ph cm$^{-2}$ s$^{-1}$ and 91$\pm$10 eV,
2243: respectively, consistent with an origin in material with a column density
2244: $10^{\sim22.5-23}$ cm$^{-2}$. The line is resolved in the pn spectrum, with
2245: a FWHM velocity 7000$\pm$1500 km s$^{-1}$. Assuming purely Keplerian motion,
2246: we estimate the radius of the line-emitting material to be
2247: $7.2^{+12.7}_{-4.9}$ light-days.
2248: The FWHM velocity and estimated radius are consistent with the
2249: BLR (as mapped by H$\beta$ emission)
2250: as well as with the $\sim$5--20 light-day
2251: inner radius of the dust as reverberation-mapped by Suganuma \et\ (2006).
2252:
2253:
2254: Time-resolved spectral fitting to the 1999--2005 {\it RXTE}-PCA monitoring
2255: data reveal tentative evidence for a significant fraction of the
2256: Fe K line photons to track variations seen in the continuum, with a
2257: light travel time delay which is not tightly constrained, but is
2258: $<$ 700 days. We thus rule out the
2259: possibility that the bulk of the (variable) Fe line photons
2260: originate at distances of 1--2 pc and more, e.g., in a multi-pc-scale molecular torus.
2261: However, more intensive sampling on time scales of days to weeks is required
2262: to beter constrain the lower limit on the continuum--line lag.
2263:
2264:
2265: Emission near 6.0 keV is detected in both the pn and MOS 1+2 spectra.
2266: The fact that it is detected in both instruments and the relatively
2267: large associated value of $\Delta\chi^2$ both argue against this feature being
2268: an artifact due to photon noise.
2269: It is inconsistent with being a Compton shoulder to the 6.4 keV
2270: Fe K$\alpha$ line.
2271: We modeled the emission using a narrow Gaussian component, with
2272: energy centroid (in the pn spectrum) 6.04$^{+0.18}_{-0.04}$ keV,
2273: width $\sigma < 200$ eV, intensity
2274: $9^{+8}_{-4} \times 10^{-6}$ ph cm $^{-2}$ s$^{-1}$ and
2275: $EW$ of 21$^{+19}_{-9}$ eV. A possible origin is a "hot-spot" in the
2276: accretion disk, originating at radii of tens to hundreds of
2277: $R_{\rm g}$. The emission feature is modeled equally well as the red wing
2278: to a relativistically broadened "diskline" component, with
2279: inner radius $<$22 $R_{\rm g}$, inclination $<$25$\degr$,
2280: and $EW$ = 81$^{+42}_{-30}$ eV.
2281:
2282:
2283:
2284:
2285: Broadband (0.2-10 keV) EPIC spectral modeling reveals a strong soft
2286: excess dominating below 1 keV. In the pn spectrum, the soft excess
2287: is fit well by a steep power law, with $\Gamma_{\rm SX}$ $\gtrsim$ 3.
2288: A blackbody component and a low-temperature Comptonization component
2289: each fit the data well, though the physical plausibility these latter two
2290: components in Seyferts is not generally accepted.
2291: The normalization of the soft excess increases by about 20$\%$ during the
2292: first $\sim$20 ks of the {\it XMM-Newton} observation, and by
2293: $\sim$40$\%$ over $\sim50$ ks. The soft X-ray band is more strongly
2294: variable than the hard X-ray band on time scales of tens of ks:
2295: unlike the $F_{\rm var}$ spectra of many other Seyferts,
2296: the $F_{\rm var}$ spectrum of NGC 3227 below 1-2 keV continues to increase as photon
2297: energy decreases. Such relatively rapid variability in the
2298: soft excess, independent of rapid variability in the hard X-ray band
2299: is very rare for Seyferts.
2300:
2301:
2302: The OM shows the UV continuum flux (near 260 nm) to
2303: increase by 10$\%$ during the {\it XMM-Newton} observation.
2304: Such $\lesssim$1 day variability is relatively strong compared to
2305: similar-duration UV continuum light curves obtained with the OM
2306: (Smith \& Vaughan 2007).
2307:
2308:
2309: We present the highest-quality gratings spectrum obtained for NGC 3227 to date.
2310: In the RGS spectrum, we model the ionized absorption using two zones.
2311: The higher-ionization zone has log$\xi_{\rm hi}$ = $2.90^{+0.21}_{-0.26}$,
2312: and an ouflow velocity relative to systemic of --(2060$^{+240}_{-170}$) km s$^{-1}$.
2313: This indicates a minimum radial distance from the black hole of 40 light-days,
2314: placing it outside the BLR radius.
2315: The lower-ionization zone has log$\xi_{\rm lo}$ = $1.21^{+0.18}_{-0.08}$
2316: (corresponding to a factor of roughly 3--10 higher than the
2317: dusty lukewarm absorber modeled by Kraemer \et\ 2000 and
2318: Crenshaw \et\ 2001). Its main signature is a Fe M shell UTA near 740--780 eV.
2319: The best estimate for its outflow velocity relative to systemic is
2320: --(420$^{+430}_{-190}$) km s$^{-1}$.
2321: We find no evidence for non-solar abundances.
2322: Both absorbing zones have column densities $N_{\rm H,WA}$
2323: near $ 1-2 \times 10^{21}$ cm$^{-2}$. $N_{\rm H,WA}$
2324: and the column densities of local, cold absorption measured from the EPIC spectra
2325: are too low to be directly associated with the
2326: obscuring BLR cloud in 2000-1 studied by Lamer \et\ (2003).
2327: However $N_{\rm H,WA}$ is similar to that of the 100-pc scale DLWA,
2328: consistent with the notion that the outflowing X-ray absorbing material
2329: may supply the UV-absorbing DLWA material.
2330:
2331: In the RGS spectrum, we detect five narrow emission lines detected at the systemic redshift; we
2332: identify these lines as being due to \ion{N}{6} {\it (r)}, \ion{N}{7} , and
2333: the {\it (f)}, {\it (i)} and {\it (r)} lines of \ion{O}{8}.
2334: No emission due to \ion{Fe}{17} L line (3d--2p) at 826 eV
2335: (a signature of collisionally-ionized plasma) is detected
2336: (upper limit of 1 eV).
2337: No strong RRC features due to H- or He-like ions
2338: (signatures of photo-ionized plasma) are detected
2339: (upper limits range from 5 to 50 eV).
2340:
2341: We presented the total spectrum derived from
2342: {\it RXTE}-PCA and HEXTE archival monitoring data taken in 1996 and from 1999--2005,
2343: and combined them with a 4-channel Swift-BAT spectrum
2344: from the 9-month survey data (obtained in 2005).
2345: The HEXTE and BAT spectral data reported here represent the first detailed
2346: spectrum of NGC 3227 above $\sim$ 20 keV and up to almost 200 keV.
2347: The hard X-ray continuum photon index in both the {\it RXTE}+BAT
2348: spectra and {\it XMM-Newton} EPIC spectra is rather flat, $\Gamma_{\rm HX} = 1.5-1.6$.
2349: The strength of the Compton reflection hump is rather low,
2350: $R \lesssim 0.5$, consistent with the $EW$ of the diskline component
2351: modeled in the pn spectrum (assuming solar abundances).
2352: We also find evidence for a high-energy continuum cutoff at 90$\pm$20 keV.
2353:
2354: The low values of $\Gamma_{\rm HX}$ and $R$, combined with the
2355: low value of $L_{\rm Bol}/L_{\rm Edd} \sim 1\%$, are consistent with the
2356: notion that NGC 3227 may harbor an optically-thin, radiatively-inefficient flow
2357: such as an ADAF, in addition to or instead of
2358: a standard geometrically-thin, radiatively-efficient disk.
2359:
2360:
2361:
2362:
2363:
2364:
2365: %%%% Crenshaw \et\ (2001):
2366: %%% We note that at soft X-ray energies, absorption by \ion{He}{2} is
2367: %%%% dominant/significant The inferred column densities thus depend on the assumed
2368: %%%%% abundance ratio of He to H.
2369:
2370:
2371: %%%% Model lukewarm gas in all: 2e21 ?????
2372:
2373:
2374:
2375:
2376: \acknowledgements
2377: A.M.\ thanks M.\ Elvis, A.\ Marscher and D.\ Evans for
2378: helpful suggestions. This work is based on an observation
2379: obtained with {\it XMM-Newton}, an ESA science mission, and
2380: has made use of HEASARC online services, supported by
2381: NASA/GSFC, the NASA/IPAC Extragalactic Database,
2382: operated by JPL/California Institute of Technology under
2383: contract with NASA, and the NIST Atomic Spectra Database.
2384:
2385:
2386: \begin{references} %%%%%%%% copied from file ``refs3'' on Apr 20
2387:
2388: \reference{A08} Ar\'{e}valo, P., M$^{\rm c}$Hardy, I.M., Markowitz, A., Papadakis, I.E., Turner, T.J., Miller, L.\ \& Reeves, J.N. 2008, MNRAS, 387, 279
2389: \reference{AU06} Ar\'{e}valo, P.\ \& Uttley, P. 2006, MNRAS, 367, 801
2390: \reference{Ar96} Arnaud, K. 1996, in {\it Astronomical Data Analysis Software and Systems}, Jacoby, G., Barnes, J., eds., ASP Conf.\ Series Vol.\ 101, p.\ 17
2391: \reference{BB67} Bearden, J.A.\ \& Burr, A.F. 1967, Rev.\ Mod.\ Phys., 39, 125
2392: \reference{Be87} Bechtold, J., Czerny, B., Elvis, M., Fabbiano, G.\ \& Green, R.F., 1987, ApJ, 314, 699
2393: %%%%\reference{B01} Behar, E., Sako, M.\ \& Kahn, S.M. 2001, ApJ, 563, 497
2394: \reference{B99} Blandford, R.\ \& Begelman, M.C. 1999, MNRAS, 303, L1
2395: \reference{B82} Blandford, R.D.\ \& Payne, D.G. 1982, MNRAS, 199, 883
2396: \reference{Bl05} Blustin, A.J., Page, M.J., Fuerst, S.V., Branduardi-Raymont, G.\ \& Ashton, C.E. 2005, A\&A, 431, 111
2397: \reference{B96} Brandt, W.N., Fabian, A.C., \& Pounds, K.A. 1996, MNRAS, 278, 326
2398: \reference{B07} Braito, V., Reeves, J.N., Dewangen, G.C.\ \et\ 2007, ApJ, 670, 978
2399: \reference{C83} Cohen, R.D. 1983, ApJ, 273, 489
2400: \reference{C92} Courvoisier, T.J.\ \& Paltani, S. 1992, IUE-Uniform Low Dispersion Archive Access Guide No.\ 4 (ESA-SP 1153A/B) (Noordwijk: ESA)
2401: \reference{C01} Crenshaw, D.M., Kraemer, S.B., Bruhweiler, F.C., Ruiz, J.R. 2001, ApJ, 555, 633
2402: \reference{CK01} Crenshaw, D.M.\ \& Kraemer, S.B. 2001, ApJ, 562, L29
2403: \reference{C95} Crocombette, J.P., Pollak, M., Jollet, F., Thromat, N., \& Gautier-Soyer, M.\ 1995, Phys.\ Rev.\ B, 52, 3143
2404: \reference{Cr06} Crummy, J., Fabian, A.C., Gallo, L.C.\ \& Ross, R.R. 2006, MNRAS, 365, 1067
2405: \reference{D91} de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H.G., Jr., Buta, R.J., Paturel, G.\ \& Fouqu\'{e}, P. 1991, Third Reference Catalogue of Bright Galaxies (New York: Springer)
2406: \reference{DL90} Dickey, J.\ \& Lockman, F. 1990, ARAA, 28, 215
2407: \reference{Dv04} Dov\u{c}iak, M., Bianchi, S., Guainazzi, M., Karas, V.\ \& Matt, G. 2004, MNRAS, 350, 745
2408: \reference{E88} Edelson, R.\ \& Krolik, J.H. 1988, ApJ, 333, 646
2409: \reference{E02} Edelson, R., Turner, T.J., Pounds, K.A., Vaughan, S.A., Markowitz, A., Marshall, H., Dobbie, P.\ \& Warwick, R. 2002, ApJ, 568, 610
2410: \reference{E94} Elvis, M.\ et al.\ 1994, ApJS, 95, 1
2411: \reference{F00} Fabian, A.C., Iwasawa, K., Reynolds, C.S.\ \& Young, A.J. 2000, PASP, 112, 1145
2412: \reference{G06} Gallimore, J., et al. 2006, AJ, 132, 546
2413: \reference{G07} Gallo, L.C., Brandt, W.N., Costantini, E., Fabian, A.C., Iwasawa, K.\ \& Papadakis, I.E.\ 2007, MNRAS, 377, 391
2414: \reference{GP87} Gaskell, M.C. \& Peterson, B.M. 1987, ApJS, 65, 1
2415: \reference{GF91} George, I.M.\ \& Fabian, A.C. 1991, MNRAS, 249, 352
2416: \reference{G98a} George, I.M., Turner, T.J., Netzer, H., Nandra, K., Mushotzky, R.F., \& Yaqoob, T. 1998a, ApJS, 114, 73
2417: \reference{G98b} George, I.M., Mushotzky, R.F., Turner, T.J., Yaqoob, T., Ptak, A., Nandra, K., \& Netzer, H. 1998b, ApJ, 509, 146
2418: \reference{G01} George, I.M., Mushotzky, R.F., Yaqoob, T., Turner, T.J., Kraemer, S., Ptak, A.\ \& Nandra, K. 2001, ApJ, 559, 167
2419: \reference{GD04} Gierli\'{n}ski, M., Done, C.\ 2004, MNRAS, 349, L7
2420: \reference{GD06} Gierli\'{n}ski, M., Done, C.\ 2006, MNRAS, 371, L16
2421: \reference{Go06} Gon\c{c}alves, A.C., Collin, S., Dumont, A.-M., Mouchet, M. R\'{o}zanska, A., Chevallier, L., \& Goosmann, R.W. 2006, A\&A, 451, L23
2422: \reference{G03} Gondoin, P., Orr, A., Lumb, D.\ \& Siddiqui, H. 2003, A\&A, 397, 883
2423: \reference{G97} Gonzalez Delgado, R.M.\ \& Perez, E. 1997, MNRAS, 284, 931
2424: \reference{G07} Goosmann, R.W., Mouchet, M., Czerny, B., Dov\u{c}iak, M., Karas, V., R\'{o}\v{z}a\'{n}ska, A.\ \& Dumont, A.-M. 2007, A\&A, 475, 155
2425: \reference{G06} Gu, M., Holczer, T., Behar, E.\ \& Kahn, S.M. 2006, ApJ, 641, 1227
2426: \reference{H01} Hubeny, I., Blaes, O., Krolik, J.\ \& Agol, E. 2001, ApJ, 559, 680
2427: \reference{J01} Janiuk, A., Czerny, B.\ \& Madejski, G. 2001, ApJ, 557, 408
2428: \reference{K04} Kallman, T.R., Palmeri, P., Bautista, M.A., Mendoza, C., Krolik, J.H.\ 2004, ApJS, 155, 675
2429: \reference{K07} Kataoka, J., Reeves, J.N., Iwasawa, K.\ \et\ 2007, PASJ, 59, 279
2430: \reference{K01} Kim, C. 2001, JApA, 22, 283
2431: \reference{K91} Kinney, A., Antonucci, R.R.J., Ward, M.J., Wilson, A.S., Whittle, M. 1991, ApJ, 377, 100
2432: \reference{KF97a} Komossa, S.\ \& Fink, H.\ 1997a, A\&A, 322, 719
2433: \reference{KF97b} Komossa, S.\ \& Fink, H.\ 1997b, A\&A, 327, 483
2434: \reference{K01} Komossa, S. 2002, Workshop on X-ray Spectroscopy of AGN with Chandra and XMM-Newton, held at MPE Garching, December 3-6, 2001, MPE Report 279, p.\ 113
2435: \reference{KK00} Kortright, J.B., \& Kim, S.-K. 2000, Phys. Rev. B, 62, 12216
2436: \reference{K00} Kraemer, S.\ \et\ 2000, ApJ, 535, 53
2437: \reference{Kr07} Krongold, Y., Nicastro, F., Elvis, M., Brickhouse, N., Binette, L., Mathur, S.\ \& Jim\'{e}nez-Bail\'{o}n, E. 2007, ApJ, 659, 1022
2438: \reference{K95} Kukula, M.J.\ \et\ 1995, MNRAS, 276, 1262
2439: \reference{L91} Laor, A.\ 1991, ApJ, 376, 90
2440: \reference{La03} Lamer, G., Uttley, P.\ \& M$^{\rm c}$Hardy, I.M. 2003, MNRAS, 342, L41
2441: \reference{L01} Lee, J.C., Ogle, P.M., Canizares, C., Marshall, H.,Schulz, N.S., Morales, R., Fabian, A.C.\ \& Iwasawa, K. 2001, ApJ, 544, L13 %%%%% Chan MCG-6 DWA,
2442: \reference{L07} Liu, B.F., Taam, R., Meyer-Hofmeister, E.\ \& Meyer, F. 2007, ApJ, 671, 695
2443: \reference{Lo03} Lodders, K.\ 2003, ApJ, 591, 1220
2444: \reference{Ly97} Lyubarskii, Y.E. 1997, MNRAS, 292, 679
2445: \reference{MZ95} Magdziarz, P.\ \& Zdziarski, A. 1995, MNRAS, 273, 837
2446: \reference{M98} Magdziarz, P., Blaes, O., Zdziarski, A.A., Johnson, W.N.\ \& Smith, D.A. 1998, MNRAS, 301, 179
2447: \reference{MEV03} Markowitz, A., Edelson, R.\ \& Vaughan, S. 2003, ApJ, 598, 935
2448: \reference{M06} Markowitz, A., Reeves, J.N.\ \& Braito, V.\ 2006, ApJ, 646, 783
2449: \reference{M07} Markowitz, A., Takahashi, T., Watanabe, S.\ \et\ 2007, ApJ, 665, 209
2450: \reference{M08} Markowitz, A., Reeves, J.N., Miniutti, G., Serlemitsos, P., Kunieda, H., Yaqoob, T., Fabian, A.C., Fukazawa, Y., et al.\ 2008, PASJ, 60S, 277
2451: \reference{Mars08} Marscher, A.\ \et\ 2008, Nature, 452, 966
2452: \reference{MM83} Morrison, R.\ \& McCammon, D. 1983, ApJ, 270, 119
2453: \reference{M00} Mould, J.R., Huchra, J.P., Freedman, W.\ \et\ 2000, ApJ, 529, 786
2454: \reference{M95b} Mundell, C., Pedlar, A., Axon, D.J., Meaburn, J.\ \& Unger, S.W. 1995b, MNRAS, 277, 641
2455: \reference{M95a} Mundell, C., Holloway, A., Pedlar, A., Meaburn, J., Kukula, M.J.\ \& Axon, D.J., 1995a, MNRAS, 275, 67
2456: \reference{M96} Murphy, E.M., Lockman, F.J., Laor, A.\ \& Elvis, M. 1996, ApJS, 105, 369
2457: \reference{N00} Narayan, R., Igumenshchev, I.V.\ \& Abramowicz, M.A. 2000, ApJ, 539, 798
2458: \reference{NY94} Narayan, R.\ \& Yi, I. 1994, ApJ, 428, L13
2459: \reference{NY95} Narayan, R.\ \& Yi, I. 1995, ApJ, 452, 710
2460: \reference{N90} Netzer, H. 1990, in Active Galactic Nuclei, ed.\ R.D.\ Blandford, H.\ Netzer, \& L.\ Woltjer (Berlin: Springer), 107
2461: \reference{N96} Netzer, H. 1996, ApJ, 473, 781
2462: \reference{NM90} Netzer, H., Maoz, D., Laor, A.\ \et\ 1990, ApJ, 353, 108
2463: \reference{N94} Netzer, H., Turner, T.J.\ \& George, I.M. 1994, ApJ, 435, 106
2464: \reference{P98} Peterson, B. \et\ 1998, \pasp, 110, 660
2465: \reference{P04} Peterson, B.M., Ferrarese, L., Gilbert, K.M., Kaspi, S., Malkan, M., Maoz, D., Merritt, D., Netzer, H.\ et al.\ 2004, ApJ, 613, 682
2466: \reference{P07} Pollack, A.M.T.\ \et\ 2007, "Status of the RGS Calibration," XMM-SOC-CAL-TN-0030
2467: \reference{PD00} Porquet, D.\ \& Dubau, J. 2000, A\&AS, 143, 495
2468: \reference{P04} Porquet, D., Reeves, J.N., Uttley, P., Turner, T.J.\ 2004, A\&A, 427, 101
2469: \reference{P07} Porter, R.\ \& Ferland, G. 2007, ApJ, 664, 586
2470: \reference{P02} Protassov, R., van Dyk, D.A., Connors, A., Kashyap, V.L.\ \& Siemiginowska, A. 2002, ApJ, 571, 545
2471: \reference{P94} Ptak, A., Yaqoob, T., Serlemitsos, P., Mushotzky, R.\ \& Otani, C. 1994, ApJ, 436, L31
2472: \reference{P07} Puccetti, S., Fiore, F., Risaliti, G., Capalbi, M., Elvis, M.\ \& Nicastro, F. 2007, MNRAS, 377, 607
2473: \reference{R97} Reynolds, C.S. 1997, MNRAS, 286, 513
2474: \reference{R05} Risaliti, G., Elvis, M., Fabbiano, G., Baldi, M., Zezas, A.\ 2005, ApJ, 623, L93
2475: \reference{REN02} Risaliti, G., Elvis, M., Nicastro, F.\ 2002, ApJ, 571, 234
2476: \reference{RR05} Ross, R.R., Fabian, A.C.\ 2005, MNRAS, 358, 211
2477: \reference{R98} Rothschild, R.\ \et\ 1998, ApJ, 496, 538
2478: \reference{RF68} Rubin, V.C.\ \& Ford, W.K. 1968, ApJ, 154, 431
2479: \reference{SK96} Schmitt, H.\ \& Kinney, A. 1996, ApJ, 463, 498
2480: \reference{S02} Schulz, N.S., Cui, W., Canizares, C.R., Marshall, H.L., Lee, J.C., Miller, J.M.\ \& Lewin, W.H.G. 2002, ApJ, 565, 1141
2481: \reference{SS73} Shakura, N.I.\ \& Sunyaev, R.A. 1973, A\&A, 24, 337
2482: \reference{S85} Shull, J.M.\ \& van Steenburg, M.E. 1985, ApJ, 294, 599
2483: \reference{SV07} Smith, R.\ \& Vaughan, S. 2007, MNRAS, 375, 1479
2484: \reference{S95} Snow, T.P.\ \& Witt, A.N. 1996, ApJ, 468, L65
2485: \reference{S01} Str\"{u}der, L.\ \et\ 2001, A\&A, 365, L18
2486: \reference{Suga} Suganuma, M., Yoshii, Y., Kobayashi, Y., Minezaki, T., Enya, K., Tomita, H., Aoki, T., Koshida, K.\ \& Peterson, B.A. 2006, ApJ, 639, 46
2487: \reference{ST80} Sunyaev, R.\ \& Titarchuk, L. 1980, A\&A, 86, 121
2488: \reference{Sw98} Swank, J.\ 1998, in Nuclear Phys.\ B (Proc.\ Suppl.): The Active X-ray Sky: Results From BeppoSAX and Rossi-XTE, Rome, Italy, 1997 October 21-24, eds.\ L.\ Scarsi, H.\ Bradt, P.\ Giommi \& F.\ Fiore, Nucl.\ Phys.\ B Suppl.\ Proc.\ (The Netherlands: Elsevier Science B.V.), 69, 12
2489: \reference{T95} Tanaka, Y.\ \et\ 1995, Nature, 375, 659
2490: \reference{Ti94} Titarchuk, L., 1994, ApJ, 434, 313
2491: \reference{Tm07} Tombesi, F., de Marco, B., Iwasawa, K., Cappi, M., Dadina, M., Ponti, G., Miniutti, G.\ \& Palumbo, G.G.C., 2007, A\&A, 467, 1057
2492: \reference{T01} Turner, M.J.L.\ \et\ 2001, A\&A, 365, L27
2493: \reference{T08} Turner, T.J., Reeves, J.N., Kraemer, S.B.\ \& Miller, L. 2008, A\&A, 483, 161
2494: \reference{UP95} Urry, C.M.\ \& Padovani, P. 1995, PASP, 107, 803
2495: \reference{U05} Uttley, P.\ \& M$^{\rm c}$Hardy, I.M. 2005, MNRAS, 363, 586
2496: \reference{V02} Vaughan, S., Boller, Th., Fabian, A.C., Ballantyne, D.R., Brandt, W.N.\ \& Tr\"{u}mper, J. 2002, MNRAS, 337, 247
2497: \reference{V03} Vaughan, S., Edelson, R., Warwick, R.\ \& Uttley, P. 2003, MNRAS, 345, 1271
2498: \reference{VF04} Vaughan, S.\ \& Fabian, A.C. 2004, MNRAS, 348, 1415
2499: %%%%%\reference{VFN03} Vaughan, S., Fabian, A.C.\ \& Nandra, K. 2003b, MNRAS, 339, 1237
2500: \reference{V08} Vaughan, S.\ \& Uttley, P. 2008, MNRAS, submitted
2501: \reference{V73} Veigele, W.M.\ 1973, Atomic Data Tables, 5, 51
2502: \reference{WPM} Wandel, A., Peterson, B.M.\ \& Malkan, M.A. 1999, ApJ, 526, 579
2503: \reference{W99} Welsh, W. 1999, PASP, 111, 1347
2504: \reference{W94} White, R.\ \& Peterson, B.M.\ 1994, PASP, 106, 879
2505: \reference{W95} Winge, C., Peterson, B.M., Horne, K., Pogge, R.W., Pastoriza, M.G.\ \& Storchi-Bergmann, T. 1995, ApJ, 445, 680
2506: \reference{W02} Woo, J.-H.\ \& Urry, C.M.\ 2002, ApJ, 579, 530
2507: \reference{W98} Wozniak, P.R., Zdziarski, A.A., Smith, D., Madejski, G.\ \& Johnson, W.N. 1998, MNRAS, 299, 449
2508: \reference{WG08} Wu, Q.\ \& Gu, M. 2008, ApJ, 682, 212
2509: %%%%%%\reference{Y03} Yaqoob, T., George, I.M., Kallman, T.., Padmanabhan, U., Weaver, K.\ \& Turner, T.J. 2003, ApJ, 596, 85
2510: \end{references}
2511:
2512:
2513: %%%%%\reference{DG} DeZotti \& Gaskell '85
2514: %%%%%% \reference{PS96} Poutanen, J.\ \& Svensson, R. 1996, ApJ, 470, 249
2515: %%%%%% \reference{Tk08} Takahashi, H., Fukazawa, Y., Mizuno, T.\ \et\ 2008, PASJ, 60S, 69
2516: %%%%%\reference{YP04} Yaqoob, T.\ \& Padmanabhan, U. 2004, ApJ, 604, 63
2517:
2518: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% TABLES
2519:
2520: \clearpage
2521:
2522: \begin{deluxetable}{llc}
2523: \tabletypesize{\footnotesize}
2524: \tablewidth{5.0in}
2525: \tablenum{1}
2526: \tablecaption{Fe K bandpass components
2527: in the time-averaged EPIC pn spectrum\label{tab1}}
2528: \tablehead{
2529: \colhead{Component} & \colhead{Parameter} & \colhead{Value}}
2530: \startdata
2531: & $\chi^2$/$dof$ & 983.2/990 \\
2532: Fe K$\alpha$ emission line & Energy (keV) & 6.403 $\pm$ 0.009 \\
2533: & Width $\sigma$ (eV) & 65 $\pm$ 14 \\
2534: & Intensity (ph cm$^{-2}$ s$^{-1}$)& $3.5 \pm 0.4 \times 10^{-5}$ \\
2535: & $EW$ (eV) & 91 $\pm$ 10 \\
2536: Ni K$\alpha$ emission line & Energy (keV) & $7.41^{+0.08}_{-0.07}$ \\
2537: & Intensity (ph cm$^{-2}$ s$^{-1}$) & $4^{+4}_{-3} \times 10^{-6}$ \\
2538: & $EW$ (eV) & $13^{+13}_{-10}$ \\
2539: 6.04 keV emission line & Energy (keV) & $6.04^{+0.20}_{-0.06}$ \\
2540: & Width $\sigma$ (eV) & $<$ 200 \\
2541: & Intensity (ph cm$^{-2}$ s$^{-1}$) & $9.1 \pm 2.5 \times 10^{-6}$ \\
2542: & $EW$ (eV) & 22 $\pm$ 6 \\
2543: Fe K Edge & Energy (keV) & 7.11 (fixed) \\
2544: & Optical Depth $\tau$ & 0.05 $\pm$ 0.03 \\
2545: \enddata
2546: \tablecomments{Parameters for the best-fitting
2547: ``GA'' model (wherein the emission near 6.0 keV is modeled with a narrow Gaussian component),
2548: fit to $>$4 keV data.
2549: The model includes a Fe K$\beta$ line, with energy fixed at 7.056 keV
2550: (rest frame), width $\sigma$ tied to that of the K$\alpha$
2551: line, and intensity equal to 0.13 times that of the K$\alpha$ line.
2552: The width $\sigma$ for the Ni K$\alpha$ line was also
2553: tied to that for the Fe K$\alpha$ line.
2554: Equivalent widths $EW$ were determined relative to a locally-fit continuum. }
2555: \end{deluxetable}
2556:
2557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2558:
2559: \begin{deluxetable}{llccc}
2560: \tabletypesize{\footnotesize}
2561: \tablewidth{6.5in}
2562: \tablenum{2}
2563: \tablecaption{Model fits to the 0.2--10 keV time-averaged EPIC pn spectrum\label{tab2}}
2564: \tablehead{
2565: \colhead{Component} & \colhead{Parameter} & \colhead{SXPL Model} & \colhead{BB Model} & \colhead{COMPST Model} }
2566: \startdata
2567: & $\chi^2$/$dof$ & 1866.9/1724 & 1870.4/1724 & 1863.4/1723 \\
2568: Local cold absorption & $N_{\rm H,local}$ (cm$^{-2}$) & $8.7^{+0.6}_{-0.5} \times 10^{20}$ & $5.1^{+1.2}_{-0.2} \times 10^{20}$ & $7.7^{+0.7}_{-0.3} \times 10^{20}$ \\
2569: Hard X-ray power-law & $\Gamma_{\rm HX}$ & 1.57 $\pm$ 0.02 & 1.61$^{+0.01}_{-0.02}$ & 1.65$^{+0.02}_{-0.03}$ \\
2570: & Norm.\ (1 keV) & $6.7^{+0.3}_{-0.2} \times 10^{-3}$ & $7.4 \pm 0.1 \times 10^{-3}$ & $8.1^{+0.2}_{-0.5} \times 10^{-3}$ \\
2571: Hard X-ray Absorption & $N_{\rm H,HX}$ (cm$^{-2}$) & $2.9^{+0.3}_{-0.8} \times 10^{21}$ & $< 6.1 \times 10^{20}$ & $4.2^{+0.3}_{-1.1} \times 10^{21}$ \\
2572: Low-ioniz.\ X-ray Warm Abs.\ & $N_{\rm H,lo}$ (cm$^{-2}$) & $1.0^{+0.3}_{-0.1} \times 10^{21}$ & $7.9^{+2.7}_{-0.7} \times 10^{20}$ & $9.3^{+2.1}_{-1.9} \times 10^{20}$ \\
2573: & log$\xi_{\rm lo}$ (erg cm s$^{-1}$) & $1.45^{+0.16}_{-0.07}$ & 1.38 $\pm$ 0.19 & $1.43^{+0.20}_{-0.43}$ \\
2574: High-ioniz.\ X-ray Warm Abs.\ & $N_{\rm H,hi}$ (cm$^{-2}$) & $1.8^{+1.2}_{-0.6} \times 10^{21}$ & $1.8^{+1.3}_{-0.4}\times 10^{21}$ & $1.5^{+1.3}_{-0.6} \times 10^{21}$ \\
2575: & log$\xi_{\rm hi}$ (erg cm s$^{-1}$) & $2.93^{+0.15}_{-0.09}$ & 2.91$^{+0.15}_{-0.07}$ & $2.89^{+0.14}_{-0.20}$ \\
2576: %%%%%%%%%Fe K edge & optical depth $\tau$ & 0.06$\pm$0.02 & $<$0.06 & $<$0.04 \\
2577: Soft X-ray power-law & $\Gamma_{\rm SX}$ & $3.35^{+0.27}_{-0.10}$ & & \\
2578: & Norm.\ (1 keV) & $4.0^{+0.2}_{-0.7} \times 10^{-3}$ & & \\
2579: Blackbody & Temperature $k_{\rm B}T$ (eV) & & $83^{+1}_{-4}$ & \\
2580: & Norm.\ & & $1.72^{+0.37}_{-0.11} \times 10^{-4}$ & \\
2581: Comptonized component & Temperature $k_{\rm B}T$ (keV) & & & $0.35^{+0.02}_{-0.03}$ \\
2582: & Optical Depth $\tau$ & & & 24$^{+2}_{-4}$ \\
2583: & Norm.\ & & & $4.41^{+0.22}_{-0.33} \times 10^{-3}$ \\
2584: \ion{O}{7} emission line & Energy (keV) & $0.58 \pm 0.01$ & $0.58^{+0.02}_{-0.01}$ & $0.58 \pm 0.01$ \\
2585: & Intensity (ph cm$^{-2}$s $^{-1}$) & $2.4^{+0.5}_{-0.4} \times 10^{-4}$ & $1.2^{+0.6}_{-0.5} \times 10^{-4}$ & $2.1 \pm 0.4 \times 10^{-4}$ \\
2586: \enddata
2587: \tablecomments{SXPL,
2588: BB and COMPST refer to the models wherein the soft excess is modeled with
2589: a steep power-law component, a blackbody component, and a Comptonized component
2590: ({\sc CompST}), respectively. The units of normalization of the power-law
2591: components are ph keV$^{-1}$ cm$^{-2}$ s$^{-1}$ at 1 keV.
2592: The blackbody normalization is $L_{39}D_{10}^{-2}$, where $L_{39}$ is the source luminosity in
2593: units of 10$^{39}$ erg s$^{-1}$ and $D_{10}$ is the distance to
2594: the source in units of 10 kpc.
2595: See the {\sc xspec} user manual for units of the {\sc CompST} component normalization.
2596: The reader is reminded that the BB and COMPST models
2597: are not generally accepted as physically plausible descriptions
2598: of soft excesses in Seyferts.
2599: The uncertainties listed on the intensity of the \ion{O}{7} emission line
2600: are statistical only and do not
2601: include likely systematic uncertainties due
2602: to blending with narrow absorption lines.
2603: The width $\sigma$ of the \ion{O}{7} emission line was kept fixed at 0.5 eV.}
2604: \end{deluxetable}
2605: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2606:
2607:
2608: %%%%\clearpage
2609:
2610:
2611: \begin{deluxetable}{lcccc}
2612: \tabletypesize{\footnotesize}
2613: \tablewidth{5.0in}
2614: \tablenum{3}
2615: \tablecaption{Emission lines fit to the RGS spectrum\label{tab3}}
2616: \tablehead{
2617: \colhead{Line} & \colhead{Energy} & \colhead{Intensity} & \colhead{} & \colhead{$F$-test} \\
2618: \colhead{Identification} & \colhead{(eV)} & \colhead{(ph cm$^{-2}$ s$^{-1}$)} & \colhead{$\Delta\chi^2$} & \colhead{Probability} }
2619: \startdata
2620: \ion{N}{6} {\it (r)} & 431 $\pm$ 1 & $9.9^{+6.7}_{-5.0} \times 10^{-5}$ & --12.0 & 0.042 \\
2621: \ion{N}{7} & $499^{+1}_{-2}$ & $8.2^{+6.0}_{-5.2} \times 10^{-5}$ & --10.4 & 0.064 \\
2622: \ion{O}{7} {\it (f)} & $561^{+1}_{-2}$ & $1.33^{+0.13}_{-0.75} \times 10^{-4}$ & --25.2 & $1.4 \times 10^{-3}$ \\
2623: \ion{O}{7} {\it (i)} & 567 $\pm$ 1 & $9.5^{+3.2}_{-3.5} \times 10^{-5}$ & --18.5 & $7.8 \times 10^{-3}$ \\
2624: \ion{O}{7} {\it (r)} & $572^{+1}_{-2}$ & $4.9^{+12.3}_{-2.9} \times 10^{-5}$ & --7.9 & 0.12 \\
2625: \enddata
2626: \tablecomments{Best-fit line emission parameters using Gaussian components.
2627: All line widths $\sigma$ were kept fixed at 0.5 eV.
2628: The uncertainties listed on the intensities (Col.\ [3])
2629: are statistical only and do not include likely systematic uncertainties due
2630: to blending with narrow absorption lines.
2631: The $F$-test probability values $P$ (Col.\ [5]) denote the
2632: changes that the null hypothesis (line not included in model) is valid;
2633: the line detection significance is 1 -- $P$.}
2634: \end{deluxetable}
2635:
2636: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% Table 4 Best-fit model parameters for the RGS spectrum
2637:
2638:
2639:
2640: \begin{deluxetable}{llc}
2641: \tabletypesize{\footnotesize}
2642: \tablewidth{5.0in}
2643: \tablenum{4}
2644: \tablecaption{Best-fit model parameters for the RGS spectrum\label{tab4}}
2645: \tablehead{
2646: \colhead{Component} & \colhead{Parameter} & \colhead{Value}}
2647: \startdata
2648: & $\chi^2$/$dof$ & 785.7/417 \\
2649: Local cold absorption & $N_{\rm H,local}$ (cm$^{-2}$) & $10.5^{+1.2}_{-1.9} \times 10^{20}$ \\
2650: Soft X-ray power-law & $\Gamma_{\rm SX}$ & 3.00$\pm$0.25 \\
2651: & Norm.\ (1 keV) & $6.2^{+0.3}_{-0.7} \times 10^{-3}$ \\
2652: Low-ioniz.\ X-ray Warm Abs.\ & $N_{\rm H,lo}$ (cm$^{-2}$) & $1.1^{+0.1}_{-0.2} \times 10^{21}$ \\
2653: & log$\xi_{\rm lo}$ (erg cm s$^{-1}$) & $1.21^{+0.18}_{-0.08}$ \\
2654: & $z_{\rm lo}$ (absolute) & $+0.00246^{+0.00144}_{-0.00064}$ \\
2655: & $z_{\rm lo}$ (rel.\ to systemic) & $-0.00140^{+0.00144}_{-0.00064}$ \\
2656: High-ioniz.\ X-ray Warm Abs.\ & $N_{\rm H,hi}$ (cm$^{-2}$) & $2.4^{+2.0}_{-1.2} \times 10^{21}$ \\
2657: & log$\xi_{\rm hi}$ (erg cm s$^{-1}$) & $2.90^{+0.21}_{-0.26}$ \\
2658: & $z_{\rm hi}$ (absolute) & $-0.00302^{+0.00057}_{-0.00080}$ \\
2659: & $z_{\rm hi}$ (rel.\ to systemic) & $-0.00688^{+0.00057}_{-0.00080}$ \\
2660: \enddata
2661: \tablecomments{Results are shown for the best-fit soft X-ray power-law model, including
2662: five emission lines due to He-like O and H- and He-like N; see Table~3
2663: for emission line parameters. The units of normalization of the power-law
2664: components are ph keV$^{-1}$ cm$^{-2}$ s$^{-1}$ at 1 keV.
2665: See text for details regarding uncertainty range in $z$.
2666: The model also included a hard X-ray power-law component
2667: absorbed by a column $N_{\rm H,HX} \sim 8 \times 10^{21}$ cm$^{-2}$ and
2668: emerging only above $\sim$ 1.5 keV.}
2669: \end{deluxetable}
2670:
2671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIGURES
2672:
2673:
2674: %%%%%%%%%%%%%%%%%%%%%%% Figure 1 = light curves
2675: \begin{figure}
2676: \epsscale{0.60}
2677: \plotone{f1.eps}
2678: \caption{The top three panels show the EPIC-pn count rate light curves
2679: for the 0.2--10, 0.2--1, and 3--10 keV bandpasses, binned to 600 s. The bottom
2680: panel shows the OM UVW1 light curve, binned to 1400 s.}
2681: \end{figure}
2682:
2683:
2684: %%%%%%%%%%%%%%%%%%%%%% Figure 2 = Data/model residuals PN FeK bandpass
2685: \begin{figure}
2686: \epsscale{0.75}
2687: \plotone{f2.eps}
2688: \caption{Data/model residuals to spectral fits to the Fe K bandpass
2689: of the EPIC-pn spectrum. Data are rebinned by a factor of 12. Vertical dashed lines
2690: denote energies of 6.04, 6.40, 7.06, and 7.37 keV.
2691: Panel {\it a)} shows residuals to a simple power-law model.
2692: In panel {\it b)}, the Fe K$\alpha$ and Fe K$\beta$ emission lines have been
2693: modeled. The 7.11 keV edge and Ni K$\alpha$ line
2694: have been modeled in panels {\it c)} and {\it d)}, respectively.
2695: In panel {\it e)}, a Gaussian with an energy centroid of 6.24 keV has
2696: been added. However, as shown in panel {\it f)}, the best-fit ``GA'' model,
2697: with a Gaussian with an
2698: energy centroid at 6.04 keV, does a superior job in modeling the remaining
2699: $\sim$6 keV residuals. The small dip near 6.8 keV is not
2700: significant and is likely an artifact of fitting.
2701: In panel {\it g)}, the best-fit ``DL'' model,
2702: the 6.04 keV line has been
2703: removed and replaced with relativistically broadened Fe K$\alpha$ and
2704: Fe K$\beta$ diskline profiles; the narrow Gaussian at 6.04 keV
2705: (panel {\it f)}) seems to model the residuals better. }
2706: \end{figure}
2707:
2708: %%%%%%% Fig 3 = results of sliding gaussian for Fe K bandpass. (pn)
2709:
2710: \begin{figure}
2711: \epsscale{0.50}
2712: \plotone{f3.eps}
2713: \caption{Contour plots showing the results of applying a
2714: ``sliding Gaussian'' (with width $\sigma$ fixed at 10 eV)
2715: to a model to the EPIC-pn data
2716: consisting of a power-law plus a narrow Gaussian
2717: at 6.40 keV to model Fe K$\alpha$ emission.
2718: 1-, 2-, 3- and 4-$\sigma$
2719: confidence levels for two interesting parameters are denoted by solid, dashed, dot-dashed, and dotted lines,
2720: respectively.
2721: Residuals near 6.0 keV are clear; additional residuals
2722: at 7.0 and 7.4 keV (above the Fe K edge at 7.11 keV)
2723: are investigated further in $\S$3.}
2724: \end{figure}
2725:
2726: %%%% Fig 4 = contour plot of 6.04 keV intensity versus energy (pn)
2727: \begin{figure}
2728: \epsscale{0.50}
2729: \plotone{f4.eps}
2730: \caption{Contour plot of intensity versus line centroid energy
2731: for the Gaussian used to model the narrow 6.0 keV emission feature in the pn spectrum.
2732: The width $\sigma$ was left as a free parameter. 1-, 2-, 3- and 4-$\sigma$
2733: confidence levels for two interesting parameters are denoted by solid, dashed, dot-dashed, and dotted lines,
2734: respectively.}
2735: \end{figure}
2736:
2737:
2738:
2739: %%%%%%%%%%%%%%%%%%%%%%%%%% Fig 5 MOS data/model residuals
2740: \begin{figure}
2741: \epsscale{0.75}
2742: \plotone{f5.eps}
2743: \caption{Data/model residuals to various model fits to the MOS 1+2 spectrum.
2744: Data are rebinned by a factor of 8.
2745: Panel {\it a)} shows residuals to a simple power-law model.
2746: In panel {\it b)}, the Fe K$\alpha$ and Fe K$\beta$ emission lines have been
2747: modeled. In panels {\it c)} and {\it d)}, respectively, the 6.11 keV emission
2748: line and the Fe K edge have been added.}
2749: \end{figure}
2750:
2751:
2752: %%%%%%%%%%%%%%%%%%%%%%%%%% Fig 6 MOS sliding Gaussian
2753: \begin{figure}
2754: \epsscale{0.60}
2755: \plotone{f6.eps}
2756: \caption{Same as Figure 3, but for the MOS 1+2 spectral fitting.}
2757: \end{figure}
2758:
2759:
2760: %%%%%%%%%%%%%%%%%%%%%%%%%% Fig 7: MOS 6.11 keV line: Contour of Intensity vs Energy
2761: \begin{figure}
2762: \epsscale{0.60}
2763: \plotone{f7.eps}
2764: \caption{Contour plot of intensity versus line centroid energy
2765: for the Gaussian used to model the narrow 6.1 keV emission feature in the MOS 1+2 spectrum.
2766: The width $\sigma$ was left as a free parameter. 1-, 2-, 3- and 4-$\sigma$
2767: confidence levels for two interesting parameters are denoted by solid, dashed, dot-dashed, and dotted lines,
2768: respectively.}
2769: \end{figure}
2770:
2771:
2772: %%%%% Figure 8: PN Broadband residuals %%%%%%%%%%%%%%%%%%%%%%
2773: \begin{figure}
2774: \epsscale{0.65}
2775: \plotone{f8.eps}
2776: \caption{Data/model residuals to spectral fits to the 0.2--10 keV
2777: EPIC-pn spectrum. All models include narrow Gaussian components
2778: to model the Fe K$\alpha$, Fe K$\beta$, Ni K$\alpha$ and 6.04 keV
2779: emission lines, an Fe K edge at 7.11 keV, and a column of neutral gas to model
2780: Galactic absorption.
2781: Panel {\it a)} shows the results when a simple hard X-ray
2782: power-law component (HXPL) is used. In Panel {\it b)}, a
2783: neutral absorbing column has
2784: been added to the HXPL, and a steep, soft X-ray power-law
2785: component (SXPL) has been added.
2786: In panels {\it c)} and {\it d)}, the low- and hi-ionization
2787: warm absorbers, respectively, have been been included.
2788: In panel {\it e)}, a narrow Gaussian to model \ion{O}{7} emission
2789: has been included; this is the best-fitting ``SXPL'' model.
2790: In panel {\it f)}, the soft power-law has been replaced with a
2791: blackbody component; this is the best-fitting ``BB'' model.
2792: Panel {\it g)} shows the results from modeling the soft excess
2793: as blurred reflection from an ionized disk.
2794: (The small dip near 6.8 keV is not
2795: significant and is likely an artifact of fitting.)}
2796: \end{figure}
2797:
2798: \clearpage
2799:
2800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% Fig 9 = Plot euf for EPIC models, + hi/lo
2801: \begin{figure}
2802: \epsscale{0.75}
2803: \plotone{f9.eps}
2804: \caption{Unfolded spectra for the best-fitting SXPL, BB and COMPST models.
2805: Panel {\it d)} shows the spectra for three selected time-resolved
2806: segments, further illustrating the increase in soft X-ray flux
2807: between segments 1, 2 and 6.}
2808: \end{figure}
2809:
2810:
2811:
2812: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% Fig 10 = RGS plot with Fe M UTA inset
2813:
2814: \begin{figure}
2815: \epsscale{0.75}
2816: \plotone{f10.eps}
2817: \caption{The RGS spectrum for NGC 3227.
2818: The solid line shows the best-fitting unfolded model,
2819: described in $\S$5. Data have been rebinned by a factor of 2, and are plotted
2820: in the rest frame. Black and red data points denote RGS 1 and 2, respectively.
2821: The vertical lines, which indicate identified lines and edges, denote
2822: rest-frame (systemic) energies.
2823: %%%%
2824: %%%%368 eV = C 6 Ly $\alpha$;
2825: %%%%420 eV = N VI (f);
2826: %426 eV = N VI (i);
2827: %431 eV = N VI (r);
2828: %436 eV = C VI Ly$\beta$;
2829: %500 eV = N VII Ly$\alpha$;
2830: %561 eV = O VII (f);
2831: %568 eV = O VII (i);
2832: %574 eV = O VII (r);
2833: %653 eV = O VIII Ly$\alpha$;
2834: %666 eV = O VII He$\beta$;
2835: %698 eV = O VII He$\gamma$ (?) ;
2836: %774 eV = O VIII Ly$\beta$;
2837: %812 eV = Fe L XVII 3d--2p;
2838: %905 eV = Ne IX (f);
2839: %915 eV = Ne IX (i);
2840: %922 eV = Ne IX (r);
2841: %1022 eV = Ne X Ly$\alpha$;
2842: %1053 eV = Fe L XXII;
2843: %1331 eV = Mg XI (f);
2844: %1352 eV = Mg XI (r).
2845: %Orange vertical dashed lines denote the
2846: %rest-frame (systemic) energies of these
2847: %expected absorption edges, none of which are detected in excess of
2848: %the best-fit continuum model:
2849: %543 eV = O I edge (for oxiding O);
2850: %703 eV = Fe${\rm X}$O$_{\rm Y}$ edge;
2851: %707 eV = Fe I L3 edge;
2852: %720 eV = Fe I L2 edge;
2853: %739 eV = O VII edge;
2854: %845 eV = Fe I L1 edge;
2855: %871 eV = O VIII edge.
2856: The residuals near 687 and 850 eV are
2857: instrumental artifacts.}
2858: \end{figure}
2859:
2860: %%%%%%%%%%%%%%%%%%%%%%% Figure 11 = PCA + HEXTE + BAT
2861: \begin{figure}
2862: \epsscale{0.60}
2863: \plotone{f11.eps}
2864: \caption{The upper panel shows the {\it RXTE}-PCA (black),
2865: {\it RXTE}-HEXTE cluster A (red) and B (purple), and
2866: {\it Swift}-BAT data (blue). The middle and lower panels show
2867: $\chi$ residuals to the best-fit models with and without a
2868: high-energy cutoff in the power-law component, respectively.
2869: In each case, the model
2870: contains a Compton reflection component and Fe K$\alpha$ emission.}
2871: \end{figure}
2872:
2873:
2874: %%%%%%%%%%%%%%%%%%%%%%% Figure 12 RECUT for RXTE+BAT spectrum.
2875: \begin{figure}
2876: \epsscale{0.40}
2877: \plotone{f12.eps}
2878: \caption{Contour plots showing the strength of the Compton reflection
2879: component, $R$ (measured via a {\sc pexrav} component; top panel)
2880: and cutoff energy (bottom) versus hard X-ray photon index from spectral
2881: fits to {\it RXTE} PCA and HEXTE and {\it Swift}-BAT data.
2882: 1-, 2-, 3- and 4-$\sigma$
2883: confidence levels for two interesting parameters
2884: are denoted by solid, dashed, dot-dashed, and dotted lines,
2885: respectively.}
2886: \end{figure}
2887:
2888:
2889:
2890:
2891: %%%%%%%%%%%%%%%%%%%%%%% Figure 13 = Results of TR-fitting with SXPL
2892: \begin{figure}
2893: \epsscale{0.50}
2894: \plotone{f13.eps}
2895: \caption{The results of fitting the SXPL model to the ten time-resolved spectral segments.
2896: From top to bottom: Normalization of the soft X-ray power-law component,
2897: $\Gamma_{\rm SX}$, normalization of the hard X-ray power-law component, and
2898: $\Gamma_{\rm HX}$.}
2899: \end{figure}
2900:
2901:
2902:
2903: %%%%%%%%%%%%%%%%% Fig 14 = RMS spectra
2904: \begin{figure}
2905: \epsscale{0.50}
2906: \plotone{f14.eps}
2907: \caption{$F_{\rm var}$ spectra derived from EPIC-pn data
2908: for the entire duration (black), the first 20 ks (red) and the
2909: final 80 ks (blue).}
2910: \end{figure}
2911:
2912: %%%%%%%%%%%%%%%%%%%%%%% Figure 15 = XTE medium- and long-term light curves
2913: \begin{figure}
2914: \epsscale{0.75}
2915: \plotone{f15.eps}
2916: \caption{The results of time-resolved spectral fits to the medium-term (top) and long-term
2917: (bottom) {\it RXTE} data, showing 2--10 keV continuum flux in units of 10$^{-11}$ erg cm$^{-2}$ s$^{-1}$,
2918: Fe K line intensity $I_{\rm Fe K\alpha}$ in units of 10$^{-5}$ ph cm$^{-2}$ s$^{-1}$, and
2919: $\Gamma_{\rm HX}$.}
2920: \end{figure}
2921:
2922: %%%%%%%%%%%%%%%%%%%%%%% Figure 16 = RXTE zero-lag correlation diagrams
2923: \begin{figure}
2924: \epsscale{0.40}
2925: \plotone{f16.eps}
2926: \caption{Zero-lag correlation diagrams in which Fe line intensity $I_{\rm Fe K\alpha}$,
2927: in units of 10$^{-5}$ ph cm$^{-2}$ s$^{-1}$, is plotted against
2928: 2--10 keV continuum flux in units of 10$^{-11}$ erg cm$^{-2}$ s$^{-1}$.
2929: Dashed lines indicate the best-fitting linear model.}
2930: \end{figure}
2931:
2932: %%%%%%%%%%%%%%%%%%%%%%%%%%%% Fig 17 = CCF plot from Rebecca
2933:
2934: \begin{figure}
2935: \epsscale{0.40}
2936: \plotone{f17.eps}
2937: \caption{ICF (red dashed line) and DCF (solid black line)
2938: functions for the 0.2--1 keV soft X-ray light curve
2939: versus the OM UV continuum light curve (positive lag indicates soft X-ray leading
2940: the UV.)}
2941: \end{figure}
2942:
2943:
2944:
2945: \end{document}
2946:
2947:
2948:
2949:
2950: