1: %Version 18, March. 17, 2008
2: \documentstyle{ar}
3:
4: \newbox\grsign \setbox\grsign=\hbox{$>$} \newdimen\grdimen \grdimen=\ht\grsign
5: \newbox\simlessbox \newbox\simgreatbox
6: \setbox\simgreatbox=\hbox{\raise.5ex\hbox{$>$}\llap
7: {\lower.5ex\hbox{$\sim$}}}\ht1=\grdimen\dp1=0pt
8: \setbox\simlessbox=\hbox{\raise.5ex\hbox{$<$}\llap
9: {\lower.5ex\hbox{$\sim$}}}\ht2=\grdimen\dp2=0pt
10: \def\simgreat{\mathrel{\copy\simgreatbox}}
11: \def\simless{\mathrel{\copy\simlessbox}}
12: \def\nhi{\noindent \hangindent=0.3cm}
13:
14: \input psfig.sty
15:
16: %Luis's definitions
17: \def\aa{{A\&A}}
18: \def\aas{{ A\&AS}}
19: \def\aj{{AJ}}
20: \def\al{$\alpha$}
21: \def\alphaox{$\alpha_{\rm ox}$}
22: \def\bet{$\beta$}
23: \def\amin{$^\prime$}
24: \def\annrev{{ARA\&A}}
25: \def\apj{{ApJ}}
26: \def\apjs{{ApJS}}
27: \def\asec{$^{\prime\prime}$}
28: \def\baas{{BAAS}}
29: \def\cc{cm$^{-3}$}
30: \def\deg{$^{\circ}$}
31: \def\ddeg{{\rlap.}$^{\circ}$}
32: \def\dsec{{\rlap.}$^{\prime\prime}$}
33: \def\cc{cm$^{-3}$}
34: \def\e#1{$\times$10$^{#1}$}
35: \def\etal{{et al. }}
36: \def\farcm{\hbox{$.\mkern-4mu^\prime$}}
37: \def\farcs{\hbox{$.\mkern-4mu^{\prime\prime}$}}
38: \def\flamb{ergs s$^{-1}$ cm$^{-2}$ \AA$^{-1}$}
39: \def\flux{ergs s$^{-1}$ cm$^{-2}$}
40: \def\fnu{ergs s$^{-1}$ cm$^{-2}$ Hz$^{-1}$}
41: \def\hal{H$\alpha$}
42: \def\hst{{\it HST}}
43: \def\kms{km s$^{-1}$}
44: \def\lamb{$\lambda$}
45: \def\lax{{$\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}$}}
46: \def\gax{{$\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}$}}
47: \def\simlt{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}
48: \def\simgt{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}
49: \def\lum{ergs s$^{-1}$}
50: %\def\mbh{{$M_{\bullet}$}}
51: \def\mbh{{$M_{\rm BH}$}}
52: \def\micron{{$\mu$m}}
53: \def\mnras{{MNRAS}}
54: \def\nat{{Nature}}
55: \def\pasj{{PASJ}}
56: \def\pasp{{PASP}}
57: \def\perang{\AA$^{-1}$}
58: \def\percm2{cm$^{-2}$}
59: \def\peryr{yr$^{-1}$}
60: \def\pp{\parshape 2 0truein 6.1truein .3truein 5.5truein}
61: \def\reference{\noindent\pp}
62: \def\refindent{\par\noindent\parskip=2pt\hangindent=3pc\hangafter=1 }
63: \def\solum{$L_\odot$}
64: \def\solmass{$M_\odot$}
65: \def\civ{C~{\sc IV}}
66: \def\fei{Fe~{\sc I}}
67: \def\feii{Fe~{\sc II}}
68: \def\feiii{Fe~{\sc III}}
69: \def\hei{He~{\sc I}}
70: \def\heii{He~{\sc II}}
71: \def\hi{H~{\sc I}}
72: \def\hii{H~{\sc II}}
73: \def\mgii{Mg~{\sc II}}
74: \def\oi{[O~{\sc I}]}
75: \def\oii{[O~{\sc II}]}
76: \def\oiii{[O~{\sc III}]}
77: \def\oiv{[O~{\sc IV}]}
78: \def\ni{[N~{\sc I}]}
79: \def\nii{[N~{\sc II}]}
80: \def\nv{[N~{\sc V}]}
81: \def\neii{[Ne~{\sc II}]}
82: \def\neiii{[Ne~{\sc III}]}
83: \def\neiv{[Ne~{\sc IV}]}
84: \def\nev{[Ne~{\sc V}]}
85: \def\sii{[S~{\sc II}]}
86: \def\siii{[S~{\sc III}]}
87:
88: \def\lhal{$L_{{\rm H}\alpha}$}
89: \def\lbol{$L_{{\rm bol}}$}
90: \def\ledd{$L_{{\rm Edd}}$}
91:
92: \def\mhi{$M_{{\rm H~I}}$}
93: \def\lb{$L_B$}
94: \def\vc{${\upsilon_c}$}
95: \def\vm{${\upsilon_m}$}
96: \def\vrot{$\upsilon_{{\rm rot}}$}
97: \def\sig{$\sigma_0$}
98:
99: \begin{document}
100: \title{Nuclear Activity in Nearby Galaxies}
101: \markboth{Ho}{Nuclear Activity in Nearby Galaxies}
102: \author{Luis C. Ho
103: \affiliation{The Observatories of the Carnegie
104: Institution of Washington, 813 Santa Barbara St., Pasadena, CA 91101;
105: e-mail: lho@ociw.edu}}
106: \begin{keywords}
107: accretion disks, active galactic nuclei, black holes, LINERs, radio galaxies,
108: Seyfert galaxies
109: \end{keywords}
110:
111: %\noindent{Send Proofs to:}
112: %
113: %\noindent{Luis C. Ho} \\
114: %The Observatories of the Carnegie Institution of Washington\\
115: %813 Santa Barbara St. \\
116: %Pasadena, CA 91101 \\
117: %Phone: (626)-304-0248 \\
118: %Fax: (626)-795-8136 \\
119: %e-mail: lho@ociw.edu \\
120:
121: \begin{abstract}
122: A significant fraction of nearby galaxies show evidence of weak nuclear
123: activity unrelated to normal stellar processes. Recent high-resolution,
124: multiwavelength observations indicate that the bulk of this activity derives
125: from black hole accretion with a wide range of accretion rates. The low
126: accretion rates that typify most low-luminosity active galactic nuclei induce
127: significant modifications to their central engine. The broad-line region and
128: obscuring torus disappear in some of the faintest sources, and the optically
129: thick accretion disk transforms into a three-component structure consisting of
130: an inner radiatively inefficient accretion flow, a truncated outer thin disk,
131: and a jet or outflow. The local census of nuclear activity supports the
132: notion that most, perhaps all, bulges host a central supermassive black hole,
133: although the existence of active nuclei in at least some late-type
134: galaxies suggests that a classical bulge is not a prerequisite to seed a
135: nuclear black hole.
136: \end{abstract}
137:
138: \maketitle
139:
140: \section{INTRODUCTION}
141:
142: Far from being rare, exotic entities that inhabit only a tiny fraction of
143: galaxies, central black holes (BHs) are now believed to be basic constituents
144: of most, if not all, massive galaxies (Magorrian et al. 1998; Kormendy 2004).
145: Although less common in low-mass systems, central BHs also have been
146: identified in some late-type, even dwarf, galaxies (Filippenko \& Ho
147: 2003; Barth et al. 2004; Greene \& Ho 2004, 2007b; Dong et al. 2007; Greene,
148: Ho \& Barth 2008). The realization that BH mass correlates strongly with the
149: properties of the host galaxy (Kormendy 1993; Kormendy \& Richstone 1995;
150: Magorrian et al. 1998; Gebhardt et al. 2000; Ferrarese \& Merritt 2000;
151: Barth, Greene \& Ho 2005; Greene \& Ho 2006) has generated intense interest in
152: linking BH growth
153: with galaxy formation, as attested by the increasing number of conferences
154: focusing on this theme (e.g., Schmitt, Kinney \& Ho 1999; Ho 2004a;
155: Storchi-Bergmann, Ho \& Schmitt 2004; Merloni, Nayakshin \& Sunyaev 2005;
156: Fiore 2006). As a direct manifestation of BH accretion, and therefore BH
157: growth, active galactic nuclei (AGNs) and the consequences of their energy
158: feedback have figured prominently in most current ideas of structure formation
159: (e.g., Granato et al. 2004; Springel, Di~Matteo \& Hernquist 2005; Hopkins et
160: al. 2006). At the same time, the community's heightened awareness of the
161: importance of BHs has galvanized broad interest in the study of
162: the AGN phenomenon itself. With the BH mass known---arguably the most
163: fundamental parameter of the system---what once rested on phenomenological
164: analysis can now be put on a more secure physical basis.
165:
166: This review focuses on nuclear activity in nearby galaxies. By selection,
167: most of the objects occupy the faintest end of the AGN luminosity function
168: and have very low accretion rates. While energetically
169: unimpressive, low-luminosity AGNs (LLAGNs) deserve scrutiny for several
170: reasons. By virtue of the short duty cycle of BH accretion ($\sim 10^{-2}$;
171: Greene \& Ho 2007a), most AGNs spend their lives in a low state, such that the
172: bulk of the population has relatively modest luminosities. Over the past
173: several decades, this attribute has led to considerable controversy regarding
174: the physical origin of LLAGNs. As absolute luminosity can no longer be used
175: as a defining metric of nonstellar activity, many alternative excitation
176: mechanisms have been proposed to explain LLAGNs. Fortunately, the advent of
177: new telescopes and new analysis techniques have yielded many fresh insights
178: into this thorny old problem. A major goal of this review is to summarize
179: these recent developments. Along the way, I will emphasize how the collective
180: properties of LLAGNs can shed light on a poorly understood regime of the
181: central engine, namely that governed by low mass accretion rate. A key point
182: I will stress is that LLAGNs are not simply scaled-down versions of their more
183: familiar cousins, the classical Seyfert galaxies and quasars.
184:
185: Despite the impressive progress made in the direct detection of central BHs in
186: nearby inactive galaxies, our knowledge of the
187: demographics of BHs remains highly incomplete. Direct measurements of BH
188: masses based on resolved gas or stellar kinematics are still far from routine
189: and are available only for about three dozen galaxies. Certainly
190: nothing approaching a ``complete'' sample exists yet. More importantly, it is
191: not obvious that the current statistics are unbiased. As discussed by Barth
192: (2004), most nearby galaxies possess chaotic gas velocity fields that defy
193: simple analysis. Stellar kinematics provide a powerful alternative, but in
194: practice this technique has been limited to relatively dust-free systems and,
195: for practical reasons, to galaxies of relatively high central surface
196: brightness. The latter restriction selects against the most luminous, giant
197: ellipticals. Present surveys also severely underrepresent disk-dominated
198: galaxies, because the bulge component in these systems is inconspicuous and
199: star formation tends to perturb the velocity field of the gas. Finally, apart
200: from galaxies within the Local Group, even the highest angular resolution
201: currently achieved is inadequate to directly detect BHs with masses \lax\
202: $10^6$ \solmass. Consequently we are nearly completely ignorant about the low
203: end of the BH mass function. Given the above limitations, it is
204: desirable to consider alternative constraints on BH demography. The commonly
205: held and now well-substantiated premise that AGNs derive their energy
206: output from BH accretion implies that an AGN signifies the presence of a
207: central BH in a galaxy. The AGN signature in and of itself provides little
208: direct information on BH masses, but AGN statistics can inform us, effectively
209: and efficiently, of some key aspects of BH demography. For example, what
210: fraction of all galaxies contain BHs? Do BHs exist preferentially in galaxies
211: of certain types? Does environment matter? I will discuss how studies of
212: nearby AGNs have begun to answer some of these important questions.
213:
214: This review is structured as follows. I begin with an overview of the basic
215: methodology of the spectral classification of emission-line nuclei (\S~2) by
216: describing the currently adopted system, its physical motivation, the
217: complications of starlight subtraction, and some practical examples. Section
218: 3 summarizes past and current spectroscopic surveys and introduces the
219: Palomar survey, covering detection rates, measurement of weak broad emission
220: lines, and issues of robustness and completeless. Host galaxy properties are
221: the subject of \S~4, where in addition to global and environmental effects
222: I also cover results on nuclear stellar populations. In Section 5, I devote
223: considerable attention to the nuclear properties of LLAGNs in general and
224: LINERs in particular, focusing on modern results obtained from
225: high-resolution, multiwavelength observations from radio to hard-X-ray
226: energies. I use these data to draw inferences concerning the broad-line
227: region (BLR), torus, narrow-line region (NLR), spectral energy distribution
228: (SED), luminosity function, bolometric luminosities, and Eddington ratios.
229: This section contains many technical details, but these will be
230: essential ingredients for formulating the big picture at the end. Section 6
231: covers the controversial subject of the excitation mechanism of LINERs, the
232: growing puzzle concerning the energy budget in these systems, and the nature
233: of narrow-line nuclei. The implications of LLAGNs for BH demographics are
234: discussed in Section 7. Section 8 attempts to synthesize the disparate lines
235: of evidence into a coherent physical framework for LLAGNs and their relation
236: to other classes of objects. Finally, Section 9 concludes with some personal
237: perspectives and suggestions for future directions.
238:
239: \section{SPECTRAL CLASSIFICATION OF GALACTIC NUCLEI}
240:
241: \subsection{Physical Motivation}
242:
243: AGNs can be identified by a variety of methods. Most AGN surveys rely on some
244: aspect of the distinctive AGN spectrum, such as the presence of strong or
245: broad emission lines, an unusually blue continuum, or strong radio or X-ray
246: emission. All of these techniques are effective, but none is free from
247: selection effects. To search for AGNs in nearby galaxies, where the
248: nonstellar signal of the nucleus is expected to be weak relative to the host
249: galaxy, the most effective and least biased method is to conduct a
250: spectroscopic survey of a complete, optical-flux limited sample of galaxies.
251: To be sensitive to weak emission lines, the survey must be deep and of
252: sufficient spectral resolution. To obtain reliable line intensity ratios on
253: which the principal nuclear classifications are based, the data must have
254: accurate relative flux calibration, and one must devise a robust scheme to
255: correct for the starlight contamination.
256:
257: The most widely used system of spectral classification of emission-line nuclei
258: follows the method promoted by Baldwin, Phillips \& Terlevich (1981), and
259: later modified by Veilleux \& Osterbrock (1987). The basic idea is that the
260: relative strengths of certain prominent emission lines can be used to probe
261: the nebular conditions of a source. In the context of the present discussion,
262: the most important diagnostic is the source of excitation, which broadly falls
263: into two categories: stellar photoionization or photoionization by a centrally
264: located, spectrally hard radiation field, such as that produced by the
265: accretion disk of a massive BH. How does one distinguish between the two? The
266: forbidden lines of the doublet \oi\ \lamb\lamb 6300, 6364 arise from
267: collisional excitation of O$^0$ by hot electrons. Since the ionization
268: potential of O$^0$ (13.6 eV) is nearly identical to that of hydrogen, in an
269: ionization-bounded nebula \oi\ is produced predominantly in the ``partially
270: ionized zone,'' wherein both neutral oxygen and free electrons coexist. In
271: addition to O$^0$, the conditions of the partially ionized zone are also
272: favorable for S$^+$ and N$^+$, whose ionization potentials are 23.3 eV and
273: 29.6 eV, respectively. Hence, in the absence of abundance anomalies, \nii\
274: \lamb\lamb 6548, 6583 and \sii\ \lamb\lamb 6716, 6731 are strong (relative to,
275: say, H\al) whenever \oi\ is strong, and {\it vice versa}.
276:
277: In a nebula photoionized by young, massive stars, the partially ionized zone
278: is very thin because the ionizing spectrum of OB stars contains few photons
279: with energies greater than 13.6 eV. Hence, in the optical spectra of \hii\
280: regions and starburst nuclei (hereinafter \hii\ nuclei\footnote{As originally
281: defined (Weedman et al. 1981), a star{\it burst}\ nucleus is one whose current
282: star formation rate is much higher than its past average rate. Since in
283: general we do not know the star formation history of any individual object, I
284: will adopt the more general designation of ``\hii\ nucleus.''}) the
285: low-ionization transitions \nii, \sii, and especially \oi\ are very weak. By
286: contrast, a harder radiation field, such as that of an AGN power-law continuum
287: that extends into the extreme-ultraviolet (UV) and X-rays, penetrates much
288: deeper into an optically thick cloud, creating an extensive partially ionized
289: zone and hence strong low-ionization forbidden lines. A hard AGN radiation
290: field also boosts the production of collisionally excited forbidden line
291: emission because its high thermal energy deposition rate enhances the gas
292: temperature.
293:
294: \begin{figure}
295: %%BoundingBox: 60 50 580 750
296: \psfig{figure=fig1.ps,width=14cm,angle=270}
297: \caption{Sample optical spectra of the various classes of emission-line
298: nuclei. Prominent emission lines are identified. (Based on Ho, Filippenko \&
299: Sargent 1993 and unpublished data.)
300: }
301: \end{figure}
302:
303: \subsection{Sample Spectra}
304: The spectra shown in Figure~1 illustrate the empirical distinction between
305: AGNs and \hii\ nuclei. In NGC~7714, which has a well-known starburst nucleus
306: (Weedman et al. 1981), \oi, \nii, and \sii\ are weak relative to H\al. The
307: \oiii\ \lamb\lamb 4959, 5007 doublet is quite strong compared to \oii\ \lamb
308: 3727 or H\bet\ because the metal abundance of NGC~7714's nucleus is rather
309: low, although the ionization level of \hii\ nuclei can span a wide range,
310: depending on metallicity (Ho, Filippenko \& Sargent 1997c; Kewley et al.
311: 2001; Groves, Heckman \& Kauffmann 2006). On the other hand, the
312: low-ionization lines are markedly stronger in the other two objects shown,
313: both of which qualify as AGNs. NGC~1358 has a
314: ``high-ionization'' AGN or ``Seyfert'' nucleus. NGC~1052 is the prototype of
315: the class known as ``low-ionization nuclear emission-line regions'' or
316: LINERs (Heckman 1980b). The ionization level can be judged by the relative
317: strengths of the oxygen lines, but in practice is most easily gauged by the
318: \oiii/H\bet\ ratio. In the commonly adopted system of Veilleux \& Osterbrock
319: (1987), the division between Seyferts and LINERs occurs at \oiii\ \lamb
320: 5007/H\bet\ = 3.0. Ho, Filippenko \& Sargent (2003) stress, however, that
321: this boundary has no strict physical significance. The ionization level of the
322: NLR in large, homogeneous samples of AGNs spans a wide and
323: apparently continuous range; there is no evidence for any clear-cut transition
324: between Seyferts and LINERs (Ho, Filippenko \& Sargent 2003), although with
325: sufficient numbers, the two classes do delineate two distinct loci in optical
326: diagnostic diagrams (Kewley et al. 2006).
327:
328: The classification system discussed above makes no reference to the profiles
329: of the emission lines. Luminous AGNs such as quasars and many classical
330: Seyfert galaxies exhibit permitted lines with a characteristically broad
331: component, with full width at half-maximum (FWHM) widths of $\sim 1000$ to
332: $10,000$ \kms. This component arises from the BLR, which is thought to be
333: physically distinct from the NLR responsible for the narrow lines. Following
334: Khachikian \& Weedman (1974), it is customary to refer to Seyferts with and
335: without (directly) detectable broad lines as ``type~1'' and ``type~2''
336: sources, respectively. As discussed in \S~3.4, this nomenclature can also be
337: extended to include LINERs.
338:
339: \begin{figure}
340: %%BoundingBox: 200 30 480 750
341: \psfig{figure=fig2.ps,width=14cm,angle=270}
342: \caption{Diagnostic diagrams plotting ({\it a}) log \oiii\ \lamb 5007/H\bet\
343: versus log \nii\ \lamb 6583/H\al, ({\it b}) log \oiii\ \lamb 5007/H\bet\
344: versus log \sii\ \lamb\lamb 6716, 6731/H\al, and ({\it c}) log \oiii\ \lamb
345: 5007/H\bet\ versus log \oi\ \lamb 6300/H\al.
346: (Adapted from Ho, Filippenko \& Sargent 1997a.)
347: }
348: \end{figure}
349:
350: \subsection{Diagnostic Diagrams}
351:
352: The classification system of Veilleux \& Osterbrock (1987), which I
353: adopt throughout this paper, is based on two-dimensional line-intensity ratios
354: constructed from \oiii\ \lamb 5007, H\bet\ \lamb 4861, \oi\ \lamb 6300, H\al\
355: \lamb6563, \nii\ \lamb 6583, and \sii\ \lamb\lamb 6716, 6731 (here H\bet\ and
356: H\al\ refer only to the narrow component of the line). The main virtues of
357: this system, shown in Figure~2, are (1) that it uses relatively strong lines,
358: (2) that the lines lie in an easily accessible region of the optical spectrum,
359: and (3) that the line ratios are relatively insensitive to reddening
360: corrections because of the close separation of the lines. The definitions of
361: the various classes of emission-line objects are given in Ho, Filippenko \&
362: Sargent (1997a)\footnote{The classification criteria adopted here differ
363: slightly from those proposed by Kewley et al. (2001), Kauffmann et al. (2003),
364: or Stasi\'nska et al. (2006), but this difference has little effect on the
365: general conclusions.}. In addition to the three main classes
366: discussed thus far---\hii\ nuclei, Seyferts, and LINERs---Ho, Filippenko \&
367: Sargent (1993) identified a group of ``transition objects'' whose \oi\
368: strengths are intermediate between those of \hii\ nuclei and LINERs. Since
369: they tend to emit weaker \oi\ emission than classical LINERs, previous authors
370: have called them ``weak-\oi\ LINERs'' (Filippenko \& Terlevich 1992;
371: Ho \& Filippenko 1993). Ho, Filippenko \& Sargent (1993) postulated
372: that transition objects are composite systems having both an \hii\ region and
373: a LINER component; I will return to the nature of these sources in \S~6.5.
374:
375: Note that my definition of LINERs differs from that originally proposed by
376: Heckman (1980b), who used solely the oxygen lines: \oii\ \lamb 3727 $>$
377: \oiii\ \lamb 5007 and \oi\ \lamb 6300 $>$ 0.33 \oiii\ \lamb 5007. The two
378: definitions, however, are nearly equivalent. Inspection of the full optical
379: spectra of Ho, Filippenko \& Sargent (1993), for example, reveals that
380: emission-line nuclei classified as LINERs based on the Veilleux \& Osterbrock
381: diagrams almost always also satisfy Heckman's criteria. This is a consequence
382: of the inverse correlation between \oiii/H\bet\ and \oii/\oiii\ in
383: photoionized gas with fairly low excitation.
384:
385: \subsection{Starlight Subtraction}
386:
387: The scheme described above, while conceptually simple, overlooks one key
388: practical complication. The integrated spectra of galactic nuclei include
389: starlight, which in most nearby systems overwhelms the nebular line
390: emission (Figure~1). Any reliable measurement of the emission-line spectrum of
391: galactic nuclei, therefore, {\it must}\ properly account for the starlight
392: contamination.
393:
394: An effective strategy for removing the starlight from an integrated spectrum
395: is that of ``template subtraction,'' whereby a template spectrum devoid of
396: emission lines is suitably scaled to and subtracted from the spectrum of
397: interest to yield a continuum-subtracted, pure emission-line spectrum. A
398: number of approaches have been adopted to construct the template. These
399: include using (1) the spectrum of an off-nuclear position within the same
400: galaxy (e.g., Storchi-Bergmann, Baldwin \& Wilson 1993); (2) the spectrum of
401: a different galaxy devoid of emission lines (e.g., Costero \& Osterbrock 1977;
402: Filippenko \& Halpern 1984; Ho, Filippenko \& Sargent 1993); (3) a weighted
403: linear combination of the spectra of a number different galaxies, chosen to
404: best match the stellar population and velocity dispersion (Ho, Filippenko \&
405: Sargent 1997a); (4) a variant of (3), but employing a stellar library and
406: simultaneously fitting for the emission lines and accounting for dust
407: reddening (Sarzi et al. 2007); (5) a mean spectrum derived from a
408: principal-component analysis of a large set of galaxies (Hao et al. 2005a);
409: and (6) a model spectrum constructed from population synthesis techniques,
410: using as input a library of spectra of either individual stars (e.g., Keel
411: 1983c), synthesis models (e.g., Tremonti et al. 2004; Sarzi et al. 2005), or
412: star clusters (e.g., Bonatto, Bica \& Alloin 1989). Some studies (e.g., Kim
413: et al. 1995) implicitly assume that only the hydrogen Balmer lines are
414: contaminated by starlight and that the absorption-line component can be
415: removed by assuming a constant equivalent width (EW = $2-3$ \AA). This
416: procedure is inadequate for a number of reasons. First, the stellar
417: population of nearby galactic nuclei, although relatively
418:
419: \clearpage
420: \begin{figure}
421: %%BoundingBox: 60 50 580 750
422: \psfig{figure=fig3.ps,width=13.7cm,angle=270}
423: \caption{Illustration of starlight removal for NGC~3998 using the template
424: galaxy NGC~3115. Prominent emission lines are labeled. The insert shows an
425: expanded view of the H\al+\nii\ region and a multi-Gaussian decomposition
426: leading to the detection of a broad H\al\ component. (Adapted from Ho,
427: Filippenko \& Sargent 1993 and Ho et al. 1997e.)
428: }
429: \end{figure}
430:
431: \noindent
432: uniform, is by no
433: means invariant (Ho, Filippenko \& Sargent 2003). Second, the equivalent
434: widths of the different Balmer absorption lines within each galaxy are
435: generally not constant. Third, the Balmer absorption lines affect not only
436: the strength but also the shape of the Balmer emission lines. And finally,
437: starlight contaminates more than just the Balmer lines.
438:
439: Figure~3 illustrates the starlight subtraction process for the LINER NGC~3998. Note that in the original observed spectrum, many of the weaker emission lines
440: were hardly visible, whereas after starlight subtraction, they can easily be
441: measured. The intensities of even strong lines such H\bet\ and \oiii\
442: \lamb\lamb4959, 5007 are modified. Importantly, starlight correction is
443: essential for properly identifying the weak broad H\al\ component in NGC~3998.
444:
445: \subsection{Other Classification Criteria}
446:
447: Although the traditional optical classification system described above is
448: the most widely used, there are instances when features in other spectral
449: regions may be more practical or useful. Surveys of intermediate-redshift
450: galaxies, for example, cannot routinely access the H\al\ region, and under
451: such circumstances it is desirable to devise a classification system based
452: only on the blue part of the spectrum. Diagnostic diagrams proposed by
453: Rola, Terlevich \& Terlevich (1997) based on the strong lines \oii\
454: \lamb3727, \neiii\ \lamb\lamb 3869, 3968, H\bet, and \oiii\ \lamb\lamb 4959,
455: 5007 provide moderately effective discrimination between starbursts and
456: AGNs. A number of redshift surveys have searched for narrow-line AGNs based
457: on the presence of \nev\ \lamb\lamb 3346, 3426 (e.g., Hall et al. 2000; Barger
458: et al. 2001; Szokoly et al. 2004). With an ionization potential of 97 eV,
459: \nev\ unambiguously arises from nonstellar excitation, but the practical
460: difficulty is that these lines are quite weak (strength $\sim$10\% of
461: \oiii\ \lamb5007) and often can only be detected with confidence in
462: stacked spectra (e.g., Zakamska et al. 2003).
463:
464: The long-standing controversy over the relevance of shock excitation also
465: has led to the development of line diagnostics outside of the traditional
466: optical window. D\'\i az, Pagel \& Wilson (1985; see also D\'\i az, Pagel
467: \& Terlevich 1985; Kirhakos \& Phillips 1989) suggested that \siii\
468: \lamb\lamb9069, 9532, in combination with the optical lines of \oii, \oiii,
469: and \sii, are effective in identifying shock-excited nebula. Since shock
470: heating on average achieves higher equilibrium electron temperatures than
471: photoionization, high-ionization UV lines such as \nv\ \lamb 1240 and
472: \civ\ \lamb 1549 can serve as a powerful discriminant between these two
473: processes (e.g., Allen, Dopita \& Tsvetanov 1998). The limited availability
474: of UV spectra, however, has restricted the wide use of these diagnostics.
475:
476: Rapid progress in infrared (IR) technology has offered an important new window
477: that is not only less affected by dust but also potentially has distinctive
478: diagnostic power. Alonso-Herrero et al. (1997) show that
479: \feii\ 1.644/Br$\gamma$ can serve as an effective substitute for the
480: conventional \oi\ \lamb6300/H\al\ ratio. Unfortunately, other strong near-IR
481: features, notably the vibrational lines of H$_2$, are less useful because they
482: can be excited by multiple mechanisms (Larkin et al. 1998). The mid-IR regime
483: is much more promising, particularly with the sensitivity and wide bandpass
484: afforded by {\it Spitzer}. For the first time, many of the diagnostic lines
485: previously discussed in a theoretical context (Spinoglio \& Malkan 1992; Voit
486: 1992) actually can now be measured (e.g., Bendo et al. 2006; Dale et al. 2006;
487: Sturm et al. 2005, 2006; Rupke et al. 2007). In addition to high-ionization
488: lines such as \neiii\ \lamb15.5 \micron, \nev\ \lamb14.3 \micron, and \oiv\
489: \lamb 25.9 \micron, the low-ionization transitions of \feii\ \lamb26.0
490: \micron\ and [Si~{\sc II}] \lamb34.8 \micron\ may prove to be especially useful as
491: they can constrain models of photo-dissociation and X-ray dissociation
492: regions. The hard radiation field of AGNs, even of low-luminosity objects
493: such as LINERs, appears to leave an imprint on the detailed emission spectrum
494: of polycyclic aromatic hydrocarbons (Sturm et al. 2006; Smith et al. 2007).
495:
496: Finally, a comment on nomenclature. It is important to stress that the
497: classification scheme outlined above, physically motivated by the desire to
498: separate objects by their source of excitation, is based strictly on the
499: characteristics of the narrow emission lines and not on ancillary attributes
500: such as luminosity, presence of broad emission lines, galaxy morphology, or
501: radio properties. Although one still customarily draws a quaint distinction
502: between quasars and Seyferts based on luminosity, it is widely acknowledged
503: that this division is largely historical. In terms of their position on the
504: line-ratio diagrams, quasars fall on the high-ionization branch and thus can be
505: classified as Seyferts. The same holds for many broad-line and narrow-line
506: radio galaxies, including most Fanaroff \& Riley (FR; 1974) type~II radio
507: sources, whose high luminosities generally translate directly into a high
508: degree of ionization. By the same token, most FR~I sources, because of their
509: low luminosity, typically have fairly low-ionization spectra, and hence
510: technically qualify as LINERs. FR~I radio galaxies and LINERs are not
511: separate beasts (cf. Falcke, K\"ording \& Markoff 2004; Chiaberge, Capetti
512: \& Macchetto 2005). Strong historical prejudice also compels many to regard
513: Seyfert nuclei as invariably radio-quiet sources that reside exclusively in
514: spiral hosts, when, in fact, neither rule strictly holds (Ho \& Peng 2001).
515: Despite claims to the contrary (Krolik 1998; Sulentic, Marziani \&
516: Dultzin-Hacyan 2000), broad emission lines emphatically are {\it not}\
517: solely confined to Seyfert nuclei (\S~3.4). This misconception has led
518: some people to define the Seyfert and LINER classes by their presence or
519: absence of broad emission lines.
520:
521: \section{SURVEYS OF NEARBY GALACTIC NUCLEI}
522:
523: \subsection{The Palomar Survey}
524:
525: The earliest redshift surveys already indicated that the spectra of galaxy
526: centers often show strong emission lines (e.g., Humason, Mayall \& Sandage
527: 1956). In many instances, the spectra revealed abnormal line-intensity
528: ratios, most notably the unusually great strength of \nii\ relative to H\al\
529: (Burbidge \& Burbidge 1962, 1965; Rubin, Ford \& Thonnard 1980; Rose \& Searle
530: 1982). That the optical emission-line spectra of some nuclei show patterns of
531: low ionization was noticed from time to time, primarily by Osterbrock and his
532: colleagues (e.g., Osterbrock \& Dufour 1973; Osterbrock \& Miller 1975; Koski
533: \& Osterbrock 1976; Costero \& Osterbrock 1977; Grandi \& Osterbrock 1978;
534: Phillips 1979), but also by others (e.g., Disney \& Cromwell 1971; Danziger,
535: Fosbury \& Penston 1977; Fosbury \etal 1977, 1978; Penston \& Fosbury 1978;
536: Stauffer \& Spinrad 1979).
537:
538: The activity in this field culminated in the 1980s, beginning with the
539: recognition (Heckman, Balick \& Crane 1980; Heckman 1980b) of LINERs as a
540: major constituent of the extragalactic population, and then followed by further
541: systematic studies of larger samples of galaxies (Stauffer 1982a, 1982b; Keel
542: 1983b, 1983c; Phillips \etal 1986; V\'eron \& V\'eron-Cetty 1986;
543: V\'eron-Cetty \& V\'eron 1986; see Ho 1996 for more details). These surveys
544: established three important results. (1) A large fraction of local galaxies
545: contain emission-line nuclei. (2) Many of these sources are LINERs. And (3)
546: LINERs may be accretion-powered systems.
547:
548: Despite the successes of these seminal studies, there was room for improvement.
549: Although most of the surveys attempted some form of starlight subtraction, the
550: accuracy of the methods used was limited (see discussion in
551: Ho, Filippenko \& Sargent 1997a), the procedure was sometimes inconsistently
552: applied, and in some of the surveys, starlight subtraction was altogether
553: neglected. The problem is exacerbated by the fact that the apertures used for
554: the observations were quite large, thereby admitting an unnecessarily large
555: amount of starlight. Furthermore, most of the data were collected with rather
556: low spectral resolution (FWHM $\approx$ 10 \AA). Besides losing useful
557: kinematic information, blending between the emission and absorption
558: components further compromises the ability to separate the two.
559:
560: Thus, it was clear that much would be gained from a survey having greater
561: sensitivity to the detection of emission lines. The sensitivity could be
562: improved in at least four ways: by taking spectra with higher signal-to-noise
563: ratio and spectral resolution, by using a narrower slit to better isolate the
564: nucleus, and by employing more effective methods to handle the starlight
565: correction.
566:
567: The Palomar spectroscopic survey of nearby galaxies (Filippenko \& Sargent
568: 1985; Ho, Filippenko \& Sargent 1995, 1997a, 1997b, 1997c, 1997d, 2003; Ho
569: et al. 1997e) was designed with these goals in mind. Using a double CCD
570: spectrograph mounted on the Hale 5-m reflector, high-quality,
571: moderate-resolution, long-slit spectra were obtained for a magnitude-limited
572: ($B_T\,\leq$ 12.5 mag) sample of 486 northern ($\delta$ $>$ 0\deg) galaxies.
573: Drawn from the Revised Shapley-Ames (RSA) Catalog of Bright Galaxies (Sandage
574: \& Tammann 1981), the bright magnitude limit ensured that the sample had a
575: high degree of completeness. The spectra simultaneously cover the wavelength
576: ranges $6210-6860$ \AA\ with $\sim$2.5 \AA\ resolution (FWHM) and $4230-5110$
577: \AA\ with $\sim$4 \AA\ resolution. Most of the observations were obtained
578: with a narrow (1\asec$-$2\asec) slit, and relatively long exposure times
579: gave high signal-to-noise ratios. This survey still contains
580: the largest database to date of homogeneous and high-quality optical spectra
581: of nearby galaxies. It is also the most sensitive; the detection limit for
582: emission lines is EW $\approx$ 0.25 \AA, roughly an order-of-magnitude
583: improvement compared to previous or subsequent work. The selection criteria
584: ensure that the sample gives a fair representation of the local
585: ($z\,\approx\,0$) galaxy population, and the proximity of the objects (median
586: distance = 17 Mpc) results in relatively good spatial resolution (typically
587: \lax\ 200 pc)\footnote{A distance scale based on $H_0$ = 75 \kms~Mpc$^{-1}$ is
588: assumed throughout}. These properties of the Palomar survey make it
589: ideally suited to address issues on the demographics and physical properties
590: of nearby, and especially low-luminosity, AGNs. Unless otherwise noted, most
591: of the results presented in this paper will be taken from the Palomar survey.
592:
593: The Palomar survey has one other virtue that is not widely appreciated.
594: Because the sample is large and essentially unbiased with respect to
595: nuclear or global properties, it is ideally suited for comparative studies
596: of various subpopulations. Examples include efforts to discern differences
597: between type~1 versus type~2 sources to test AGN unification, to ascertain the
598: influence of bars or environment on nuclear activity, or to test for subtle
599: differences between the different AGN classes. The robustness of these and
600: similar studies almost always hinges on the availability of proper control
601: samples. With the Palomar survey, there is no need to construct a separate
602: control sample, which is always a difficult and somewhat dubious undertaking,
603: because the control sample is built into the survey.
604:
605: \subsection{Other Surveys}
606:
607: For completeness, I mention several other sources of nearby AGNs that have
608: been widely used by the community. The AGN sample culled from the CfA
609: Redshift Survey (Huchra \& Burg 1992) has been an important resource for a
610: long time. Comprising 47 relatively bright Seyferts and a handful of LINERs,
611: the CfA sample in many ways complements the Palomar sample at the bright end
612: of the luminosity function. However, as discussed in Ho \& Ulvestad (2001),
613: the selection effects of the CfA sample are not easy to quantify because of
614: the subjective and somewhat nonstandard manner in which AGNs were picked from
615: the parent survey. Prior to the full publication of the Palomar survey,
616: Maiolino \& Rieke (1995) assembled a compilation of 91 Seyferts from a
617: literature search of the galaxies in the RSA. These ``RSA Seyferts'' have
618: subsequently been used in a number of follow-up studies. The substantial
619: improvement in the data quality and analysis of the Palomar survey has
620: resulted in many revised classifications of the RSA galaxies. Lastly, a
621: cautionary note. Many investigators rely on literature compilations, such as
622: those assembled in V\'eron-Cetty \& V\'eron's (2006) catalog or the
623: NASA/IPAC Extragalactic Database, as their source for AGN classifications.
624: This is dangerous. The classifications in these compilations are
625: highly heterogeneous and in some cases wrong.
626:
627: The sample of nearby AGNs emerging from the Sloan Digital Sky Survey (SDSS)
628: (Kauffmann et al. 2003; Hao et al. 2005b; Kewley et al. 2006) far surpasses
629: that of the Palomar survey in number but not in sensitivity. Moreover,
630: because SDSS samples more distant galaxies, the 3\asec-diameter fibers used in
631: the survey subtend a physical scale of $\sim$5.5 kpc at the typical redshift
632: $z \approx 0.1$, 30 times larger than in the Palomar survey. The SDSS spectra
633: therefore include substantial contamination from off-nuclear emission, which
634: dilutes and, in some cases, inevitably confuses the signal from the nucleus.
635: Contamination by host galaxy emission has two consequences. First, only
636: relatively bright nuclei have enough contrast to be detected. But second,
637: contamination can introduce a more pernicious systematic effect
638: that can be hard to quantify. Apart from normal \hii\ regions, galactic disks
639: are known to contain {\it extended}\ emission-line regions that exhibit
640: low-ionization, LINER-like spectra. They can be confused with genuine {\it
641: nuclear}\ LINERs. Examples include gas shocked by supernova remnants (e.g.,
642: Dopita \& Sutherland 1995), ejecta from starburst-driven winds (Armus, Heckman
643: \& Miley 1990), and diffuse, warm ionized plasma (e.g., Collins \&
644: Rand 2001). Massive, early-type galaxies, though generally lacking in ongoing
645: star formation, also often possess X-ray emitting atmospheres that exhibit
646: extended, low-ionization emission-line nebulae (e.g., Fabian et al. 1986;
647: Heckman et al. 1989). These physical processes, while interesting in their
648: own right, are not directly related, and thus irrelevant, to the AGN
649: phenomenon. Thus, LINERs selected from samples of distant galaxies should be
650: regarded with considerable caution. This comment applies also
651: to LINERs selected from samples of IR-bright galaxies (e.g., Kim et
652: al. 1995; Kewley et al. 2001; Corbett et al. 2003), which, in addition to being
653: relatively distant and maximally confused with starburst processes, have the
654: additional disadvantage of often being merging or interacting systems, wherein
655: shocks undoubtedly generate extended LINER-like emission. I strongly
656: recommend that researchers avoid IR-selected samples if they are interested in
657: investigating LINERs as an accretion phenomenon. Many of the objects in the
658: catalog of LINERs compiled by Carrillo et al. (1999), which has been the basis
659: of several recent studies (Satyapal, Sambruna \& Dudik 2004; Dudik et al.
660: 2005; Gonz\'alez-Mart\'\i n et al. 2006), suffer precisely from this
661: complication and should be used judiciously.
662:
663: \begin{figure}
664: %%BoundingBox: 30 200 570 620
665: \hskip 0.0truein
666: \psfig{figure=fig4a.ps,width=6.9cm,angle=0}
667: \hskip -0.2truein
668: \psfig{figure=fig4b.ps,width=6.9cm,angle=0}
669: \caption{{\it Left}: Detection rate of emission-line nuclei as a function of
670: Hubble type. The different classes of nuclei are marked as follows: Seyferts =
671: blue squares, LINERs = red circles, transition objects = green triangles,
672: LINERs + transition objects + Seyferts = black stars, \hii\ nuclei = magenta
673: crosses. {\it Right}: Distribution of bulge-to-total ($B/T$) light ratios,
674: derived from the morphological type and its statistical dependence on $B/T$.
675: The histograms have been shifted vertically for clarity.
676: (Adapted from Ho, Filippenko \& Sargent 1997a, 1997b.)
677: }
678: \end{figure}
679:
680: \subsection{Detection Rates}
681:
682: In qualitative agreement with early work, the Palomar survey shows that a
683: substantial fraction (86\%) of all galaxies contain detectable emission-line
684: nuclei (Ho, Filippenko \& Sargent 1997b). The detection rate is essentially
685: 100\% for all disk (S0 and spiral) galaxies, and $>$50\% for elliptical
686: galaxies. One of the most surprising results is the large fraction of objects
687: classified as AGNs or AGN candidates, as summarized in Figure~4.
688: Summed over all Hubble types, 43\% of all galaxies that fall in the survey
689: limits can be considered ``active.'' This percentage becomes even more
690: remarkable for galaxies with an obvious bulge component, rising to
691: $\sim$50\%$-$70\% for Hubble types E$-$Sb. By contrast, the detection rate
692: of AGNs drops dramatically toward later Hubble types (Sc and later), which
693: almost invariably (80\%) host \hii\ nuclei. This strong dependence of nuclear
694: spectral class on Hubble type has been noticed in earlier studies (Heckman
695: 1980a; Keel 1983b), and further confirmed in SDSS (Kauffmann et al. 2003;
696: Miller et al. 2003). A qualitatively similar conclusion, cast in terms of
697: host galaxy stellar mass rather than Hubble type, is reached by Gallo et al.
698: (2008) and Decarli et al. (2007). Decarli et al. also claim that the
699: occurrence of AGN activity in Virgo cluster spirals does not depend on
700: morphological type, but it must be noted that the sources of spectroscopy and
701: nuclear classification employed in that study are very heterogeneous.
702:
703: Within the parent galaxy sample, 11\% have Seyfert nuclei, at least doubling
704: estimates based on older (Stauffer 1982b; Keel 1983b; Phillips, Charles \&
705: Baldwin 1983) or shallower (Huchra \& Burg 1992; Maia, Machado \& Willmer
706: 2003; Gronwall et al. 2004; Hao et al. 2005a) surveys. LINERs constitute the
707: dominant population of AGNs. ``Pure'' LINERs are present in $\sim$20\% of all
708: galaxies, whereas transition objects, which by assumption also contain a LINER
709: component, account for another $\sim$13\%. Thus, if all LINERs can be regarded
710: as genuine AGNs (see \S~6), they truly are the most populous constituents,
711: making up 1/3 of all galaxies and 2/3 of the AGN population (here taken to
712: mean all objects classified as Seyferts, LINERs, and transition objects).
713:
714: Within the magnitude range $14.5 < r < 17.7$ in SDSS, Kauffmann et al. (2003)
715: report an overall AGN fraction (for narrow-line sources) of $\sim$40\%, of
716: which $\sim$10\% are Seyferts. The rest are LINERs and transition objects.
717: Using a different method of starlight subtraction, Hao et al. (2005b) obtain
718: very similar statistics for their sample of Seyfert galaxies. Roughly 30\% of
719: the galaxies on the red sequence in SDSS exhibit LINER-like emission (Yan et
720: al. 2006). Although these detection rates broadly resemble those of the
721: Palomar survey, one should recognize important differences between the two
722: surveys. The Palomar objects extend much farther down the luminosity function
723: than the SDSS. The emission-line detection limit of the Palomar survey, EW =
724: 0.25 \AA, is roughly 10 times fainter than the cutoff chosen by Hao et al.
725: (2005b). The faint end of the Palomar H\al\ luminosity function reaches
726: $\sim$1\e{37} \lum\ (\S~5.9), again a factor of 10 lower than the SDSS
727: counterpart. Given that LINERs selected from SDSS are highly susceptible to
728: extranuclear contamination, as discussed earlier, it is in fact quite
729: surprising---and perhaps fortuitous---that the detection rates of these
730: objects agree so well between the two surveys.
731:
732: \subsection{Broad Emission Lines}
733:
734: Broad emission lines, a defining attribute of classical Seyferts and quasars,
735: are also found in nuclei of much lower luminosities. The well-known case of
736: the nucleus of M81 (Peimbert \& Torres-Peimbert 1981; Filippenko \& Sargent
737: 1988), for example, has a broad (FWHM $\approx$ 3000 \kms) H\al\ line with a
738: luminosity of $2 \times\,10^{39}$ \lum\ (Ho, Filippenko \& Sargent
739: 1996), and many other less conspicuous cases have been discovered in the
740: Palomar survey (Ho et al. 1997e). Searching for broad H\al\ emission in
741: nearby nuclei is nontrivial, because it entails measurement of a (generally)
742: weak, low-contrast, broad emission feature superposed on a complicated stellar
743: background. Thus, the importance of careful starlight subtraction cannot be
744: overemphasized. Moreover, even if this could be accomplished perfectly,
745: one still has to contend with deblending the H\al +\nii\ complex. The narrow
746: lines in this complex are often heavily blended together, and rarely do the
747: lines have simple profiles. The strategy adopted by Ho \etal (1997e) is to
748: use the empirical line profile of the \sii\ lines to model \nii\ and the
749: narrow component of H\al. The case of NGC~3998 is shown in Figure~3.
750:
751: Of the 221 emission-line nuclei in the Palomar survey classified as LINERs,
752: transition objects, and Seyferts, 33 (15\%) definitely have broad H\al\ and an
753: additional 16 (7\%) probably do. Questionable detections were found in
754: another 8 objects (4\%). Thus, approximately 20\%$-$25\% of all nearby AGNs
755: are type~1 sources. These numbers, of course, should be regarded as lower
756: limits, since undoubtedly there must exist AGNs with even weaker broad-line
757: emission that fall below the detection threshold. Although the numbers are
758: meager, direct comparison with small-aperture {\it Hubble Space Telescope
759: (HST)}\ spectra (e.g., Nicholson et al. 1998; Barth et al. 2001b; Shields et
760: al. 2007) reveals that the Palomar statistics on broad H\al\ detections seem
761: to be quite robust. The type~1.9 classification of almost every
762: object with overlapping \hst\ data turns out to survive. More surprising
763: still, no new cases of broad H\al\ emission have turned up from \hst\
764: observations. Given the difficulty of measuring the weak broad H\al\ feature
765: on top of the dominant stellar continuum, it is likely that in general the
766: line widths may have been systematically underestimated from the ground-based
767: spectra. Circumstantial support for this conjecture comes from Zhang,
768: Dultzin-Hacyan \& Wang (2007), who find that Palomar LLAGNs tend to have
769: smaller BH virial masses (estimated from the H\al\ linewidth and a BLR
770: size-luminosity relation) than predicted from their bulge stellar velocity
771: dispersion. They conclude that the BLR size in LLAGNs may be larger than
772: normal given their luminosity, but an equally plausible explanation is that
773: the H\al\ linewidths have been systematically underestimated.
774:
775: It is illuminating to consider the incidence of broad H\al\ emission as
776: a function of spectral class. Among objects formally classified as Seyferts,
777: $\sim$40\% are Seyfert~1s. The implied ratio of Seyfert~1s to Seyfert~2s
778: (1:1.6) has important consequences for some models concerning the evolution
779: and small-scale geometry of AGNs (e.g., Osterbrock \& Shaw 1988). Within the
780: Palomar sample, nearly 25\% of the ``pure'' LINERs have detectable broad H\al\
781: emission. By direct analogy with the historical nomenclature established for
782: Seyferts, LINERs can be divided into ``type~1'' and ``type~2'' sources
783: according to the presence or absence of broad-line emission, respectively (Ho,
784: Filippenko \& Sargent 1997a; Ho et al. 1997e). The detection rate of broad
785: H\al, however, drops drastically for transition objects. The cause for this
786: dramatic change is unclear. In these objects the broad-line component may
787: simply be too weak to be detected in the presence of substantial contamination
788: from the \hii\ region component, or it may be intrinsically absent (\S\S~5.5,
789: 6.5).
790:
791: \subsection{Robustness and Completeness}
792:
793: To gain confidence in the AGN statistics based on optical spectroscopy, one
794: must have some handle on whether the existing AGN detections are trustworthy
795: and whether there are many AGNs that have been missed. The robustness issue
796: hinges on the question of whether the weak, nearby sources classified as AGNs
797: are truly accretion-powered. As I argue in \S~6, this appears largely to be the
798: case. The completeness issue can be examined in two regimes. Among galaxies
799: with prominent bulges (Sbc and earlier), for which the spectroscopic AGN
800: fractions are already very high ($\sim$50\%$-$75\%), there is not much room
801: for a large fraction of missing AGNs, although it is almost certain that some
802: have indeed eluded detection in the optical (e.g., Tzanavaris \&
803: Georgantopoulos 2007). The same does not necessarily hold for galaxies of
804: Hubble types Sc and later. While the majority of these systems are
805: spectroscopically classified as \hii\ nuclei, one must be wary that weak AGNs,
806: if present, may be masked by brighter off-nuclear \hii\ regions or \hii\
807: regions projected along the line of sight. After all, some very late-type
808: galaxies {\it do}\ host {\it bona fide}\ AGNs (see \S~7).
809:
810: The AGN content of late-type galaxies can independently be assessed by using
811: a diagnostic less prone to confusion by star-forming regions, namely by
812: looking for compact, nuclear radio or X-ray cores. Ulvestad \& Ho (2002)
813: performed a Very Large Array (VLA) survey for radio cores in a
814: distance-limited sample of 40 Palomar Sc galaxies classified as hosting \hii\
815: nuclei. To a sensitivity limit of $P_{\rm rad} \approx 10^{18}-10^{20}$
816: W~Hz$^{-1}$ at 5~GHz, and a resolution of $\Delta \theta$ = 1\asec, they found
817: that {\it none} of the galaxies contains a compact radio core. The VLA study
818: of Filho, Barthel \& Ho (2000) also failed to detect radio cores in a more
819: heterogeneous sample of 12 \hii\ nuclei.
820:
821: Information on nuclear X-ray cores in late-type galaxies is much more limited
822: because to date there has been no systematic investigation of these systems
823: with {\it Chandra}. A few studies, however, have exploited the High
824: Resolution Imager (HRI) on {\it ROSAT} to resolve the soft X-ray (0.5$-$2 keV)
825: emission in late-type galaxies (Colbert \& Mushotzky 1999; Lira, Lawrence \&
826: Johnson 2000; Roberts \& Warwick 2000). Although the resolution of the HRI
827: ($\sim$5\asec) is not ideal, it is nonetheless quite effective for identifying
828: point sources given the relatively diffuse morphologies of late-type galaxies.
829: Compact X-ray sources, often quite luminous (\gax\ 10$^{38}$ \lum), are
830: frequently found, but generally they do {\it not} coincide with the galaxy
831: nucleus; the nature of these off-nuclear ``ultraluminous X-ray sources'' is
832: discussed by Fabbiano (2006).
833:
834: To summarize: unless \hii\ nuclei in late-type galaxies contain radio and
835: X-ray cores far weaker than the current survey limits---a possibility worth
836: exploring---they do not appear to conceal a significant population of
837: undetected AGNs.
838:
839: \section{HOST GALAXY PROPERTIES}
840:
841: \subsection{Global Parameters}
842:
843: The near dichotomy in the distribution of Hubble types for galaxies hosting
844: active versus inactive nuclei (Figure~4) leads to the expectation that the two
845: populations ought to have fairly distinctive global, and perhaps even nuclear,
846: properties. Moreover, a detailed examination of the host galaxies of AGNs
847: may shed light on the origin of their spectral diversity. These issues were
848: examined by Ho, Filippenko \& Sargent (2003) using the database from the
849: Palomar survey. The host galaxies of Seyferts, LINERs, and transition objects
850: display a remarkable degree of homogeneity in their large-scale properties,
851: {\it after} factoring out spurious differences arising from small mismatches
852: in Hubble type distribution. The various nuclear types have slightly different
853: Hubble distributions, which largely control many of the statistical properties
854: of the host galaxies. Unless this effect is taken into account, one can
855: arrive at erroneous conclusions about the intrinsic differences of the AGN
856: populations. This is a crucial step, one that is often not appreciated.
857: All three classes have essentially identical total luminosities ($\sim L^*$),
858: bulge luminosities, sizes, and neutral hydrogen content. Moreover, no obvious
859: differences are found in terms of integrated optical colors or far-IR
860: luminosities and colors, which implies very similar global stellar content and
861: current star formation rates. No clear differences in environment can be
862: seen either. The only exception is that, relative to LINERs, transition
863: objects may show a mild enhancement in the level of global star formation, and
864: they may be preferentially more inclined. This is consistent with the
865: hypothesis that the transition class arises from spatial blending of emission
866: from a LINER and \hii\ regions. The velocity field of the ionized gas within
867: the nuclear region, as measured by the width and asymmetry of the narrow
868: emission lines, is crudely similar among the three AGN classes, an observation
869: that argues against the proposition that fast shocks primarily drive the
870: spectral variations (\S~6.2).
871:
872: \subsection{Nuclear Stellar Populations}
873:
874: The uniformity in the global stellar populations among the three AGN classes
875: extends to circumnuclear and even nuclear scales. The Palomar spectra cover a
876: suite of stellar absorption-line indices and nuclear continuum colors, which
877: collectively can be used to obtain crude constraints on the age and
878: metallicity of the stars within the central 2\asec\ ($\sim 200$ pc). After
879: isolating a subsample that mitigates the Hubble type dependence, Ho, Filippenko,
880: \& Sargent (2003) find that Seyferts, LINERs, and transition objects have very
881: similar stellar content. The same holds true when comparing type~1 and type~2
882: objects, both for LINERs and Seyferts. With a few notable exceptions such as
883: NGC~404 and NGC~4569 (Maoz et al. 1998; Barth \& Shields 2000; Gabel \&
884: Bruhweiler 2002) or NGC~4303 (Colina et al. 2002), the stellar population
885: always appears evolved. Similar findings have been reported for smaller
886: samples of LINERs, based on both optical (Boisson et al. 2000; Serote-Roos \&
887: Gon\c{c}alves 2004; Zhang, Gu \& Ho 2008) and near-IR spectroscopy (Larkin et
888: al. 1998; Bendo \& Joseph 2004). The optical regime is not well suited to
889: detect very young, ionizing stars. However, the Palomar spectra do cover the
890: broad \heii\ \lamb4686 emission bump, a feature indicative of Wolf-Rayet stars
891: commonly seen in very young ($3-6$ Myr) starbursts. Notwithstanding the
892: difficulty of detecting this feature on top of a dominant old population, it
893: is noteworthy that not a {\it single}\ case has been seen among the sample of
894: over 200 Palomar LLAGNs. By contrast, the Wolf-Rayet bump has been found in a
895: number of the \hii\ nuclei (Sargent \& Filippenko 1991; Ho, Filippenko \&
896: Sargent 1995), which, as a class compared to the LLAGNs, exhibit markedly
897: younger stars, as evidenced by their blue continuum, strong H\bet\ and
898: H$\delta$ absorption, and weak metal lines (Ho, Filippenko \& Sargent 2003).
899: The general dearth of young, massive stars in LLAGNs presents a serious
900: challenge to proposals that seek to account for the excitation of their line
901: emission in terms of starburst models.
902:
903: Closer in, on scales \lax\ 10 pc accessible by {\it HST},
904: Sarzi et al. (2005) studied the nuclear stellar population for a
905: distance-limited subsample of 18 Palomar LLAGNs. Their population synthesis
906: analysis shows that the majority (80\%) of the objects have predominantly old
907: (\gax\ 5 Gyr), mildly reddened stars of near-solar metallicity, the only
908: exceptions being 3 out of 6 transition objects and 1 out of 4 LINER~2s that
909: require a younger (\lax\ 1 Gyr) component. In no case, however, is the younger
910: component ever energetically dominant, falling far short of being able to
911: account for the ionization budget for the central region.
912:
913: The results of Ho, Filippenko \& Sargent (2003) have been disputed by
914: Cid~Fernandes et al. (2004), who obtained new ground-based spectra for a
915: subset of the Palomar LINERs and transition objects. Gonz\'alez~Delgado et
916: al. (2004), in a study similar to that of Sarzi et al. (2005), further
917: analyzed STIS spectra of some of these. An important improvement of their
918: ground-based data is that they extend down to $\sim 3500$ \AA, covering the
919: 4000~\AA\ break and the higher-order Balmer lines, which are sensitive probes
920: of intermediate-age ($\sim 10^7-10^9$ yr) stars. While LINERs are
921: predominantly old, roughly half of the transition nuclei show significant
922: higher-order Balmer lines. Again, there are virtually no traces of Wolf-Rayet
923: features. These authors propose that the ionization mechanism of transition
924: sources must be somehow linked to the intermediate-age stellar population.
925:
926: I disagree with their assessment. Figure~1 of Cid~Fernandes et al. (2004)
927: clearly shows that, as in the parent Palomar sample, the Hubble type
928: distribution of the LINERs is skewed toward much earlier types than that of the
929: transition objects. Any meaningful comparison of the stellar population, which
930: strongly depends on Hubble type, must take this into account. The strategy
931: employed by Ho, Filippenko \& Sargent (2003) is to restrict the two-sample
932: comparisons to a relatively narrow range of Hubble types (Sab$-$Sbc). Within
933: this domain, LINERs and transition objects (as well as Seyferts) have
934: statistically indistinguishable stellar indices and continuum colors. Given
935: the limited size of Cid~Fernandes et al.'s sample, it is not possible to adopt
936: the same strategy. To minimize the Hubble type bias, I examined the subsample
937: of 30 spiral galaxies in their study. Out of 17 transition objects, 15 (88\%)
938: contain intermediate-age stars according to their Table~3; but so do 10 out of
939: the 13 (77\%) LINERs in this subgroup. This simple exercise underscores the
940: importance of sample selection effects, and leaves me unconvinced
941: that transition objects have a younger stellar population than LINERs. True,
942: both classes evidently do contain detectable amounts of intermediate-age
943: stars---a qualitatively different conclusion than was reached in the Palomar
944: survey, whose spectral coverage was not well-suited to detect this
945: population---but the fact remains that in a {\it relative}\ sense all three
946: LLAGN classes in the Palomar survey have statistically similar populations.
947: If the poststarburst component is responsible for the nebular emission in
948: LLAGNs, we might expect the intensity of the two to be correlated, by analogy
949: with what has been found for higher luminosity AGNs (Kauffmann et al. 2003).
950: I searched for this effect in the final sample presented in Gonz\'alez~Delgado
951: et al. (2004), but did not find any correlation. Clearly we wish to know what
952: factors drive the spectral variations in LLAGNs; whatever they are (\S~6),
953: they are unlikely to be related to stellar population.
954:
955: \subsection{Influence of Bars and Environment}
956:
957: Numerical simulations (e.g., Heller \& Shlosman 1994; see review in Kormendy
958: \& Kennicutt 2004) suggest that large-scale
959: stellar bars can be highly effective in delivering gas to the central few
960: hundred parsecs of a spiral galaxy, thereby potentially leading to rapid star
961: formation. Further instabilities result in additional inflow to smaller
962: scales, which may lead to increased BH fueling and hence elevated nonstellar
963: activity in barred galaxies compared to unbarred galaxies. As discussed in
964: \S~3.1, the Palomar sample is ideally suited for statistical comparisons of
965: this nature, which depend delicately on issues of sample
966: completeness and the choice of control sample. Ho, Filippenko \& Sargent
967: (1997d) find that while the presence of a bar indeed does enhance both the
968: probability and rate of star formation in galaxy nuclei, it appears to have no
969: impact on either the frequency or strength of AGN activity. Bearing in mind
970: the substantial uncertainties associated with sample selection, as well as the
971: method and wavelength used to identify bars (Laurikainen, Salo \& Buta 2004),
972: other studies broadly come to a similar conclusion (see review by Combes 2003),
973: although Maia, Machado \& Willmer (2003) claim, on the basis of a
974: significantly larger and somewhat more luminous sample drawn from the Southern
975: Sky Redshift Survey, that Seyfert galaxies are preferentially more barred.
976:
977: In the same vein, dynamical interactions with neighboring companions should
978: lead to gas dissipation, enhanced nuclear star formation, and perhaps central
979: fueling (e.g., Hernquist 1989). Schmitt (2001) and Ho, Filippenko \& Sargent
980: (2003) studied this issue using the Palomar data, parameterizing the nearby
981: environment of each object by its local galaxy density and the distance to its
982: nearest sizable neighbor. After accounting for the well-known
983: morphology-density relation, it was found that the local environment, like
984: bars, has little impact on AGNs, at least in the low-luminosity regime sampled
985: locally. These findings broadly agree with the results from SDSS (Miller et
986: al. 2003; Li et al. 2008). Kauffmann et al. (2004), Wake et al. (2004), and
987: Constantin \& Vogeley (2006), in fact, report a drop in the fraction of
988: high-luminosity AGNs for dense environments.
989:
990: \section{NUCLEAR PROPERTIES}
991:
992: \subsection{Ionizing Continuum Radiation}
993:
994: AGNs, at least when unobscured, reveal themselves as pointlike nuclear sources
995: with power-law spectra at optical and UV wavelengths, typically described
996: by a continuum flux density $f_\nu \propto \nu^{\alpha}$, with $\alpha \approx
997: -0.5$ (e.g., Vanden~Berk et al. 2001). In unbeamed sources, this featureless
998: continuum traces the low-frequency tail of the ``big blue bump'' (Shields
999: 1978; Malkan \& Sargent 1982), which supplies the bulk of the ionizing
1000: photons. This feature is extremely difficult to detect in LLAGNs, both
1001: because the big blue bump is weak or absent (\S~5.8) and because the sources
1002: are exceedingly faint. The optical nuclei of LINERs can have $M_B$ \gax\
1003: $-10$ mag (Ho 2004b), at least $10^{4}$ times fainter than their (usually
1004: bulge-dominated) hosts ($M_B \simeq M^* \approx -20$ mag). To overcome this
1005: contrast problem, searches for nuclear point sources in the optical and
1006: near-IR have relied on \hst\ images (e.g., Chiaberge, Capetti \& Celotti 1999;
1007: Quillen et al. 2001; Verdoes~Kleijn et al. 2002; Chiaberge, Capetti \&
1008: Macchetto 2005; Balmaverde \& Capetti 2006; Gonz\'alez-Mart\'\i n et al.
1009: 2006). But resolution alone is not enough. Given the extreme faintness of
1010: the nucleus, the intrinsic cuspiness of the underlying bulge profile,
1011: complexities of the point-spread function, and the often irregular background
1012: marred by circumnuclear dust features, one must pay very close attention to
1013: {\it how}\ the measurements are made. Simple aperture photometry or searching
1014: for central excess emission can yield very misleading results. The most robust
1015: technique to extract faint nuclei in the presence of these complications
1016: employs two-dimensional, multi-component fitting (Ho \& Peng 2001;
1017: Ravindranath et al. 2001; Peng et al. 2002). Using this method, nuclear
1018: sources with optical magnitudes as faint as $\sim 20$ have been measured, with
1019: limits down to $\sim 22-23$ mag possible for nearby galaxies. Due to the
1020: computational requirements of two-dimensional fitting, however, not many
1021: LLAGNs have yet been analyzed in this manner, and fewer still have enough
1022: photometric points to define even a crude spectral slope.
1023:
1024: In a few cases, the optical featureless continuum has been detected
1025: spectroscopically. From the ground, this was only possible for a couple of
1026: the brightest sources. The stellar features of NGC~7213 (Halpern \&
1027: Filippenko 1984) and Pictor~A (Carswell et al. 1984; Filippenko 1985) show
1028: dilution by a featureless continuum, which can be described approximately by a
1029: power law with a spectral index of $\alpha \approx -1.5$. The
1030: nuclear continuum is much more readily seen in small-aperture spectra that
1031: help to reject the bulge starlight. \hst\ spectra have isolated the optical
1032: continuum in several LINERs (Ho, Filippenko \& Sargent 1996; Nicholson et al.
1033: 1998; Ho et al. 2000; Shields et al. 2000; Barth et al. 2001a; Sabra et al.
1034: 2003), although in most objects it remains too faint to be detected
1035: spectroscopically (Sarzi et al. 2005). In all well-studied cases, the optical
1036: continuum is quite steep, with $\alpha \approx -1$ to $-2$. This range
1037: in spectral slopes is consistent with the broad-band optical (Verdoes~Kleijn
1038: et al. 2002) and optical-UV colors (Chiaberge et al. 2002) of the cores
1039: frequently detected in the LINER nuclei of FR~I radio galaxies.
1040:
1041: The predominantly old population of present-day bulges ensures that the
1042: stellar contamination largely disappears in the UV, especially at high
1043: resolution. A number of attempts have been made to detect UV emission in
1044: LINERs using {\it IUE}, but most of these efforts yielded ambiguous results
1045: (see review in Filippenko 1996), and real progress had to await the {\it HST}.
1046: Two dedicated \hst\ UV ($\sim 2300$ \AA) imaging studies have been completed.
1047: Using the pre-COSTAR FOC, Maoz et al. (1996) surveyed a complete sample of 110
1048: large, nearby galaxies, and among the subset with spectral classifications
1049: from Palomar, Maoz et al. (1995) discovered that $\sim$25\% of the LINERs show
1050: an unresolved UV core. Barth et al. (1998) found similar statistics in a more
1051: targeted WFPC2 study. They also made the suggestion, later
1052: confirmed by Pogge et al. (2000), that dust obscuration is probably the main
1053: culprit for the nondetection of UV emission in the majority of LINERs. The
1054: implication is that UV emission is significantly more common in LINERs than
1055: indicated by the detection rates. In some type~2 objects (e.g., NGC~4569
1056: and NGC~6500), the UV emission is spatially extended and presumably not
1057: related to the nuclear source. Second-epoch UV observations with the
1058: ACS/HRC revealed that nearly all of the UV-bright sources exhibit long-term
1059: variability (Maoz et al. 2005), an important result that helps assuage fears
1060: that the UV emission might arise mainly from young stars (Maoz et al. 1998).
1061: Importantly, both type~1 {\it and}\ type~2 LINERs vary. UV variability has
1062: also been discovered serendipitously in a few other sources (Renzini et al.
1063: 1995; O'Connell et al. 2005).
1064:
1065: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1066: %%BoundingBox: 80 150 350 770
1067: \begin{figure*}[t]
1068: \centerline{\psfig{file=table1.ps,width=14.0cm,angle=90}}
1069: \end{figure*}
1070: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1071:
1072: \subsection{Radio Cores}
1073:
1074: AGNs, no matter how weak, are almost never silent in the radio. Barring
1075: chance superposition with a supernova remnant, the presence of a compact radio
1076: core is therefore a good AGN indicator. Because of the
1077: expected faintness of the nuclei, however, any search for core emission must
1078: be conducted at high sensitivity, and arcsecond-scale angular resolution or
1079: better is generally needed to isolate the nucleus from the surrounding host,
1080: which emits copious diffuse synchrotron radiation. In practice, this
1081: requires an interferometer such as the VLA.
1082:
1083: The prevalence of weak AGNs in nearby early-type galaxies has been established
1084: from the VLA radio continuum studies of Sadler, Jenkins \& Kotanyi (1989) and
1085: Wrobel \& Heeschen (1991), whose 5~GHz surveys with $\Delta \theta \approx$
1086: 5\asec\ report a high incidence ($\sim30$\%--40\%) of radio cores in complete,
1087: optical flux-limited samples of elliptical and S0 galaxies. Interestingly,
1088: the radio detection rate is similar to the detection rate of optical
1089: emission lines (Figure~4), and the optical counterparts of the radio cores are
1090: mostly classified as LINERs (Phillips et al. 1986; Ho 1999a). Conversely,
1091: Heckman (1980b) showed that LINERs tend to be associated with compact radio
1092: sources. The radio powers are quite modest, generally in the range of
1093: $10^{19}-10^{21}$ W Hz$^{-1}$ at 5 GHz. When available, the spectral indices
1094: tend to be relatively flat (e.g., Wrobel 1991; Slee et al. 1994). With the
1095: exception of a handful of well-known radio galaxies with extended jets (Wrobel
1096: 1991), most of the radio emission is centrally concentrated.
1097:
1098: No comparable radio survey has been done for spiral galaxies. Over the last
1099: few years, however, a number of studies, mostly using the VLA, have
1100: systematically targeted sizable subsets of the Palomar galaxies, to the point
1101: that by now effectively the entire Palomar AGN sample has been surveyed at
1102: arcsecond ($\Delta \theta \approx$ 0\farcs15$-$2\farcs5) resolution (Filho,
1103: Barthel \& Ho 2000, 2002a, 2006; Nagar et al. 2000, 2002; Ho \& Ulvestad
1104: 2001; Filho et al. 2004; Nagar, Falcke \& Wilson 2005; Krips et al. 2007).
1105: Because the sensitivity, resolution, and observing frequency varied from study
1106: to study, each concentrating on different subclasses of objects, it is
1107: nontrivial to combine the literature results. The only survey that samples a
1108: significant fraction of the three LLAGN classes at a uniform sensitivity and
1109: resolution is that by Nagar et al. (2000, 2002; Nagar, Falcke \& Wilson
1110: 2005), which was conducted at 15~GHz and $\Delta \theta$ = 0\farcs15. The
1111: main drawback is that the sensitivity of this survey (rms $\approx$ 0.2 mJy)
1112: is rather modest, and mJy-level sources can be missed if they possess
1113: relatively steep spectra. Despite these limitations, Nagar et al. detected a
1114: compact core, to a high level of completeness, in 44\% of the LINERs.
1115: Importantly, to the same level of completeness, the Seyferts exhibit a very
1116: similar detection rate (47\%). LINER~2s have a lower detection rate than
1117: LINER~1s (38\% versus 63\%; see Table~1), but the same pattern is reflected
1118: almost exactly within the Seyfert population (detection rate 30\% for type~2s
1119: versus 72\% for type~1s). Transition objects, on the other hand, clearly
1120: differ, showing a markedly lower detection rate of only 16\%, consistent with
1121: the 8.4~GHz survey of Filho, Barthel \& Ho (2000, 2002a, 2006), where the
1122: detection rate is $\sim 25$\%. The statistical differences in the Hubble type
1123: distributions of the three AGN classes (Ho, Filippenko \& Sargent 2003)
1124: slightly complicate the interpretation of these results. To the extent that
1125: radio power shows a mild dependence on bulge strength or BH mass (Nagar,
1126: Falcke \& Wilson 2005; see Ho 2002a), the detection rates, strictly speaking,
1127: should be renormalized to account for the differences in morphological types
1128: among the three classes. This effect, however, will not qualitatively change
1129: the central conclusion: if a compact radio core guarantees AGN pedigree, then
1130: LINERs, of either type~1 or type~2, are just as AGN-like as Seyferts, whereas
1131: a significant fraction of transition objects (roughly half) may be unrelated
1132: to AGNs.
1133:
1134: The detection rates from the Nagar et al. survey can be viewed as firm lower
1135: limits. At $\Delta \theta$ = 1\asec\ and rms = 0.04 mJy at 1.4 and 5 GHz, for
1136: example, the detection rate for the Palomar Seyferts rises to 75\% (Ho \&
1137: Ulvestad 2001). Although no lower frequency survey of LINERs has been
1138: completed so far (apart from the lower resolution studies of Sadler, Jenkins,
1139: \& Kotanyi 1989 and Wrobel \& Heeschen 1991 confined to early-type galaxies),
1140: the preliminary study by Van~Dyk \& Ho (1998) of 29 LINERs at 5 and 3.6 GHz
1141: ($\Delta \theta$ = 0\farcs5; rms = 0.05$-$0.1 mJy) yielded a detection rate of
1142: over 80\%, again suggesting that LINERs and Seyferts have a comparably high
1143: incidence of radio cores.
1144:
1145: Importantly, a sizable, flux-limit subset of the 15 GHz detections has been
1146: reobserved with the Very Long Baseline Array at 5 GHz, and essentially
1147: {\it all}\ of them have been detected at milliarcsecond resolution (Nagar,
1148: Falcke \& Wilson 2005). The high brightness temperatures (\gax\
1149: $10^6-10^{11}$~K) leaves no doubt that the radio cores are nonthermal and
1150: genuinely associated with AGN activity.
1151:
1152: Where multifrequency data exist, their spectra tend to be flat or even mildly
1153: inverted ($\alpha \approx -0.2$ to $+0.2$; Ho et al. 1999b; Falcke et al. 2000;
1154: Nagar et al. 2000; Nagar, Wilson \& Falcke 2001; Ulvestad \& Ho 2001b;
1155: Anderson, Ulvestad \& Ho 2004; Doi et al. 2005; Krips et al. 2007), seemingly
1156: more optically thick than Seyferts (median $\alpha = -0.4$; Ulvestad \& Ho
1157: 2001a), and variability on timescales of months is common (Nagar et al. 2002;
1158: Anderson \& Ulvestad 2005). Both of these characteristics suggest that the
1159: radio emission in LINERs is mainly confined to a compact core or base of a
1160: jet. Seyfert galaxies contain radio cores as well, but they are often
1161: accompanied by linear, jetlike features resolved on arcsecond scales (e.g.,
1162: Ulvestad \& Wilson 1989; Kukula et al. 1995; Ho \& Ulvestad 2001; Gallimore et
1163: al. 2006). This extended component appears to be less prevalent in LINERs,
1164: although a definitive comparison must await a survey matched in resolution,
1165: sensitivity, and wavelength with that performed for the Seyferts (Ho \&
1166: Ulvestad 2001). Higher resolution images on milliarcsecond scales
1167: do resolve elongated structures akin to subparsec-scale jets, but most of the
1168: power is concentrated in a compact, high-brightness temperature core
1169: (Bietenholz, Bartel \& Rupen 2000; Falcke et al. 2000; Ulvestad \& Ho 2001b;
1170: Filho, Barthel \& Ho 2002b; Anderson, Ulvestad \& Ho 2004; Filho et al. 2004;
1171: Krips et al. 2007). The comprehensive summary presented in Nagar, Falcke \&
1172: Wilson (2005) indicates that the incidence of milliarcsecond-scale radio cores
1173: is similar for LINERs and Seyferts, but that subparsec-scale jets occur more
1174: frequently in LINERs.
1175:
1176: \subsection{X-ray Cores}
1177:
1178: X-ray observations provide another very effective tool to isolate LLAGNs and
1179: to diagnose their physical properties. Ultra-faint LLAGNs can be identified
1180: where none was previously known in the optical (e.g., Loewenstein et al.
1181: 2001; Ho, Terashima \& Ulvestad 2003; Fabbiano et al. 2004; Pellegrini et al.
1182: 2007; Wrobel, Terashima \& Ho 2008). Here, too, sensitivity and resolution
1183: are critical, as the central regions of galaxies contain a plethora of
1184: discrete nonnuclear sources, often suffused with a diffuse thermal plasma.
1185: {\it Chandra}, whose ACIS camera delivers $\sim$0\farcs5 images, is the
1186: instrument of choice, although in some instances even data at $\sim$5\asec\
1187: resolution (e.g., {\it ROSAT} HRI) can still provide meaningful constraints,
1188: especially if accompanied by spectral information (e.g., {\it XMM-Newton}).
1189:
1190: As in the radio, no truly unbiased high-resolution X-ray survey has yet been
1191: performed of an optical flux-limited sample of nearby galaxies. The closest
1192: attempt was made by Roberts \& Warwick (2000), who searched for X-ray nuclear
1193: sources in 83 Palomar galaxies ($\sim 20$\% of the total sample) having
1194: archival {\it ROSAT}\ HRI data. This subset is probably not unbiased, but it
1195: does encapsulate all the nuclear spectral classes in the Palomar survey.
1196: In total, X-ray cores were detected in 54\% of the sample, with Seyferts and
1197: LINERs (including transition objects) both showing a higher detection rate
1198: ($\sim$70\%) than absorption (30\%) or \hii\ nuclei (40\%). The high detection
1199: rate among the optically classified LLAGNs agrees well with other {\it ROSAT}\
1200: studies of Palomar sources (Koratkar et al. 1995; Komossa, B\"ohringer \&
1201: Huchra 1999; Halderson et al. 2001; Roberts, Schurch \& Warwick 2001), but
1202: the nonnegligible detection rate among the inactive members suggests that
1203: a significant fraction of the ``core'' flux may be nonnuclear emission [X-ray
1204: binaries (XRBs) and diffuse gas] insufficiently resolved by {\it ROSAT}.
1205:
1206: Observations with {\it Chandra}\ (e.g., Ho et al. 2001; Eracleous et al. 2002)
1207: confirm the suspicion that earlier X-ray studies may have suffered from
1208: confusion with extranuclear sources (Figure~5). Importantly, the sharp
1209: resolution and low background noise of ACIS enable faint point sources to be
1210: detected with brief (few ks) exposures. This makes feasible, for the first
1211: time, X-ray surveys of large samples of galaxies selected at non-X-ray
1212: wavelengths. In a snapshot survey of a distance-limited sample of Palomar
1213: LLAGNs, Ho et al. (2001) find that $\sim$75\% of LINERs, both type~1 and
1214: type~2, contain X-ray cores, some as faint as $\sim 10^{38}$ \lum\ in the
1215: $2-10$ keV band. Terashima \& Wilson (2003b) report an even
1216:
1217: \begin{figure}
1218: %%BoundingBox: 250 55 500 785
1219: \psfig{figure=fig5.ps,width=13.7cm,angle=0}
1220: \caption{{\it Chandra}/ACIS images of three LLAGNs, illustrating the diversity
1221: and complexity of the X-ray morphologies of their circumnuclear regions.
1222: The cross marks the near-IR position of the nucleus. (Courtesy of H.M.L.G.
1223: Flohic and M. Eracleous.)
1224: }
1225: \end{figure}
1226:
1227: \noindent
1228: higher detection
1229: rate (100\%) for a sample of LINERs chosen for having a flat-spectrum radio
1230: core. To date, roughly 50\% of the entire Palomar sample, among them 40\% of
1231: the AGNs, have been observed by {\it Chandra}. This rich archival resource
1232: has been the basis of a number of recent investigations focused on quantifying
1233: the AGN content of LINERs, chief among them Satyapal, Sambruna \& Dudik
1234: (2004), Dudik et al. (2005), Pellegrini (2005), Satyapal et al. (2005),
1235: Flohic et al. (2006), and Gonz\'alez-Mart\'\i n et al. (2006). A common
1236: conclusion that can be distilled from these studies is that the incidence of
1237: X-ray cores among LINERs is quite high, ranging from $\sim$50\% to 70\%, down
1238: to luminosity limits of $\sim 10^{38}$ \lum. The incidence of X-ray cores in
1239: LINERs is somewhat lower than, but still compares favorably to, that found in
1240: Palomar Seyferts ($\sim 90$\%), the vast majority of which now have suitable
1241: X-ray observations, as summarized in Cappi et al. (2006) and Panessa et al.
1242: (2006). While the impact of selection biases cannot be assessed easily,
1243: they are probably not very severe because most of the observations
1244: were not originally intended to study LINERs, nor were they targeting famous
1245: X-ray sources.
1246:
1247: It is of interest to ask whether the incidence of X-ray cores in LINERs
1248: depends on the presence of broad H\al\ emission. The moderate-resolution
1249: {\it ROSAT}/HRI studies of Roberts \& Warwick (2000) and Halderson et al.
1250: (2001) showed roughly comparable detection rates for type~1 and type~2
1251: LINERs, suggesting that the two classes are intrinsically similar and that
1252: obscuration plays a minor role in differentiating them. On the other hand,
1253: detailed X-ray spectral analysis has raised the suspicion that LINER~2s may be
1254: a highly heterogeneous class, with the bulk of the X-ray emission possibly
1255: arising from stellar processes. An important caveat is that these studies
1256: were based on large-beam observations, mostly using {\it ASCA}\ (Terashima,
1257: Ho \& Ptak 2000; Terashima et al. 2000a, 2002; Roberts, Schurch \& Warwick
1258: 2001) and the rest using {\it BeppoSAX}\ (Georgantopoulos et al. 2002;
1259: Pellegrini et al. 2002). A clearer, more consistent picture emerges from the
1260: recent {\it Chandra}\ work cited above. Although the individual samples
1261: remain small, most {\it Chandra}\ surveys detect LINER~2s with roughly similar
1262: frequency as LINER~1s, $\sim$50\%$-$60\%. To gain a more comprehensive
1263: census, I have assembled {\it Chandra}\ measurements for all Palomar LINERs
1264: from the literature, along with unpublished material for a significant number
1265: of additional objects in public archives, which were analyzed following Ho
1266: et al. (2001). Although clearly heterogeneous and incomplete, the final
1267: collection of 64 LINERs (20 type~1, 44 type~2) does constitute 70\% of the
1268: entire Palomar sample. The detection rate among all LINERs is 86\%, broken
1269: down into 95\% for LINER~1s and 74\% for LINER~2s. For completeness, note
1270: that a similar exercise for 36 transition objects (55\% of the parent sample)
1271: yields a detection rate of 74\%, identical to that of LINER~2s and only
1272: marginally lower than that of Seyfert~2s (86\%; Table~1).
1273:
1274: The X-ray spectral properties of LLAGNs, particularly LINERs, have most
1275: thoroughly been investigated using {\it ASCA}\ (Yaqoob et al. 1995; Ishisaki et
1276: al. 1996; Iyomoto et al. 1996, 1997, 1998a, 1998b; Ptak et al. 1996, 1999;
1277: Terashima et al. 1998a, 1998b, 2000a, 2000b, 2002; Ho et al. 1999a; Terashima,
1278: Ho \& Ptak 2000; Roberts, Schurch \& Warwick 2001), with important
1279: contributions from {\it BeppoSAX} (Pellegrini et al. 2000a, 2000b, 2002;
1280: Iyomoto et al. 2001; Georgantopoulos et al. 2002; Ptak et al. 2004). A
1281: seminal study on M81 was done using {\it BBXRT}\ (Petre et al. 1993). Although
1282: the nuclear
1283: component was not spatially isolated because of the poor angular resolution of
1284: these telescopes, they had sufficient effective area to yield good photon
1285: statistics over the energy range $\sim 0.5-10$ keV to {\it spectrally}\
1286: isolate the hard, power-law AGN signal. The most salient properties are the
1287: following. (1) Over the region $\sim 0.5-10$ keV, the continuum can be fit
1288: with a power law with an energy index of $\alpha \approx -0.4$ to $-1.2$.
1289: Although this range overlaps with that seen in more luminous sources, the
1290: typical value of $\sim -0.8$ in LLAGNs may be marginally flatter than in
1291: Seyfert~1s ($\langle\alpha\rangle = -0.87\pm0.22$; Nandra et al. 1997b) or
1292: radio-quiet quasars ($\langle\alpha\rangle = -0.93\pm0.22$; Reeves \& Turner
1293: 2000), perhaps being more in line with radio-loud quasars ($\langle \alpha
1294: \rangle = -0.6\pm0.16$; Reeves \& Turner 2000). (2) With a few notable
1295: exceptions (e.g., M51: Fukazawa et al. 2001, Terashima \& Wilson 2003a;
1296: NGC~1052: Weaver et al. 1999, Guainazzi et al. 2000; NGC~4258: Makishima et
1297: al. 1994, Fiore et al. 2001; NGC~4261: Matsumoto et al. 2001), the power-law
1298: component shows very little intrinsic absorption. This trend conflicts with
1299: the tendency for the degree of obscuration to increase with decreasing
1300: luminosity (e.g., Lawrence \& Elvis 1982). (3) Signatures of X-ray
1301: reprocessing by material from an optically thick accretion disk, in the form
1302: of Fe~K$\alpha$ emission or Compton reflection (Lightman \& White 1988; George
1303: \& Fabian 1991), are weak or absent; the weakness of the Fe~K$\alpha$ line in
1304: LLAGNs runs counter to the inverse correlation between iron line strength and
1305: luminosity observed in higher luminosity AGNs (Nandra et al. 1997b). (4) In
1306: the few cases where Fe~K$\alpha$ emission has been detected, it is always
1307: narrow. (5) Apart from the hard power law, most objects require an extra soft
1308: component at energies \lax\ 2 keV that can be fit by a thermal plasma model
1309: with a temperature of $kT \approx 0.4-0.8$ keV and near-solar abundances. (6)
1310: Contrary to the trend established for luminous sources (Nandra et al. 1997a),
1311: short-term, large-amplitude X-ray variability is rare in LLAGNs (Ptak et al.
1312: 1998).
1313:
1314: More recent observations with {\it Chandra}\ and {\it XMM-Newton}\ have
1315: refined, but not qualitatively altered, the above results. Where detailed
1316: spectral analysis is possible (e.g., B\"ohringer et al. 2001; Kim \& Fabbiano
1317: 2003; Pellegrini et al. 2003a; Terashima \& Wilson 2003b; Filho et al. 2004;
1318: Page et al. 2004; Starling et al. 2005; Flohic et al. 2006;
1319: Gonz\'alez-Mart\'\i n et al. 2006; Soria et al. 2006), the hard power-law
1320: component (except in objects previously known to be heavily absorbed)
1321: continues to be relatively unabsorbed, even among many type~2 sources, and to
1322: show little signs of reflection. No convincing case of a relativistic
1323: Fe~K$\alpha$ line has yet surfaced in an LLAGN. The marginally broad iron
1324: lines discovered with {\it ASCA}\ in M81 (Ishisaki et al. 1996) and NGC~4579
1325: (Terashima et al. 1998a) has now been resolved into multiple components
1326: (Dewangan et al. 2004; Page et al. 2004; Young et al. 2007), none of which can
1327: be associated with a canonical disk. At the same time, the equivalent width
1328: limits for even the narrow component have become impressively low (e.g., Ptak
1329: et al. 2004). Interestingly, a soft thermal component is still required in
1330: many objects (\S~5.4), but there is no evidence for blackbody-like soft excess
1331: emission commonly seen in Seyferts and quasars (e.g., Turner \& Pounds 1989;
1332: Inoue, Terashima \& Ho 2007).
1333:
1334: \subsection{Circumnuclear Thermal Plasma}
1335:
1336: Early X-ray observations of LLAGNs using {\it ASCA}\ have consistently revealed
1337: the presence of a diffuse, thermal component, typically with a temperature of
1338: $kT \approx 0.5$ keV (Ptak et al. 1999; Terashima et al. 2002). The uniform
1339: analysis of {\it ROSAT}\ data by Halderson et al. (2001) concluded that $\sim
1340: 80$\% of the Palomar sources contain an extended component. However, without
1341: better resolution, it was impossible to know the extent of confusion with
1342: point sources, how much of the gas is truly associated with the nuclear region
1343: of the galaxy, or the density and temperature profile of the gas.
1344:
1345: Our view of the diffuse component in the nuclear region has been dramatically
1346: sharpened with {\it Chandra}\ and {\it XMM-Newton}. Not only has the near
1347: ubiquity of diffuse gas been confirmed in many nearby galaxies (Ho et al.
1348: 2001; Eracleous et al. 2002; Terashima \& Wilson 2003b; Pellegrini 2005;
1349: Rinn, Sambruna \& Gliozzi 2005; Cappi et al. 2006; Gonz\'alez-Mart\'\i n et
1350: al. 2006; Soria et al. 2006), including
1351: our own (Muno et al. 2004) and our close neighbor M31 (Garcia et al.
1352: 2005), but quantitative, statistical properties of the gas are now becoming
1353: available. In the comprehensive investigation of 19 LINERs by Flohic et al.
1354: (2006), the diffuse emission, detected in 70\% of the sample, is concentrated
1355: within the central few hundred pc. With an average 0.5$-$2 keV luminosity of
1356: $\sim 10^{38}$ \lum, it accounts for more than half of the total central
1357: luminosity in most cases. The average spectrum is similar to that seen in
1358: normal galaxies: it can be described by a thermal plasma with $kT = 0.5$ keV
1359: plus a power-law component with $\alpha = -0.3$ to $-0.5$. I will return to
1360: the nature of the hard component in \S~6.5. What is the origin of the thermal
1361: plasma? Given what we know about the stellar populations (\S~4.2), a starburst
1362: origin, as suggested by Gonz\'alez-Mart\'\i n et al. (2006), seems improbable.
1363: In normal elliptical galaxies, the X-ray--emitting gas represents the
1364: repository of thermalized stellar ejecta generated from mass loss from evolved
1365: stars and Type~Ia supernovae (e.g., Awaki et al. 1994). There is no reason
1366: not to adopt the same picture to explain the hot plasma in LINERs and other
1367: LLAGNs. High-resolution X-ray spectroscopy of the highly ionized gas around
1368: the nucleus of M81 (Page et al. 2003; Young et al. 2007) and NGC~7213
1369: (Starling et al. 2005) reveals that the plasma is collisionally ionized.
1370: Starling et al. note that this may be a property unique to LINERs, as thermal
1371: gas in luminous Seyferts is usually photoionized rather than collisionally
1372: ionized (e.g., Kinkhabwala et al. 2002).
1373:
1374: \subsection{Broad-line Region}
1375:
1376: Luminous, unobscured AGNs distinguish themselves unambiguously by their
1377: characteristic broad permitted lines. The detection of broad H\al\ emission
1378: in
1379:
1380: \clearpage
1381: \begin{figure}
1382: %%BoundingBox: 50 10 590 760
1383: \psfig{figure=fig6.ps,width=13.7cm,angle=270}
1384: \caption{LINERs with broad, double-peaked H\al\ emission discovered with
1385: {\it HST}. A model fit for the disk profile in NGC 4450 is shown for
1386: illustration (green curve). (Adapted from Ho et al. 2000, Shields et al.
1387: 2000, and Barth et al. 2001a.)
1388: }
1389: \end{figure}
1390:
1391: \noindent
1392: $\sim$25\% of LINERs (Ho et al. 1997e) thus constitutes strong evidence in
1393: favor of the AGN interpretation of these sources. LINERs, like Seyferts,
1394: come in two flavors---some have a visible BLR (type~1), and others do not
1395: (type~2). The broad component becomes progressively more difficult to detect
1396: in ground-based spectra for permitted lines weaker than H\al. However, \hst\
1397: spectra of LINERs, when available, show broad higher-order Balmer lines as
1398: well as UV lines such as Ly$\alpha$, \civ\ \lamb1549, \mgii\ \lamb2800, and
1399: \feii\ multiplets (Barth \etal 1996; Ho, Filippenko \& Sargent 1996).
1400: A subset of LINERs contain broad lines with {\it double-peaked}\ profiles
1401: (Figure~6), analogous to those seen in a minority of radio galaxies (Eracleous
1402: \& Halpern 1994), where they are often interpreted as a kinematic signature of
1403: a relativistic accretion disk (Chen \& Halpern 1989). Most
1404: of the nearby cases have been discovered serendipitously, either as a result
1405: of the broad component being variable (NGC~1097: Storchi-Bergmann, Baldwin \&
1406: Wilson 1993; M81: Bower et al. 1996; NGC~3065: Eracleous \& Halpern 2001) or
1407: because of the increased sensitivity to weak, broad features afforded by
1408: small-aperture measurements made with {\it HST}\ (NGC~4450: Ho et al. 2000;
1409: NGC~4203: Shields et al. 2000; NGC~4579: Barth et al. 2001a). Double-peaked
1410: broad-line AGNs may be more common than previously thought, especially among
1411: LLAGNs, perhaps as a consequence of their accretion disk structure (\S~8).
1412:
1413: A pressing question, however, is: What fraction of the more numerous LINER~2s
1414: are AGNs? By analogy with the Seyfert~2 class, do LINER~2s contain a hidden
1415: LINER~1 nucleus? At first sight, it might seem that there is no
1416: {\it a priori}\ reason why the orientation-dependent unification model, which
1417: has enjoyed much success in the context of Seyfert galaxies, should not apply
1418: equally to LINERs. If we suppose that the ratio of LINER~2s to LINER~1s is
1419: similar to the ratio of Seyfert~2s to Seyfert~1s---1.6:1 in the Palomar
1420: survey---we can reasonably surmise that the AGN fraction in LINERs may be as
1421: high as $\sim$60\%. That at least some LINERs {\it do}\ indeed contain a
1422: hidden BLR was demonstrated by the deep Keck spectropolarimetric observations
1423: of Barth, Filippenko \& Moran (1999a, 1999b). In a survey of 14 LLAGNs,
1424: mostly LINERs, these authors detected broad H\al\ emission in three objects
1425: ($\sim 20$\%) polarized at a level of 1\%$-$3\%. Interestingly, all three
1426: objects are elliptical galaxies with double-sided radio jets. NGC~315 and
1427: NGC~1052 technically qualify as type~1.9 LINERs (Ho et al. 1997e), whereas
1428: NGC~4261 is a LINER~2. Although the sample is small, these observations
1429: prove two important points: (1) the weak broad H\al\ features detected in
1430: direct light is not always scattered emission (Antonucci 1993), since
1431: polarized emission was not detected in several other LINER~1.9s included in
1432: Barth, Filippenko \& Moran's survey; (2) an obscured nucleus does lurk in
1433: some LINER~2s.
1434:
1435: At the same time, other bright LINER~2s have resisted detection by
1436: spectropolarimetry. As in the case of Seyferts (Tran 2001), however, the
1437: nondetection of polarized broad lines does not necessarily imply that there is
1438: no hidden BLR. Nevertheless, the BLR in some type~2 AGNs, especially LINERs
1439: but also Seyferts, may be intrinsically absent, not obscured. In the case of
1440: some Seyferts, mostly weak sources, the evidence comes from low absorbing
1441: X-ray column densities (Bassani et al. 1999; Pappa et al. 2001; Panessa \&
1442: Bassani 2002; Gliozzi et al. 2004; Cappi et al. 2006; Gliozzi, Sambruna \&
1443: Foschini 2007; Bianchi et al. 2008; but see Ghosh et al. 2007) as well as
1444: optical variability (Hawkins 2004). LINERs, as a class, very much conform to
1445: this picture. As discussed further below, LINERs of either type generally
1446: show very little sign of absorbing or reprocessing material, and UV
1447: variability is common. A few exceptions exist (e.g., NGC~1052: Guainazzi et
1448: al. 2000; NGC~4261: Sambruna et al. 2003, Zezas et al. 2005), but,
1449: interestingly, these are precisely the very ones for which Barth, Filippenko,
1450: \& Moran discovered hidden BLRs. NGC~4258, also highly absorbed in the X-rays
1451: (Fiore et al. 2001), shows polarized narrow lines rather than broad lines
1452: (Barth et al. 1999).
1453:
1454: An excellent of a LINER with a naked type~2 nucleus is the Sombrero galaxy.
1455: Although clearly an AGN, it shows no trace of a broad-line component, neither
1456: in direct light (Ho et al. 1997e), not even when very well isolated with a
1457: small \hst\ aperture (Nicholson et al. 1998), nor in polarized light (Barth,
1458: Filippenko \& Moran 1999b). Its Balmer decrement indicates little reddening
1459: to the NLR. For all practical purposes, the continuum emission from the
1460: nucleus looks unobscured. It is detected as a variable UV source (Maoz et al.
1461: 1995, 2005) and in the soft and hard X-rays (Nicholson et al. 1998; Ho et al.
1462: 2001). The X-ray spectrum is only very mildly absorbed (Nicholson et al.
1463: 1998; Pellegrini et al. 2002, 2003a; Terashima et al. 2002), with no signs of
1464: Fe~K$\alpha$ emission expected from reprocessed material, consistent with the
1465: modest mid-IR emission reported by Bendo et al. (2006). In short, there is no
1466: sign of anything being hidden or much doing the hiding. So where is the BLR?
1467: It is just not there.
1468:
1469: The lack of a BLR in very low-luminosity sources may be related to a physical
1470: upper limit in the broad-line width (Laor 2003). If LLAGNs obey the same
1471: BLR-luminosity relation as in higher luminosity systems, their BLR velocity
1472: depends on the BH mass and luminosity. At a limiting bolometric luminosity of
1473: $L_{\rm bol} \approx 10^{41.8} (M_{\rm BH}/10^8\,M_\odot)^2$ \lum, $\Delta
1474: v\approx 25,000$ \kms, above which clouds may not survive due to excessive
1475: shear or tidal forces. Alternatively, if BLR clouds arise from condensations
1476: in a radiation-driven, outflowing wind (Murray \& Chiang 1997), a viewpoint
1477: now much espoused, then it is reasonable to expect that very low-luminosity
1478: sources would be incapable of generating a wind, and hence of sustaining a
1479: BLR. For example, the clumpy torus model of Elitzur \& Shlosman (2006)
1480: predicts that the BLR can no longer be sustained for $L_{\rm bol}$ \lax\
1481: $10^{42}$ \lum. In the scenario of Nicastro (2000), the BLR originates from a
1482: disk outflow formed at the transition radius between regions dominated by gas
1483: and radiation pressure. As this radius shrinks with decreasing
1484: $L_{\rm bol}/L_{\rm Edd}$, where $L_{\rm Edd} = 1.3\times 10^{38}\,(M_{\rm BH}/
1485: M_{\odot})$ \lum, the BLR is expected to disappear for
1486: $L_{\rm bol}/L_{\rm Edd}$ \lax\ $10^{-3}$. The apparent correlation between
1487: BLR line width and $L_{\rm bol}/L_{\rm Edd}$ qualitatively supports this
1488: picture (Xu \& Cao 2007). Although the existing data are sparse, they
1489: indicate that LINERs generally lack UV resonance absorption features
1490: indicative of nuclear outflows (Shields et al. 2002). The models by Elitzur
1491: \& Shlosman and Nicastro are probably correct in spirit but not in detail,
1492: because many of the Palomar LLAGNs plainly violate their proposed thresholds
1493: (\S~5.10).
1494:
1495: Nonetheless, the statistics within the Palomar survey already provide tentative
1496: support to the thesis that the BLR vanishes at the lowest luminosities or
1497: Eddington ratios. Which of the two is the controlling variable is
1498: still difficult to say. For both Seyferts and LINERs, type~1 sources are
1499: almost a factor of 10 more luminous than type~2 sources in terms of their
1500: median total H\al\ luminosity (Table~1). (The statistical differences between
1501: type~1 and type~2 sources cannot be ascribed to sensitivity differences in the
1502: detectability of broad H\al\ emission. Type~1 objects do have stronger line
1503: emission compared to the type~2s, but on average their narrow H\al\ flux and
1504: equivalent width are only $\sim 50$\% higher, and the two types overlap
1505: significantly. Moreover, as noted in \S 3.4, the broad H\al\ detection rates
1506: turn out to be quite robust even in light of the much higher sensitivity
1507: afforded by \hst.) The differences persist after normalizing by the Eddington
1508: luminosities: adopting a bolometric correction of $L_{\rm bol} \approx 16
1509: L_{\rm X}$, $L_{\rm bol}/L_{\rm Edd}=1.1\times10^{-3}$ and $5.9\times10^{-6}$
1510: for Seyfert~1s and Seyfert~2s, respectively, whereas the corresponding values
1511: for LINER~1s and LINER~2s are $1.0\times10^{-5}$ and $4.8\times10^{-6}$. Two
1512: caveats are in order. First, while most of the type~1 sources have X-ray
1513: data, only 60\% of the LINER~2s and 70\% of the Seyfert~2s do. Second, the
1514: X-ray luminosities, which pertain to the $2-10$ keV band, have been corrected
1515: for intrinsic absorption whenever possible, but many sources are too faint for
1516: spectral analysis. The lower X-ray luminosities for the type~2 sources must
1517: be partly due to absorption, but considering the generally low absorbing
1518: columns, particularly among the LINERs (Georgantopoulous et al. 2002;
1519: Terashima et al. 2002), it is unclear if absorption alone can erase the
1520: statistical difference between the two types. The tendency for Seyfert~2s to
1521: have lower Eddington ratios than Seyfert~1s has previously been noted, for the
1522: Palomar sample (Panessa et al. 2006) and others (Middleton, Done \& Schurch
1523: 2008).
1524:
1525: Several authors have raised the suspicion that LINER~2s may not be
1526: accretion-powered. Large-aperture X-ray spectra of LINER~2s, like those of
1527: LINER~1s, can be fit with a soft thermal component plus a power law with
1528: $\alpha \approx -0.7$ to $-1.5$ (Georgantopoulos et al. 2002; Terashima et al.
1529: 2002). But this alone does not provide enough leverage to distinguish AGNs from
1530: starburst galaxies, many of which look qualitatively similar over the limited
1531: energy range covered by these observations. We cannot turn to the iron K\al\
1532: line or variability for guidance, because LLAGNs generally exhibit neither
1533: (\S~5.3). The hard X-ray emission in LINER~2s is partly extended (Terashima
1534: et al. 2000a; Georgantopoulos et al. 2002), but the implications of this
1535: finding are unclear. Just because the X-ray emission surrounding the LLAGN is
1536: morphologically complex and there is evidence for circumnuclear star formation
1537: (e.g., NGC~4736; Pellegrini et al. 2002) does not necessarily imply that there
1538: is a causal connection between the starburst and the LLAGN. Roberts, Schurch
1539: \& Warwick (2001) advocate a starburst connection from the observation that
1540: LINER~2s have a mean flux ratio in the soft and hard X-ray band ($\sim 0.7$)
1541: similar to that found in NGC~253. This interpretation, however, conflicts
1542: with the stellar population constraints discussed in \S~4.2. It is also not
1543: unique. Luminous, AGN-dominated type~1 sources themselves exhibit a tight
1544: correlation between soft and hard X-ray luminosity, with a ratio not
1545: dissimilar from the quoted value (Miniutti et al. 2008).
1546:
1547: An important clue comes from the fact that many LINER~2s have a lower
1548: $L_{\rm X}/L_{{\rm H}\alpha}$ ratio than LINER~1s (Ho et al. 2001). In
1549: particular, the observed X-ray luminosity from the nucleus, when extrapolated
1550: to the UV, does not have enough ionizing photons to power the H\al\ emission
1551: (Terashima et al. 2000a). This implies that (1) the X-rays are heavily
1552: absorbed, (2) nonnuclear processes power much of the optical line emission,
1553: or (3) the ionizing SED is different than assumed. As discussed in \S~6.4, this
1554: energy budget discrepancy appears to be symptomatic of all LLAGNs in general,
1555: not just LINER~2s, and most likely results from a combination of the second and
1556: third effect. There are some indications that the SEDs of LINER~2s indeed
1557: differ systematically from those of LINER~1s (e.g., Maoz et al. 2005; Sturm et
1558: al. 2006). In light of the evidence given in \S\S~5.3, 5.6, I consider the
1559: first solution to be no longer tenable. One can point to objects such as
1560: NGC~4261 (Zezas et al. 2005) as examples of LINER~2s with strong obscuration,
1561: but such cases are rare.
1562:
1563: From the point of view of BH demographics, the most pressing issue is what
1564: fraction of the LINER~2s should be included in the AGN tally. Some cases are
1565: beyond dispute (M84, M87, Sombrero). What about the rest? The strongest
1566: argument that the majority of LINER~2s are AGN-related comes from the
1567: detection frequency of radio (\S~5.2) and X-ray (\S~5.3) cores, which is
1568: roughly 60\% of that of LINER~1s. On the other hand, the detection rate of
1569: Seyfert~2s are similarly lower compared to Seyfert~1s, most likely reflecting
1570: the overall reduction of nuclear emission across all bands in type~2 LLAGNs
1571: as a consequence of their lower accretion rates. In summary, the AGN fraction
1572: among LINER~2s is at least 60\%, and possibly as high as 100\%.
1573:
1574: \subsection{Torus}
1575:
1576: In line with the absence of a BLR discussed above and using very much the
1577: same set of evidence, a convincing case can be made that the torus also
1578: disappears at very low luminosities. In a large fraction of nearby LINERs,
1579: the low absorbing column densities and weak or undetected Fe~K$\alpha$ emission
1580: (\S~5.3) strongly indicate that we have a direct, unobstructed view of the
1581: nucleus. Ghosh et al. (2007) warn that absorbing columns can be underestimated
1582: in the presence of extended soft emission, especially when working with
1583: spectra of low signal-to-noise ratio. While this bias no doubt enters at some
1584: level, cases like the Sombrero (\S~5.3) cannot be so readily dismissed. By
1585: analogy with situation in luminous AGNs (e.g., Inoue, Terashima \& Ho 2007;
1586: Nandra et al. 2007), type~1 LLAGNs, if they possess tori, should also show
1587: strong, narrow fluorescent Fe~K$\alpha$ emission. This expectation is not
1588: borne out by observations. NGC~3998, which has excellent X-ray data, offers
1589: perhaps the most dramatic example. Apart from showing no signs whatsoever for
1590: intrinsic photoelectric absorption, it also possesses one of the tightest
1591: upper limits to date on Fe~K$\alpha$ emission: EW $<$ 25 eV (Ptak et al.
1592: 2004). Our sight line to the nucleus is as clean as a whistle. Satyapal,
1593: Sambruna \& Dudik (2004) claim that many LINERs have obscured nuclei, but
1594: this conclusion is based on IR-bright, dusty objects chosen from Carrillo et
1595: al. (1999); as I have discussed in \S~3.2, I regard these objects not only as
1596: biased, but also confusing with respect to their nuclear properties.
1597:
1598: Palomar Seyferts, whose luminosities and Eddington ratios are about an order
1599: of magnitude higher than those of LINERs (\S~5.10), show markedly larger
1600: absorbing column densities and stronger Fe~K$\alpha$ lines. In an
1601: {\it XMM-Newton}\ study of a
1602: distance-limited sample of 27 Palomar Seyferts, Cappi et al. (2006) detect
1603: strong Fe~K$\alpha$ emission in over half of objects. The distribution of
1604: absorbing columns is nearly continuous, from $N_{\rm H} \approx 10^{20}$ to
1605: $10^{25}~{\rm cm}^{-2}$, with 30\%$-$50\% of the type~2 sources being
1606: Compton-thick (Panessa et al. 2006). This seems consistent with the tendency
1607: for Seyferts to be more gas-rich than LINERs, to the extent that this is
1608: reflected in their higher NLR densities (Ho, Filippenko \& Sargent 2003).
1609:
1610: The trend of increasing absorption with increasing luminosity or Eddington
1611: ratio observed in Palomar LLAGNs has an interesting parallel among radio
1612: galaxies. A substantial body of recent work indicates that the nuclei of FR~I
1613: sources, most of which are, in fact, LINERs, are largely unobscured (e.g.,
1614: Chiaberge, Capetti \& Celotti 1999; Donato, Sambruna \& Gliozzi 2004;
1615: Balmaverde \& Capetti 2006). In contrast, FR~II systems, especially those with
1616: broad or high-excitation lines (analogs of Seyferts), show clear signs of
1617: absorption and Fe~K$\alpha$ emission (Evans et al. 2006).
1618:
1619: Even if we are fooled by the X-ray observations, substantial absorption
1620: must result in strong thermal reemission of ``waste heat'' in the IR. While
1621: sources such as Cen~A provide a clear reminder that every rule has its
1622: exception (Whysong \& Antonucci 2004), the existing data do suggest that,
1623: as a class, FR~I radio galaxies tend to be weak mid-IR or far-IR sources
1624: (Haas et al. 2004; M\"uller et al. 2004). The same holds for more nearby
1625: LINERs. Their SEDs do show a pronounced mid-IR peak (\S~5.8), but as I will
1626: argue later, it is due to emission from the accretion flow rather than from
1627: dust reemission.
1628:
1629: \subsection{Narrow-line Region Kinematics}
1630:
1631: The kinematics of the NLR are complex. At the smallest scales probed by \hst,
1632: Verdoes~Kleijn, van~der~Marel \& Noel-Storr (2006) find that the velocity
1633: widths of the ionized gas in the LINER nuclei of early-type galaxies can be
1634: modeled as unresolved rotation of a thin disk in the gravitational potential
1635: of the central BH. The subset of objects with FR~I radio morphologies, on the
1636: other hand, exhibit line broadening in excess of that expected from purely
1637: gravitational motions; these authors surmise that the super-virial motions may
1638: be related to an extra source of energy injection by the radio jet.
1639: Walsh et al. (2008) use multiple-slit STIS observations to map the
1640: kinematics of the inner $\sim 100$ pc of the NLR in a sample of 14 LLAGNs,
1641: mostly LINERs. Consistent with earlier findings (Ho et al. 2002; Atkinson et
1642: al. 2005), the velocity fields are generally quite disorganized, rarely
1643: showing clean signatures of dynamically cold disks undergoing circular rotation.
1644: Nevertheless, two interesting trends can be discerned. The emission line
1645: widths tend to be largest within the sphere of influence of the BH,
1646: progressively decreasing toward large radii to values that roughly match the
1647: stellar velocity dispersion of the bulge. The luminous members of the sample,
1648: on the other hand, show more chaotic kinematics, as evidenced by large velocity
1649: splittings and asymmetric line profiles, reminiscent of the pattern observed
1650: by Rice et al. (2006) in their sample of Seyfert galaxies. Walsh et al.
1651: suggest that above a certain luminosity threshold---one that perhaps coincides
1652: with the LINER/Seyfert division---AGN outflows and radio jets strongly perturb
1653: the kinematics of the NLR.
1654:
1655: A large fraction ($\sim 90$\%) of the Palomar LLAGNs have robust measurements
1656: of integrated \nii\ \lamb6583 line widths, which enable a crude assessment of
1657: the dynamical state of the NLR and its relation to the bulge. Consistent with
1658: what has been established for more powerful systems (Nelson \& Whittle 1996;
1659: Greene \& Ho 2005a), the kinematics of the ionized gas are dominated by random
1660: motions that, to first order, trace the gravitational potential of the stars
1661: in the bulge. Among the objects with available central stellar velocity
1662: dispersions, $\sigma_{\rm NLR}/\sigma_* \approx 0.7-0.8$ for the weakest
1663: sources ($L_{{\rm H}\alpha} \approx 10^{38}$ \lum), systematically rising to
1664: $\sigma_{\rm NLR}/\sigma_* \approx 1.2$ in the more luminous members
1665: ($L_{{\rm H}\alpha} \approx 10^{41.5}$ \lum). L.C. Ho (in preparation)
1666: speculates that the central AGN injects a source of dynamical heating of
1667: nongravitational origin to the NLR, either in the form of radiation pressure
1668: from the central continuum or mechanical interaction from radio jets. Given
1669: the empirical correlation between optical line luminosity and radio power
1670: (e.g., Ho \& Peng 2001; Ulvestad \& Ho 2001a; Nagar, Falcke \& Wilson 2005),
1671: and the near ubiquity of compact radio sources, it is {\it a priori}\
1672: difficult to determine which of these two sources acts as the primary driver.
1673: The tendency for extended radio emission to be more prevalent in Seyferts
1674: (\S~5.2) suggests that jets may be more important.
1675:
1676: \subsection{Spectral Energy Distribution}
1677:
1678: The broad-band SED provides one of the most fundamental probes of the physical
1679: processes in AGNs. Both thermal and nonthermal emission contribute to the
1680: broad-band spectrum of luminous AGNs such as quasars and classical Seyfert
1681: galaxies. In objects whose intrinsic spectrum has not been modified severely
1682: by relativistic beaming or absorption, the SED can be separated into several
1683: distinctive components (e.g., Elvis et al. 1994): radio synchrotron emission
1684: from a jet, which may be strong (``radio-loud'') or weak (``radio-quiet''); an
1685: IR excess, now generally considered to be predominantly thermal
1686: reradiation by dust grains; a prominent optical to UV ``big blue bump,''
1687: usually interpreted to be pseudo-blackbody emission from an optically thick,
1688: geometrically thin accretion disk (Shields 1978; Malkan \& Sargent 1982); a
1689: soft X-ray excess, whose origin is still highly controversial (Done et al.
1690: 2007; Miniutti et al. 2008); and an underlying power law, which is most
1691: conspicuous at hard X-ray energies but is thought to extend down to IR
1692: wavelengths, that can be attributed to Comptonization of softer seed photons.
1693:
1694: Within this backdrop, there were already early indications that the SEDs of
1695: LINERs may deviate from the canonical form. Halpern \& Filippenko (1984)
1696: succeeded in detecting the featureless optical continuum in NGC~7213, and
1697: while these authors suggested that a big blue bump may be present in this
1698: object, they also noted that it possesses an exceptionally high
1699: X-ray--to--optical flux ratio, although perhaps one not inconsistent with the
1700: extrapolation of the trend of increasing X-ray--to--optical flux ratio with
1701: decreasing luminosity seen in luminous sources (Zamorani et al. 1981; Avni \&
1702: Tananbaum 1982). A more explicit suggestion that LINERs may possess a weak UV
1703: continuum was made in the context of double-peaked broad-line AGNs such as
1704: Arp~102B and Pictor~A, whose narrow-line spectra share many characteristics
1705: with LINERs (Chen \& Halpern 1989; Halpern \& Eracleous 1994). The {\it HST}\
1706: spectrum of Arp~102B, in fact, shows an exceptionally steep optical-UV
1707: nonstellar continuum ($\alpha \approx -2.1$ to $-2.4$; Halpern et al. 1996).
1708: Halpern \& Eracleous (1994) further suggested that the SEDs are flat in the
1709: far-IR. In an important study of M81, Petre et al. (1993) proposed that the
1710: relative weakness of the UV continuum compared to the X-rays is a consequence
1711: of a change in the structure of the central accretion flow, from a standard
1712: thin disk to an ion-supported torus (see \S~8.3). Parameterizing the
1713: two-point spectral index between 2500 \AA\ and 2 keV by \alphaox\ $\equiv\,
1714: [\log L_\nu({\rm 2500~\AA}) - \log L_\nu({\rm 2~keV})]/
1715: [\log \nu({\rm 2500~\AA}) - \log \nu({\rm 2~keV})]$, M81 and possibly other
1716: LINERs (Mushotzky 1993) have \alphaox\ \gax\ $-1$, to be compared with
1717: \alphaox\ $\approx -1.4$ for quasars and \alphaox\ $\approx -1.2$ for Seyferts
1718: (Mushotzky \& Wandel 1989).
1719:
1720: The full scope of the spectral uniqueness of LLAGNs only became evident once
1721: the modern, albeit still fragmentary, multiwavelength data could be assembled.
1722: The initial studies concentrated on individual objects, emphasizing the
1723: weakness of the UV bump (M81: Ho, Filippenko \& Sargent 1996; Sombrero:
1724: Nicholson et al. 1998) and the overall consistency of the SED with spectral
1725: models generated from advection-dominated accretion flows (ADAFs; see Narayan
1726: 2002 and Yuan 2007 for reviews) as unique attributes of systems with low
1727: Eddington ratios (NGC~4258: Lasota et al. 1996, Chary et al. 2000; M87:
1728: Reynolds et al. 1996; M60: Di~Matteo \& Fabian 1997). Ho (1999b)
1729: systematically investigated the SEDs of a small sample of seven LLAGNs with
1730: available BH mass estimates and reliable small-aperture fluxes from radio to
1731: X-ray wavelengths. This was followed by a study of another
1732: five similar objects, which have the additional distinction of having
1733: double-peaked broad emission lines (Ho et al. 2000; Ho 2002b). Figure~7 gives
1734: the latest update from a comprehensive analysis of the SEDs of 150 nearby
1735: type~1 AGNs spanning 4 dex in BH mass ($M_{\rm BH}\approx 10^5-10^9$ \solmass)
1736: and 6.5 dex in Eddington ratio ($L_{\rm bol}/L_{\rm Edd} \approx 10^{-6} -
1737: 10^{0.5}$). Let us focus on two regimes: $L_{\rm bol}/L_{\rm Edd} = 0.1$ to
1738: 1, typical of classical, luminous AGNs, and $L_{\rm bol}/L_{\rm Edd} <
1739: 10^{-3.0}$, which characterizes most nearby LLAGNs (\S~5.10). I defer the
1740: discussion of the physical implications until \S~8, but for now list the most
1741: notable features concerning the LLAGN SED, some of which are also apparent in
1742: the composite LINER SED assembled by Eracleous, Hwang \& Flohic (2008a). (1)
1743: The big blue bump is conspicuously absent. (2) Instead, a broad excess is
1744: shifted to the mid-IR, forming a ``big red bump''; this component is probably
1745: related to the mid-IR excess previously noted by Lawrence et al. (1985),
1746: Willner et al. (1985), and Chen \& Halpern (1989), and more recently from
1747: {\it Spitzer}\ observations (e.g., Willner et al. 2004; Bendo et al. 2006; Gu
1748: et al. 2007). (3) As a consequence of this shift, the optical-UV slope is
1749: exceptionally steep, generally in the range $\alpha_{\rm ou} \approx -1$ to
1750: $-2.5$, to be compared with $\alpha_{\rm ou} \approx -0.5$ to $-0.7$ for
1751: luminous AGNs
1752: (Vanden~Berk et al. 2001; Shang et al. 2005); the X-ray--to--optical ratio is
1753: large, resulting in \alphaox\ \gax\ $-1$. (4) There is no evidence for a soft
1754: X-ray excess. (5) Lastly, the overall SED can be considered radio-loud,
1755: defined here by the convention that the radio-to-optical luminosity ratio
1756: exceeds a value
1757:
1758: \clearpage
1759: \begin{figure}
1760: %%BoundingBox: 60 5 550 760
1761: \hskip -0.2cm
1762: \psfig{figure=fig7.ps,width=14cm,angle=270}
1763: \caption{Composite SEDs for radio-quiet AGNs binned by Eddington
1764: ratio. The SEDs are normalized at 1 \micron. (Adapted from L.C. Ho, in
1765: preparation.)
1766: }
1767: \end{figure}
1768:
1769: \noindent
1770: of $R_o \equiv\,L_{\nu}({\rm 5~GHz})/L_{\nu}(B) = 10$.
1771: Radio-loudness, in fact, seems to be a property common to essentially
1772: {\it all}\ nearby weakly active nuclei (Ho 1999b, 2002a; Ho et al. 2000) and a
1773: substantial fraction of Seyfert nuclei (Ho \& Peng 2001). Defining
1774: radio-loudness based on the relative strength of the radio and X-ray emission,
1775: $R_{\rm X} \equiv\,\nu L_{\nu}({\rm 5~GHz})/ L_{\rm X}$, Terashima \& Wilson
1776: (2003b) also find that LINERs tend to be radio-loud, here taken to be
1777: $R_{\rm X} > 10^{-4.5}$. Moreover, the degree of radio-loudness scales
1778: inversely with $L_{\rm bol}/L_{\rm Edd}$ (Ho 2002a; Terashima \& Wilson 2003b;
1779: Wang, Luo \& Ho 2004; Greene, Ho \& Ulvestad 2006; Panessa et al. 2007;
1780: Sikora, Stawarz \& Lasota 2007; L.C. Ho, in preparation; see Figure~10{\it b}).
1781:
1782: In a parallel development, studies of the low-luminosity, often LINER-like
1783: nuclei of FR~I radio galaxies also support the notion that they lack a UV
1784: bump. M84 (Bower et al. 2000) and M87 (Sabra et al. 2003) are two familiar
1785: examples, but it has been well documented that FR~I nuclei tend to exhibit
1786: flat \alphaox\ (Donato, Sambruna \& Gliozzi 2004; Balmaverde, Capetti \&
1787: Grandi 2006; Gliozzi et al. 2008) and steep slopes in the optical (Chiaberge,
1788: Capetti \& Celotti 1999; Verdoes~Kleijn et al. 2002) and optical-UV (Chiaberge
1789: et al. 2002).
1790:
1791: Finally, I note that the UV spectral slope can be indirectly constrained
1792: from considering the strength of the \heii\ \lamb4686 line. While this
1793: line is clearly detected in Pictor~A (Carswell et al. 1984; Filippenko 1985),
1794: its weakness in NGC~1052 prompted P\'equignot (1984) to deduce that the
1795: ionizing spectrum must show a sharp cutoff above the He$^+$ ionization limit
1796: (54.4 eV). In this respect, NGC~1052 is quite representative of LINERs in
1797: general. \heii\ \lamb4686 was not detected convincingly in a {\it single}\
1798: case among a sample of 159 LINERs in the entire Palomar survey (Ho,
1799: Filippenko \& Sargent 1997a). Starlight contamination surely contributes
1800: partly to this, but the line has also eluded detection in {\it HST}\ spectra
1801: (e.g., Ho, Filippenko \& Sargent 1996; Nicholson et al. 1998; Barth et al.
1802: 2001b; Sabra et al. 2003; Sarzi et al. 2005; Shields et al. 2007), which
1803: indicates that it is truly intrinsically very weak. To a first
1804: approximation, the ratio of \heii\ \lamb4686 to H\bet\ reflects the relative
1805: intensity of the ionizing continuum between 1 and 4 Ryd. For an ionizing
1806: spectrum $f_\nu \propto \nu^{\alpha}$, case~B recombination predicts
1807: \heii\ \lamb4686/H\bet\ = $1.99 \times 4^\alpha$ (Penston \& Fosbury 1978).
1808: The current observational limits of \heii\ \lamb4686/H\bet\ \lax\ 0.1 thus
1809: imply $\alpha$ \lax\ $-2$, qualitatively consistent with the evidence from the
1810: SED studies.
1811:
1812: Maoz (2007) has offered an alternative viewpoint to the one presented above.
1813: Using a sample of 13 LINERs with variable UV nuclei, he argues that their
1814: SEDs do not differ appreciably from those of more luminous AGNs, and hence that
1815: LINERs inherently have very similar accretion disks compared to powerful AGNs.
1816: Maoz does not disagree that LINERs have large X-ray--to-UV flux ratios or that
1817: they tend to be radio-loud; his data show both trends. Rather, he contends
1818: that because LINERs lie on the low-luminosity extrapolation of the well-known
1819: relation between \alphaox\ and luminosity (Zamorani et al. 1981; Avni \&
1820: Tananbaum 1982; Strateva et al. 2005) they do not form a distinct population.
1821: And while LINERs do have large values of $R_o$, they nonetheless occupy the
1822: ``radio-quiet'' branch of the $R_o$ versus $L_{\rm bol}/L_{\rm Edd}$ plane
1823: (Sikora, Stawarz \& Lasota 2007). In my estimation, the key point is not,
1824: and has never been, whether LINERs constitute a disjoint class of AGNs, but
1825: whether they fit into a physically plausible framework in which their
1826: distinctive SEDs, among other properties, find a natural, coherent
1827: explanation. Section 8 attempts to offer such a framework.
1828:
1829: It should be noted that Maoz's results strongly depend on his decision to
1830: exclude all optical and near-IR data from the SEDs, on the grounds that they
1831: may be confused by starlight. I think this step is too draconian, as it
1832: throws away valuable information. While stellar contamination is certainly a
1833: concern, one can take necessary precautions to try to isolate the nuclear
1834: emission as much as possible, either through high-resolution imaging (e.g., Ho
1835: \& Peng 2001; Ravindranath et al. 2001; Peng et al. 2002) or spectral
1836: decomposition. In well-studied sources, there is little doubt that the
1837: optical continuum is truly both featureless and nonstellar (e.g., Halpern \&
1838: Filippenko 1984; Ho, Filippenko \& Sargent 1996; Ho et al. 2000; Bower et al.
1839: 2000; Sabra et al. 2003). Given what we know about the nuclear stellar
1840: population, we cannot assign the featureless continuum to young stars. In a
1841: few cases, the nonstellar nature of the nucleus can even be established
1842: through variability in the optical (Bower et al. 2000; Sabra et al. 2003;
1843: O'Connell et al. 2005) and mid-IR (Rieke, Lebofsky \& Kemp 1982; Grossan et
1844: al. 2001; Willner et al. 2004).
1845:
1846: While the SEDs of LINERs differ from those of traditional AGNs, it is
1847: important to recognize that they are decidedly {\it nonstellar}\ and approximate
1848: the form predicted for radiatively inefficient accretion flows (RIAFs)
1849: onto BHs, often coupled to a jet (Quataert et al. 1999; Yuan, Markoff \&
1850: Falcke 2002; Yuan et al. 2002; Fabbiano et al. 2003; Pellegrini et al. 2003b;
1851: Ptak et al. 2004; Nemmen et al. 2006; Wu, Yuan \& Cao 2007). They bear
1852: little resemblance to SEDs characteristic of normal stellar systems.
1853: Inactive galaxies or starburst systems not strongly affected by dust
1854: extinction emit the bulk of their radiation in the optical--UV and in the
1855: thermal IR regions, with only an energetically miniscule contribution from
1856: X-rays.
1857:
1858: \clearpage
1859: \begin{figure}
1860: %%BoundingBox: 20 10 600 760
1861: \psfig{figure=fig8.ps,width=13.5cm,angle=270}
1862: \caption{The H\al\ nuclear luminosity function of nearby AGNs derived from the
1863: Palomar survey. The top axis gives an approximate conversion to absolute
1864: magnitudes in the $B$ band, using the H\al-continuum correlation of Greene \&
1865: Ho (2005b). The unfilled circles include only type~1 sources, while the
1866: filled circles represent both type~2 and type~1 sources. The luminosities
1867: have been corrected for extinction, and in the case of type~1 nuclei, they
1868: include both the narrow and broad components of the line. For
1869: comparison, I show the $z < 0.35$ luminosity function for SDSS Seyfert
1870: galaxies (types 1 and 2; dashed line; Hao et al. 2005b). (Adapted from
1871: L.C. Ho, A.V. Filippenko \& W.L.W. Sargent, in preparation.)
1872: }
1873: \end{figure}
1874:
1875: \subsection{Luminosity Function}
1876:
1877: Many astrophysical applications of AGN demographics benefit from knowing
1878: the AGN luminosity function, $\Phi(L,z)$. Whereas $\Phi(L,z)$ has been
1879: reasonably well charted at high $L$ and high $z$ using quasars, it is very
1880: poorly known at low $L$ and low $z$. Indeed, until very recently there has
1881: been no reliable determination of $\Phi(L,0)$. The difficulty in determining
1882: $\Phi(L,0)$ can be ascribed to a number of factors, as discussed in Huchra \&
1883: Burg (1992). First and foremost is the challenge of securing a reliable,
1884: spectroscopically selected sample. Since nearby AGNs are expected to be faint
1885: relative to their host galaxies, most of the traditional techniques used to
1886: identify quasars cannot be applied without introducing large biases. The
1887: faintness of nearby AGNs presents another obstacle, namely how to disentangle
1888: the nuclear emission---the only component relevant to the AGN---from the
1889: usually much brighter contribution from the host galaxy. Finally, most
1890: optical luminosity functions of bright, more distant AGNs are specified in
1891: terms of the nonstellar optical continuum (usually the $B$ band), whereas
1892: spectroscopic surveys of nearby galaxies generally only reliably measure
1893: optical line emission (e.g., H\al) because the featureless nuclear continuum
1894: is often impossible to detect in ground-based, seeing-limited apertures.
1895:
1896: A different strategy can be explored by taking advantage of the fact that
1897: H\al\ luminosities are now available for nearly all of the AGNs in the Palomar
1898: survey. Figure~8 shows the H\al\ luminosity function for the Palomar sources,
1899: computed using the $V/V_{\rm max}$ method (L.C. Ho, A.V. Filippenko \& W.L.W.
1900: Sargent, in preparation). Two versions are shown, each representing an
1901: extreme view of what kind of sources should be regarded as {\it bona fide}\
1902: AGNs. The open symbols include only type~1 nuclei, whose AGN status is
1903: incontrovertible. This may be regarded as the most conservative assumption
1904: and a lower bound, since we know that genuine narrow-line AGNs do exist. The
1905: filled symbols lump together all sources classified as LINERs, transition
1906: objects, and Seyferts, both type~1 and type~2. This represents the most
1907: optimistic view and an upper bound, if some type~2 sources are in fact AGN
1908: impostors, although, as I argue in \S~6.5, this is likely to
1909: be a small effect. The true space density of local AGNs
1910: lies between these two possibilities. In either case, the differential
1911: luminosity function can be approximated by a single power law from
1912: $L_{{\rm H}\alpha} \approx 10^{38}$ to $3\times 10^{41}$ \lum, roughly of the
1913: form $\Phi \propto L^{-1.2\pm0.2}$. The slope seems to flatten below
1914: $L_{{\rm H}\alpha} \approx 10^{38}$ \lum, but the luminosity function is
1915: highly uncertain at the faint end because of density fluctuations in our local
1916: volume. Nevertheless, it is remarkable that the Palomar luminosity function
1917: formally begins at $L_{{\rm H}\alpha} \approx 6 \times 10^{36}$ \lum,
1918: roughly the luminosity of the Orion nebula (Kennicutt 1984). In units more
1919: familiar to the AGN community, this corresponds to an absolute $B$-band
1920: magnitude of roughly $-8$ (using the H\al-optical continuum conversion of
1921: Greene \& Ho 2005b), no brighter than a single supergiant star.
1922:
1923: For comparison, I have overlaid the H\al\ luminosity function of $z$ \lax\
1924: 0.35 Seyfert galaxies derived from the SDSS by Hao et al. (2005b). The
1925: Palomar survey reaches $\sim 2$ orders of magnitude fainter in H\al\
1926: luminosity than SDSS, but the latter extends a factor of 10 higher at the
1927: bright end. Over the region of overlap, the two surveys show reasonably good
1928: agreement, especially considering the small number statistics of the Palomar
1929: survey and the fact that Hao et al.'s sample only includes Seyferts.
1930:
1931: \subsection{Bolometric Luminosities and Eddington Ratios}
1932:
1933: \begin{figure}
1934: %%BoundingBox: 20 150 550 670
1935: \vbox{\hbox{
1936: \psfig{figure=fig9a.ps,width=6.7cm,angle=0}
1937: \hskip -0.2 cm
1938: \psfig{figure=fig9b.ps,width=6.7cm,angle=0}
1939: }}
1940: \caption{Distribution of ({\it a}) bolometric luminosity, $L_{\rm bol}$, and
1941: ({\it b}) ratio of bolometric luminosity to the Eddington luminosity,
1942: $L_{\rm bol}/L_{\rm Edd}$, for all objects, Seyferts (S), LINERs (L),
1943: transition objects (T), and absorption-line nuclei (A). $L_{\rm bol}$ is
1944: based on the X-ray ($2-10$ keV) luminosity. The hatched and open histograms
1945: denote detections and upper limits, respectively; type~1 objects are plotted
1946: in blue, type~2 objects in red. (Adapted from L.C. Ho, in preparation.)
1947: }
1948: \end{figure}
1949:
1950:
1951: To gain further insight into the physical nature of LLAGNs, it is more
1952: instructive to examine their bolometric luminosities rather than their
1953: luminosities in a specific band or emission line. Because AGNs emit a very
1954: broad spectrum, their bolometric luminosities ideally should be measured
1955: directly from their full SEDs. In practice, however, complete SEDs are not
1956: readily available for most AGNs, and one commonly estimates $L_{\rm bol}$
1957: by applying bolometric corrections derived from a set of well-observed
1958: calibrators. As discussed in \S~5.8, the SEDs of LLAGNs differ quite markedly
1959: from those of conventionally studied AGNs. Nonetheless, they do exhibit a
1960: characteristic shape, which enables bolometric corrections to be calculated.
1961: The usual practice of choosing the optical $B$ band as the reference point
1962: should be abandoned for LLAGNs, not only because reliable optical continuum
1963: measurements are scarce but also because the optical/UV region of the SED
1964: shows the maximal variance with respect to accretion rate (\S~5.8) and depends
1965: sensitively on extinction. What is available, by selection, is nuclear
1966: emission-line fluxes, and upper limits thereon. Although the H\al\ luminosity
1967: comprises only a small percentage of the total power, its fractional
1968: contribution to $L_{\rm bol}$ turns out to be fairly well defined: from the
1969: SED study of L.C. Ho (in preparation), $L_{\rm bol} \approx 220
1970: L_{{\rm H}\alpha}$, with an rms scatter of $\sim 0.4$ dex, consistent with the
1971: calibration given in Greene \& Ho (2005b, 2007a). Because of the wide range
1972: of ionization levels among LLAGNs, a bolometric correction based on H\al\
1973: should be more stable than one tied to \oiii\ \lamb 5007 (e.g., Heckman et al.
1974: 2005). Nevertheless, in light of the nonnuclear component of the nebular
1975: flux in LLAGNs (\S~6.4), the luminosity of the narrow H\al\ line will tend to
1976: overestimate $L_{\rm bol}$. I recommend that, whenever possible, $L_{\rm bol}$
1977: should be based on the hard X-ray ($2-10$ keV) luminosity, bearing in mind the
1978: added complication that the bolometric correction in this band is
1979: luminosity-dependent. Making use again of the database from L.C. Ho (in
1980: preparation), I estimate $L_{\rm bol}/L_{\rm X} \approx 83$, 28, and 16 for
1981: quasars, luminous Seyferts, and LLAGNs, respectively.
1982:
1983: Figure~9 shows the distributions of $L_{\rm bol}$ and their values normalized
1984: with respect to the Eddington luminosity for Palomar galaxies with
1985: measurements of $L_{\rm X}$ and central stellar velocity dispersion. The
1986: $M_{\rm BH} - \sigma$ relation of Tremaine et al. (2002) was used to obtain
1987: \ledd. Although there is substantial overlap, the four spectral classes
1988: clearly delineate a luminosity sequence, with $L_{\rm bol}$ decreasing
1989: systematically as S$\rightarrow$L$\rightarrow$T$\rightarrow$A. The
1990: differences become even more pronounced in terms of $L_{\rm bol}/L_{\rm Edd}$,
1991: with Seyferts having a median value (1.3\e{-4}) 20 times higher than in LINERs
1992: (5.9\e{-6}), which in turn are higher than transition objects by a factor of
1993: $\sim 5$. Among Seyferts and LINERs, type~1 sources are systematically more
1994: luminous than type~2s. Notably, the vast majority of nearby nuclei have
1995: highly sub-Eddington luminosities. The total distribution of Eddington ratios
1996: is characterized by a prominent peak at $L_{\rm bol}/L_{\rm Edd} \approx
1997: 10^{-5}$ dominated by Seyfert~2s, LINERs, and transition objects, and a
1998: precipitous drop toward larger Eddington ratios. Contrary to previous claims
1999: (Wu \& Cao 2005; Hopkins \& Hernquist 2006) based on the smaller sample of Ho
2000: (2002a), the distribution of Eddington ratios shows no bimodality. The
2001: systematic difference in Eddington ratios between LINERs and Seyferts has been
2002: noticed before in the Palomar survey (Ho 2002b, 2003, 2005) and in SDSS
2003: (Kewley et al. 2006), but this is the first time that the more subtle
2004: differences among the different subclasses can be discerned.
2005:
2006: \section{EXCITATION MECHANISMS}
2007:
2008: \subsection{Nonstellar Photoionization}
2009:
2010: The origin and excitation of the ionized gas in the central regions of nearby
2011: galaxies has been a longstanding problem (Minkowski \& Osterbrock 1959;
2012: Osterbrock 1960). Ever since the early suggestion of Ferland \& Netzer (1983)
2013: and Halpern \& Steiner (1983), photoionization by a central AGN has surfaced
2014: as the leading candidate for the excitation mechanism of LINERs. Given the
2015: success with which more luminous sources have been explained within this
2016: framework, and the growing realization that BHs are commonplace, extending
2017: photoionization models to LINERs is both natural and appealing. The requisite
2018: relative strengths of the low-ionization lines in LINERs can be achieved by
2019: lowering the ionization parameter, commonly defined as $U = Q_{\rm ion}/4\pi
2020: r^2 c n$, where $Q_{\rm ion}$ is the number of photons s$^{-1}$ capable of
2021: ionizing hydrogen, $r$ is the distance of the inner face of the cloud from the
2022: central continuum, $n$ the hydrogen density, and $c$ the speed of light. While
2023: Seyfert line ratios can be matched with $\log U \approx -2\pm0.5$ (e.g.,
2024: Ferland \& Netzer 1983; Stasi\'nska 1984; Ho, Shields \& Filippenko 1993),
2025: LINERs require $\log U \approx -3.5\pm0.5$ (Ferland \& Netzer 1983; Halpern \&
2026: Steiner 1983; P\'equignot 1984; Binette 1985; Ho, Filippenko \& Sargent 1993;
2027: Groves, Dopita \& Sutherland 2004).
2028:
2029: An issue that has not been properly addressed is which of the primary variables
2030: --- $Q_{\rm ion}$, $r$, or $n$ --- conspire to reduce $U$ in LINERs to the
2031: degree required by the models. The answer seems to be all three. The dominant
2032: factor comes from the luminosity, as LINERs emit an order of magnitude less
2033: ionizing luminosity than do Seyferts: $\langle L_{{\rm H}\alpha} \rangle = 3
2034: \times 10^{39}$ \lum\ versus $29 \times 10^{39}$ \lum\ (Ho, Filippenko \&
2035: Sargent 2003). Given the gas-poor environments of LINERs (see \S~5.6), we can
2036: now say with some certainty that the reduction in ionizing luminosity is {\it
2037: intrinsic}\ and not due to obscuration (as proposed by Halpern \& Steiner
2038: 1983). But this is unlikely to be the end of the story. The electron density
2039: of LINERs ($\langle n_e \rangle \approx 280$ \cc), at least as probed by the
2040: relatively low-density tracer \sii\ \lamb\lamb6716, 6731, is $\sim 50$\% lower
2041: than in Seyferts ($\langle n_e \rangle \approx 470$ \cc). Very little is
2042: known about the detailed morphology and spatial distribution of the NLR in
2043: LINERs, or for that matter in low-luminosity Seyferts. Whereas the NLR in
2044: luminous Seyferts span $\sim 50-1000$ pc in radius, scaling roughly as
2045: $L^{0.5}$ (Bennert et al. 2006), LINERs seem to be significantly more compact.
2046: This is perhaps not surprising, if the NLR size-luminosity relation extends to
2047: LLAGNs. At typical ground-based resolution, narrow-band imaging studies find
2048: that the ionized gas in LINERs tends to be quite centrally peaked, with typical
2049: dimensions of $r$ \lax\ $50-100$ pc (Keel 1983a; Pogge 1989). In the
2050: instances where narrow-band images or slit spatial profiles are available from
2051: \hst\ (Bower et al. 2000; Pogge et al. 2000; Cappellari et al. 2001;
2052: Verdoes~Kleijn et al. 2002; Gonz\'alez Delgado et al. 2004; Walsh et al. 2008),
2053: the line-emitting gas appears even more concentrated still,
2054: with scales \lax\ tens pc, although some of it clearly extends to scales of at
2055: least $\sim 200$ pc (Shields et al. 2007). The covering factor is high, on
2056: average $\sim 0.3$ for the LINER nuclei of the radio galaxies studied by
2057: Capetti, Verdoes~Kleijn \& Chiaberge (2005). With very few exceptions (e.g.,
2058: NGC~1052; Pogge et al. 2000), extended, elongated structures analogous to
2059: classical ionization cones in Seyferts do not exist in LINERs. Interestingly,
2060: both of these trends (smaller $n_e$ and $r$) would naively drive $U$ in the
2061: opposite direction needed to explain LINERs. On the other hand, we know that
2062: the NLRs of AGNs in general, and perhaps of LINERs especially, contain a wide
2063: range of densities not probed by \sii\ and that this material is highly
2064: stratified radially (Wilson 1979; Pelat, Alloin \& Fosbury 1981; Carswell et
2065: al. 1984; Filippenko \& Halpern 1984; P\'equignot 1984; Binette 1985;
2066: Filippenko 1985; Ho, Filippenko \& Sargent 1993, 1996; Barth et al. 2001b;
2067: Laor 2003; Shields et al. 2007; Walsh et al. 2008). The
2068: effective ionization parameter, therefore, depends on the detailed spatial
2069: distribution of the gas.
2070:
2071: Whereas basic single-zone photoionization models can match many of the strong
2072: lines, it is well-known that more complex, multi-component models, especially
2073: ones that incorporate a range of densities, are required to achieve a
2074: satisfactory fit (Binette 1985; Gabel et al. 2000; Sabra et al. 2003). A
2075: notorious deficiency of single-zone models has been their inability to
2076: reproduce the high values of the temperature-sensitive ratio
2077: \oiii\ \lamb4363/\oiii\ \lamb5007. However, because of the different critical
2078: densities of these two transitions, they need not originate cospatially, making
2079: \oiii\ \lamb4363/\oiii\ \lamb5007 no longer a valid thermometer. The
2080: discovery that these two lines indeed have different line widths removed one
2081: of the principal objections to photoionization models (Filippenko \& Halpern
2082: 1984; Filippenko 1985).
2083:
2084: The observed weakness of \heii\ \lamb4686/H\bet\ (see also \S~5.8) has also
2085: been a thorny problem. P\'equignot (1984) achieved a consistent fit to the
2086: spectrum of NGC~1052 by invoking a modified ionizing spectrum consisting of an
2087: 80,000 K blackbody coupled with an X-ray tail extending to higher energies.
2088: The main difficulty with this proposal is that the observed SEDs of LINERs do
2089: not have a blackbody component peaked in the UV, nor is P\'equignot's
2090: specific model unique because Gabel et al. (2000) achieved an equally good---if
2091: not better---fit using a simple power-law continuum with $\alpha = -1.2$.
2092: Since we now have ample evidence that the SEDs of LINERs are {\it not}\ the
2093: same as those of more luminous AGNs, future photoionization calculations
2094: should adopt empirically motivated input spectra. Important steps in these
2095: directions have been taken (e.g., Nicholson et al. 1998; Gabel et al. 2000;
2096: Nagao et al. 2002; Lewis, Eracleous \& Sambruna 2003), but much more can be
2097: done. One fruitful avenue to pursue is to incorporate the {\it full}\ observed
2098: SED, which is noteworthy not only because of its hard ionizing spectrum but,
2099: due to its radio-loudness, also because it presents a copious supply of
2100: relativistic, synchrotron-emitting particles, which can dramatically alter the
2101: excitation of the NLR (Aldrovandi \& P\'equignot 1973; Ferland \& Mushotzky
2102: 1984; Gruenwald \& Viegas-Aldrovandi 1987). ``Cosmic ray heating'' boosts the
2103: strengths of low-ionization lines such as \nii\ \lamb\lamb6548, 6583 and
2104: \sii\ \lamb\lamb6716, 6731 (Viegas-Aldrovandi \& Gruenwald 1990), which are
2105: normally underpredicted (e.g., Ho, Filippenko \& Sargent 1993; Lewis,
2106: Eracleous \& Sambruna 2003), and thereby help to constrain alternative
2107: solutions that invoke selective abundance enhancement of N and S
2108: (Storchi-Bergmann \& Pastoriza 1990) or dust grain depletion (Gabel et al.
2109: 2000).
2110:
2111: \subsection{Contribution from Fast Shocks}
2112:
2113: Despite the natural appeal of AGN photoionization, alternative excitation
2114: mechanisms for LINERs have been advanced. Collisional ionization by shocks
2115: has been a popular contender from the outset (Burbidge, Gould \& Pottasch 1963;
2116: Osterbrock 1971; Osterbrock \& Dufour 1973; Koski \& Osterbrock 1976;
2117: Danziger, Fosbury \& Penston 1977; Fosbury et al. 1978; Ford \& Butcher 1979;
2118: Heckman 1980b; Baldwin, Phillips \& Terlevich 1981). Shocks continue to be
2119: invoked (Bonatto, Bica \& Alloin 1989; Dopita \& Sutherland 1995;
2120: Alonso-Herrero et al. 2000; Sugai \& Malkan 2000) even after concerns over the
2121: \oiii\ \lamb 4363 temperature problem had been dispelled as a result of either
2122: revised measurements (Keel \& Miller 1983; Rose \& Tripicco 1984) or
2123: complications arising from density stratification (Filippenko \& Halpern 1984).
2124: Dopita \& Sutherland (1995) showed that the diffuse radiation field
2125: generated by fast ($\upsilon\, \approx 150-500$ \kms) shocks can reproduce the
2126: optical narrow emission lines seen in both LINERs and Seyferts. In their
2127: models, LINER-like spectra are realized under conditions in which the
2128: precursor \hii\ region of the shock is absent, as might be the case in
2129: gas-poor environments. The postshock cooling zone attains a much higher
2130: equilibrium electron temperature than a photoionized plasma; consequently, a
2131: robust prediction of shock models is that shocked gas should produce a
2132: higher excitation spectrum, most readily discernible in the UV, than
2133: photoionized gas. In all the cases studied so far, however, the UV spectra
2134: are inconsistent with the fast-shock scenario because the observed intensities
2135: of high-excitation lines such as \civ\ \lamb1549 and \heii\ \lamb1640 are much
2136: weaker than predicted (Barth \etal 1996, 1997; Maoz \etal 1998; Nicholson
2137: \etal 1998; Gabel et al. 2000). Dopita \etal (1997) used the spectrum of the
2138: circumnuclear {\it disk} of M87 to advance the view that LINERs are
2139: shock-excited. This argument is misleading because their analysis deliberately
2140: avoids the nucleus. Sabra et al. (2003) demonstrate that the UV--optical
2141: spectrum of the {\it nucleus}\ of M87 is best explained by a multi-component
2142: photoionization model.
2143:
2144: Analysis of the emission-line profiles of the Palomar nuclei further casts
2145: doubt on the viability of the fast-shock scenario (Ho, Filippenko \& Sargent
2146: 2003). The velocity dispersions of the nuclear gas generally fall short of
2147: the values required for fast-shock excitation to be important. Furthermore,
2148: the close similarity between the velocity field of LINERs and Seyferts as
2149: deduced from their line profiles contradicts the basic premise that shocks
2150: are primarily responsible for the spectral differences between the two classes
2151: of objects. For a given bulge potential, LINERs, if anything, have {\it
2152: smaller}, not larger, gas velocity dispersions than Seyferts (L.C. Ho, in
2153: preparation). And as discussed in \S~5.2, the incidence of extended radio
2154: jets, the most likely source of kinetic energy injection into the NLR, is
2155: actually lower in LINERs than in Seyferts, again contrary to naive expectations.
2156:
2157: Notwithstanding these complications, it is inconceivable that mechanical
2158: heating, especially by lower velocity ($\sim 50-100$ \kms) shocks, does not
2159: play some role in the overall excitation budget of LINERs. The velocities
2160: of the line-emitting gas are, after all, highly supersonic, turbulent, and
2161: most likely pressure-dominated (L.C. Ho, in preparation). The trick is to
2162: figure out what is the balance between shocks and photoionization, and what
2163: physical insights can be gained from knowing the answer. It would be
2164: worthwhile to revisit composite shock plus photoionization models, such as
2165: those developed by Viegas-Aldrovandi \& Gruenwald (1990) and Contini (1997)
2166: with the latter component maximally constrained by observations so that
2167: robust, quantitative estimates can be placed on the former. Such an approach
2168: might yield meaningful measurements of the amount of mechanical energy
2169: deposited into the host galaxy by AGN feedback.
2170:
2171: \subsection{Contribution from Stellar Photoionization}
2172:
2173: Another widely discussed class of models invokes hot stars formed in a
2174: short-duration burst of star formation to supply the primary ionizing
2175: photons. Ordinary O-type stars with effective temperatures typical of those
2176: found in giant \hii\ regions in galactic disks do not produce sufficiently
2177: strong low-ionization lines to account for the spectra of LINERs. The
2178: physical conditions in the centers of galaxies, on the other hand, may be more
2179: favorable for generating LINER-like spectra. For example, Terlevich \&
2180: Melnick (1985) postulate that the high-metallicity environment of galactic
2181: nuclei may be particularly conducive to forming very hot, $T \approx (1-2)
2182: \times 10^5$ K, luminous Wolf-Rayet stars whose ionizing spectrum would
2183: effectively mimic the power-law continuum of an AGN. The models of Filippenko
2184: \& Terlevich (1992) and Shields (1992) appeal to less extreme conditions.
2185: These authors show that photoionization by ordinary O stars, albeit of
2186: somewhat higher effective temperature than normal (but see Schulz \& Fritsch
2187: 1994), embedded in an environment with high density and low ionization
2188: parameter can explain the spectral properties of transition objects. Barth \&
2189: Shields (2000) extended this work by modeling the ionizing source not as single
2190: O-type stars but as a more realistic evolving young star cluster. They
2191: confirm that young, massive stars can indeed generate optical emission-line
2192: spectra that match those of transition objects, and, under some plausible
2193: conditions, even those of {\it bona fide} LINERs. But there is an important
2194: caveat: the star cluster must be formed in an instantaneous burst, and its age
2195: must coincide with the brief phase ($\sim$3$-$5 Myr after the burst) during
2196: which sufficient Wolf-Rayet stars are present to supply the extreme-UV photons
2197: necessary to boost the low-ionization lines. The necessity of a sizable
2198: population of Wolf-Rayet stars is also emphasized in the study by Gabel
2199: \& Bruhweiler (2002). As discussed in Ho, Filippenko \& Sargent (2003), the
2200: main difficulty with this scenario, and indeed with all models that appeal to
2201: young or intermediate-age stars (e.g., Engelbracht et al. 1998;
2202: Alonso-Herrero et al. 2000; Taniguchi, Shioya \& Murayama 2000), is that the
2203: nuclear stellar population of the host galaxies of the majority of nearby
2204: AGNs, irrespective of spectral class, is demonstrably {\it old}\ (\S~4.2).
2205: Stellar absorption indices indicative of young or intermediate-age stars are
2206: seldom seen, and the telltale emission features of Wolf-Rayet stars are
2207: notably absent, both in ground-based and {\it HST}\ spectra. Sarzi et al.
2208: (2005) find that young stars can account for at most a few percent of the blue
2209: light within the central few parsecs of nearby LLAGNs, in most cases incapable
2210: of providing enough ionizing photons to account for the observed H\al\
2211: emission. Post-starburst scenarios face another serious dilemma: if most
2212: LLAGNs, which constitute the majority of nearby AGNs and a large percentage of
2213: all galaxies, are described by this scenario, then where are their precursors?
2214: They do not exist. These empirical facts seriously undermine the viability of
2215: starburst or post-starburst models for LLAGNs.
2216:
2217: Evolved, low-mass stars, on the other hand, probably contribute at some level
2218: to the ionization. This idea was advocated by Binette et al. (1994), who
2219: proposed that post-asymptotic giant branch (post-AGB) stars, which can attain
2220: effective temperatures as high as $\sim 10^5$ K, might be responsible for
2221: photoionizing the extended ionized gas often observed in elliptical galaxies.
2222: The emission-line spectrum of these nebulae, in fact, tend to be of relatively
2223: low ionization (Demoulin-Ulrich, Butcher \& Boksenberg 1984; Phillips et al.
2224: 1986). Invoking evolved stars has the obvious appeal of not violating the
2225: stellar population constraints discussed above. This mechanism, however,
2226: cannot be the dominant contributor to compact LINERs. The line emission tends
2227: to be very centrally concentrated (\S~6.1), much more so than the underlying
2228: stellar density profile. Moreover, the line strengths in most LLAGNs are
2229: simply too high. The calculations of Binette et al. predict H\al\ equivalent
2230: widths of EW $\approx$ 1 \AA, whereas the LINERs and transition objects in the
2231: Palomar survey have an average EW 3$-$4 times higher (Ho, Filippenko \&
2232: Sargent 2003), with over 70\% of the sample having EW $>$ 1 \AA.
2233:
2234: To obtain a quantitative estimate of the contribution of post-AGB stars to the
2235: ionization budget of the weaker emission-line nuclei, I convert the nuclear
2236: stellar magnitudes ($m_{44}$) given in Ho, Filippenko \& Sargent (1997a) to
2237: stellar masses assuming a mass-to-light ratio of $M/L_B = 8(M/L_B)_\odot$ and
2238: that post-AGB stars have a specific ionization rate of $Q_{\rm ion} = 7.3
2239: \times 10^{40} (M/M_\odot)\, {\rm s}^{-1}$ (Binette et al. 1994). Within the
2240: $100 \times 200$ pc aperture of the Palomar spectra, the integrated stellar
2241: mass is $\sim 10^7$ to $10^{10}$ \solmass, with a median
2242: value of $2\times10^9$ \solmass, which corresponds to an ionizing photon rate
2243: of $Q_{\rm ion} = 1.5 \times 10^{50}\,{\rm s}^{-1}$. These estimates depend
2244: on the choice of the stellar initial mass function, and, most importantly, on
2245: detailed processes during the late stages of stellar evolution that are still
2246: not fully understood (see, e.g., O'Connell 1999). Nevertheless, taking the
2247: fiducial estimates as a rough guide, the predicted values of $Q_{\rm ion}$ can
2248: be compared to the actually observed, extinction-corrected H\al\ luminosity.
2249: Assuming complete reprocessing of the ionizing continuum and that on average it
2250: takes 2.2 Lyman continuum photons to produce one H\al\ photon, I estimate that
2251: the nebular line emission in roughly one-third of the Palomar sources can be
2252: powered by photoionization from post-AGB stars. The fraction is higher for
2253: LINERs (39\%) than Seyferts (16\%), being most prevalent for LINER~2s (44\%)
2254: and transition objects (33\%). Eracleous, Hwang \& Flohic (2008b) performed a
2255: similar analysis for a handful of LINERs with central stellar luminosity
2256: profiles available from \hst, concluding also that post-AGB stars can
2257: alleviate the ionizing photon deficit in some objects.
2258:
2259: \begin{figure}
2260: %%BoundingBox: 20 150 550 670
2261: \vbox{\hbox{
2262: \psfig{figure=fig10a.ps,width=6.7cm,angle=0}
2263: \hskip -0.2 cm
2264: \psfig{figure=fig10b.ps,width=6.7cm,angle=0}
2265: }}
2266: \caption{({\it a}) Correlation between (total) H\al\ luminosity and X-ray
2267: ($2-10$ keV) luminosity. The dotted lines mark $L_{\rm X}/L_{{\rm H}\alpha}$
2268: = 1, 5, and 25. ({\it b}) Distribution of radio-loudless parameter versus
2269: Eddington ratio, with the bolometric luminosity estimated assuming
2270: $L_{\rm bol} = 16 L_{\rm X}$.
2271: }
2272: \end{figure}
2273:
2274: \subsection{Energy Budget}
2275:
2276: In luminous AGNs, important confirmation of the basic photoionization
2277: paradigm comes from the strong empirical scaling and correlated variability
2278: between line flux and the strength of the ionizing continuum. Although very
2279: little information exists in terms of line-continuum variability for LLAGNs,
2280: enough X-ray observations have now been amassed to examine the correlation
2281: between optical line luminosity and X-ray luminosity. The X-ray band
2282: is only indirectly coupled to the bulk of the Lyman continuum, but
2283: in LLAGNs, it offers the most reliable probe of the high-energy spectrum.
2284: Studies in the soft X-ray band suggest that LLAGNs roughly follow the
2285: extrapolation of the $L_{{\rm H}\alpha}-L_{\rm X}$ correlation established for
2286: higher luminosity sources (Koratkar et al. 1995; Roberts \& Warwick 2000;
2287: Halderson et al. 2001). Intriguingly, no clear differences could be discerned
2288: between LINERs and Seyferts, confirming preliminary evidence presented in
2289: Halpern \& Steiner (1983). A more complex picture, however, emerges at higher
2290: energies ($2-10$ keV). Terashima, Ho \& Ptak (2000) and Terashima et al.
2291: (2000) note that LINER~2s, unlike LINER~1s, suffer from a deficit of ionizing
2292: photons: the X-ray emission of the nucleus, when extrapolated to the UV,
2293: underpredicts the observed H\al\ luminosity by a factor of $\sim 10-100$. This
2294: trend persists in more recent {\it Chandra}\ observations (Terashima \& Wilson
2295: 2003b; Flohic et al. 2006; Eracleous, Hwang \& Flohic 2008b), one that seems
2296: to be especially endemic to transition objects (Ho et al. 2001; Filho et al.
2297: 2004).
2298:
2299: Figure~10{\it a}\ shows an update of the $L_{{\rm H}\alpha}-L_{\rm X}$ relation
2300: for all Palomar sources with high-resolution (\lax\ 5\asec) X-ray measurements.
2301: For comparison, I also include $z < 0.5$ Palomar-Green (PG; Schmidt \& Green
2302: 1983) quasars and luminous Seyfert~1s with
2303: well-determined SEDs (Figure~7; L.C. Ho, in preparation). All broad-line
2304: sources follow an approximately linear relation over nearly 7 orders of
2305: magnitude in luminosity. In detail, $L_{{\rm H}\alpha}\propto L_{\rm X}^{1.1}$,
2306: such that quasars and luminous Seyferts have a median ratio
2307: $L_{\rm X}/L_{{\rm H}\alpha} \approx 2$, to be compared with
2308: $L_{\rm X}/L_{{\rm H}\alpha} \approx 7$ for Palomar Seyfert~1s. Ignoring the
2309: possible effect of a luminosity-dependent covering factor, this can be
2310: interpreted as the consequence of a decrease in the ratio of UV radiation to
2311: X-rays with decreasing luminosity, reflecting a pattern familiar from samples
2312: of bright AGNs (e.g., Strateva et al. 2005), now extended down to lower
2313: luminosities by $\sim 3$ orders of magnitude. Since LINER~1s are weaker than
2314: Seyfert~1s and their SEDs lack a UV bump, it is surprising that they actually
2315: have a somewhat {\it lower}\ $L_{\rm X}/L_{{\rm H}\alpha}$ ratio ($\sim 5$)
2316: than Seyfert~1s. This suggests that at least some of the H\al\ emission in
2317: LINER~1s, presumably in the narrow component, is not powered by AGN
2318: photoionization. In fact, this turns out to be a property symptomatic of
2319: {\it all}\ type~2 LLAGNs (Table~1). The most extreme manifestation can be
2320: seen among transition objects, with $L_{\rm X}/L_{{\rm H}\alpha} \approx 0.4$,
2321: but both LINER~2s and Seyfert~2s also exhibit an ionizing photon deficit.
2322: For conditions typical of LLAGNs, Eracleous, Hwang \& Flohic (2008b) use
2323: photoionization calculations to infer $L_{\rm ion} = 18 L_{{\rm H}\alpha}\,
2324: f_c^{-1}$, where $L_{\rm ion}$ is the ionizing luminosity between 1 Ryd and 10
2325: keV and $f_c$ is the covering factor of the line-emitting gas. Since
2326: $L_{\rm ion} \approx 3 L_{\rm X}$ for a power-law spectrum with $\alpha =
2327: -0.1$ to $-0.9$, $L_{\rm X}/L_{{\rm H}\alpha} \approx 6 f_c^{-1}$. It is
2328: clear that most narrow-line LLAGNs violate this energy balance condition, even
2329: for the optimistic assumption of $f_c = 1$.
2330:
2331: There are several possible solutions to the ionization deficit problem. (1) The
2332: X-rays could be highly obscured, perhaps even Compton-thick. In light of the
2333: evidence given in \S\S~5.3, 5.6, I consider this solution to be untenable for
2334: LINER~2s; highly absorbed sources do exist (e.g., NGC~4261; Zezas et al.
2335: 2005), but they are in the minority. Moreover, many of the X-ray measurements
2336: have already been corrected for absorption. The situation is more complex for
2337: Seyfert~2s. Some of the sources with low $L_{\rm X}/L_{{\rm H}\alpha}$ indeed
2338: show direct evidence for Compton thickness from their X-ray spectra (Cappi et
2339: al. 2006). Others, however, are too faint for spectral analysis, and for
2340: these, their status as Compton-thick sources was based on the observed ratio of
2341: 2$-$10 keV flux to \oiii\ \lamb5007 flux (Bassani et al. 1999; Panessa \&
2342: Bassani 2002). Applying an average correction factor of 60 to the X-ray
2343: luminosity would bring the Seyfert~2s into agreement with the Seyfert~1s on the
2344: $L_{{\rm H}\alpha}-L_{\rm X}$ relation (Panessa et al. 2006). But this
2345: procedure {\it assumes}\ that the low values of $L_{\rm X}/L_{{\rm H}\alpha}$
2346: are due to a reduction of the X-rays by absorption rather than an enhancement
2347: of H\al\ (see below). (2) The SED could be drastically different,
2348: specifically in having a much more prominent UV component. This proposition
2349: can be promptly dismissed because the SEDs of LLAGNs generically lack a UV
2350: bump. There is certainly no indication that type~2 sources are preferentially
2351: brighter in the UV; in the case of LINERs, type~2 sources, if anything, tend
2352: to be redder than type~1 sources (Maoz et al. 2005). (3) Lastly, and most
2353: plausibly, a significant fraction of the ionization for the narrow-line gas
2354: comes from nonnuclear sources. As discussed in connection with the preceding
2355: two subsections, young, massive stars and fast shocks are generally not viable
2356: options. There are a number of candidate sources of ``extra'' ionization,
2357: including hot, evolved stars, turbulent mixing layers, diffuse X-ray emitting
2358: plasma, low-mass XRBs, cosmic ray heating, and mechanical heating
2359: from radio jets. As all of these sources probably contribute at some level,
2360: efforts to single out any dominant mechanism may be hopelessly challenging.
2361: Nevertheless, as discussed in \S~6.3, post-AGB stars appear especially
2362: promising. Taking the calculations of Binette et al. (1994) as a guide, the
2363: stellar mass within the central $100-200$ pc region generates sufficient Lyman
2364: continuum photons to account for the H\al\ emission in $\sim 30\%-40\%$ of the
2365: LINER~2s and transition objects. This estimate is crude and admittedly
2366: optimistic, as it assumes a covering factor of unity for the NLR, but it
2367: serves as a useful illustration of the types of effects that should be
2368: included in any complete treatment of the energy budget problem in LLAGNs.
2369:
2370: \subsection{The Nature of Transition Objects and a Unified View of LLAGNs}
2371:
2372: The physical origin of transition nuclei continues to be a thorny, unresolved
2373: problem. In standard line-ratio diagrams (Figure~3), transition nuclei are
2374: empirically defined to be those sources that lie sandwiched between the loci
2375: of ``normal'' \hii\ regions and LINERs. This motivated Ho, Filippenko \&
2376: Sargent (1993) to propose that transition objects may be composite systems
2377: consisting of a LINER nucleus plus an \hii\ region component. The latter
2378: could arise from neighboring circumnuclear \hii\ regions or from \hii\ regions
2379: randomly projected along the line of sight. A similar argument, based on
2380: decomposition of line profiles, was made by V\'eron, Gon\c{c}alves \&
2381: V\'eron-Cetty (1997). If transition objects truly are LINERs sprinkled with a
2382: frosting of star formation, one would expect that their host galaxies should
2383: be similar to those of LINERs, modulo minor differences due to extra
2384: contaminating star formation. The study of Ho, Filippenko \& Sargent (2003)
2385: tentatively supports this picture. The host galaxies of transition nuclei
2386: exhibit systematically higher levels of recent star formation, as indicated by
2387: their far-IR emission and broad-band optical colors, compared to LINERs of
2388: matched morphological types. Moreover, the hosts of transition nuclei tend to
2389: be slightly more inclined than LINERs. Thus, all else being equal,
2390: transition-type spectra seem to be found precisely in those galaxies whose
2391: nuclei have a higher probability of being contaminated by extra-nuclear
2392: emission from star-forming regions.
2393:
2394: This story, however, has some holes. If spatial blending of circumnuclear
2395: \hii\ regions is sufficient to transform a regular LINER into a transition
2396: object, the LINER nucleus should reveal itself unambiguously in spectra taken
2397: with angular resolution sufficiently high to isolate it. This test was
2398: performed by Barth, Ho \& Filippenko (2003), who obtained {\it HST}/STIS
2399: spectra, taken with a 0\farcs2-wide slit, of a well-defined subsample of 15
2400: transition objects selected from the Palomar catalog. To their surprise, the
2401: small-aperture spectra of the nuclei, for the most part, look very similar to
2402: the ground-based spectra; they are {\it not} more LINER-like. Shields et al.
2403: (2007) reached the same conclusion from their STIS study of Palomar nuclei,
2404: which included six transition objects, showing that even at \hst\
2405: resolution these objects do not reveal the expected excitation gradients.
2406:
2407: The ``masqueraded-LINER'' hypothesis can be further tested by searching for
2408: compact radio and X-ray cores using high-resolution images. Recall that this
2409: is a highly effective method to filter out weak AGNs (\S\S~5.2, 5.3). Filho,
2410: Barthel \& Ho (2000, 2002a; Filho et al. 2004) have systematically surveyed
2411: the full sample of Palomar transition objects using the VLA at 8.4~GHz. They
2412: find that $\sim$25\% of the population contains arcsecond-scale radio cores.
2413: These cores appear to be largely nonstellar in nature. The brighter subset of
2414: these sources that are amenable to follow-up Very Long Baseline Interferometry
2415: (VLBI) observations (Filho et al. 2004) all reveal more compact
2416: (milliarcsecond-scale) cores with flat radio
2417: spectra and high brightness temperatures (\gax\ $10^7$ K). These radio
2418: statistics are hard to interpret, however, in the absence of a control sample
2419: of other LLAGNs surveyed to the same depth, resolution, and wavelength. The
2420: Nagar, Falcke \& Wilson (2005) 15~GHz survey satisfies these criteria. As
2421: Table~1 shows, the frequency of radio cores in transition objects is roughly
2422: half of that in Seyfert~2s and LINER~2s. On the other hand, the detection
2423: rate of X-ray cores is actually remarkably high---74\%---identical to that of
2424: LINER~2s and similar to that of Seyfert~2s. This observation strongly
2425: suggests that the majority of transition objects indeed {\it do}\ harbor AGNs.
2426:
2427: In light of these recent developments, the basic picture for the physical
2428: nature of transition objects needs to be revised. Inspection of the
2429: statistical properties in Table~1 offers the following clues, which help not
2430: only to explain transition objects but provide a unified view to relate the
2431: different classes of LLAGNs.
2432:
2433: \begin{enumerate}
2434:
2435: {\item
2436: Seyferts, LINERs, and transition objects define a sequence of decreasing
2437: accretion rate. This is most evident from $L_{\rm X}$ and
2438: $L_{\rm bol}/L_{\rm Edd}$, but it is also seen in $L_{{\rm H}\alpha}$
2439: and $P_{\rm rad}$.
2440: }
2441:
2442: {\item
2443: As noted in \S~5.10, type~1 sources have significantly higher luminosities and
2444: Eddington ratios than type~2 systems. The basic premise of the conventional
2445: orientation-based unification scenario does not hold for LLAGNs. The
2446: systematic reduction in accretion rate along the sequence
2447: S$\rightarrow$L$\rightarrow$T also provides a viable explanation for the
2448: systematic decrease in the detection rate of broad H\al\ emission, especially
2449: the precipitous drop among transition sources ($f_b$ in Table~1).
2450: }
2451:
2452: {\item
2453: Transition objects appear to be anomalously strong in their H\al\ emission.
2454: In light of the \hst\ evidence for a distributed source of ionization, I
2455: suspect that a significant fraction of the H\al\ emission in these objects in
2456: fact is {\it not}\ photoionized by the central AGN. This leads to misleading
2457: values of $L_{\rm X}/L_{{\rm H}\alpha}$ and $R_o$ (which is based on
2458: $L_{{\rm H}\alpha}$). For this class either the X-ray or radio core provides
2459: a cleaner measure of the AGN power.
2460: }
2461:
2462: {\item
2463: The loose inverse correlation between radio-loudness and accretion rate, best
2464: seen by comparing $R_{\rm X}$ versus either $L_{\rm X}$ or
2465: $L_{\rm X}/L_{\rm Edd}$, mirrors the trends found by Ho (2002a) and
2466: Terashima \& Wilson (2003b).
2467: }
2468:
2469: {\item
2470: Focusing on the type~2 sources, note that LINER~2s are very similar to
2471: Seyfert~2s, the former being $\sim 1/3-1/2$ as strong as the latter in terms
2472: of H\al\ luminosity and radio power. The two groups have almost identical
2473: $L_{\rm X}$ and $L_{\rm X}/L_{\rm Edd}$, although this may be an artifact of
2474: incomplete absorption correction for Seyfert~2s, some of which are highly
2475: absorbed (Panessa et al. 2006). In the same vein, I propose that transition
2476: objects represent the next step in the luminosity sequence. Judging by their
2477: X-ray luminosity, radio power, $L_{\rm X}/L_{\rm Edd}$, and radio detection
2478: fraction, the AGN component in transition objects is $\sim$1/4 to 1/2 as
2479: strong as that in LINER~2s.
2480: }
2481: \end{enumerate}
2482:
2483: According to the picture just outlined, most, if not all, type~2 sources are
2484: genuinely accretion powered. Using the accretion rate as the metric for
2485: the level of AGN activity, Seyfert~1s rank at the top of the scale, followed
2486: by Seyfert~2s, LINER~1s, LINER~2s, and finally ending with transition objects.
2487: This scenario, which in broad-brush terms explains a wide range of data
2488: summarized in Table~1, has the virtue of simplicity. It is also physically
2489: appealing, given the broad spectrum of accretion rates anticipated in nearby
2490: galaxies.
2491:
2492: There is, however, one loose end that needs to be tied. What powers the
2493: spatially extended, ``excess'' optical line emission in transition objects?
2494: For the reasons explained before, the source of the ionization is unlikely to
2495: be shock heating or photoionization by hot, massive stars, notwithstanding the
2496: success with which such models have been applied to some individual objects
2497: (Engelbracht et al. 1998; Barth \& Shields 2000; Gabel \& Bruhweiler 2002).
2498: Shields et al. (2007) suggest two candidates for a spatially extended source
2499: of ionization: hot, evolved stars and turbulent mixing layers in the
2500: interstellar medium (Begelman \& Fabian 1990). In \S~6.3, I showed that
2501: the stellar mass in the central 100--200 pc indeed seems to provide enough
2502: post-AGB stars to account for the correct level of H\al\ emission in a
2503: significant fraction of the transition objects. I would like to suggest two
2504: other sources, ones that have the advantage of being empirically well
2505: motivated by recent observations. These processes probably operate in all
2506: nuclear environments all the time, maintaining a ``baseline'' level of weak
2507: optical line emission that is only noticeable after the AGN has subsided to a
2508: very low level.
2509:
2510: As discussed in \S\S~5.3, 5.4, the X-ray morphology of the central few hundred
2511: parsecs of galaxies can be quite complex. The nucleus, if present, is often
2512: encircled by other point sources, mostly XRBs (Fabbiano 2006).
2513: With X-ray luminosities ranging from $\sim 10^{37}$ to $10^{39}$ \lum\ (Flohic
2514: et al. 2006), XRBs individually or collectively can outshine the nucleus
2515: itself (Ho et al. 2001; Eracleous et al. 2002; Ho, Terashima \& Ulvestad 2003;
2516: see Figure~5). The discrete sources themselves are embedded in extended
2517: emission, consisting of an optically thin thermal plasma with $kT \approx 0.5$
2518: keV and a spectrally harder diffuse component, which contributes a luminosity
2519: of $\sim (5-9)\times 10^{38}$ \lum\ in the 0.5$-$10 keV band (Flohic et al.
2520: 2006). The hard diffuse component most likely represents the cumulative
2521: emission from faint, unresolved low-mass XRBs, although this interpretation
2522: seems somewhat at odds with the spectrum derived by Flohic et al. (power
2523: law with $\alpha = -0.3$ to $-0.5$). Low-mass XRBs typically can be fit by a
2524: thermal bremsstrahlung model with $kT = 5-10$ keV or a power law with $\alpha
2525: = -0.6$ to $-0.9$ (e.g., Makishima et al. 1989). High-mass XRBs would provide
2526: a better match to the observed spectrum, but in view of what we know about the
2527: stellar populations (\S~4.2), they are probably untenable. A possible
2528: solution is to invoke a multi-temperature plasma (M. Eracleous, private
2529: communications); a hotter component ($kT \approx$ few keV), when added to the
2530: cooler $kT = 0.5$ keV component, would presumably permit a significant
2531: contribution from low-mass XRBs without violating the spectral constraints.
2532: We can estimate the expected X-ray output from low-mass XRBs from the
2533: correlation between optical and X-ray luminosity established for normal
2534: galaxies (Fabbiano \& Trinchieri 1985). Using again the nuclear stellar
2535: magnitudes from the Palomar survey, I obtain a median $L_{\rm X}$($2-10$ keV)
2536: = $(3\pm1)\times10^{38}$ \lum\ within the central 2\asec$\times$4\asec\
2537: aperture, which falls within the ballpark of the value measured by Flohic et
2538: al. (2006). The combination of hot gas and XRB emission, therefore, supplies
2539: $\sim10^{39}$ \lum\ in X-rays, comparable to the amount coming from the
2540: nucleus alone for Seyfert~2s and LINER~2s, and double the amount from
2541: transition nuclei (Table~1). Voit \& Donahue (1997) suggest that hot plasma
2542: additionally may transfer heat conductively to the line-emitting gas in
2543: LINERs, in a manner analogous to the situation in cooling flow filaments in
2544: galaxy clusters.
2545:
2546: Lastly, cosmic ray heating (\S~6.1) by the central radio core will further
2547: enhance the optical line luminosity (Ferland \& Mushotzky 1984). The very
2548: source of the fast particles, namely compact radio jets, itself probably
2549: injects an additional source of mechanical heating, although this is more
2550: difficult to model concretely. Both processes---photoionization by off-nuclear
2551: X-rays and cosmic ray heating---have a convenient virtue: they will tend to
2552: produce low-ionization spectra and therefore provide a natural match to the
2553: spectral characteristics of nearby LLAGNs.
2554:
2555: \section{IMPLICATIONS FOR BLACK HOLE DEMOGRAPHICS}
2556:
2557: As summarized by Kormendy (2004), spatially resolved kinematical observations
2558: have convincingly measured BH masses in a sizable number of nearby galaxies,
2559: to the point that important inferences on their demographics can be drawn
2560: (Richstone 2004). Following an argument first due to So\l tan (1982),
2561: comparison of the integrated radiation density from quasars to the integrated
2562: mass density in local BHs shows that BHs have grown mostly via a radiatively
2563: efficient mode of accretion during their bright AGN phase (e.g., Yu \&
2564: Tremaine 2002). Nearby galaxies, therefore, should be home to AGN relics.
2565:
2566: The LLAGNs summarized in this review provide an important confirmation of this
2567: basic prediction. Not only are weakly accreting BHs found in great abundance
2568: in the local Universe, but they are found where prevailing wisdom says that
2569: they should be found, namely in the centers of galaxies that contain bulges.
2570: Among E, S0, and Sb galaxies, the AGN detection rate is $\sim 50$\%,
2571: increasing to over 70\% among Sa galaxies. Since sensitivity and confusion
2572: impact the detection rates, these statistics are not inconsistent with the
2573: notion that BHs are ubiquitous in essentially {\it all}\ E$-$Sb galaxies.
2574:
2575: Notably, the incidence of AGNs plummets for galaxies with Hubble types
2576: Sc and later, precisely at the point where classical ($r^{1/4}$ profile)
2577: bulges effectively cease to exist and secular dissipation processes begin
2578: to kick in (Kormendy \& Kennicutt 2004). Although the AGN fraction is low for
2579: late-type spirals, it is also not zero. Careful scrutiny of this minority
2580: population addresses two important questions: (1) Are there
2581: central (nonstellar) BHs with masses below $10^6$ \solmass? (2) Must central
2582: BHs always be encased in a bulge?
2583:
2584: \begin{figure}
2585: \vbox{\hbox{
2586: \psfig{figure=fig11a.ps,width=0.465\columnwidth,angle=0}
2587: \hskip +0.2 cm
2588: \psfig{figure=fig11b.ps,width=0.465\columnwidth,angle=0}
2589: }}
2590: \caption{
2591: Two examples of AGNs in late-type galaxies. The {\it left}\ panel shows an
2592: optical image of NGC 4395, adapted from the {\it Carnegie Atlas of Galaxies}\
2593: (Sandage \& Bedke 1994); the image is $\sim$15\amin\ (17 kpc) on a side.
2594: The {\it right}\ panel shows an {\it HST}\ $I$-band image of POX 52, adapted
2595: from C.E. Thornton, A.J. Barth, L.C. Ho, R.E. Rutledge, J.E. Greene (in
2596: preparation); the image is $\sim$11\asec\ (5 kpc) on a side.
2597: }
2598: \end{figure}
2599:
2600: Two remarkable galaxies give the clearest testimony that low-mass BHs do, in
2601: fact, exist. Within the Palomar survey, the nearby ($\sim$4 Mpc) galaxy
2602: NGC~4395 contains all the usual attributes of a self-respecting AGN: broad
2603: optical and UV emission lines (Filippenko \& Sargent 1989; Filippenko, Ho \&
2604: Sargent 1993), a compact radio core (Ho \& Ulvestad 2001) of high brightness
2605: temperature (Wrobel, Fassnacht \& Ho 2001; Wrobel \& Ho 2006), and rapidly
2606: variable hard X-ray emission (Shih, Iwasawa \& Fabian 2003; Moran et al. 2005).
2607: Contrary to expectations, however, NGC~4395 is an extremely late-type
2608: (Sdm) spiral (Figure~11, {\it left}), whose central stellar velocity
2609: dispersion does not exceed $\sim 30$ \kms\ (Filippenko \& Ho 2003). If
2610: NGC~4395 obeys the $M_{\rm BH}-\sigma$ relation, its central BH should have a
2611: mass \lax\ $10^5$ \solmass. This limit agrees surprisingly well with the value
2612: of \mbh\ estimated from its broad H\bet\ line width or X-ray variability
2613: properties ($\sim 10^4-10^5$ \solmass; Filippenko \& Ho 2003), or from
2614: reverberation mapping (3.6\e{5} \solmass; Peterson et al. 2005). POX~52
2615: (Figure~11, {\it right}) presents another interesting case. As first noted by
2616: Kunth, Sargent \& Bothun (1987), the presence of a Seyfert-like nucleus in
2617: POX~52 is unusual because of the low luminosity of the host galaxy. Barth et
2618: al. (2004) show that POX~52 bears a close spectroscopic resemblance to
2619: NGC~4395. Based on the broad profile of H\bet, these authors derive a virial
2620: BH mass of 1.6\e{5} \solmass\ for POX~52, again surprisingly close to the
2621: value of 1.3\e{5} \solmass\ predicted from the $M_{\rm BH} - \sigma$ relation
2622: given the measured central stellar velocity dispersion of 35 \kms.
2623:
2624: The two objects highlighted above demonstrate that the mass spectrum of nuclear
2625: BHs indeed does indeed extend below $10^6$ \solmass, providing great leverage
2626: for anchoring the $M_{\rm BH} - \sigma$ at the low end. Furthermore, they
2627: shed light on the variety of environments in which nuclear BHs can form,
2628: providing much-needed empirical clues to the conditions that fostered the
2629: formation of the seeds for supermassive BHs. NGC~4395 has little or no bulge,
2630: but it does have a compact, central star cluster in which the BH is embedded
2631: (Filippenko \& Ho 2003). Interestingly, G1, a massive star cluster in M31
2632: that to date contains the best direct detection of an intermediate-mass
2633: ($\sim 20,000$ \solmass) BH (Gebhardt, Rich \& Ho 2002, 2005; Ulvestad, Greene
2634: \& Ho 2007), is thought to be the stripped nucleus of a once small galaxy. The
2635: same holds for the Galactic cluster $\omega$ Cen, for which Noyola, Gebhardt
2636: \& Bergmann (2008) report a central dark mass of $5\times10^4$ \solmass.
2637: POX~52 is equally striking. Deep images reveal POX~52 to be most akin to a
2638: spheroidal galaxy (Barth et al. 2004; C.E. Thornton, A.J. Barth, L.C.
2639: Ho, R.E. Rutledge, J.E. Greene, in preparation), to date an unprecedented
2640: morphology for an AGN host galaxy. This is quite unexpected because
2641: spheroidal galaxies, while technically hot stellar systems, bear little
2642: physical resemblance to classical bulges. Spheroidals occupy a distinct
2643: locus on the fundamental plane (e.g., Geha, Guhathakurta \& van~der~Marel
2644: 2002; Kormendy et al. 2008), and they may originate from harassment
2645: and tidal stripping of late-type disk galaxies (e.g., Moore et al. 1996).
2646: Thus, like NGC~4395, POX~52 stands as testimony that a classical bulge is not
2647: a prerequisite for the formation of central BH.
2648:
2649: But how common are such objects? AGNs hosted in high-surface brightness,
2650: late-type spirals appear to be quite rare in the nearby Universe. Within the
2651: comprehensive Palomar survey, NGC~4395 emerges as a unique case of an
2652: unambiguous broad-line AGN hosted in a late-type system. The majority of
2653: late-type spirals do possess compact, photometrically distinct nuclei (B\"oker
2654: et al. 2002), morphologically not dissimilar from that in NGC~4395, but, with
2655: few exceptions (Shields et al. 2008), these nuclei are compact star clusters
2656: with no compelling evidence for an accompanying accreting central BH (Walcher
2657: et al. 2006). Nuclear star clusters do not appear to directly impact a
2658: galaxy's ability to host an AGN (Seth et al. 2008). Several serendipitous
2659: cases of AGNs in late-type galaxies have recently been found from analysis of
2660: {\it Spitzer}\ mid-IR spectra (Satyapal et al. 2007, 2008), as well as a
2661: number of AGN candidates from inspection of {\it Chandra}\ images (Desroches
2662: \& Ho 2008). Among earlier Hubble types, Gallo et al. (2008) report the
2663: detection of X-ray nuclei in two low-luminosity early-type galaxies.
2664:
2665: To assess the true incidence of AGNs like NGC~4395 and POX~52 requires a
2666: spectroscopic survey much larger than Palomar. Greene \& Ho (2007b)
2667: performed a systematic spectral analysis of over 500,000 SDSS spectra with $z <
2668: 0.35$ to search for broad-line AGNs, producing not only a detailed BH mass
2669: function for low-redshift AGNs (Greene \& Ho 2007a) but also a comprehensive
2670: catalog of $\sim 200$ low-mass ($M_{\rm BH} < 10^6$ \solmass) objects. Not
2671: much is known yet about the host galaxies, except that on average they are
2672: about 1 mag fainter than $L^*$. {\it HST}\ imaging of the initial sample of
2673: 19 objects discovered by Greene \& Ho (2004) reveal that the host galaxies are
2674: either mid- to late-type spirals (although none seems as late-type as
2675: NGC~4395) or compact, spheroidal-looking systems not unlike POX~52 (Greene, Ho
2676: \& Barth 2008). When projected onto the galaxy fundamental plane, the
2677: ``bulge'' component in some systems resides within the locus of spheroidal
2678: galaxies. Follow-up high-dispersion spectroscopy shows that these
2679: objects approximately follow the $M_{\rm BH} - \sigma$ relation (Barth, Greene
2680: \& Ho 2005).
2681:
2682: \section{IMPLICATIONS FOR ACCRETION PHYSICS}
2683:
2684: \subsection{Why Are LLAGNs So Dim?}
2685:
2686: This review highlights a class of galactic nuclei that are extraordinary
2687: for being so ordinary. At their most extreme manifestation, LLAGNs emit a
2688: billion times less light than the most powerful known quasars. When quasars
2689: were first discovered, the challenge then was to explain their tremendous
2690: luminosities. Ironically, more than four decades later, the problem has been
2691: reversed: the challenge now is to explain how dead quasars can remain so
2692: dormant. The luminosity deficit problem was noted by Fabian \& Canizares
2693: (1988), who drew attention to the fact that elliptical galaxies, despite being
2694: suffused with a ready supply of hot gas capable of undergoing spherical
2695: accretion, have very dim nuclei. We can no longer speculate that these
2696: systems lack supermassive BHs, for we now know that they are there, at least
2697: in galaxies with bulges. And as I have shown in this review, the problem is
2698: by no means confined to ellipticals either.
2699:
2700: Explanations of the luminosity paradox fall in several categories.
2701:
2702: \begin{itemize}
2703:
2704: {\item
2705: {\underbar{\it Obscuration}} \ \ \ This trivial possibility can be summarily
2706: dismissed as a general solution in light of the evidence presented in
2707: \S\S~5.3, 5.6.
2708: }
2709:
2710: {\item
2711: {\underbar{\it Low accretion rate}} \ \ \ A more plausible strategy is to
2712: starve the BH. Present-day massive galaxies, after all, should be relatively
2713: gas-poor, especially in their central regions, which are largely devoid of
2714: significant ongoing star formation. This argument quickly falls apart when
2715: one realizes just how little material is needed to light up the nuclei. The
2716: bolometric luminosities of nearby nuclei span $\sim 10^{38}-10^{44}$ \lum,
2717: with a median value of $L_{\rm bol} = 3\times 10^{40}$ \lum\ and half of the
2718: sample lying between $3\times 10^{39}$ and $3\times 10^{41}$ \lum. For a
2719: canonical radiative efficiency of $\eta = 0.1$, the required accretion rate is
2720: merely $\dot M=L_{\rm bol}/\eta c^2=5\times10^{-6\pm1}$ \solmass\ \peryr.
2721: This is a pitifully miniscule amount, in comparison with the amount of fuel
2722: actually available to be accreted. Galactic nuclei unavoidably receive fuel
2723: from two sources: ordinary mass loss from evolved stars and gravitational
2724: capture of gas from the hot interstellar medium.
2725:
2726: During the normal course of stellar evolution, evolved stars return a
2727: significant fraction of their mass to the ISM through mass loss. For a
2728: Salpeter stellar initial mass function with a lower-mass cutoff of 0.1
2729: \solmass, an upper-mass cutoff of 100 \solmass, solar metallicities, and an
2730: age of 15 Gyr, Padovani \& Matteucci (1993) estimate that
2731:
2732: \begin{displaymath}
2733: \dot M_*\,\approx\,3\times10^{-11} \, \left(\frac{L}{L_{\odot, V}}\right)
2734: \,\,\,\,\,\,\, M_{\odot}\,{\rm yr}^{-1}.
2735: \end{displaymath}
2736:
2737: \noindent
2738: This result is consistent, within a factor of $\sim$2, with the work of Ciotti
2739: et al. (1991) and Jungwiert, Combes \& Palous (2001). {\it HST}\ images reveal
2740: that galaxies contain central density concentrations, either in the form of
2741: nuclear cusps or photometrically distinct, compact stellar nuclei. The
2742: cusp profiles continue to rise to the resolution limit of \hst\ (0\farcs1),
2743: which is $r \approx 10$ pc at a distance of 20 Mpc, where
2744: $\rho\, \approx\, 10-10^3$ $L_{\odot, V}$ pc$^{-3}$ for the ``core''
2745: ellipticals and $\rho\, \approx\, 10^2-10^4$ $L_{\odot, V}$ pc$^{-3}$ for the
2746: ``power-law'' ellipticals and bulges of early-type spirals and S0s (e.g.,
2747: Faber et al. 1997). Within a spherical region of $r = 10$ pc, the diffuse
2748: cores have $L \approx 4\times 10^4 - 4\times 10^6\, L_{\odot, V}$, which
2749: yields $\dot M_*\,\approx\,1\times10^{-6} - 1\times10^{-4}$ \solmass\
2750: \peryr; for the denser power-law cusps, $L \approx 4\times 10^5 -
2751: 4\times 10^7\, L_{\odot, V}$, or $\dot M_*\,\approx\,1\times10^{-5} -
2752: 1\times10^{-3}$ \solmass\ \peryr. Centrally dominant nuclear star clusters,
2753: present in a large fraction of disk galaxies, typically have luminosities
2754: $L \approx 10^7$ \solum\ (Carollo et al. 1997; B\"oker et al. 2002), and hence
2755: $\dot M_*\,\approx\,10^{-3}$ \solmass\ \peryr.
2756:
2757: Diffuse, hot gas in the central regions of galaxies holds another potential
2758: fuel reservoir. Low-angular momentum gas sufficiently close to a BH can
2759: accrete spherically in the manner described by Bondi (1952). To estimate the
2760: Bondi accretion rate, one needs to know the gas density and temperature at the
2761: accretion radius, $R_a \approx GM_{\rm BH}/c_s^2$, where $c_s \approx 0.1
2762: T^{1/2}$ \kms\ is the sound speed of the gas at temperature $T$. The mass
2763: accretion rate follows from the continuity equation, $\dot M_{\rm B}\,=\,4\pi
2764: R_a^2 \rho_a c_s$, where $\rho_a$ is the gas density at $R_a$. Expressed in
2765: terms of typical observed parameters (see below),
2766:
2767: \begin{displaymath}
2768: \dot M_{\rm B}\,\approx\,7.3\times 10^{-4}\,
2769: \left(\frac{M_{\rm BH}}{10^8\, M_{\odot}}\right)^2\,
2770: \left(\frac{n}{0.1\, {\rm cm}^{-3}}\right)\,
2771: \left(\frac{200\, {\rm km\,s}^{-1}}{c_s}\right)^3\,
2772: \,M_{\odot}\,{\rm yr}^{-1}.
2773: \end{displaymath}
2774:
2775: {\it Chandra}\ observations with sufficient resolution to resolve $R_a$ find
2776: that the diffuse gas in the central regions of elliptical galaxies typically
2777: has temperatures of $kT\,\approx\,0.3-1$ keV and densities of $n\,\approx\,
2778: 0.1-0.5$ \cc\ (e.g., Di~Matteo et al. 2001, 2003; Loewenstein et al. 2001;
2779: Pellegrini 2005). Our knowledge of the hot gas
2780: content in the central regions of the bulges of spiral and S0 galaxies is more
2781: fragmentary. {\it Chandra}\ has so far resolved the hot gas in the centers of
2782: a handful of bulges (Milky Way: Baganoff et al. 2003; M31: Garcia et al. 2005;
2783: M81: Swartz et al. 2002; NGC~1291: Irwin, Sarazin \& Bregman 2002; NGC~1553:
2784: Blanton, Sarazin \& Irwin 2001; Sombrero: Pellegrini et al. 2003a). These
2785: studies suggest that bulges typically have gas temperatures of $kT\,\approx\,
2786: 0.3-0.6$ keV. Information on gas densities is sketchier, but judging from the
2787: work on M81 and the Sombrero, a fiducial value might be $n\,\approx\,0.1$ \cc.
2788:
2789: If, for simplicity, we assume that the hot gas in most bulges is characterized
2790: by $n\,=\,0.1$ \cc\ and $kT\,=\,0.3$ keV, then $\dot M_{\rm B}\,\approx\,
2791: 10^{-5}-10^{-3}$ \solmass\ \peryr\ for \mbh\ = $10^7-10^8$ \solmass. In
2792: elliptical galaxies \mbh\ $\approx\,10^8-10^9$ \solmass, and for $n\,=\,0.2$
2793: \cc\ and $kT\,=\,0.5$ keV, $\dot M_{\rm B}\,\approx\, 10^{-4}-10^{-2}$
2794: \solmass\ \peryr. We note that these estimates of the Bondi accretion rates,
2795: which fall within the range given in recent compilations (e.g., Donato,
2796: Sambruna \& Gliozzi 2004; Pellegrini 2005; Soria et al. 2006), are probably
2797: lower limits because the actual densities near $R_a$ are likely to be higher
2798: than we assumed. For the above fiducial temperatures and BH masses,
2799: $R_a\,\approx\,1-10$ pc for bulges and $\sim 10-100$ pc for ellipticals,
2800: roughly an order of magnitude smaller than the typical linear resolution
2801: achieved by {\it Chandra}\ for nearby galaxies. In well-resolved cases, the
2802: gas temperature profile generally remains constant to within $\sim 50$\%
2803: whereas the density typically increases by a factor of a few toward the center
2804: (e.g., Milky Way: Baganoff et al. 2003; M31: Garcia et al. 2005; M87:
2805: Di~Matteo et al. 2003; NGC~1316: Kim \& Fabbiano 2003).
2806:
2807: Although the above estimates are very rough, and they are valid only in a
2808: statistical sense, one cannot escape the conclusion that in general in the
2809: central few parsecs of nearby galaxies $\dot M_{\rm B} + \dot M_* \gg \dot M$.
2810: Although meager, the joint contributions from stellar mass loss and Bondi
2811: accretion, if converted to radiation with $\eta = 0.1$, would generate nuclei
2812: far more luminous than actually observed. The net accretion from the Bondi
2813: flow would be considerably reduced if the gas possesses some angular momentum
2814: at large radii (Proga \& Begelman 2003)---as inevitably it must---but even so
2815: it seems likely that the BH still has plenty of food at its disposal. LLAGNs
2816: are by no means fuel-starved. Moreover, the above estimates have erred on
2817: the conservative side. For example, I have assumed that all of the stars are
2818: evolved, although in reality most nuclei have composite populations and hence
2819: larger mass loss rates. Furthermore, I have neglected additional dissipation
2820: from larger scales (e.g., due to nuclear bars or spirals), as well as
2821: discrete, episodic events such as stellar tidal disruptions, which can provide
2822: a significant source of fuel, especially for BHs with masses \lax\ $10^{7}$
2823: \solmass\ (Milosavljevi\'c, Merritt \& Ho 2006). All of these additional
2824: sources will only exacerbate the fuel surplus crisis.
2825: }
2826:
2827: {\item
2828: {\underbar{\it Low radiative efficiency}} \ \ \ If accretion does proceed
2829: at a reasonable fraction of the supply rate, then one has no option but to
2830: conclude that the radiative efficiency is much less than $\eta = 0.1$, the
2831: standard value for an optically thick, geometrically thin disk. This type of
2832: argument has been invoked to explain the apparent conflict between the nuclear
2833: luminosities and Bondi accretion rates in many early-type galaxies (e.g.,
2834: Fabian \& Rees 1995; Reynolds et al. 1996; Di~Matteo \& Fabian 1997; Mahadevan
2835: 1997; Di~Matteo et al. 2001, 2003; Loewenstein et
2836: al. 2001; Ho, Terashima \& Ulvestad 2003; Pellegrini et al. 2003a; Donato,
2837: Sambruna \& Gliozzi 2004; Evans et al. 2006). Accretion flows with low
2838: radiative efficiency, of which the most popular version is the ADAF (see
2839: reviews in Narayan 2002; Yuan 2007), arise when the accreting medium is
2840: sufficiently tenuous that its cooling time exceeds the accretion timescale.
2841: RIAFs are predicted to exist for accretion rates below a critical threshold of
2842: $\dot M_{\rm crit} \approx 0.3\alpha^2 \dot M_{\rm Edd} \approx 0.01 \dot
2843: M_{\rm Edd}$, where the Shakura \& Sunyaev (1973) viscosity parameter is taken
2844: to be $\alpha \approx 0.1-0.3$ and $\dot M_{\rm Edd} \equiv L_{\rm Edd}/\eta
2845: c^2 = 0.22 \left(\eta/0.1\right) \left(M_{\rm BH}/10^8\, M_{\odot}\right)$
2846: \solmass\ \peryr. LLAGNs lie comfortably below this threshold.
2847: }
2848:
2849: {\item
2850: {\underbar{\it Inefficient accretion/jet feedback}} \ \ \ Precisely how low
2851: $\eta$ can become depends on how much of the native fuel supply actually gets
2852: accreted. In the presence of some rotation in the ambient medium, numerical
2853: simulations find that the amount of material accreted is much less than is
2854: available at large radii (e.g., Stone \& Pringle 2001; Igumenshchev, Narayan
2855: \& Abramowicz 2003). Since the gravitational binding energy in a RIAF cannot
2856: be radiated efficiently, it must be lost by nonradiative means (Blandford \&
2857: Begelman 1999), either through convective transport of energy and angular
2858: momentum to large radii or through a global outflowing wind (see review by
2859: Quataert 2003). The net effect of either process is to flatten the density
2860: profile near the center and to dramatically reduce the accretion rate. At
2861: very low accretion rates \lax\ $(10^{-5}-10^{-6}) \,\dot M_{\rm Edd}$, such as
2862: in Sgr~A$^*$ and some giant elliptical galaxies, electron heat conduction will
2863: further suppress the accretion rate (Johnson \& Quataert 2007). While these
2864: effects will ease the burden of invoking extremely small and perhaps
2865: physically unrealistic radiative efficiencies, it is important to note that
2866: these more recent models are {\it still}\ radiatively inefficient.
2867:
2868: Whether the outflows generated in RIAFs can lead directly to relativistic jets
2869: is unclear, but what observations have established is the tendency for lowly
2870: accreting systems to become increasingly jet-dominated. We see this not only
2871: in FR~I radio galaxies (Chiaberge, Capetti \& Celotti 1999; Donato, Sambruna
2872: \& Gliozzi 2004; Kharb \& Shastri 2004; Chiaberge, Capetti \& Macchetto 2005;
2873: Balmaverde \& Capetti 2006; Wu, Yuan \& Cao 2007), but it is also reflected in
2874: the nuclear properties of more run-of-the-mill LLAGNs (\S\S~5.3, 5.8), as well
2875: as in nearly quiescent nuclei (Pellegrini et al. 2007; Wrobel, Terashima \& Ho
2876: 2008). Detailed analysis of some sources, in fact, indicates that most of the
2877: accretion power is not radiated but instead channeled into the kinetic energy
2878: of relativistic jets (M87: Di~Matteo et al. 2003; IC~1459: Fabbiano et al.
2879: 2003; IC~4296: Pellegrini et al. 2003b). By analogy with the situation in
2880: cooling flows in galaxy clusters, the propensity for LLAGNs to become
2881: radio-loud opens up the possibility that the kinetic energy from small-scale
2882: jets or collimated outflows provides a major source of ``feedback'' into the
2883: circumnuclear environment, perhaps to the extent that accretion can be
2884: significantly interrupted or curtailed (Binney \& Tabor 1995; Di~Matteo et al.
2885: 2003; Pellegrini et al. 2003a; Omma et al. 2004). Indeed, calculations by
2886: K\"ording, Jester \& Fender (2008; see also Heinz, Merloni \& Schwab 2007)
2887: show that the total kinetic energy injected by LLAGN jets is very substantial.
2888: Given the prominent hard X-ray component in LLAGN spectra, inverse-Compton
2889: scattering of the hard photons might also provide another avenue to heat the
2890: ambient medium (Ciotti \& Ostriker 2001). Either form of energy
2891: injection---mechanical or radiative---can lead to unsteady, intermittent
2892: accretion with a short duty cycle.
2893: }
2894:
2895: {\item
2896: {\underbar{\it Subluminous disk}} \ \ \ A thin disk can be tolerated if it
2897: can be made extremely subluminous during periods of intermittent activity
2898: (Shields \& Wheeler 1978). This situation would arise if accretion disks in
2899: AGNs undergo the thermal-viscous ionization instability (Lin \& Shields 1986;
2900: Siemiginowska, Czerny \& Kostyunin 1996), in which case they spend most of
2901: their time in quiescence, punctuated by brief episodes of intense outbursts.
2902: Menou \& Quataert (2001) questioned the applicability of the ionization
2903: instability in AGNs, but they suggested that low-luminosity systems (with
2904: $\dot M$ \lax\ $10^{-3}$ \solmass\ \peryr) may contain disks in which mass
2905: accumulates in a stable, nonaccreting ``dead zone.'' Others have managed
2906: to stall accretion by condensing the hot flow into an inner cold, inert disk
2907: (Nayakshin 2003; Tan \& Blackman 2005; Jolley \& Kuncic 2007), which may form
2908: naturally from Compton cooling of the corona (Liu et al. 2007). Finally,
2909: Merloni \& Fabian (2002) proposed that LLAGNs do contain a cold thin disk, but
2910: because of their low mass accretion rates, they liberate a large fraction of
2911: their gravitational energy in a strongly magnetized, unbound corona. Since a
2912: cold disk component is present in all these models, they face a serious, and
2913: in my opinion insurmountable, challenge because LLAGNs generally do not show
2914: fluorescent Fe~K\al\ emission or reflection features in their X-ray spectra.
2915: The Merloni \& Fabian model may be spared of this criticism, as the disk may
2916: be highly ionized, but it does predict strong and rapid X-ray variability,
2917: which is generally {\it not}\ observed in LLAGNs (\S~5.3; Ptak et al. 1998).
2918: }
2919: \end{itemize}
2920:
2921: \subsection{The Disk-Jet Connection}
2922:
2923: As the mass accretion rate drops and the radiative efficiency declines, an
2924: increasing fraction of the accretion power gets channeled into a relativistic
2925: jet whose energy release is mainly kinetic rather than radiative. The
2926: principal evidence for the growing importance of jets in LLAGNs comes from the
2927: broad-band SEDs, which invariably are prominent in the radio, with the degree
2928: of radio-loudness rising systematically (albeit with significant scatter) with
2929: decreasing Eddington ratio (\S~5.8; Figure~10{\it b}). Where available, VLBI
2930: imaging on milliarcsecond scales reveals unresolved cores with nonthermal
2931: brightness temperatures and a flat or slightly inverted spectrum---classical
2932: signposts of a relativistic jet (Blandford \& K\"onigl 1979). Detailed
2933: modeling of the SEDs of individual sources often shows that the accretion flow
2934: itself does not produce enough radio emission to match the data: that extra
2935: ``something else'' is most plausibly attributed to the jet component (Quataert
2936: et al. 1999; Ulvestad \& Ho 2001b; Fabbiano et al. 2003; Pellegrini et al.
2937: 2003b; Anderson, Ulvestad \& Ho 2004; Ptak et al. 2004; Wu \& Cao 2005; Nemmen
2938: et al. 2006; Wu, Yuan \& Cao 2007). Moreover, RIAF models predict radio
2939: spectral indices of $\alpha \approx +0.4$ (Mahadevan 1997), whereas the
2940: observed values more typically fall in the range $\alpha \approx -0.2$ to
2941: $+0.2$.
2942:
2943: The jet may contribute substantially outside of the radio band, especially
2944: in the optical and X-rays. Some advocate that the jet, in fact, accounts
2945: for most or even all of the emission across the broad-band SED. For example,
2946: Yuan et al. (2002) successfully fitted the multiwavelength data of NGC~4258
2947: with effectively a jet-only model. In their picture, a radiative shock at the
2948: base of the jet gives rise to synchrotron emission in the near-IR and optical
2949: regions, whose self-Compton component then explains the X-rays; the
2950: flat-spectrum radio emission comes from further out in the jet. Similar
2951: models have been devised for the Galactic Center source Sgr~A$^*$ (Falcke \&
2952: Markoff 2000; Yuan, Markoff \& Falcke 2002). The gross similarity between the
2953: SEDs of some FR~I nuclei and BL Lac objects, which are jet-dominated sources
2954: but otherwise also low-accretion rate systems (Wang, Staubert \& Ho 2002), has
2955: also been noted (e.g., Bower et al. 2000; Capetti et al. 2000; Chiaberge
2956: et al. 2003; Meisenheimer et al. 2007).
2957:
2958: Statistical samples that are larger but more limited in spectral coverage have
2959: come from combining radio data with high-resolution optical or X-ray
2960: observations.
2961: Studies that specifically target radio galaxies, particularly FR~I sources
2962: and weak-line FR~IIs, report that the core radio power scales tightly with the
2963: optical and/or X-ray continuum luminosity, a finding often taken to support a
2964: common nonthermal, jet origin for the broad-band emission (Worrall \&
2965: Birkinshaw 1994; Canosa et al. 1999; Chiaberge, Capetti \& Celotti 1999, 2000;
2966: Capetti et al. 2002; Verdoes~Kleijn et al. 2002; Donato, Sambruna \& Gliozzi
2967: 2004; Balmaverde \& Capetti 2006; Balmaverde, Capetti \& Grandi 2006; Evans
2968: et al. 2006; for a counterargument, see Rinn, Sambruna \& Gliozzi 2005 and
2969: Gliozzi et al. 2008). A similar radio-optical correlation, after correcting
2970: for Doppler boosting, is also seen among BL Lac objects (Giroletti et al.
2971: 2006), strengthening the case that FR~I radio galaxies and BL Lac objects are
2972: intrinsically the same but misoriented siblings. Many FR~II systems, on the
2973: other hand, especially those with broad lines, deviate systematically from the
2974: baseline FR~I correlations, by exhibiting stronger optical and
2975: X-ray emission for a given level of radio emission (Chiaberge, Capetti \&
2976: Celotti 2000, 2002; Varano et al. 2004). In concordance with the frequent
2977: detection of X-ray absorption and Fe~K$\alpha$ emission (Evans et al. 2006),
2978: this suggests that FR~IIs have higher accretion rates and a much more dominant
2979: accretion flow component, relative to the jet, than FR~Is.
2980:
2981: Any attempt to explain the broad-band spectrum of LLAGNs with either just a
2982: RIAF or just a jet runs the risk of oversimplification. Clearly both are
2983: required. The trick is to figure out a reliable way to divvy up the two
2984: contributions to the SED. We cannot deny that there is a jet, because we see
2985: it directly in the radio at a strength far greater than can be attributed to
2986: the RIAF. The jet emission must contribute at some level outside of the radio
2987: band. At the same time, the jet cannot exist in isolation; it is anchored to
2988: and fed by some kind of accretion flow, of which a promising configuration is
2989: a vertically thick RIAF (Livio, Ogilvie \& Pringle 1999; Meier 1999). An
2990: outstanding problem is that the interpretation of the data is not unique.
2991: Because many of the model parameters are poorly constrained and the broad-band
2992: data remain largely fragmentary and incomplete, the SEDs often can be fit with
2993: pure jet models, pure accretion flow models, or some combination of the two.
2994: The recent detection of high levels of polarization in the optical nuclei of
2995: FR~Is (Capetti et al. 2007) strongly points toward a synchrotron origin in the
2996: jet for the optical continuum, but even this observation cannot be considered
2997: definitive, because a RIAF can also produce nonthermal flares (e.g., in
2998: Sgr~A$^*$; Quataert 2003).
2999:
3000: The so-called BH fundamental plane---a nonlinear correlation among radio
3001: luminosity, X-ray luminosity, and BH mass---offers a promising framework to
3002: unify accreting BHs over a wide range in mass and accretion rates. Merloni,
3003: Heinz \& Di~Matteo (2003) first demonstrated that the correlation between
3004: $L_{\rm rad}$ and $L_{\rm X}$ tightens considerably after including \mbh\ as a
3005: third variable. Combining observational material for several Galactic stellar
3006: BHs and a large sample of nearby LLAGNs, they find that
3007:
3008: \begin{displaymath}
3009: \log L_{\rm rad} = 0.60 \log L_{\rm X} + 0.78 \log M_{\rm BH}.
3010: \end{displaymath}
3011:
3012: \noindent
3013: This empirical correlation agrees well with the theoretical relations between
3014: radio flux, BH mass, and accretion rate derived from the scale-invariant
3015: disk-jet model of Heinz \& Sunyaev (2003). The BH fundamental plane, however,
3016: appears to be a very blunt tool. In an independent analysis, Falcke,
3017: K\"ording \& Markoff (2004) obtained a similar empirical relation, but unlike
3018: Merloni, Heinz \& Di~Matteo these authors explained the scaling coefficients
3019: entirely in terms of a jet-dominated model. Moreover, as emphasized by
3020: K\"ording, Falcke \& Corbel (2006), objects with very different emission
3021: processes, including luminous quasars and BL Lac objects, sit on the same
3022: correlation, albeit with larger scatter.
3023:
3024: \begin{figure}
3025: %%BoundingBox: 20 150 550 670
3026: \vbox{\hbox{
3027: \hskip -0.2 cm
3028: \psfig{figure=fig12a.ps,width=6.7cm,angle=0}
3029: \hskip -0.0 cm
3030: \psfig{figure=fig12b.ps,width=6.7cm,angle=0}
3031: }}
3032: \caption{({\it a}) Fundamental plane correlation among core radio luminosity,
3033: X-ray luminosity, and BH mass. ({\it b}) Deviations from the fundamental
3034: plane as a function of Eddington ratio.
3035: }
3036: \end{figure}
3037:
3038: I illustrate this point in Figure~12{\it a}, which includes all Palomar LLAGNs
3039: with suitable data, along with the collection of high-luminosity sources from
3040: L.C. Ho (in preparation). With the exception of a handful of radio-loud
3041: quasars, the vast
3042: majority of the objects fall on a well-defined swath spanning $\sim 10$ orders
3043: of magnitude in luminosity. There are no obvious differences among the various
3044: subclasses of LLAGNs, except that the type~1 sources appear more tightly
3045: correlated. Plotting the residuals of the fundamental plane relation versus
3046: the Eddington ratio reveals two interesting points (Figure~12{\it b}). First,
3047: although the intrinsic scatter of the relation is quite large, it markedly
3048: increases for objects with high Eddington ratios, at $L_{\rm bol}/L_{\rm Edd}
3049: \approx 10^{-1\pm1}$, as already noted by Maccarone, Gallo \& Fender (2003)
3050: and Merloni, Heinz \& Di~Matteo (2003). The scatter flares up because the
3051: radio-loud quasars lie offset above the relation and the radio-quiet quasars
3052: on average lie offset below the relation. At the opposite extreme, sources
3053: with $L_{\rm bol}/L_{\rm Edd}$ \lax\ $10^{-6.5}$ may also show a systematic
3054: downturn, in possible agreement with the proposal by Yuan \& Cui (2005) that
3055: below a critical threshold, $L_{\rm X} \approx 10^{-5.5} L_{\rm Edd}$, both
3056: the radio {\it and}\ the X-rays should be dominated by emission from the jet.
3057: M31 (Garcia et al. 2005), NGC~821 (Pellegrini et al. 2007), and NGC~4621 and
3058: NGC~4697 (Wrobel, Terashima \& Ho 2008) seem to conform to Yuan \& Cui's
3059: prediction, but M32 and especially Sgr~A$^*$ clearly do not. Additional deep
3060: radio and X-ray observations of ultra-low-luminosity nuclei would be very
3061: valuable to clarify the situation in this regime.
3062:
3063: If, as surmised, the relative proportions between jet and disk output depend
3064: on accretion rate, with the bulk of the radiated power, even in the X-rays,
3065: originating from the former in the lowest accretion rate systems, two
3066: important consequences ensue. With respect to the microphysics of RIAFs, it
3067: implies that the radiative efficiencies are even lower than previously
3068: inferred under the assumption that the X-rays emanate solely from the accretion
3069: flow. On a more global, environmental scale, shifting the emphasis from
3070: the disk to the jet changes the balance between kinetic versus radiative
3071: output, with important implications for prescriptions of AGN feedback in models
3072: of galaxy formation because BHs spend most of their lives in a low-state.
3073: From empirical and theoretical considerations (Heinz, Merloni \& Schwab 2007;
3074: K\"ording, Jester \& Fender 2008), the jet carries a substantial fraction
3075: of the accreted rest mass energy:
3076: $P_{\rm jet} \approx 0.2 \eta \dot M c^2 \approx 7.2\times10^{36}
3077: (L_{\rm rad}/10^{30}\,{\rm ergs\,\,s}^{-1})^{12/17}$ \lum. In
3078: fact, the total kinetic energy injected by LLAGN jets is comparable to or
3079: perhaps even greater than the contribution from supernovae. At low
3080: redshifts, radiative feedback from quasars, which is commonly assumed to
3081: operate with an efficiency of $\sim 5$\%, may
3082:
3083: \clearpage
3084: \begin{figure}
3085: \psfig{figure=fig13.ps,width=13.5cm,angle=0}
3086: \caption{
3087: A cartoon of the central engine of LLAGNs, consisting of three
3088: components: an inner, radiatively inefficient accretion flow (RIAF), an outer,
3089: truncated thin disk, and a jet or outflow. (Courtesy of S. Ho.)
3090: }
3091: \end{figure}
3092:
3093: \noindent
3094: be less important then
3095: jet-driven feedback from LLAGNs (K\"ording, Jester \& Fender 2008).
3096:
3097: \subsection{The Central Engine of LLAGNs}
3098:
3099: The preceding sections argue that the weak nuclear activity seen in the
3100: majority of nearby galaxies traces low-level BH accretion akin to the more
3101: familiar form observed in powerful AGNs. However, multiple lines of evidence
3102: indicate that LLAGNs are not simply scaled-down versions of their more luminous
3103: cousins. They are qualitatively different. From the somewhat fragmentary clues
3104: presented in this review, we can piece together a schematic view of the
3105: structure of the central engine in LLAGNs (Ho 2002b, 2003, 2005). As sketched
3106: in Figure~13, it has three components.
3107:
3108: \begin{enumerate}
3109: \item{{\underbar{\it Radiatively inefficient accretion flow}} \ \ \
3110: In the present-day Universe, and especially in the centers of big bulges, the
3111: amount of material available for accretion is small, resulting in mass
3112: accretion rates that fall far below $10^{-2} \dot M_{\rm Edd}$. In such a
3113: regime, the low-density, tenuous material is optically thin and cannot cool
3114: efficiently. Rather than settling into a classical optically thick,
3115: geometrically thin, radiatively efficient disk---the normal configuration for
3116: luminous AGNs---the accretion flow puffs up into a hot, quasi-spherical,
3117: radiatively inefficient distribution, whose dynamics may be dominated by
3118: advection, convection, or outflows. This is an area of active ongoing
3119: theoretical research. In the interest of brevity, I will gloss over the
3120: technical details and simply follow Quataert (2003) by calling these RIAFs.
3121: The existence of RIAFs, or conversely the absence of a standard thin disk
3122: extending all the way to small radii (a few Schwarzschild radii $R_{\rm S}$),
3123: is suggested by the feeble luminosities of LLAGNs, by their low Eddington
3124: ratios, and especially by their low inferred radiative efficiencies. The great disparity between the available fuel supply and the actually observed accretion
3125: luminosity demands that the radiative efficiency of the accretion flow be much
3126: less than $\eta = 0.1$ (\S~8.1). Additional support for RIAFs comes from
3127: considerations of the SED, particularly the absence of the big blue bump, a
3128: classical signature of the thin disk, and the preponderance of intrinsically
3129: hard X-ray spectra.
3130: }
3131:
3132: \item{{\underbar{\it Truncated thin disk}} \ \ \
3133: Beyond a transition radius $R_{\rm tr} \approx 100-1000 R_{\rm S}$, the RIAF
3134: switches to a truncated optically thick, geometrically thin disk. The
3135: observational evidence for this component comes in three forms. First, the
3136: SEDs of some well-studied LLAGNs require a truncated thin disk to explain the
3137: big red bump---the prominent mid-IR peak and the steep fall-off of the
3138: spectrum in the optical--UV region (\S~5.8). The thermal disk emission is
3139: cool (red) not only because of a low accretion rate (Lawrence 2005) but also
3140: because the inner radius of the disk does not extend all the way in to a few
3141: $R_{\rm S}$ as in luminous AGNs. Second, the very same truncated disk
3142: structure employed to model the SED simultaneously accounts for the
3143: relativistically broadened, double-peaked emission-line profiles observed in
3144: some sources (\S~5.5). Indeed, in the case of NGC~1097 (Nemmen et al. 2006),
3145: the transition radius derived from modeling the SED ($R_{\rm tr} = 225
3146: R_{\rm S}$) agrees remarkably well with the inner radius of the disk obtained
3147: from fitting the double-peaked broad H\al\ profile. Ho et al. (2000) suggested
3148: that double-peaked broad emission lines are commonplace in LLAGNs. By
3149: implication, the truncated disk configuration inferred from this special class
3150: of line profiles must be commonplace too. Lastly, the striking absence of
3151: broad Fe~K\al\ emission in the X-ray spectra of LLAGNs (\S~5.3), a feature
3152: commonly attributed to X-ray fluorescence off of a cold accretion disk
3153: extending inward to a few $R_{\rm S}$ in bright Seyfert~1 nuclei (e.g., Nandra
3154: et al. 1997b, 2007), strongly suggests that in low-luminosity sources such a
3155: structure is either absent or truncated interior to some radius, such that it
3156: subtends a significantly smaller solid angle. Similar lines of reasoning have
3157: been advanced for broad-line radio galaxies that show weak Fe~K\al\ emission
3158: and weak Compton reflection (Wo\'zniak et al. 1998; Eracleous, Sambruna \&
3159: Mushotzky 2000; Lewis et al. 2005), although these characteristics can be
3160: mimicked by an ionized but otherwise untruncated disk (Ballantyne, Ross \&
3161: Fabian 2002).
3162: }
3163:
3164: \item{{\underbar{\it Jet/outflow}} \ \ \ The empirical connection between
3165: LLAGNs and jets has been established unequivocally from radio observations.
3166: Not only are the SEDs of LLAGNs generically radio-loud, but the strength of the
3167: radio emission generally cannot be fit without recourse to a jet component,
3168: which in many cases can be seen directly from VLBI-scale radio images. From a
3169: theoretical point of view, jets may share a close physical connection with
3170: RIAFs. As emphasized by Narayan \& Yi (1995) and Blandford \& Begelman (1999),
3171: RIAFs have a strong tendency to drive bipolar outflows due to the high thermal
3172: energy content of the hot gas. Whether such outflows can develop into highly
3173: collimated, relativistic ejections remains to be seen, but they at least
3174: provide a promising starting point. RIAFs may be additionally conducive to
3175: jet formation because its vertically thick structure enhances the large-scale
3176: poloidal component of the magnetic field, which plays a critical role in
3177: launching jets (Livio, Ogilvie \& Pringle 1999; Meier 1999; Ballantyne \&
3178: Fabian 2005; Ballantyne 2007). It is interesting to recall that the original
3179: motivation for ion-supported tori (Rees et al. 1982), an early incarnation of
3180: RIAFs, was to explain the low luminosity of radio galaxies. Rees et al.
3181: postulated that the puffed-up structure of the ion torus may help facilitate
3182: the collimation of the jet.
3183: }
3184: \end{enumerate}
3185:
3186: The above-described three-component structure has been applied to model the
3187: broad-band spectrum of a number of LLAGNs, including NGC~4258 (Lasota et al.
3188: 1996; Gammie, Narayan \& Blandford 1999), M81 and NGC~4579 (Quataert et al.
3189: 1999), NGC~3998 (Ptak et al. 2004), and NGC~1097 (Nemmen et al. 2006). For the
3190: handful of LLAGNs with available estimates of the transition radii,
3191: $R_{\rm tr}$ seems to scale roughly inversely with $L_{\rm bol}/L_{\rm Edd}$
3192: (Yuan \& Narayan 2004). This trend may be in agreement with models
3193: for disk evaporation (Liu \& Meyer-Hofmeister 2001). As the latter authors
3194: note, however, disks attain their maximum evaporation efficiency at
3195: $R_{\rm tr} \approx 300 R_{\rm S}$, making sources such as M81 and NGC~4579,
3196: both with $R_{\rm tr} \approx 100 R_{\rm S}$ (Quataert et al. 1999), difficult
3197: to explain. At a qualitative level, at least, the general idea that the thin
3198: disk recedes to larger and larger radii as the accretion rate drops is
3199: probably correct. In an analysis of 33 PG quasars with Fe~K\al\ emission
3200: detected in {\it XMM-Newton}\ spectra, Inoue, Terashima \& Ho (2007) find
3201: that the iron line profile varies systematically with Eddington ratio.
3202: Specifically, the Fe~K\al\ profile becomes narrower with decreasing
3203: $L_{\rm bol}/L_{\rm Edd}$, a result that can be interpreted as a systematic
3204: increase in the inner radius of the accretion disk at low accretion rates.
3205:
3206: The basic schematic proposed in Figure~13 is hardly new. To my knowledge, a
3207: hybrid model consisting of a RIAF---then called an ion-supported torus---plus
3208: a truncated thin disk was most clearly articulated in a prescient paper by
3209: Chen \& Halpern (1989) in their description of Arp~102B, later elaborated by
3210: Eracleous \& Halpern (1994) in the general context of double-peaked broad-line
3211: radio galaxies. Chen \& Halpern identified the 25 \micron\ peak in the SED
3212: with the turnover frequency of the synchrotron peak from the RIAF, whose
3213: elevated structure illuminates an outer thin disk that emits the double-peaked
3214: broad optical lines. The overall weakness of the UV continuum in Arp~102B
3215: (Halpern et al. 1996) further corroborates a truncated thin disk structure
3216: and potentially provides an explanation for the low-ionization state of the
3217: emission-line spectrum. As for the jet component, it was assumed to be
3218: present, at least implicitly, insofar as the double-peaked broad-line AGNs
3219: were thought to reside preferentially in radio-loud AGNs, and the very concept
3220: of ion-supported tori was invented with reference to radio galaxies (Rees et
3221: al. 1982).
3222:
3223: Recent developments add important refinements and modifications to Chen \&
3224: Halpern's original picture. First, the mid-IR peak in most objects is
3225: dominated by thermal emission from the truncated thin disk rather than by the
3226: synchrotron peak of the RIAF. Second, the jet component, which was once
3227: regarded as somewhat incidental, has emerged as a natural and perhaps inevitable
3228: outgrowth of the inner accretion flow itself. Third, although the original
3229: model was invented to explain a small minority of the AGN population
3230: (double-peaked radio-loud sources), now we have good reason to believe that
3231: similar physical conditions prevail in LINERs as a class (Ho et al. 2000),
3232: and, by extension, in the majority of nearby accreting BHs.
3233:
3234: The physical picture outlined above for LLAGNs shares strong similarities with
3235: that developed for X-ray binaries in their hard state (see Maccarone, Fender
3236: \& Ho 2005 and references therein), suggesting that the basic architecture
3237: of the central engine around accreting BHs---across 10 orders of magnitude in
3238: mass---is essentially scale-invariant (Meier 2001; Maccarone, Gallo \&
3239: Fender 2003; Merloni, Heinz \& Di~Matteo 2003; Falcke, K\"ording \& Markoff
3240: 2004; Ho 2005; K\"ording, Jester \& Fender 2006).
3241:
3242: \section{CONCLUDING REMARKS}
3243:
3244: The topics covered in this article are both very old and very new. It has been
3245: known for over three decades that a large segment of the galaxy population
3246: exhibits signs of unusual activity in their nuclei, and for nearly as long
3247: people have puzzled over the physical origin of this activity. Over time the
3248: observational material at optical wavelengths has improved markedly,
3249: especially with the completion of the Palomar survey, but the debate has only
3250: intensified. Given their abundance, most of the controversy has centered, not
3251: surprisingly, on LINERs. While the nonstellar nature of a sizable fraction of
3252: LINERs is now incontrovertible (e.g., those with broad H\al\ emission), the
3253: AGN content in the majority of the class remains unsettled. Determining the
3254: physical origin of these systems is more than of mere phenomenological
3255: interest. Because LINERs are so numerous---being the dominant constituent of
3256: the local LLAGN population and a sizable fraction of all galaxies---they have
3257: repercussions on virtually every issue related to AGN and BH demographics.
3258:
3259: The main source of contention stems from the fact that Mother Nature knows too
3260: many ways of generating nebular conditions that qualitatively look similar to
3261: the low-ionization characteristics of LINERs at optical wavelengths. A
3262: dizzying array of excitation mechanisms has been proposed to explain LINERs,
3263: ranging from variants of conventional AGN photoionization, to shocks of
3264: various flavors, to interstellar processes such as cooling flows and turbulent
3265: mixing layers, to stellar-based photoionization by populations both young and
3266: old, prosaic and exotic. This field has suffered not from a shortage of
3267: ideas, but from too many. As a consequence, whenever LINERs are discussed, it
3268: is customary to end with the pessimistic mantra that they are a mixed-bag,
3269: heterogeneous class of objects, a statement with somewhat dismissive
3270: connotations that is often taken to mean that we have no idea what they are
3271: and that they are too messy to deal with. This is an unfair characterization
3272: of the progress that has been made, and I think that there is good reason to
3273: sound a more positive note.
3274:
3275: As summarized in this review, a number of developments during the last few
3276: years shed considerable light on the physical origin of LLAGNs in general and
3277: LINERs in particular. The key advances have come from the broader
3278: perspective afforded by observations outside of the traditional optical
3279: window, especially in the radio and X-rays, although important
3280: insights can also be credited to optical and UV data taken with \hst. In all
3281: instances, high angular resolution has been a critical factor to disentangle
3282: the weak nuclear emission from the blinding host galaxy background.
3283:
3284: Other developments have been instrumental in forging a coherent view of nuclear
3285: activity in the nearby Universe. On the theoretical side, rapid advances
3286: in the study of radiatively inefficient accretion flows, originally primarily
3287: motivated by applications to X-ray binaries and to the Galactic Center source
3288: Sgr~A$^*$, has led to a growing appreciation that they are also relevant to
3289: LLAGNs in general. Many investigators have sharpened the physical analogy
3290: between the spectral states of X-ray binaries and certain classes of AGNs, an
3291: effort that has resulted in a more holistic picture of BH accretion,
3292: especially as it concerns the evolution of the accretion flow in response to
3293: variations in mass accretion rate and the mechanism for generating jets or
3294: outflows. Meanwhile, the dynamical detection of supermassive BHs, their
3295: ubiquity, and the discovery of scaling relations between BHs and their host
3296: galaxies have given a major boom to studies of AGNs in all their multi-faceted
3297: manifestations. More than ever, in the grand scheme of things, AGNs are no
3298: longer viewed as rare and exotic oddities but as natural episodes during the
3299: life cycle of galaxies during which their BHs accrete, grow, and shine. The
3300: impact of BH growth and AGN feedback have emerged forcefully as major new
3301: themes in galaxy formation. LLAGNs gain an even greater prominence within
3302: this context. Although the bulk of the mass density of BHs was accreted in a
3303: luminous, radiatively efficient mode, it behooves us to understand how BHs
3304: spend most of their lives. The detection of supermassive BHs has also
3305: fundamentally altered the character of the discourse on LLAGNs. We can now
3306: shift our attention from the question of {\it whether}\ LLAGNs contain
3307: BHs---an implicit or explicit motivation for much of the past discussion on
3308: the nature of these sources---to {\it why}\ these BHs have the properties that
3309: they do. Among other things, LLAGNs can be used as an effective platform for
3310: exploring accretion physics in highly sub-Eddington systems and for
3311: investigating physical processes in the circumnuclear regions of galaxies that
3312: are normally masked by brighter nuclei.
3313:
3314: The following is a list of the ``top ten'' results from this paper.
3315:
3316: \begin{enumerate}
3317:
3318: \item{Approximately 2/3 of local E--Sb galaxies exhibit weak nuclear activity
3319: incompatible with normal stellar processes; in contrast, only about 15\% of
3320: Sc--Sm galaxies show AGN activity (\S 3). }
3321:
3322: \item{The vast majority of LINERs, and, by implication, most nearby weakly
3323: active nuclei, are genuine, accretion-powered AGNs (\S\S 6.1, 6.5).}
3324:
3325: \item{The ubiquity of LLAGNs in galaxies with bulges strongly supports the
3326: current paradigm derived from dynamical studies that all bulges contain BHs.
3327: However, the detection of AGNs in some bulgeless, even dwarf, galaxies proves
3328: that bulges are not necessary for the formation of central BHs (\S 7).}
3329:
3330: \item{The luminosity function of nearby LLAGNs follows $\Phi \propto
3331: L^{-1.2\pm0.2}$ from $L_{{\rm H}\alpha} \approx 3\times 10^{41}$ to $10^{38}$
3332: \lum, below which it appears to flatten down to $L_{{\rm H}\alpha} \approx 6
3333: \times 10^{36}$ \lum\ or $M_B \approx -8$ mag (\S 5.9).}
3334:
3335: \item{Stellar photoionization by young or intermediate-stars and shock heating
3336: can be ruled out as the excitation mechanisms for LLAGNs (\S\S 6.2, 6.3).}
3337:
3338: \item{Despite the overall success of AGN photoionization models, many LLAGNs,
3339: especially type~2 sources, have a shortage of ionizing photons. The energy
3340: deficit problem could be solved with cosmic ray heating and extra ionization
3341: from evolved (post-AGB) stars, diffuse thermal plasma, and the cumulative
3342: X-ray emission from low-mass X-ray binaries (\S 6.4).}
3343:
3344: \item{Variations in the mass accretion rate give rise to the different classes
3345: of emission-line nuclei. LINERs are the low-luminosity, low-accretion rate
3346: extension of Seyferts, followed by transition nuclei, and ending with
3347: absorption-line nuclei at the end of the BH starvation sequence (\S 6.5).}
3348:
3349: \item{LLAGNs are not simply scaled-down versions of powerful AGNs. Their
3350: central engines undergo fundamental changes when the accretion rate drops to
3351: extremely sub-Eddington values. In this regime, the BLR and obscuring torus
3352: disappear (\S\S 5.5, 5.6). LLAGNs do not follow the standard AGN unification
3353: model.}
3354:
3355: \item{Below a characteristic luminosity of $\sim$1\% Eddington, the
3356: canonical optically thick, geometrically thin accretion disk transforms
3357: into a three-component structure consisting of an inner vertically thick and
3358: radiatively inefficient accretion flow, a truncated outer thin disk, and a jet
3359: or outflow (\S\S 5.8, 8.3).}
3360:
3361: \item{At the lowest accretion rates, an increasing fraction of the accretion
3362: energy gets channeled into a relativistic jet. The emitted energy is mainly
3363: kinetic rather than radiative. Since radiation and kinetic jets interact
3364: differently with the surrounding gas, this has important implications for
3365: AGN feedback into galaxy formation (\S 8.2).}
3366:
3367: \end{enumerate}
3368:
3369: \vskip 0.2cm
3370: My research is supported by the Carnegie Institution of Washington and by NASA
3371: grants from the Space Telescope Science Institute (operated by AURA, Inc.,
3372: under NASA contract NAS5-26555). I would like to recognize my collaborators
3373: who have contributed to the work covered in this review, especially A.J. Barth,
3374: M. Eracleous, M.E. Filho, A.V. Filippenko, J.E. Greene, D. Maoz, E.C. Moran,
3375: C.Y. Peng, A. Ptak, E. Quataert, H.-W. Rix, W.L.W. Sargent, M. Sarzi, J.C.
3376: Shields, Y. Terashima, J.S. Ulvestad, and J.M. Wrobel. Several of them
3377: (A.J. Barth, M. Eracleous, J.E. Greene, M. Sarzi, J.C. Shields, Y. Terashima)
3378: read an early draft of the manuscript and provided useful feedback that helped
3379: to improve it. Some of the concepts expressed in the review were sharpened
3380: after correspondence with M. Eracleous, G. Ferland, J.C. Shields, and Y.
3381: Terashima. I thank A.J. Barth, A.V. Filippenko, and W.L.W.
3382: Sargent for permission to cite material in advance of publication, H.M.L.G.
3383: Flohic and M. Eracleous for providing the images for Figure 5, and Salvador Ho
3384: for drafting Figure 13. I am grateful to J. Kormendy for his steadfast
3385: encouragement, wise counsel, and meticulous editing.
3386:
3387: \vskip 0.5cm
3388:
3389: %\begin{thebibliography}{99}
3390:
3391: \noindent {LITERATURE CITED}
3392:
3393: \frenchspacing
3394:
3395: \vskip 0.5cm
3396:
3397: \nhi
3398: Aldrovandi SMV, P\'equignot D. 1973. {\it Astron. Astrophys.} 26:33
3399:
3400: \nhi
3401: Allen MG, Dopita, MA, Tsvetanov ZI. 1998. {\it Ap. J.} 493:571
3402:
3403: \nhi
3404: Alonso-Herrero A, Rieke MJ, Rieke GH, Ruiz M. 1997. {\it Ap. J.} 482:74
3405:
3406: \nhi
3407: Alonso-Herrero A, Rieke MJ, Rieke GH, Shields JC. 2000. {\it Ap. J.} 530:688
3408:
3409: \nhi
3410: Anderson JM, Ulvestad JS. 2005. {\it Ap. J.} 627:674
3411:
3412: \nhi
3413: Anderson JM, Ulvestad JS, Ho LC. 2004. {\it Ap. J.} 603:42
3414:
3415: \nhi
3416: Antonucci R. 1993. {\it Annu. Rev. Astron. Astrophys.} 31:473
3417:
3418: \nhi
3419: Armus L, Heckman TM, Miley GK. 1990. {\it Ap. J.} 364:471
3420:
3421: \nhi
3422: Atkinson JW, Collett JL, Marconi A, Axon DJ, Alonso-Herrero A, et al. 2005.
3423: {\it MNRAS} 359:504
3424:
3425: \nhi
3426: Avni Y, Tananbaum H. 1982. {\it Ap. J.} 262:L17
3427:
3428: \nhi
3429: Awaki H, Mushotzky R, Tsuru T, Fabian AC, Fukazawa Y, et al. 1994. {\it PASJ}
3430: 46:L65
3431:
3432: \nhi
3433: Baganoff FK, Maeda Y, Morris M, Bautz MW, Brandt WN, et al. 2003. {\it Ap. J.}
3434: 591:891
3435:
3436: \nhi
3437: Baldwin JA, Phillips MM, Terlevich R. 1981. {\it PASP} 93:5
3438:
3439: \nhi
3440: Ballantyne DR. 2007. {\it Mod. Phys. Lett. A} 22:2397
3441:
3442: \nhi
3443: Ballantyne DR, Fabian AC. 2005. {\it Ap. J.} 622:L97
3444:
3445: \nhi
3446: Ballantyne DR, Ross RR, Fabian AC. 2002. {\it MNRAS} 332:L45
3447:
3448: \nhi
3449: Balmaverde B, Capetti A. 2006. {\it Astron. Astrophys.} 447:97
3450:
3451: \nhi
3452: Balmaverde B, Capetti A, Grandi P. 2006. {\it Astron. Astrophys.} 451:35
3453:
3454: \nhi
3455: Barger AJ, Cowie LL, Mushotzky RF, Richards EA. 2001. {\it Astron. J.} 121:662
3456:
3457: \nhi
3458: Barth AJ. 2004. In {\it Carnegie Observatories Astrophysics Series, Vol 1:
3459: Coevolution of Black Holes and Galaxies}, ed. LC Ho, p. 21. Cambridge:
3460: Cambridge Univ. Press
3461:
3462:
3463: \nhi
3464: Barth AJ, Filippenko AV, Moran EC. 1999a. {\it Ap. J.} 515:L61
3465:
3466: \nhi
3467: Barth AJ, Filippenko AV, Moran EC. 1999b. {\it Ap. J.} 525:673
3468:
3469: \nhi
3470: Barth AJ, Greene JE, Ho LC. 2005. {\it Ap. J.} 619:L151
3471:
3472: \nhi
3473: Barth AJ, Ho LC, Filippenko AV. 2003. In {\it Active Galactic Nuclei: from
3474: Central Engine to Host Galaxy}, ed. S Collin, F Combes, I Shlosman, p. 387.
3475: San Francisco: ASP
3476:
3477: \nhi
3478: Barth AJ, Ho LC, Filippenko AV, Rix H-W, Sargent WLW 2001a. {\it Ap. J.} 546:205
3479:
3480: \nhi
3481: Barth AJ, Ho LC, Filippenko AV, Sargent WLW. 1998. {\it Ap. J.} 496:133
3482:
3483: \nhi
3484: Barth AJ, Ho LC, Rutledge RE, Sargent WLW. 2004. {\it Ap. J.} 607:90
3485:
3486: \nhi
3487: Barth AJ, Reichert GA, Filippenko AV, Ho LC, Shields JC, Mushotzky RF,
3488: Puchnarewicz EM. 1996. {\it Astron. J.} 112:1829
3489:
3490: \nhi
3491: Barth AJ, Reichert GA, Ho LC, Shields JC, Filippenko AV, Puchnarewicz EM.
3492: 1997. {\it Astron. J.} 114:2313
3493:
3494: \nhi
3495: Barth AJ, Sarzi M, Rix H-W, Ho LC, Filippenko AV, Sargent WLW. 2001b.
3496: {\it Ap. J.} 555:685
3497:
3498: \nhi
3499: Barth AJ, Shields JC. 2000. {\it PASP} 112:753
3500:
3501: \nhi
3502: Barth AJ, Tran H, Brotherton MS, Filippenko AV, Ho LC, et al. 1999.
3503: {\it Astron. J.} 118:1609
3504:
3505: \nhi
3506: Bassani L, Dadina M, Maiolino R, Salvati M, Risaliti G, et al.
3507: 1999. {\it Ap.~J.~Suppl.} 121:473
3508:
3509: \nhi
3510: Begelman MC, Fabian AC. 1990. {\it MNRAS} 244:36P
3511:
3512: \nhi
3513: Bendo GJ, Dale DA, Draine BT, Engelbracht CW, Kennicutt RC Jr, et al. 2006.
3514: {\it Ap. J.} 645:134
3515:
3516: \nhi
3517: Bendo GJ, Joseph RD. 2004. {\it Astron. J.} 127:3338
3518:
3519: \nhi
3520: Bennert N, Jungwiert B, Komossa S, Haas M, Chini R. 2006.
3521: {\it Astron. Astrophys.} 456:953
3522:
3523: \nhi
3524: Bianchi S, Corral A, Panessa F, Barcons X, Matt G, et al. 2008. {\it MNRAS}
3525: 385:195
3526:
3527: \nhi
3528: Bietenholz MF, Bartel N, Rupen NP. 2000. {\it Ap. J.} 532:895
3529:
3530: \nhi
3531: Binette L. 1985. {\it Astron. Astrophys.} 143:334
3532:
3533: \nhi
3534: Binette L, Magris CG, Stasi\'nska G, Bruzual AG. 1994.
3535: {\it Astron. Astrophys.} 292:13
3536:
3537: \nhi
3538: Binney J, Tabor G. 1995. {\it MNRAS} 276:663
3539:
3540: \nhi
3541: Blandford RD, Begelman MC. 1999. {\it MNRAS} 303:L1
3542:
3543: \nhi
3544: Blandford RD, K\"onigl A. 1979. {\it Ap. J.} 232:34
3545:
3546: \nhi
3547: Blanton EL, Sarazin CL, Irwin JA. 2001. {\it Ap. J.} 552:106
3548:
3549: \nhi
3550: B\"ohringer H, Belsole E, Kennea J, Matsushita K, Molendi S, et al. 2001.
3551: {\it Astron. Astrophys.} 365::181
3552:
3553: \nhi
3554: Boisson C, Joly M, Moultaka J, Pelat D, Serote-Roos M. 2000.
3555: {\it Astron. Astrophys.} 357:850
3556:
3557: \nhi
3558: B\"{o}ker T, van~der~Marel RP, Laine S, Rix H-W, Sarzi M, Ho LC, Shields JC.
3559: 2002. {\it Astron. J.} 123:1389
3560:
3561: \nhi
3562: Bonatto C, Bica E, Alloin D. 1989. {\it Astron. Astrophys.} 226:23
3563:
3564: \nhi
3565: Bondi H. 1952. {\it MNRAS} 112:195
3566:
3567: \nhi
3568: Bower GA, Green RF, Quillen AC, Danks A, Gull T, et al. 2000. {\it Ap. J.}
3569: 534:189
3570:
3571: \nhi
3572: Bower GA, Wilson AS, Heckman TM, Richstone DO. 1996. {\it Astron. J.} 111:1901
3573:
3574: \nhi
3575: Burbidge EM, Burbidge G. 1962. {\it Ap. J.} 135:694
3576:
3577: \nhi
3578: Burbidge EM, Burbidge G. 1965. {\it Ap. J.} 142:634
3579:
3580: \nhi
3581: Burbidge G, Gould RJ, Pottasch SR. 1963. {\it Ap. J.} 138:945
3582:
3583: \nhi
3584: Canosa CM, Worrall DM, Hardcastle MJ, Birkinshaw M. 1999. {\it MNRAS} 310:30
3585:
3586: \nhi
3587: Capetti A, Axon DJ, Chiaberge M, Sparks WB, Macchetto FD, et al.
3588: 2007. {\it Astron. Astrophys.} 471:137
3589:
3590: \nhi
3591: Capetti A, Celotti A, Chiaberge M, de Ruiter HR, Fanti R, et al.
3592: 2002. {\it Astron. Astrophys.} 383:104
3593:
3594: \nhi
3595: Capetti A, Trussoni E, Celotti A, Feretti L, Chiaberge M. 2000. {\it MNRAS}
3596: 318:493
3597:
3598: \nhi
3599: Capetti A, Verdoes Kleijn G, Chiaberge M. 2005. {\it Astron. Astrophys.}
3600: 439:935
3601:
3602: \nhi
3603: Cappellari M, Bertola F, Burstein D, Buson LM, Greggio L, Renzini A. 2001.
3604: {\it Ap. J.} 551:197
3605:
3606: \nhi
3607: Cappi M, Panessa F, Bassani L, Dadina M, Dicocco G, et al. 2006.
3608: {\it Astron. Astrophys.} 446:459
3609:
3610: \nhi
3611: Carollo CM, Stiavelli M, de Zeeuw PT, Mack, J. 1997. {\it Astron. J.} 114:2366
3612:
3613: \nhi
3614: Carrillo R, Masegosa J, Dultzin-Hacyan D, Ordo\~{n}ez R. 1999. {\it Rev. Mex.
3615: Astron. Astrof.} 35:187
3616:
3617: \nhi
3618: Carswell RF, Baldwin JA, Atwood B, Phillips MM. 1984. {\it Ap. J.} 286:464
3619:
3620: \nhi
3621: Chary R, Becklin EE, Evans AS, Neugebauer G, Scoville NZ, et al.
3622: 2000. {\it Ap. J.} 531:756
3623:
3624: \nhi
3625: Chen K, Halpern JP. 1989. {\it Ap. J.} 344:115
3626:
3627: \nhi
3628: Chiaberge M, Capetti A, Celotti A. 1999. {\it Astron. Astrophys.} 349:77
3629:
3630: \nhi
3631: Chiaberge M, Capetti A, Celotti A. 2000. {\it Astron. Astrophys.} 355:873
3632:
3633: \nhi
3634: Chiaberge M, Capetti A, Celotti A. 2002. {\it Astron. Astrophys.} 394:791
3635:
3636: \nhi
3637: Chiaberge M, Capetti A, Macchetto FD. 2005. {\it Ap. J.} 625:716
3638:
3639: \nhi
3640: Chiaberge M, Gilli R, Capetti A, Macchetto FD. 2003. {\it Ap. J.} 597:166
3641:
3642: \nhi
3643: Chiaberge M, Macchetto FD, Sparks W, Capetti A, Allen MG, Martel AR. 2002.
3644: {\it Ap. J.} 571:247
3645:
3646: \nhi
3647: Cid Fernandes R Jr, Golz\'alez Delgado RM, Schmitt H, Storchi-Bergmann T, Martins LP, et al. 2004. {\it Ap. J.} 605:105
3648:
3649: \nhi
3650: Ciotti L, D'Ercole A, Pellegrini S, Renzini A. 1991. {\it Ap. J.} 376:380
3651:
3652: \nhi
3653: Ciotti L, Ostriker JP. 2001. {\it Ap. J.} 551:131
3654:
3655: \nhi
3656: Colbert EJM, Mushotzky RF. 1999. {\it Ap. J.} 519:89
3657:
3658: \nhi
3659: Colina L, Golz\'alez Delgado RM, Mas-Hesse JM, Leitherer C. 2002. {\it Ap. J.}
3660: 579:545
3661:
3662: \nhi
3663: Collins JA, Rand RJ. 2001. {\it Ap. J.} 551:57
3664:
3665: \nhi
3666: Combes F. 2003. In {\it Active Galactic Nuclei: from Central Engine to Host
3667: Galaxy}, ed. S Collin, F Combes, I Shlosman, p. 411. San Francisco: ASP
3668:
3669: \nhi
3670: Constantin A, Vogeley MS. 2006. {\it Ap. J.} 650:727
3671:
3672: \nhi
3673: Contini M. 1997. {\it Astron. Astrophys.} 323:71
3674:
3675: \nhi
3676: Corbett EA, Kewley L, Appleton PN, Charmandaris V, Dopita MA, et al.
3677: 2003. {\it Ap. J.} 583:670
3678:
3679: \nhi
3680: Costero R, Osterbrock DE. 1977. {\it Ap. J.} 211:675
3681:
3682: \nhi
3683: Dale DA, Smith JDT, Armus L, Buckalew BA, Helou G, et al. 2006. {\it Ap. J.}
3684: 646:161
3685:
3686: \nhi
3687: Danziger IJ, Fosbury RAE, Penston MV. 1977. {\it MNRAS} 179:41P
3688:
3689: \nhi
3690: Decarli R, Gavazzi G, Arosio I, Cortese L, Boselli A, et al. 2007. {\it MNRAS}
3691: 381:136
3692:
3693: \nhi
3694: Demoulin-Ulrich M-H, Butcher HR, Boksenberg A. 1984. {\it Ap. J.} 285:527
3695:
3696: \nhi
3697: Desroches L-B, Ho LC. 2008. {\it Ap. J.}, submitted
3698:
3699: \nhi
3700: Dewangan GC, Griffiths RE, Di Matteo T, Schurch NJ. 2004. {\it Ap. J.} 607:788
3701:
3702: \nhi
3703: D\'\i az AI, Pagel BEJ, Wilson IRG. 1985. {\it MNRAS} 212:737
3704:
3705: \nhi
3706: D\'\i az AI, Pagel BEJ, Terlevich E. 1985. {\it MNRAS} 214:41P
3707:
3708: \nhi
3709: Di Matteo T, Allen SW, Fabian AC, Wilson AS, Young AJ. 2003. {\it Ap. J.}
3710: 582:133
3711:
3712: \nhi
3713: Di Matteo T, Fabian AC. 1997. {\it MNRAS} 286:L50
3714:
3715: \nhi
3716: Di Matteo T, Johnstone RM, Allen SW, Fabian AC. 2001. {\it Ap. J.} 550:L19
3717:
3718: \nhi
3719: Disney MJ, Cromwell RH. 1971. {\it Ap. J.} 164:L35
3720:
3721: \nhi
3722: Doi A, Kameno S, Kohno K, Nakanishi K, Inoue M. 2005. {\it MNRAS} 363:692
3723:
3724: \nhi
3725: Donato D, Sambruna RM, Gliozzi M. 2004. {\it Ap. J.} 617:915
3726:
3727: \nhi
3728: Done C., Gierli\'nski M, Sobolewska M, Schurch N. 2007. In {\it The Central
3729: Engine of Active Galactic Nuclei}, ed. LC Ho, J-M Wang, p. 121. San
3730: Francisco: ASP
3731:
3732: \nhi
3733: Dong X-B, Wang T, Yuan W, Shan H, Zhou H, et al. 2007. {\it Ap. J.} 657:700
3734:
3735: \nhi
3736: Dopita MA, Koratkar AP, Allen MG, Tsvetanov ZI, Ford HC, et al.
3737: 1997. {\it Ap. J.} 490:202
3738:
3739: \nhi
3740: Dopita MA, Sutherland RS. 1995. {\it Ap. J.} 455:468
3741:
3742: \nhi
3743: Dudik RP, Satyapal S, Gliozzi M, Sambruna RM. 2005. {\it Ap. J.} 620:113
3744:
3745: \nhi
3746: Elitzur M, Shlosman I. 2006. {\it Ap. J.} 648:L101
3747:
3748: \nhi
3749: Elvis M, Wilkes BJ, McDowell JC, Green RF, Bechtold J, et al. 1994.
3750: {\it Ap.~J.~Suppl.} 95:1
3751:
3752: \nhi
3753: Engelbracht CW, Rieke MJ, Rieke GH, Kelley DM, Achtermann JM. 1998.
3754: {\it Ap. J.} 505:639
3755:
3756: \nhi
3757: Eracleous M, Halpern JP. 1994. {\it Ap.~J.~Suppl.} 90:1
3758:
3759: \nhi
3760: Eracleous M, Halpern JP. 2001. {\it Ap.~J.} 554:240
3761:
3762: \nhi
3763: Eracleous M, Hwang JA, Flohic HMLG. 2008a. {\it Ap.~J.} submitted
3764:
3765: \nhi
3766: Eracleous M, Hwang JA, Flohic HMLG. 2008b. {\it Ap.~J.} submitted
3767:
3768: \nhi
3769: Eracleous M, Sambruna R, Mushotzky RF. 2000. {\it Ap. J.} 537:654
3770:
3771: \nhi
3772: Eracleous M, Shields JC, Chartas G, Moran EC. 2002. {\it Ap.~J.} 565:108
3773:
3774: \nhi
3775: Evans DA, Worrall DM, Hardcastle MJ, Kraft RP, Birkinshaw M. 2006.
3776: {\it Ap.~J.} 642:96
3777:
3778: \nhi
3779: Fabbiano G. 2006. {\it Annu. Rev. Astron. Astrophys.} 44:323
3780:
3781: \nhi
3782: Fabbiano G, Baldi A, Pellegrini S, Siemiginowska A, Elvis M, et al.
3783: 2004. {\it Ap.~J.} 616:730
3784:
3785: \nhi
3786: Fabbiano G, Elvis M, Markoff S, Siemiginowska A, Pellegrini S, et al.
3787: 2003. {\it Ap.~J.} 588:175
3788:
3789: \nhi
3790: Fabbiano G, Trinchieri G. 1985. {\it Ap. J.} 296:430
3791:
3792: \nhi
3793: Faber SM, Tremaine S, Ajhar EA, Byun Y-I, Dressler A, et al. 1997.
3794: {\it Astron. J.} 114:1365
3795:
3796: \nhi
3797: Fabian AC, Arnaud KA, Nulsen PEJ, Mushotzky RF. 1986. {\it Ap. J.} 305:9
3798:
3799: \nhi
3800: Fabian AC, Canizares CR. 1988. {\it Nature} 333:829
3801:
3802: \nhi
3803: Fabian AC, Rees MJ. 1995. {\it MNRAS} 277:L55
3804:
3805: \nhi
3806: Falcke H, K\"ording E, Markoff S. 2004. {\it Astron. Astrophys.} 414:895
3807:
3808: \nhi
3809: Falcke H, Markoff S. 2000. {\it Astron. Astrophys.} 362:113
3810:
3811: \nhi
3812: Falcke H, Nagar NM, Wilson AS, Ulvestad JS. 2000. {\it Ap. J.} 542:197
3813:
3814: \nhi
3815: Fanaroff BL, Riley JM. 1974. {\it MNRAS} 167:31P
3816:
3817: \nhi
3818: Ferrarese L, Merritt D. 2000. {\it Ap. J.} 539:L9
3819:
3820: \nhi
3821: Ferland GJ, Mushotzky RF. 1984. {\it Ap. J.} 286:42
3822:
3823: \nhi
3824: Ferland GJ, Netzer H. 1983. {\it Ap. J.} 264:105
3825:
3826: \nhi
3827: Filho ME, Barthel PD, Ho LC. 2000. {\it Ap.~J.~Suppl.} 129:93
3828:
3829: \nhi
3830: Filho ME, Barthel PD, Ho LC. 2002a. {\it Ap.~J.~Suppl.} 142:223
3831:
3832: \nhi
3833: Filho ME, Barthel PD, Ho LC. 2002b. {\it Astron. Astrophys.} 385:425
3834:
3835: \nhi
3836: Filho ME, Barthel PD, Ho LC. 2006. {\it Astron. Astrophys.} 451:71
3837:
3838: \nhi
3839: Filho ME, Fraternali F, Nagar NM, Barthel PD, Markoff S, et al. 2004.
3840: {\it Astron. Astrophys.} 418:429
3841:
3842: \nhi
3843: Filippenko AV. 1985. {\it Ap. J.} 289:475
3844:
3845: \nhi
3846: Filippenko AV. 1996. In {\it The Physics of LINERs in View of Recent
3847: Observations}, ed. M Eracleous, A Koratkar, C Leitherer, LC Ho, p. 17.
3848: San Francisco: ASP
3849:
3850: \nhi
3851: Filippenko AV, Halpern JP. 1984. {\it Ap. J.} 285:458
3852:
3853: \nhi
3854: Filippenko AV, Ho LC. 2003. {\it Ap. J.} 588:L13
3855:
3856: \nhi
3857: Filippenko AV, Ho LC, Sargent WLW. 1993. {\it Ap. J.} 410:L75
3858:
3859: \nhi
3860: Filippenko AV, Sargent WLW. 1985. {\it Ap.~J.~Suppl.} 57:503
3861:
3862: \nhi
3863: Filippenko AV, Sargent WLW. 1988. {\it Ap. J.} 324:134
3864:
3865: \nhi
3866: Filippenko AV, Sargent WLW. 1989. {\it Ap. J.} 342:L11
3867:
3868: \nhi
3869: Filippenko AV, Terlevich R. 1992. {\it Ap. J.} 397:L79
3870:
3871: \nhi
3872: Fiore F, ed. 2006. {\it AGNs and Galaxy Evolution}. {\it MmSAI} 77
3873:
3874: \nhi
3875: Fiore F, Pellegrini S, Matt G, Antonelli LA, Comastri A, et al. 2001.
3876: {\it Ap. J.} 556:150
3877:
3878: \nhi
3879: Flohic HMLG, Eracleous M, Chartas G, Shields JC, Moran EC. 2006. {\it Ap. J.}
3880: 647:140
3881:
3882: \nhi
3883: Ford HC, Butcher H. 1979. {\it Ap.~J.~Suppl.} 41:147
3884:
3885: \nhi
3886: Fosbury RAE, Melbold U, Goss WM, Dopita MA. 1978. {\it MNRAS} 183:549
3887:
3888: \nhi
3889: Fosbury RAE, Melbold U, Goss WM, van Woerden H. 1977. {\it MNRAS} 179:89
3890:
3891: \nhi
3892: Fukazawa Y, Iyomoto N, Kubota A, Matsumoto Y, Makishima K. 2001.
3893: {\it Astron. Astrophys.} 374:73
3894:
3895: \nhi
3896: Gabel JR, Bruhweiler FC. 2002. {\it Astron. J.} 124:737
3897:
3898: \nhi
3899: Gabel JR, Bruhweiler FC, Crenshaw DM, Kraemer SB, Miskey CL. 2000.
3900: {\it Ap. J.} 532:883
3901:
3902: \nhi
3903: Gallimore JF, Axon DJ, O'Dea CP, Baum SA, Pedlar A. 2006. {\it Astron. J.}
3904: 132:546
3905:
3906: \nhi
3907: Gallo E, Treu T, Jacob J, Woo J-H, Marshall RJ, Antonucci R. 2008. {\it Ap. J.}
3908: in press
3909:
3910: \nhi
3911: Gammie CF, Narayan R, Blandford RD. 1999. {\it Ap. J.} 516:177
3912:
3913: \nhi
3914: Garcia MR, Williams BF, Yuan F, Kong AKH, Primini FA, et al.
3915: 2005. {\it Ap. J.} 632:1042
3916:
3917: \nhi
3918: Gebhardt K, Bender R, Bower G, Dressler A, Faber SM, et al. 2000.
3919: {\it Ap. J.} 539:L13
3920:
3921: \nhi
3922: Gebhardt K, Rich RM, Ho LC. 2002. {\it Ap. J.} 578:L41
3923:
3924: \nhi
3925: Gebhardt K, Rich RM, Ho LC. 2005. {\it Ap. J.} 634:1093
3926:
3927: \nhi
3928: Geha M, Guhathakurta P, van~der~Marel RP. 2002. {\it Astron. J.} 124:3073
3929:
3930: \nhi
3931: Georgantopoulos I, Panessa F, Akylas A, Zezas A, Cappi M, Comastri A. 2002.
3932: {\it Astron. Astrophys.} 386:60
3933:
3934: \nhi
3935: George IM, Fabian AC. 1991. {\it MNRAS} 249:352
3936:
3937: \nhi
3938: Ghosh H, Pogge RW, Mathur S, Martini P, Shields JC. 2007. {\it Ap. J.} 656:105
3939:
3940: \nhi
3941: Giroletti M, Giovannini G, Taylor GB, Falomo R. 2006. {\it Ap. J.} 646:801
3942:
3943: \nhi
3944: Gliozzi M, Foschini L, Sambruna RM, Tavecchio F. 2008.
3945: {\it Astron. Astrophys.} 478:723
3946:
3947: \nhi
3948: Gliozzi M, Sambruna RM, Brandt WN, Mushotzky RF, Eracleous M. 2004.
3949: {\it Astron. Astrophys.} 413:139
3950:
3951: \nhi
3952: Gliozzi M, Sambruna RM, Foschini L. 2007. {\it Ap. J.} 662:878
3953:
3954: \nhi
3955: Gonz\'alez Delgado RM, Cid Fernandes R, Per\'ez E, Martins LP,
3956: Storchi-Bergmann T, et al. 2004. {\it Ap. J.} 605:127
3957:
3958: \nhi
3959: Gonz\'alez-Mart\'\i n O, Masegosa J, M\'arquez I, Guerrero MA, Dultzin-Hacyan
3960: D. 2006. {\it Astron. Astrophys.} 460:45
3961:
3962: \nhi
3963: Granato GL, De Zotti G, Silva L, Bressan A, Danese L. 2004. {\it Ap. J.} 600:580
3964:
3965: \nhi
3966: Grandi SA, Osterbrock DE. 1978. {\it Ap. J.} 220:783
3967:
3968: \nhi
3969: Greene JE, Ho LC. 2004. {\it Ap. J.} 610:722
3970:
3971: \nhi
3972: Greene JE, Ho LC. 2005a. {\it Ap. J.} 627:721
3973:
3974: \nhi
3975: Greene JE, Ho LC. 2005b. {\it Ap. J.} 630:122
3976:
3977: \nhi
3978: Greene JE, Ho LC. 2006. {\it Ap. J.} 641:L21
3979:
3980: \nhi
3981: Greene JE, Ho LC. 2007a. {\it Ap. J.} 667:131
3982:
3983: \nhi
3984: Greene JE, Ho LC. 2007b. {\it Ap. J.} 670:92
3985:
3986: \nhi
3987: Greene JE, Ho LC, Barth AJ. 2008. {\it Ap. J.} submitted
3988:
3989: \nhi
3990: Greene JE, Ho LC, Ulvestad JS. 2006. {\it Ap. J.} 636:56
3991:
3992: \nhi
3993: Gronwall C, Jangren A, Salzer JJ, Werk JK, Ciardullo R. 2004. {\it Astron. J.}
3994: 128:644
3995:
3996: \nhi
3997: Grossan B, Gorjian V, Werner M, Ressler M. 2001. {\it Ap. J.} 563:687
3998:
3999: \nhi
4000: Groves BA, Dopita MA, Sutherland RS. 2004. {\it Ap.~J.~Suppl.} 153:75
4001:
4002: \nhi
4003: Groves B, Heckman T, Kauffmann G. 2006. {\it MNRAS} 371:1559
4004:
4005: \nhi
4006: Gruenwald RB, Viegas-Aldrovandi SM. 1987. {\it Astron. Astrophys. Suppl.} 70:143
4007:
4008: \nhi
4009: Gu Q-S, Huang J-S, Wilson G, Fazio GG. 2007. {\it Ap.~J.} 671:L105
4010:
4011: \nhi
4012: Guainazzi M, Oosterbroek T, Antonelli LA, Matt G. 2000.
4013: {\it Astron. Astrophys.} 364:L80
4014:
4015: \nhi
4016: Haas M, M\"uller SAH, Bertoldi F, Chini R, Egner S, et al. 2004.
4017: {\it Astron. Astrophys.} 424:531
4018:
4019: \nhi
4020: Halderson EL, Moran EC, Filippenko AV, Ho LC. 2001. {\it Astron. J.} 122:637
4021:
4022: \nhi
4023: Hall PB, Yee HKC, Lin H, Morris SL, Patton DR, et al. 2000. {\it Astron. J.}
4024: 120:2220
4025:
4026: \nhi
4027: Halpern JP, Eracleous M. 1994. {\it Ap. J.} 433:L17
4028:
4029: \nhi
4030: Halpern JP, Eracleous M, Filippenko AV, Chen K. 1996. {\it Ap. J.} 464:704
4031:
4032: \nhi
4033: Halpern JP, Filippenko AV. 1984. {\it Ap. J.} 285:475
4034:
4035: \nhi
4036: Halpern JP, Steiner JE. 1983. {\it Ap. J.} 269:L37
4037:
4038: \nhi
4039: Hao L, Strauss MA, Fan XH, Tremonti CA, Schlegel DJ, et al. 2005a.
4040: {\it Astron. J.} 129:1783
4041:
4042: \nhi
4043: Hao L, Strauss MA, Tremonti CA, Schlegel DJ, Heckman TM, et al. 2005b.
4044: {\it Astron. J.} 129:1795
4045:
4046: \nhi
4047: Hawkins MRS. 2004. {\it Astron. Astrophys.} 424:519
4048:
4049: \nhi
4050: Heckman TM. 1980a. {\it Astron. Astrophys.} 87:142
4051:
4052: \nhi
4053: Heckman TM. 1980b. {\it Astron. Astrophys.} 87:152
4054:
4055: \nhi
4056: Heckman TM, Balick B, Crane PC. 1980. {\it Astron. Astrophys. Suppl. } 40:295
4057:
4058: \nhi
4059: Heckman TM, Baum SA, van Breugel WJM, McCarthy P. 1989. {\it Ap. J.} 338:48
4060:
4061: \nhi
4062: Heckman TM, Ptak A, Hornschemeier A, Kauffmann G. 2005. {\it Ap. J.} 634:161
4063:
4064: \nhi
4065: Heinz S, Merloni A, Schwab J. 2007. {\it Ap. J.} 658:L9
4066:
4067: \nhi
4068: Heinz S, Sunyaev R. 2003. {\it MNRAS} 343:L59
4069:
4070: \nhi
4071: Heller CH, Shlosman I. 1994. {\it Ap. J.} 424:84
4072:
4073: \nhi
4074: Hernquist L. 1989. {\it Nature}. 340;687
4075:
4076: \nhi
4077: Ho LC. 1996. In {\it The Physics of LINERs in View of Recent Observations},
4078: ed. M Eracleous, A Koratkar, C Leitherer, LC Ho, p. 103. San Francisco: ASP
4079:
4080: \nhi
4081: Ho LC. 1999a. {\it Ap. J.} 510:631
4082:
4083: \nhi
4084: Ho LC. 1999b. {\it Ap. J.} 516:672
4085:
4086: \nhi
4087: Ho LC. 2002a. {\it Ap. J.} 564:120
4088:
4089: \nhi
4090: Ho LC. 2002b. In {\it Issues in Unification of AGNs}, ed. R Maiolino,
4091: A Marconi, N Nagar, p. 165. San Francisco: ASP
4092:
4093: \nhi
4094: Ho LC. 2003. In {\it Active Galactic Nuclei: from Central Engine to Host
4095: Galaxy}, ed. S Collin, F Combes, I Shlosman, p. 379. San Francisco: ASP
4096:
4097: \nhi
4098: Ho LC, ed. 2004a. {\it Carnegie Observatories Astrophysics Series, Vol. 1:
4099: Coevolution of Black Holes and Galaxies} Cambridge: Cambridge Univ. Press
4100:
4101: \nhi
4102: Ho LC. 2004b. In {\it Carnegie Observatories Astrophysics Series, Vol. 1:
4103: Coevolution of Black Holes and Galaxies}, ed. LC Ho, p. 293. Cambridge:
4104: Cambridge Univ. Press
4105:
4106: \nhi
4107: Ho LC. 2005. In {\it From X-ray Binaries to Quasars: Black Hole Accretion on
4108: All Mass Scales}, ed. TJ Maccarone, RP Fender, LC Ho, p. 219. Dordrecht: Kluwer
4109:
4110: \nhi
4111: Ho LC, Feigelson ED, Townsley LK, Sambruna RM, Garmire GP, et al. 2001.
4112: {\it Ap. J.} 549:L51
4113:
4114: \nhi
4115: Ho LC, Filippenko AV. 1993. {\it Astrophys. Space Sci.} 205:19
4116:
4117: \nhi
4118: Ho LC, Filippenko AV, Sargent WLW. 1993. {\it Ap. J.} 417:63
4119:
4120: \nhi
4121: Ho LC, Filippenko AV, Sargent WLW. 1995. {\it Ap.~J.~Suppl.} 98:477
4122:
4123: \nhi
4124: Ho LC, Filippenko AV, Sargent WLW. 1996. {\it Ap. J.} 462:183 %M81
4125:
4126: \nhi
4127: Ho LC, Filippenko AV, Sargent WLW. 1997a. {\it Ap.~J.~Suppl.} 112:315
4128:
4129: \nhi
4130: Ho LC, Filippenko AV, Sargent WLW. 1997b. {\it Ap. J.} 487:568 %stats
4131:
4132: \nhi
4133: Ho LC, Filippenko AV, Sargent WLW. 1997c. {\it Ap. J.} 487:579 %HII
4134:
4135: \nhi
4136: Ho LC, Filippenko AV, Sargent WLW. 1997d. {\it Ap. J.} 487:591 %bars
4137:
4138: \nhi
4139: Ho LC, Filippenko AV, Sargent WLW. 2003. {\it Ap. J.} 583:159 %HFSVI
4140:
4141: \nhi
4142: Ho LC, Filippenko AV, Sargent WLW, Peng CY. 1997e. {\it Ap.~J.~Suppl.}
4143: 112:391
4144:
4145: \nhi
4146: Ho LC, Peng CY. 2001. {\it Ap. J.} 555:650
4147:
4148: \nhi
4149: Ho LC, Ptak A, Terashima Y, Kunieda H, Serlemitsos PJ, et al.
4150: 1999a. {\it Ap. J.} 525:168
4151:
4152: \nhi
4153: Ho LC, Rudnick G, Rix H-W, Shields JC, McIntosh DH, et al.
4154: 2000. {\it Ap. J.} 541:120
4155:
4156: \nhi
4157: Ho LC, Sarzi M, Rix H-W, Shields JC, Rudnick G, et al. 2002. {\it PASP} 114:137
4158:
4159: \nhi
4160: Ho LC, Shields JC, Filippenko AV. 1993. {\it Ap. J.} 410:567
4161:
4162: \nhi
4163: Ho LC, Terashima Y, Ulvestad JS. 2003. {\it Ap. J.} 589:783
4164:
4165: \nhi
4166: Ho LC, Ulvestad JS. 2001. {\it Ap.~J.~Suppl.} 133:77
4167:
4168: \nhi
4169: Ho LC, Van Dyk SD, Pooley GG, Sramek RA, Weiler KW. 1999b. {\it Astron. J}
4170: 118:843
4171:
4172: \nhi
4173: Hopkins PF, Hernquist L. 2006. {\it Ap.~J.~Suppl.} 166:1
4174:
4175: \nhi
4176: Hopkins PF, Hernquist L, Cox TJ, Di Matteo T, Martini P, et al.
4177: 2006. {\it Ap.~J.~Suppl.} 163:1
4178:
4179: \nhi
4180: Huchra JP, Burg R. 1992. {\it Ap. J.} 393:90
4181:
4182: \nhi
4183: Humason ML, Mayall NU, Sandage AR. 1956. {\it Astron. J.} 61:97
4184:
4185: \nhi
4186: Igumenshchev IV, Narayan R, Abramowicz MA. 2003. {\it Ap. J.} 592:1042
4187:
4188: \nhi
4189: Inoue H, Terashima Y, Ho LC. 2007. {\it Ap. J.} 662:860
4190:
4191: \nhi
4192: Irwin JA, Sarazin CL, Bregman JN. 2002. {\it Ap. J.} 570:152
4193:
4194: \nhi
4195: Ishisaki Y, Makishima K, Iyomoto N, Hayashida K, Inoue H, et al. 1996.
4196: {\it PASJ} 48:237
4197:
4198: \nhi
4199: Iyomoto N, Fukazawa Y, Nakai N, Ishihara Y. 2001. {\it Ap. J.} 561:L69
4200:
4201: \nhi
4202: Iyomoto N, Makishima K, Fukazawa Y, Tashiro M, Ishisaki Y, et al.
4203: 1996. {\it PASJ}, 48:231
4204:
4205: \nhi
4206: Iyomoto N, Makishima K, Fukazawa Y, Tashiro M, Ishisaki Y. 1997. {\it PASJ}
4207: 49:425
4208:
4209: \nhi
4210: Iyomoto N, Makishima K, Matsushita K, Fukazawa Y, Tashiro M, Ohashi T. 1998a.
4211: {\it Ap. J.} 503:168
4212:
4213: \nhi
4214: Iyomoto N, Makishima K, Tashiro M, Inoue S, Kaneda H, et al.
4215: 1998b. {\it Ap. J.} 503:L31
4216:
4217: \nhi
4218: Jolley EJD, Kuncic Z. 2007. {\it Ap\&SS} 310:327
4219:
4220: \nhi
4221: Johnson BM, Quataert E. 2007. {\it Ap. J.} 660:1273
4222:
4223: \nhi
4224: Jungwiert B, Combes F, Palous J. 2001. {\it Astron. Astrophys.} 376:85
4225:
4226: \nhi
4227: Kauffmann G, Heckman TM, Tremonti C, Brinchmann J, Charlot S, et al. 2003.
4228: {\it MNRAS} 346:1055
4229:
4230: \nhi
4231: Kauffmann G, White SDM, Heckman TM, Me\'nard B, Brinchmann J, et al.
4232: 2004. {\it MNRAS} 353:713
4233:
4234: \nhi
4235: Keel WC. 1983a. {\it Ap. J.} 268:632
4236:
4237: \nhi
4238: Keel WC. 1983b. {\it Ap. J.} 269:466
4239:
4240: \nhi
4241: Keel WC. 1983c. {\it Ap.~J.~Suppl.} 52:229
4242:
4243: \nhi
4244: Keel WC, Miller JS. 1983. {\it Ap. J.} 266:L89
4245:
4246: \nhi
4247: Kennicutt RC Jr. 1984. {\it Ap. J.} 287:116
4248:
4249: \nhi
4250: Kewley LJ, Groves B, Kauffmann G, Heckman T. 2006. {\it MNRAS} 372:961
4251:
4252: \nhi
4253: Kewley LJ, Heisler CA, Dopita MA, Lumsden S. 2001. {\it Ap.~J.~Suppl.} 132:37
4254:
4255: \nhi
4256: Khachikian EY, Weedman DW. 1974. {\it Ap. J.} 192:581
4257:
4258: \nhi
4259: Kharb P, Shastri P. 2004. {\it Astron. Astrophys.} 425:825
4260:
4261: \nhi
4262: Kim D-C, Sanders DB, Veilleux S, Mazzarella JM, Soifer BT. 1995.
4263: {\it Ap.~J.~Suppl.} 98:129
4264:
4265: \nhi
4266: Kim D-W, Fabbiano G. 2003. {\it Ap.~J.} 586:826
4267:
4268: \nhi
4269: Kinkhabwala A, Sako M, Behar E, Kahn SM, Paerels F, et al. 2002.
4270: {\it Ap. J.} 575:732
4271:
4272: \nhi
4273: Kirhakos S, Phillips MM. 1989. {\it PASP} 101:949
4274:
4275: \nhi
4276: Komossa S, B\"ohringer H, Huchra JP. 1999. {\it Astron. Astrophys.} 349:88
4277:
4278: \nhi
4279: Koratkar AP, Deustua S, Heckman TM, Filippenko AV, Ho LC, Rao M. 1995.
4280: {\it Ap. J.} 440:132
4281:
4282: \nhi
4283: K\"ording E, Falcke H, Corbel S. 2006. {\it Astron. Astrophys.} 456:439
4284:
4285: \nhi
4286: K\"ording E, Jester S, Fender R. 2006. {\it MNRAS} 372:1366
4287:
4288: \nhi
4289: K\"ording E, Jester S, Fender R. 2008. {\it MNRAS} 383:277
4290:
4291: \nhi
4292: Kormendy J. 1993. In {\it The Nearest Active Galaxies}, ed. J Beckman,
4293: L Colina, H Netzer, p. 197. (Madrid: CSIC)
4294:
4295: \nhi
4296: Kormendy J. 2004. In {\it Carnegie Observatories Astrophysics Series, Vol 1:
4297: Coevolution of Black Holes and Galaxies}, ed. LC Ho, p. 1. Cambridge:
4298: Cambridge Univ. Press
4299:
4300: \nhi
4301: Kormendy J, Fisher DB, Cornell ME, Bender R. 2008. {\it Ap.~J.~Suppl.} submitted
4302:
4303: \nhi
4304: Kormendy J, Kennicutt RC. 2004. {\it Annu. Rev. Astron. Astrophys.} 42:603
4305:
4306: \nhi
4307: Kormendy J, Richstone DO. 1995. {\it Annu. Rev. Astron. Astrophys.} 33:581
4308:
4309: \nhi
4310: Koski AT, Osterbrock DE. 1976. {\it Ap. J.} 203:L49
4311:
4312: \nhi
4313: Krips M, Eckart A, Krichbaum TP, Pott J-U, Leon S, et al. 2007. {\it Astron.
4314: Astrophys.} 464:553
4315:
4316: \nhi
4317: Krolik JH. 1998. {\it Active Galactic Nuclei}. Princeton: Princeton Univ. Press
4318:
4319: \nhi
4320: Kukula MJ, Pedlar A, Baum SA, O'Dea CP. 1995. {\it MNRAS} 276:1262
4321:
4322: \nhi
4323: Kunth D, Sargent WLW, Bothun GD. 1987. {\it Astron. J.} 92:29
4324:
4325: \nhi
4326: Laor A. 2003. {\it Ap. J.} 590:86
4327:
4328: \nhi
4329: Larkin JE, Armus L, Knop RA, Soifer BT, Matthews K. 1998. {\it Ap.~J.~Suppl.}
4330: 114:59
4331:
4332: \nhi
4333: Lasota J-P, Abramowicz MA, Chen X, Krolik J, Narayan R, Yi I. 1996.
4334: {\it Ap. J.} 462:142
4335:
4336: \nhi
4337: Laurikainen E, Salo H, Buta R. 2004. {\it Ap. J.} 607:103
4338:
4339: \nhi
4340: Lawrence A. 2005. {\it MNRAS} 363:57
4341:
4342: \nhi
4343: Lawrence A, Elvis M. 1982. {\it Ap. J.} 256:410
4344:
4345: \nhi
4346: Lawrence A, Ward M, Elvis M, Fabbiano G, Willner SP, et al. 1985.
4347: {\it Ap. J.} 291:117
4348:
4349: \nhi
4350: Lewis KT, Eracleous M, Gliozzi M, Sambruna RM, Mushotzky RF. 2005. {\it Ap. J.}
4351: 622:816
4352:
4353: \nhi
4354: Lewis KT, Eracleous M, Sambruna RM. 2003. {\it Ap. J.} 593:115
4355:
4356: \nhi
4357: Li C, Kauffmann G, Heckman TM, White SDM, Jing YP. 2008. {\it MNRAS} in press
4358:
4359: \nhi
4360: Lightman AP, White TR. 1988. {\it Ap. J.} 335:57
4361:
4362: \nhi
4363: Lin DC, Shields GA. 1986. {\it Ap. J.} 305:28
4364:
4365: \nhi
4366: Lira P, Lawrence A, Johnson RA. 2000. {\it MNRAS} 319:17
4367:
4368: \nhi
4369: Liu BF, Meyer-Hofmeister E. 2001. {\it Astron. Astrophys.} 372:386
4370:
4371: \nhi
4372: Liu BF, Taam RE, Meyer-Hofmeister E, Meyer F. 2007. {\it Ap. J.} 671:695
4373:
4374: \nhi
4375: Livio M, Ogilvie GI, Pringle JE. 1999. {\it Ap. J.} 512:100
4376:
4377: \nhi
4378: Loewenstein M, Mushotzky RF, Angelini L, Arnaud KA, Quataert E. 2001.
4379: {\it Ap. J.} 555:L21
4380:
4381: \nhi
4382: Maccarone TJ, Fender RP, Ho LC, eds. 2005. {\it From X-ray Binaries to
4383: Quasars: Black Hole Accretion on All Mass Scales} Dordrecht: Kluwer
4384:
4385: \nhi
4386: Maccarone TJ, Gallo E, Fender RP. 2003. {\it MNRAS} 345:L19
4387:
4388: \nhi
4389: Magorrian J, Tremaine S, Richstone D, Bender R, Bower G, et al. 1998.
4390: {\it Astron. J.} 115:2285
4391:
4392: \nhi
4393: Mahadevan R. 1997. {\it Ap. J.} 477:585
4394:
4395: \nhi
4396: Maia MAG, Machado RS, Willmer CNA. 2003. {\it Astron. J.} 126:1750
4397:
4398: \nhi
4399: Maiolino R, Rieke GH. 1995. {\it Ap. J.} 454:95
4400:
4401: \nhi
4402: Makishima K, Fujimoto R, Ishisaki Y, Kii T, Loewenstein M, et al. 1994.
4403: {\it PASJ} 46:L77
4404:
4405: \nhi
4406: Makishima K, Ohashi T, Hayashida K, Inoue H, Koyama K, et al. 1989. {\it PASJ}
4407: 41:697
4408:
4409: \nhi
4410: Malkan MA, Sargent WLW. 1982. {\it Ap. J.} 254:22
4411:
4412: \nhi
4413: Maoz D. 2007. {\it MNRAS} 377:1696
4414:
4415: \nhi
4416: Maoz D, Filippenko AV, Ho LC, Macchetto FD, Rix H-W, Schneider DP. 1996.
4417: {\it Ap.~J.~Suppl.} 107:215
4418:
4419: \nhi
4420: Maoz D, Filippenko AV, Ho LC, Rix H-W, Bahcall JN, et al. 1995.
4421: {\it Ap. J.} 440:91
4422:
4423: \nhi
4424: Maoz D, Koratkar AP, Shields JC, Ho LC, Filippenko AV, Sternberg A. 1998.
4425: {\it Astron. J.} 116:55
4426:
4427: \nhi
4428: Maoz D, Nagar NM, Falcke H, Wilson AS. 2005. {\it Ap. J.} 625:699
4429:
4430: \nhi
4431: Matsumoto Y, Fukazawa Y, Nakazawa K, Iyomoto N, Makishima K. 2001. {\it PASJ}
4432: 53:475
4433:
4434: \nhi
4435: Meier DL. 1999. {\it Ap. J.} 522:753
4436:
4437: \nhi
4438: Meier DL. 2001. {\it Ap. J.} 548:L9
4439:
4440: \nhi
4441: Meisenheimer K, Tristram KRW, Jaffe W, Israel F, Neumayer N, et al. 2007.
4442: {\it Astron. Astrophys.} 471:453
4443:
4444: \nhi
4445: Menou K, Quataert E. 2001. {\it Ap. J.} 552:204
4446:
4447: \nhi
4448: Merloni A, Fabian AC. 2002. {\it MNRAS} 332:165
4449:
4450: \nhi
4451: Merloni A, Heinz S, Di Matteo T. 2003. {\it MNRAS} 345:1057
4452:
4453: \nhi
4454: Merloni A, Nayakshin S, Sunyaev R, eds. 2005. {\it Growing Black Holes:
4455: Accretion in a Cosmological Context} Berlin: Springer-Verlag
4456:
4457: \nhi
4458: Middleton M, Done C, Schurch N. 2008. {\it MNRAS} 383:1501
4459:
4460: \nhi
4461: Miller CJ, Nichol RC, Gomez PL, Hopkins AM, Bernardi M. 2003. {\it Ap. J.}
4462: 597:142
4463:
4464: \nhi
4465: Milosavljevi\'c M, Merritt D, Ho LC. 2006. {\it Ap. J.} 652:120
4466:
4467: \nhi
4468: Minkowski R, Osterbrock DE. 1959. {\it Ap. J.} 129:583
4469:
4470: \nhi
4471: Miniutti G, Ponti G, Greene JE, Ho LC, Fabian AC, Iwasawa K. 2008. {\it MNRAS}
4472: in press
4473:
4474: \nhi
4475: Moore B, Katz N, Lake G, Dressler A, Oemler A. 1996. {\it Nature} 379:613
4476:
4477: \nhi
4478: Moran EC, Eracleous M, Leighly KM, Chartas G, Filippenko AV, et al.
4479: 2005. {\it Astron. J.} 129:2108
4480:
4481: \nhi
4482: M\"{u}ller SAH, Haas M, Siebenmorgen R, Klaas U, Meisenheimer K, et al.
4483: 2004. {\it Astron. Astrophys.} 426:L29
4484:
4485: \nhi
4486: Muno MP, Baganoff FK, Bautz MW, Feigelson ED, Garmire GP, et al. 2004.
4487: {\it Ap. J.} 613:326
4488:
4489: \nhi
4490: Murray N, Chiang J. 1997. {\it Ap. J.} 474:91
4491:
4492: \nhi
4493: Mushotzky RF. 1993. In {\it The Nearest Active Galaxies}, ed. J Beckman,
4494: L Colina, H Netzer, p. 47. Madrid: CSIC Press
4495:
4496: \nhi
4497: Mushotzky RF, Wandel A. 1989. {\it Ap. J.} 339:674
4498:
4499: \nhi
4500: Nagao T, Murayama T, Shioya Y, Taniguchi Y. 2002. {\it Ap. J.} 567:73
4501:
4502: \nhi
4503: Nagar NM, Falcke H, Wilson AS. 2005. {\it Ap. J.} 435:521
4504:
4505: \nhi
4506: Nagar NM, Falcke H, Wilson AS, Ho LC. 2000. {\it Ap. J.} 542:186
4507:
4508: \nhi
4509: Nagar NM, Falcke H, Wilson AS, Ulvestad JS. 2002. {\it Astron. Astrophys.}
4510: 392:53
4511:
4512: \nhi
4513: Nagar NM, Wilson AS, Falcke H. 2001. {\it Ap. J.} 559:L87
4514:
4515: \nhi
4516: Nandra K, George IM, Mushotzky RF, Turner TJ, Yaqoob T. 1997a. {\it Ap. J.}
4517: 476:70
4518:
4519: \nhi
4520: Nandra K, George IM, Mushotzky RF, Turner TJ, Yaqoob T. 1997b. {\it Ap. J.}
4521: 477:602
4522:
4523: \nhi
4524: Nandra K, O'Neill PM, George IM, Reeves JN. 2007. {\it MNRAS} 382:194
4525:
4526: \nhi
4527: Narayan R. 2002. In {\it Lighthouses of the Universe}, ed. M Gilfanov et al.,
4528: p. 405. Berlin: Springer
4529:
4530: \nhi
4531: Narayan R, Yi I. 1995. {\it Ap. J.} 444:231
4532:
4533: \nhi
4534: Nayakshin S. 2003. {\it Astron. Nachr. Suppl.} 324:3
4535:
4536: \nhi
4537: Nelson CH, Whittle M. 1996. {\it Ap. J.} 465:96
4538:
4539: \nhi
4540: Nemmen RS, Storchi-Bergmann T, Yuan F, Eracleous M, Terashima Y, Wilson AS.
4541: 2006. {\it Ap. J.} 643:652
4542:
4543: \nhi
4544: Nicastro F. 2000. {\it Ap. J.} 530:L65
4545:
4546: \nhi
4547: Nicholson KL, Reichert GA, Mason KO, Puchnarewicz EM, Ho LC, et al.
4548: 1998. {\it MNRAS} 300:893
4549:
4550: \nhi
4551: Noyola E, Gebhardt K, Bergmann M. 2008. {\it Ap. J.} in press
4552:
4553: \nhi
4554: O'Connell RW. 1999. {\it Annu. Rev. Astron. Astrophys.} 37:603
4555:
4556: \nhi
4557: O'Connell RW, Martin JR, Crane JD, Burstein D, Bohlin RC, et al.
4558: 2005. {\it Ap. J.} 635:305
4559:
4560: \nhi
4561: Omma H, Binney J, Bryan G, Slyz A. 2004. {\it MNRAS} 348:1105
4562:
4563: \nhi
4564: Osterbrock DE. 1960. {\it Ap. J.} 132:325
4565:
4566: \nhi
4567: Osterbrock DE. 1971. In {\it Nuclei of Galaxies}, ed. DJK O'Connell, p. 151.
4568: Amsterdam: North Holland
4569:
4570: \nhi
4571: Osterbrock DE, Dufour RJ. 1973. {\it Ap. J.} 185:441
4572:
4573: \nhi
4574: Osterbrock DE, Miller JS. 1975. {\it Ap. J.} 197:535
4575:
4576: \nhi
4577: Osterbrock DE, Shaw RA. 1988. {\it Ap. J.} 327:89
4578:
4579: \nhi
4580: Padovani P, Matteucci F. 1993. {\it Ap. J.} 416:26
4581:
4582: \nhi
4583: Page MJ, Breeveld AA, Soria R, Wu K, Branduardi-Raymont G, et al.
4584: 2003. {\it Astron. Astrophys.} 400:145
4585:
4586: \nhi
4587: Page MJ, Soria R, Zane S, Wu K, Starling R. 2004. {\it Astron. Astrophys.}
4588: 422:77
4589:
4590: \nhi
4591: Panessa F, Barcons X, Bassani L, Cappi M, Carrera FJ, et al.
4592: 2007. {\it Astron. Astrophys.} 467:519
4593:
4594: \nhi
4595: Panessa F, Bassani L. 2002. {\it Astron. Astrophys.} 394:435
4596:
4597: \nhi
4598: Panessa F, Bassani L, Cappi M, Dadina M, Barcons X, et al.
4599: 2006. {\it Astron. Astrophys.} 455:173
4600:
4601: \nhi
4602: Pappa A, Georgantopoulos I, Stewart GC, Zezas AL. 2001. {\it MNRAS} 326:995
4603:
4604: \nhi
4605: Peimbert M, Torres-Peimbert S. 1981. {\it Ap. J.} 245:845
4606:
4607: \nhi
4608: Pelat D, Alloin D, Fosbury RAE. 1981. {\it MNRAS} 195:787
4609:
4610: \nhi
4611: Pellegrini S. 2005. {\it Ap. J.} 624:155
4612:
4613: \nhi
4614: Pellegrini S, Baldi A, Fabbiano G, Kim D-W. 2003a. {\it Ap. J.} 597:175
4615:
4616: \nhi
4617: Pellegrini S, Cappi M, Bassani L, Della Ceca R, Palumbo GGC. 2000a.
4618: {\it Astron. Astrophys.} 360:878
4619:
4620: \nhi
4621: Pellegrini S, Cappi M, Bassani L, Malaguti G, Palumbo GGC, Persic M. 2000b.
4622: {\it Astron. Astrophys.} 353:447
4623:
4624: \nhi
4625: Pellegrini S, Fabbiano G, Fiore F, Trinchieri G, Antonelli A. 2002.
4626: {\it Astron. Astrophys.} 383:1
4627:
4628: \nhi
4629: Pellegrini S, Siemiginowska A, Fabbiano G, Elvis M, Greenhill L, et al. 2007.
4630: {\it Ap. J.} 667:749
4631:
4632: \nhi
4633: Pellegrini S, Venturi T, Comastri A, Fabbiano G, Fiore F, et al.
4634: 2003b. {\it Ap. J.} 585:677
4635:
4636: \nhi
4637: Peng CY, Ho LC, Impey CD, Rix H-W. 2002. {\it Astron. J.} 124:266
4638:
4639: \nhi
4640: Penston MV, Fosbury RAE. 1978. {\it MNRAS} 183:479
4641:
4642: \nhi
4643: P\'equignot D. 1984. {\it Astron. Astrophys.} 131:159
4644:
4645: \nhi
4646: Peterson BM, Bentz MC, Desroches L-B, Filippenko AV, Ho LC, et al. 2005.
4647: {\it Ap. J.} 632:799. Erratum. 2005. {\it Ap. J.} 641:638
4648:
4649: \nhi
4650: Petre R, Mushotzky RF, Serlemitsos PJ, Jahoda K, Marshall FE. 1993.
4651: {\it Ap. J.} 418:644
4652:
4653: \nhi
4654: Phillips MM. 1979. {\it Ap. J.} 227:L121
4655:
4656: \nhi
4657: Phillips MM, Charles PA, Baldwin JA. 1983. {\it Ap. J.} 266:485
4658:
4659: \nhi
4660: Phillips MM, Jenkins CR, Dopita MA, Sadler EM, Binette L. 1986.
4661: {\it Astron. J.} 91:1062
4662:
4663: \nhi
4664: Pogge RW. 1989. {\it Ap.~J.~Suppl.} 71:433
4665:
4666: \nhi
4667: Pogge RW, Maoz D, Ho LC, Eracleous M. 2000. {\it Ap. J.} 532:323
4668:
4669: \nhi
4670: Proga D, Begelman MC. 2003. {\it Ap. J.} 592:767
4671:
4672: \nhi
4673: Ptak A, Serlemitsos PJ, Yaqoob T, Mushotzky R. 1999. {\it Ap.~J.~Suppl.}
4674: 120:179
4675:
4676: \nhi
4677: Ptak A, Terashima Y, Ho LC, Quataert E. 2004. {\it Ap. J.} 606:173
4678:
4679: \nhi
4680: Ptak A, Yaqoob T, Mushotzky R, Serlemitsos P, Griffiths R. 1998. {\it Ap. J.}
4681: 501:L37
4682:
4683: \nhi
4684: Ptak A, Yaqoob T, Serlemitsos PJ, Kunieda H, Terashima Y. 1996. {\it Ap. J.}
4685: 459:542
4686:
4687: \nhi
4688: Quataert E. 2003. {\it Astron. Nachr. Suppl.} 324:435
4689:
4690: \nhi
4691: Quataert E, Di Matteo T, Narayan R, Ho LC. 1999. {\it Ap. J.} 525:L89
4692:
4693: \nhi
4694: Quillen AC, McDonald C, Alonso-Herrero A, Lee A, Shaked S, et al.
4695: 2001. {\it Ap. J.} 547:129
4696:
4697: \nhi
4698: Ravindranath S, Ho LC, Peng CY, Filippenko AV, Sargent WLW. 2001.
4699: {\it Astron. J.} 122:653
4700:
4701: \nhi
4702: Rees MJ, Phinney ES, Begelman MC, Blandford RD. 1982. {\it Nature} 295:17
4703:
4704: \nhi
4705: Reeves JN, Turner MJL. 2000. {\it MNRAS} 316:234
4706:
4707: \nhi
4708: Renzini A, Greggio L, di Serego Alighieri S, Cappellari M, Burstein
4709: D, Bertola F. 1995. {\it Nature} 378:39
4710:
4711: \nhi
4712: Reynolds CS, Di Matteo T, Fabian AC, Hwang U, Canizares CR. 1996. {\it MNRAS}
4713: 283:L111
4714:
4715: \nhi
4716: Rice MS, Martini P, Greene JE, Pogge RW, Shields JC, et al. 2006. {\it Ap. J.}
4717: 636:654
4718:
4719: \nhi
4720: Richstone DO. 2004. In {\it Carnegie Observatories Astrophysics Series, Vol 1:
4721: Coevolution of Black Holes and Galaxies}, ed. LC Ho, p. 280. Cambridge:
4722: Cambridge Univ. Press
4723:
4724: \nhi
4725: Rieke GH, Lebofsky MJ, Kemp JC. 1982. {\it Ap. J.} 252:L53
4726:
4727: \nhi
4728: Rinn AS, Sambruna RM, Gliozzi M. 2005. {\it Ap. J.} 621:167
4729:
4730: \nhi
4731: Roberts TP, Schurch NJ, Warwick RS. 2001. {\it MNRAS} 324:737
4732:
4733: \nhi
4734: Roberts TP, Warwick RS. 2000. {\it MNRAS} 315:98
4735:
4736: \nhi
4737: Rola C, Terlevich E, Terlevich R. 1997. {\it MNRAS} 289:419
4738:
4739: \nhi
4740: Rose JA, Searle L. 1982. {\it Ap. J.} 253:556
4741:
4742: \nhi
4743: Rose JA, Tripicco MJ. 1984. {\it Ap. J.} 285:55
4744:
4745: \nhi
4746: Rubin VC, Ford WK Jr, Thonnard N. 1980. {\it Ap. J.} 238:471
4747:
4748: \nhi
4749: Rupke D, Veilleux S, Kim D-C, Sturm E, Contursi A, et al. 2007. In {\it The
4750: Central Engine of Active Galactic Nuclei}, ed. LC Ho, J-M Wang, p. 525.
4751: San Francisco: ASP
4752:
4753: \nhi
4754: Sabra BM, Shields JC, Ho LC, Barth AJ, Filippenko AV 2003. {\it Ap. J.} 584:164
4755:
4756: \nhi
4757: Sadler EM, Jenkins CR, Kotanyi CG. 1989. {\it MNRAS} 240:591
4758:
4759: \nhi
4760: Sambruna RM, Gliozzi M, Eracleous M, Brandt WN, Mushotzky RF. 2003.
4761: {\it Ap. J.} 586:L37
4762:
4763: \nhi
4764: Sandage A, Bedke J. 1994. {\it The Carnegie Atlas of Galaxies}.
4765: Washington, DC: Carnegie Inst. of Washington
4766:
4767: \nhi
4768: Sandage A, Tammann, GA. 1981. {\it A Revised Shapley-Ames Catalog of
4769: Bright Galaxies}. Washington, DC: Carnegie Inst. of Washington
4770:
4771: \nhi
4772: Sargent WLW, Filippenko AV. 1991. {\it Astron. J.} 102:107
4773:
4774: \nhi
4775: Sarzi M, Shields JC, Pogge RW, Martini P. 2007. In {\it The Central Engine of
4776: Active Galactic Nuclei}, ed. LC Ho, J-M Wang, p. 643. San Francisco: ASP
4777:
4778: \nhi
4779: Sarzi M, Rix H-W, Shields JC, Ho LC, Barth AJ, et al.
4780: 2005. {\it Ap. J.} 628:169
4781:
4782: \nhi
4783: Satyapal S, Dudik RP, O'Halloran B, Gliozzi M. 2005. {\it Ap. J.} 633:86
4784:
4785: \nhi
4786: Satyapal S, Sambruna RM, Dudik RP. 2004. {\it Astron. Astrophys.} 414:825
4787:
4788: \nhi
4789: Satyapal S, Vega D, Dudik RP, Abel NP, Heckman T. 2008. {\it Ap. J.} in press
4790:
4791: \nhi
4792: Satyapal S, Vega D, Heckman T, O'Halloran B, Dudik R. 2007. {\it Ap. J.} 663:L9
4793:
4794: \nhi
4795: Schmidt M, Green RF. 1983. {\it Ap. J.} 269:352
4796:
4797: \nhi
4798: Schmitt HR. 2001. {\it Astron. J.} 122:2243
4799:
4800: \nhi
4801: Schmitt HR, Kinney AL, Ho LC, eds. 1999. {\it The AGN/Normal Galaxy
4802: Connection} Oxford: Elsevier Science Ltd.
4803:
4804: \nhi
4805: Schulz H, Fritsch C. 1994. {\it Astron. Astrophys.} 291:713
4806:
4807: \nhi
4808: Serote-Roos M, Gon\c{c}alves AC. 2004. {\it Astron. Astrophys.} 413:91
4809:
4810: \nhi
4811: Seth AC, Ag\"ueros M, Lee D, Basu-Zych A. 2008. {\it Ap. J.} in press
4812:
4813: \nhi
4814: Shakura NI, Sunyaev RA. 1973. {\it Astron. Astrophys.} 24:337
4815:
4816: \nhi
4817: Shang Z, Brotherton MS, Green RF, Kriss GA, Scott J, et al. 2005.
4818: {\it Ap. J.} 619:41
4819:
4820: \nhi
4821: Shields GA. 1978. {\it Nature}. 272:706
4822:
4823: \nhi
4824: Shields GA, Wheeler JC. 1978. {\it Ap. J.} 222:667
4825:
4826: \nhi
4827: Shields JC. 1992. {\it Ap. J.} 399:L27
4828:
4829: \nhi
4830: Shields JC, Rix H-W, McIntosh DH, Ho LC, Rudnick G, et al.
4831: 2000. {\it Ap. J.} 534:L27
4832:
4833: \nhi
4834: Shields JC, Rix H-W, Sarzi M, Barth AJ, Filippenko AV, et al. 2007.
4835: {\it Ap. J.} 654:125
4836:
4837: \nhi
4838: Shields JC, Sabra BM, Ho LC, Barth AJ, Filippenko AV. 2002. In {\it Mass
4839: Outflow in Active Galactic Nuclei: New Perspectives}, ed. DM Crenshaw, SB
4840: Kraemer, IM George, p. 105. San Francisco: ASP
4841:
4842: \nhi
4843: Shields JC, Walcher CJ, B\"oker, T, Ho LC, Rix H-W, van~der~Marel RP. 2008.
4844: {\it Ap. J.} in press
4845:
4846: \nhi
4847: Shih DC, Iwasawa K, Fabian AC. 2003. {\it MNRAS} 341:973
4848:
4849: \nhi
4850: Siemiginowska A, Czerny B, Kostyunin V. 1996. {\it Ap. J.} 458:491
4851:
4852: \nhi
4853: Sikora M, Stawarz L, Lasota J-P. 2007. {\it Ap. J.} 658:815
4854:
4855: \nhi
4856: Slee OB, Sadler EM, Reynolds JE, Ekers RD. 1994. {\it MNRAS} 269:928
4857:
4858: \nhi
4859: Smith JDT, Draine BT, Dale DA, Moustakas J, Kennicutt RC, et al. 2007.
4860: {\it Ap. J.} 656:770
4861:
4862: \nhi
4863: So\l tan A. 1982. {\it MNRAS} 200:115
4864:
4865: \nhi
4866: Soria R, Fabbiano G, Graham A, Baldi A, Elvis M, et al.
4867: 2006. {\it Ap. J.} 640:126
4868:
4869: \nhi
4870: Spinoglio L, Malkan MA. 1992. {\it Ap. J.} 399:504
4871:
4872: \nhi
4873: Springel V, Di~Matteo T, Hernquist L. 2005. {\it MNRAS} 361:776
4874:
4875: \nhi
4876: Starling RLC, Page MJ, Branduardi-Raymont G, Breeveld AA, Soria R, Wu K. 2005.
4877: {\it MNRAS} 356:727
4878:
4879: \nhi
4880: Stasi\'nska G. 1984. {\it Astron. Astrophys.} 135:341
4881:
4882: \nhi
4883: Stasi\'nska G, Cid Fernandes R, Mateus A, Sodr\'e L Jr, Asari NV. 2006.
4884: {\it MNRAS} 371:972
4885:
4886: \nhi
4887: Stauffer JR. 1982a. {\it Ap.~J.~Suppl.} 50:517
4888:
4889: \nhi
4890: Stauffer JR. 1982b. {\it Ap. J.} 262:66
4891:
4892: \nhi
4893: Stauffer JR, Spinrad H. 1979. {\it Ap. J.} 231:L51
4894:
4895: \nhi
4896: Stone JD, Pringle JE. 2001. {\it MNRAS} 322:461
4897:
4898: \nhi
4899: Storchi-Bergmann T, Baldwin JA, Wilson AS. 1993. {\it Ap. J.} 410:L11
4900:
4901: \nhi
4902: Storchi-Bergmann T, Ho LC, Schmitt HR, eds. 2004. {\it IAU Symp. 222,
4903: Interplay among Black Holes, Stars and ISM in Galactic Nuclei} Cambridge:
4904: Cambridge Univ. Press
4905:
4906: \nhi
4907: Storchi-Bergmann T, Pastoriza MG. 1990. {\it PASP} 102:1359
4908:
4909: \nhi
4910: Strateva IV, Brandt WN, Schneider DP, Vanden Berk DG, Vignali C. 2005.
4911: {\it Astron. J.} 130:387
4912:
4913: \nhi
4914: Sturm E, Rupke D, Contursi A, Kim D-C, Lutz D, et al. 2006. {\it Ap. J.}
4915: 653:L13
4916:
4917: \nhi
4918: Sturm E, Schweitzer M, Lutz D, Contursi A, Genzel R, et al. 2005.
4919: {\it Ap. J.} 629:L21
4920:
4921: \nhi
4922: Sugai H, Malkan MA. 2000. {\it Ap. J.} 529:219
4923:
4924: \nhi
4925: Sulentic JW, Marziani P, Dultzin-Hacyan D. 2000.
4926: {\it Annu. Rev. Astron. Astrophys.} 38:521
4927:
4928: \nhi
4929: Swartz DA, Ghosh KK, Suleimanov V, Tennant AF, Wu K. 2002. {\it Ap. J.} 574:382
4930:
4931: \nhi
4932: Szokoly GP, Bergeron J, Hasinger G, Lehmann I, Kewley L, et al. 2004.
4933: {\it Ap.~J.~Suppl.} 155:271
4934:
4935: \nhi
4936: Tan JC, Blackman EG. 2005. {\it MNRAS} 362:983
4937:
4938: \nhi
4939: Taniguchi Y, Shioya Y, Murayama T. 2000. {\it Astron. J.} 120:1265
4940:
4941: \nhi
4942: Terashima Y, Ho LC, Ptak AF. 2000. {\it Ap. J.} 539:161
4943:
4944: \nhi
4945: Terashima Y, Ho LC, Ptak AF, Mushotzky RF, Serlemitsos PJ, et al.
4946: 2000a. {\it Ap. J.} 533:729
4947:
4948: \nhi
4949: Terashima Y, Ho LC, Ptak AF, Yaqoob T, Kunieda H, et al.
4950: 2000b. {\it Ap. J.} 535:L79
4951:
4952: \nhi
4953: Terashima Y, Iyomoto N, Ho LC, Ptak AF. 2002. {\it Ap.~J.~Suppl.} 139:1
4954:
4955: \nhi
4956: Terashima Y, Kunieda H, Misaki K, Mushotzky RF, Ptak AF, Reichert GA. 1998a.
4957: {\it Ap. J.} 503:212
4958:
4959: \nhi
4960: Terashima Y, Ptak A, Fujimoto R, Itoh M, Kunieda H, et al.
4961: 1998b. {\it Ap. J.} 496:210
4962:
4963: \nhi
4964: Terashima Y, Wilson AS. 2003a. {\it Ap. J.} 560:139
4965:
4966: \nhi
4967: Terashima Y, Wilson AS. 2003b. {\it Ap. J.} 583:145
4968:
4969: \nhi
4970: Terlevich R, Melnick J. 1985. {\it MNRAS} 213:841
4971:
4972: \nhi
4973: Tran HD. 2001. {\it Ap. J.} 554:L19
4974:
4975: \nhi
4976: Tremaine S, Gebhardt K, Bender R, Bower G, Dressler A, et al. 2002.
4977: {\it Ap. J.} 574:740
4978:
4979: \nhi
4980: Tremonti CA, Heckman TM, Kauffmann G, Brinchmann J, Charlot S, et al.
4981: 2004. {\it Ap. J.} 613:898
4982:
4983: \nhi
4984: Turner TJ, Pounds KA. 1989. {\it MNRAS} 240:833
4985:
4986: \nhi
4987: Tzanavaris P, Georgantopoulos I. 2007. {\it Astron. Astrophys.} 468:129
4988:
4989: \nhi
4990: Ulvestad JS, Greene JE, Ho LC. 2007. {\it Ap. J.} 661:L159
4991:
4992: \nhi
4993: Ulvestad JS, Ho LC. 2001a. {\it Ap. J.} 558:561
4994:
4995: \nhi
4996: Ulvestad JS, Ho LC. 2001b. {\it Ap. J.} 562:L133
4997:
4998: \nhi
4999: Ulvestad JS, Ho LC. 2002. {\it Ap. J.} 581:925
5000:
5001: \nhi
5002: Ulvestad JS, Wilson AS. 1989. {\it Ap. J.} 343:659
5003:
5004: \nhi
5005: Vanden Berk DE, Richards GT, Bauer A, Strauss MA, Schneider DP, et al. 2001.
5006: {\it Astron. J.} 122:549
5007:
5008: \nhi
5009: Van Dyk SD, Ho LC. 1998. In {\it IAU Symp. 184, The Central Regions of the
5010: Galaxy and Galaxies}, ed. Y Sofue, p. 489. Dordrecht: Kluwer
5011:
5012: \nhi
5013: Varano S, Chiaberge M, Macchetto FD, Capetti A. 2004. {\it Astron. Astrophys.}
5014: 428:401
5015:
5016: \nhi
5017: Veilleux S, Osterbrock DE. 1987. {\it Ap.~J.~Suppl.} 63:295
5018:
5019: \nhi
5020: Verdoes Kleijn GA, Baum SA, de Zeeuw PT, O'Dea CP. 2002.
5021: {\it Astron. J.} 123:1334
5022:
5023: \nhi
5024: Verdoes Kleijn GA, van~der~Marel RP, Noel-Storr J. 2006.
5025: {\it Astron. J.} 131:1961
5026:
5027: \nhi
5028: V\'eron P, Gon\c{c}alves AC, V\'eron-Cetty M-P. 1997.
5029: {\it Astron. Astrophys.} 319:52
5030:
5031: \nhi
5032: V\'eron P, V\'eron-Cetty M-P. 1986. {\it Astron. Astrophys.} 161:145
5033:
5034: \nhi
5035: V\'eron-Cetty M-P, V\'eron P. 1986. {\it Astron. Astrophys. Suppl.} 66:335
5036:
5037: \nhi
5038: V\'eron-Cetty M-P, V\'eron P. 2006. {\it Astron. Astrophys.} 455:773
5039:
5040: \nhi
5041: Viegas-Aldrovandi SM, Gruenwald RB. 1990. {\it Ap. J.} 360:474
5042:
5043: \nhi
5044: Voit GM. 1992. {\it Ap. J.} 399:495
5045:
5046: \nhi
5047: Voit GM, Donahue M. 1997. {\it Ap. J.} 486:242
5048:
5049: \nhi
5050: Wake DA, Miller CJ, Di~Matteo T, Nichol RC, Pope A, et al. 2004.
5051: {\it Ap. J.} 610:L85
5052:
5053: \nhi
5054: Walcher CJ, B\"oker T, Charlot S, Ho LC, Rix H-W, et al.
5055: 2006. {\it Ap. J.} 649:692
5056:
5057: \nhi
5058: Walsh JL, Barth AJ, Ho LC, Filippenko AV, Rix H-W, et al. 2008. {\it Ap. J.}
5059: submitted
5060:
5061: \nhi
5062: Wang J-M, Luo B, Ho LC. 2004. {\it Ap. J.} 615:L5
5063:
5064: \nhi
5065: Wang J-M, Staubert R, Ho LC. 2002. {\it Ap. J.} 579:554
5066:
5067: \nhi
5068: Weaver KA, Wilson AS, Henkel C, Braatz JA. 1999. {\it Ap. J.} 520:130
5069:
5070: \nhi
5071: Weedman DW, Feldman FR, Balzano VA, Ramsey LW, Sramek RA, Wu C-C. 1981.
5072: {\it Ap. J.} 248:105
5073:
5074: \nhi
5075: Whysong D, Antonucci R. 2004. {\it Ap. J.} 602:116
5076:
5077: \nhi
5078: Willner SP, Ashby MLN, Barmby P, Fazio GG, Pahre M, et al. 2004.
5079: {\it Ap.~J.~Suppl.} 154:222
5080:
5081: \nhi
5082: Willner SP, Elvis M, Fabbiano G, Lawrence A, Ward MJ. 1985. {\it Ap. J.} 299:443
5083:
5084: \nhi
5085: Wilson AS 1979. {\it Proc. Roy. Soc. London} A366:461
5086:
5087: \nhi
5088: Worrall DM, Birkinshaw M. 1994. {\it Ap. J.} 427:134
5089:
5090: \nhi
5091: Wo\'zniak PR, Zdziarski AA, Smith D, Madejski GM, Johnson WN. 1998. {\it MNRAS}
5092: 299:449
5093:
5094: \nhi
5095: Wrobel JM. 1991. {\it Astron. J.} 101:127
5096:
5097: \nhi
5098: Wrobel JM, Fassnacht CD, Ho LC. 2001. {\it Ap. J.} 553:L23
5099:
5100: \nhi
5101: Wrobel JM, Heeschen DS. 1991. {\it Astron. J.} 101:148
5102:
5103: \nhi
5104: Wrobel JM, Ho LC. 2006. {\it Ap. J.} 646:L95
5105:
5106: \nhi
5107: Wrobel JM, Terashima Y, Ho LC. 2008. {\it Ap. J.} in press
5108:
5109: \nhi
5110: Wu Q, Cao X. 2005. {\it Ap. J.} 621:130
5111:
5112: \nhi
5113: Wu Q, Yuan F, Cao X. 2007. {\it Ap. J.} 669:96
5114:
5115: \nhi
5116: Xu Y, Cao X-W. 2007. {\it ChJAA} 7:63
5117:
5118: \nhi
5119: Yan R, Newman JA, Faber SM, Konidaris N, Koo D, Davis M. 2006. {\it Ap. J.}
5120: 648:281
5121:
5122: \nhi
5123: Yaqoob T, Serlemitsos PJ, Ptak A, Mushotzky RF, Kunieda H, Terashima Y. 1995.
5124: {\it Ap. J.} 455:508
5125:
5126: \nhi
5127: Young AJ, Nowak MA, Markoff S, Marshall HL, Canizares CR. 2007. {\it Ap. J.}
5128: 669:830
5129:
5130: \nhi
5131: Yu Q, Tremaine S. 2002. {\it MNRAS} 335:965
5132:
5133: \nhi
5134: Yuan F. 2007. In {\it The Central Engine of Active Galactic Nuclei}, ed. LC
5135: Ho, J-M Wang, p. 95. San Francisco: ASP
5136:
5137: \nhi
5138: Yuan F, Cui W. 2005. {\it Ap. J.} 629:408
5139:
5140: \nhi
5141: Yuan F, Markoff S, Falcke H. 2002. {\it Astron. Astrophys.} 383:854
5142:
5143: \nhi
5144: Yuan F, Markoff S, Falcke H, Biermann PL. 2002. {\it Astron. Astrophys.}
5145: 391:139
5146:
5147: \nhi
5148: Yuan F, Narayan R. 2004. {\it Ap. J.} 612:724
5149:
5150: \nhi
5151: Zakamska NL, Strauss MA, Krolik JH, Collinge MJ, Hall PB, et al. 2003.
5152: {\it Astron. J.} 126:2125
5153:
5154: \nhi
5155: Zamorani G, Henry JP, Maccacaro T, Tananbaum H, So\l tan A, et al. 1981.
5156: {\it Ap. J.} 245:357
5157:
5158: \nhi
5159: Zezas A, Birkinshaw M, Worrall DM, Peters A, Fabbiano G. 2005. {\it Ap. J.}
5160: 627:711
5161:
5162: \nhi
5163: Zhang X-G, Dultzin-Hacyan D, Wang T-G. 2007. {\it MNRAS} 374:691
5164:
5165: \nhi
5166: Zhang Y, Gu Q-S, Ho LC. 2008. {\it Astron. Astrophys.} submitted
5167:
5168: %\end{thebibliography}
5169:
5170: \vfill\eject
5171:
5172: %\listoffigures
5173:
5174: \end{document}
5175: