<<

Measurement of Masses in Active Galactic Nuclei

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the

Graduate School of The Ohio State University

By

Christopher A. Onken

*****

The Ohio State University

2005

Dissertation Committee: Approved by

Professor Bradley M. Peterson, Adviser

Professor Richard W. Pogge Adviser Graduate Program Professor Smita Mathur ABSTRACT

We investigate the calibration and application of reverberation mapping techniques for determining black hole (BH) masses in active galactic nuclei (AGNs).

We present revised BH mass determinations for several AGNs based on the use of updated methodologies with archival data, demonstrating significant reductions in the sizes of the BH mass uncertainties. Moreover, the study of the Seyfert 1 ,

NGC 3783, shows that the gas in the broad-line region of this AGN obeys the .

We use measurements of stellar velocity dispersions, σ∗, in AGNs and the assumption that AGNs follow the same relation between BH mass and σ∗ as quiescent to provide the first empirical calibration for reverberation-based

BH masses. We also attempt to determine an independent calibration of these masses by studying the reverberation-mapped AGN, NGC 4151, with ground- and space-based observations, and by trying to constrain the BH mass through modeling of the galaxy’s .

We estimate the BH masses and bolometric luminosities in 400 AGNs ∼ selected from the multi-wavelength AGN and Galaxy Evolution Survey (AGES),

ii where the BH masses are calculated from scaling relationships that have grown out of reverberation mapping. We find the distribution of Eddington ratios at fixed luminosity to be sharply peaked around a value of 1/3, with a dispersion of just

0.3 dex. The distribution of Eddington ratios at fixed mass looks to be similarly narrow, and we are able to confirm a drop in the underlying distribution at low

Eddington ratios for certain combinations of redshift and BH mass—all previous studies in these redshift mass bins are affected by selection effects at low Eddington − ratio (as are the AGES data in lower mass or higher redshift bins). The dominance of AGN at rates relatively close to the Eddington limit has important implications for the growth of BHs and the joint evolution of BHs and their host galaxies.

iii Dedicated to all the helping hands along the way...

the ones who guided and the ones who pushed.

iv ACKNOWLEDGMENTS

Innumerable people deserve to be acknowledged for their influence, personal and professional, on the incredible experience I have had. Thank you to my adviser,

Brad Peterson, for taking a fellow Minnesotan under your wing and showing me how to work as an . Thanks to Rick Pogge for your willingness to share your time and encyclopedic knowledge of topics astrophysical and otherwise. Thank you to all of the Astronomy Department faculty members, postdocs, and staff at The

Ohio State University–I’ve pestered every one of you with questions at some time over the last five years and you have all been willing to stop whatever you were doing and provide me with answers. To the OSU faculty with whom I’ve been fortunate enough work directly (in chronological order: Darren DePoy, Don Terndrup, Brad,

Jordi Miralda-Escud´e, Smita Mathur, Rick, Chris Kochanek, Andy Gould, and

David Weinberg), I thank you for fostering my development as a scientist. And to the postdocs who have given me so much guidance, Marianne Vestergaard and

Matthias Dietrich, thank you for your help.

To my fellow OSU grad students, I am grateful to call all of you my friends.

The space it would take to explain your importance to my time in Columbus would

v double the size of this document. Let me simply say that you are a magnificent bunch of people, and I look forward to seeing all of you again.

Prior to my arrival in Columbus, I benefited from the advice and instruction offered by several people. I am indebted to Evan Skillman, who kindly took the time to answer question after question in Intro Astronomy and in the years that followed; and to Ted Bowell, Bruce Koehn, and Brian Skiff, who gave me my first taste of observing; and to David Alves and Howard Bond, who sparked continuing research interests that are beyond the scope of this dissertation; and to John Dickey, who both demonstrated how to be an effective teacher and patiently advised me on my senior thesis; and to Terry Jones, who first suggested applying to OSU for grad school.

Finally, I am pleased to be able to acknowledge the role of my wonderful family.

Without your love, interest, and support, I never would have come so far. Thank you.

vi VITA

August 14, 1978 ...... Born – Springfield, IL, USA

2000 ...... B.S., Physics (magna cum laude), University of Minnesota

2000 ...... B.S., (summa cum laude), University of Minnesota

2000 – 2005 ...... Graduate Teaching and Research Associate, The Ohio State University

2003 ...... M.S., The Ohio State University

PUBLICATIONS

Research Publications

1. H. E. Bond, D. R. Alves, and C. Onken, “CCD of the NGC 5986 and Its Post-Asymptotic Giant Branch and RR Lyrae Stars”, AJ, 121, 318, (2001).

2. C. A. Onken, and B. M. Peterson, “The Mass of the Central Black Hole in the NGC 3783”, ApJ, 572, 746, (2002).

3. B. M. Peterson, et al. (41 authors, incl. C. A. Onken), “Steps toward Determination of the Size and Structure of the Broad-Line Region in Active Galactic Nuclei. XVI. A 13 Year Study of Spectral Variability in NGC 5548”, ApJ, 581, 197, (2002).

vii 4. C. A. Onken, B. M. Peterson, M. Dietrich, A. Robinson, and I. M. Salamanca, “Black Hole Masses in Three Seyfert Galaxies”, ApJ, 585, 121, (2003).

5. C. A. Onken, and J. Miralda-Escud´e, “History of Hydrogen Reionization in the Cold Model”, ApJ, 610, 1, (2004).

6. B. M. Peterson et al. (12 authors, incl. C. A. Onken), “Central Masses and Broad-Line Region Sizes of Active Galactic Nuclei. II. A Homogeneous Analysis of a Large Reverberation-Mapping Database”, ApJ, 613, 682, (2004).

7. C. A. Onken, L. Ferrarese, D. Merritt, B. M. Peterson, R. W. Pogge, M. Vestergaard, and A. Wandel, “Supermassive Black Holes in Active Galactic Nuclei. II. Calibration of the Black Hole Mass-Velocity Dispersion Relationship for Active Galactic Nuclei”, ApJ, 615, 645, (2004).

FIELDS OF STUDY

Major Field: Astronomy

viii Table of Contents

Abstract ...... ii

Dedication ...... iv

Acknowledgments ...... v

Vita ...... vii

List of Tables ...... xii

List of Figures ...... xiv

Chapter 1 Introduction 1

1.1 Active Galactic Nuclei ...... 1

1.2 Reverberation Mapping ...... 3

1.2.1 Time Delay ...... 4

1.2.2 Velocity Width ...... 9

1.2.3 BH Mass ...... 10

1.3 Stellar Dynamical BH Masses ...... 11

1.4 Scaling Relations ...... 14

1.5 Focus of This Study ...... 16

1.6 Published Work ...... 18

Chapter 2 Reverberation Mapping of NGC 3783 23

ix 2.1 Introduction ...... 23

2.2 Data ...... 24

2.2.1 UV Data ...... 24

2.2.2 Optical Data ...... 26

2.3 Methods and Techniques ...... 28

2.3.1 Spectral Analysis ...... 28

2.3.2 Light Curve Analysis ...... 29

2.4 Results ...... 30

2.4.1 Line widths ...... 30

2.4.2 Time delays ...... 30

2.5 Discussion ...... 33

Chapter 3 Additional Reverberation Mapping Studies 70

3.1 Introduction ...... 70

3.2 Observations and Data Reduction ...... 71

3.2.1 NGC 3227 ...... 72

3.2.2 NGC 3516 ...... 73

3.2.3 NGC 4593 ...... 74

3.3 Analysis ...... 75

3.4 Mass Calculations ...... 76

3.4.1 NGC 3227 ...... 77

3.4.2 NGC 3516 ...... 77

3.4.3 NGC 4593 ...... 78

3.5 The BH Mass—Stellar Velocity Dispersion Relationship ...... 78

x 3.6 Conclusion ...... 80

Chapter 4 The Black Hole Mass Stellar Velocity Dispersion Relation − for AGNs 89

4.1 Introduction ...... 89

4.2 Observations and Data Reduction ...... 91

4.2.1 Data Reduction and Analysis ...... 93

4.3 Results and Discussion ...... 94

4.3.1 Measuring Velocity Dispersions ...... 94

4.3.2 Measuring Virial Products ...... 95

4.3.3 The BH Mass–Velocity Dispersion Relation ...... 96

4.3.4 BLR Models ...... 100

4.3.5 Gravitational Redshifts ...... 101

4.3.6 Velocity Dispersion Versus FWHM([O III]λ5007 A)˚ ...... 103

4.4 Conclusion ...... 104

Chapter 5 Stellar Dynamics of NGC 4151 113

5.1 Introduction ...... 113

5.2 Stellar Dynamics Overview ...... 115

5.3 Imaging ...... 117

5.3.1 ACS/HRC ...... 118

5.3.2 MDM ...... 118

5.4 Spectroscopy ...... 119

5.4.1 HST/STIS ToO ...... 120

5.4.2 Ground-based ...... 122

xi 5.5 Analysis & Results ...... 125

5.5.1 Surface Brightness Profile Modeling ...... 125

5.5.2 Gauss-Hermite Parameter Estimation ...... 129

5.5.3 Orbit Modeling ...... 131

5.6 Discussion ...... 133

Chapter 6 Black Hole Masses and Eddington Ratios at 0.3 < z < 4 156

6.1 Introduction ...... 156

6.2 Data ...... 158

6.2.1 Completeness ...... 160

6.3 Analysis ...... 162

6.3.1 Line Width ...... 163

6.3.2 Continuum Luminosity ...... 166

6.3.3 Calibrating the Mg IIRelation ...... 167

6.3.4 Bolometric Luminosity Calculation ...... 168

6.4 Results ...... 169

6.4.1 Luminosity-Redshift Bins ...... 171

6.4.2 Mass-Redshift Bins ...... 175

6.5 Discussion ...... 179

Chapter 7 Conclusion 197

7.1 Future Work ...... 200

Bibliography 202

xii List of Tables

2.1 NGC 3783 Continuum and Emission Line Wavelength Limits . . . . . 42

2.2 NGC 3783 UV Continuum Flux Data ...... 43

2.3 NGC 3783 UV Emission Line Flux Data ...... 51

2.4 NGC 3783 UV Velocity Data ...... 64

2.5 NGC 3783 Optical Flux Data ...... 65

2.6 NGC 3783 Sampling Statistics ...... 68

2.7 NGC 3783 Cross-Correlation Results ...... 69

3.1 Sampling Statistics ...... 85

3.2 Cross-Correlation Results ...... 86

3.3 Velocity Width Results ...... 87

3.4 Reverberation Masses ...... 88

4.1 Observing Log ...... 109

4.2 Velocity Dispersion Measurements ...... 110

4.3 MBH σ∗ Data ...... 111 −

4.4 MBH σ∗ Fitting Results ...... 112 − 5.1 GALFIT Results for MDM Image of NGC 4151 ...... 153

5.2 GALFIT Results for ACS Image of NGC 4151 ...... 154

5.3 MGE Results for Combined Imaging Data of NGC 4151 ...... 155

xiii 6.1 Gaussian Parameters of Data and Fits to Data ...... 196

xiv List of Figures

1.1 AGN structure diagram...... 20

1.2 Ellipsoidal light echo...... 21

1.3 Velocity width vs. time delay...... 22

2.1 NGC 3783 mean and rms UV spectra...... 36

2.2 NGC 3783 mean and rms optical spectra...... 37

2.3 NGC 3783 UV and optical light curves...... 38

2.4 Time lag versus centroid threshold for NGC 3783...... 39

2.5 Light curve cross-correlation for NGC 3783...... 40

2.6 Line width versus time lag for NGC 3783...... 41

3.1 Mean and rms Hα spectra...... 81

3.2 Mean and rms Hβ spectra...... 82

3.3 Light curve cross-correlation...... 83

3.4 Black hole mass versus stellar velocity dispersion...... 84

4.1 Stellar absorption line spectra...... 106

4.2 Virial product versus stellar velocity dispersion...... 107

4.3 Stellar velocity dispersion versus O III line width...... 108

5.1 ACS image of NGC 4151 ...... 135

5.2 MDM image of NGC 4151 ...... 136

xv 5.3 Radial color profile in NGC 4151 ...... 137

5.4 Optical light curve of NGC 4151 ...... 138

5.5 STIS spectrum of NGC 4151 ...... 139

5.6 Residual MDM image of NGC 4151 ...... 140

5.7 Residual ACS image of NGC 4151 ...... 141

5.8 MGE fit to ACS image ...... 142

5.9 MGE fit to MDM image ...... 143

5.10 Example pPXF fit to MMT data ...... 144

5.11 Example pPXF fit to KPNO data ...... 145

5.12 GH parameter fits to MMT spectra ...... 146

5.13 GH parameter fits to KPNO spectra ...... 147

5.14 Orbit model results with BH ...... 148

5.15 Orbit model results with BH (cont.) ...... 149

5.16 Orbit model results with BH (cont.) ...... 150

5.17 Orbit model results with BH (cont.) ...... 151

5.18 Orbit model results with BH (cont.) ...... 152

6.1 AGES-I completeness ...... 184

6.2 Method of FWHM measurement ...... 185

6.3 Examples of AGES-I spectra ...... 186

6.4 Recalibration of Mg II scaling relation ...... 187

6.5 Dependence of BH masses on Mg II FWHM ...... 188

6.6 AGN luminosity versus redshift ...... 189

6.7 BH mass versus AGN luminosity ...... 190

xvi 6.8 Eddington ratio distribution in redshift-luminosity bins ...... 191

6.9 BH mass versus AGN luminosity for rejected objects ...... 192

6.10 Eddington ratio distribution in redshift-mass bins ...... 193

6.11 Eddington ratio distributions not biased by selection effects ...... 194

6.12 Eddington ratio distributions accounting for removed objects . . . . . 195

xvii Chapter 1

Introduction

The last five years have seen rapid advancement in our understanding of the role that supermassive black holes (BHs) play in the evolution of galaxies. Part of this progress has come from increased confidence in the techniques for measuring

BH masses, and part has arisen from the discovery of new correlations between

BH masses and properties of their host galaxies. Some of these developments have regarded quiescent, or inactive, BHs, and some have come from the study of active galactic nuclei (AGNs; 1.1). §

This dissertation employs several techniques used for measuring BH masses, examining the calibration of such methods and using them to provide BH mass estimates for a large survey of AGNs.

1.1. Active Galactic Nuclei

The observational characteristics of AGNs in the UV and optical bands can be summarized as follows (although certain exceptions exist to many of these generalizations). In imaging data, AGNs (as the moniker suggests) appear as bright,

1 unresolved point-sources at the centers of galaxies. The luminosity of the AGN places it into one of two categories: (brighter: MB < 21.5 mag; rarer in the − local universe) or Seyfert galaxy (fainter: MB > 21.5 mag; more common locally). − Spectroscopically, AGNs are further divided into two classes: Types 1 and 2. Type 1

AGNs typically show broad emission lines of highly ionized atomic species, narrower emission lines (sometimes of the same species as the broad lines, but also of lower ionization atoms), a power-law continuum slope, and, in some cases, absorption features at velocities indicative of fast-moving gas. Type 2 AGNs show only the narrow emission lines and typically a weaker continuum. Nearby galaxies that host

AGNs offer us the best opportunity to study these systems in detail, but often prove to be exceedingly complicated in their structure and kinematics.

The standard model for AGNs consists of a central, supermassive BH (with a

6−10 mass of 10 M¯) surrounded by highly ionized gas, some of which is accreting ∼ onto the BH (and thereby powering the system), and some of which is being driven away from the BH in a wind or outflow (by radiative, electromagnetic, or kinetic processes). A recent incarnation of such a model is diagrammed in Figure 1.1.

These kinds of models have come to be favored over earlier versions (in which the broad-line region (BLR) was made up of many small clouds) because they provide a simpler explanation for a variety of AGN phenomena (as suggested by the number of labels in Fig. 1.1) and because very high spectral resolution has yet to see the broad emission lines break into clumps of material with discrete velocities,

2 as one would expect for “clouds” of any appreciable size. The smoothness of the emission lines is easier to understand in a disk-wind model.

Whatever the precise geometry and kinematic structure of the BLR, the gas close enough to the BH to produce the emission lines of highly ionized species is strongly influenced by the gravity of the BH. This was demonstrated most convincingly with the advent of reverberation mapping.

1.2. Reverberation Mapping

The fundamental idea of reverberation mapping is that by watching a light echo propagate through an AGN, you can learn about the spatial and kinematic distribution of the BLR gas (Blandford & McKee 1982; Peterson 2001). In the simplest case, one can imagine a rapid burst of ionizing photons being emitted from very close to the BH (something akin to a flashbulb going off), and as these photons expand outward in a shell, they encounter ions of the BLR and increase the strength of their emission lines. After the shell of photons has passed, the emission lines return to their prior state (determined by the steady-state ionization conditions of the BLR).

Because the BLR is not yet spatially resolved with interferometric techniques, reverberation mapping is currently a spectroscopic method. As such, the main properties measured by such studies are the light-travel-time (i.e. the distance)

3 between the BH and the BLR (determined from the time delay between variations in the ionizing continuum and the emission line response), and the velocity of the gas in the BLR (determined by the width of the emission line). The key elements of these measurements are described below.

1.2.1. Time Delay

While the light echo propagates through the BLR in a radial direction and impinges on all gas at a given distance, R, from the BH at the same, unless the

BLR takes the form of a perfectly face-on , the 3-D geometry of the system will spread out the emission line response from a given R over a time span of up to

2R/c, where c is the speed of light. Moreover, because the emission line arises from a range in radius, the line strength at a given time will be the sum of the gas response at each point that intersects the expanding ellipsoid that characterizes the material illuminated at a fixed value of the time delay (cross-sections of such ellipsoids are shown in Fig. 1.2).

Formally, the responses of the emission line to the continuum changes are determined by the product of the ionizing continuum light curve and the “transfer function” that describes the gas kinematic and spatial distribution:

∞ L(t) = Ψ(τ)C(t τ)dτ, (1.1) Z0 − 4 where L(t) is the emission line flux as a function of time t, Ψ is the transfer function,

C(t τ) is the continuum flux at a point in time that is τ earlier than t. The transfer − function is also a function of line-of-sight velocity, but the quality of data needed to perform velocity-resolved reverberation mapping (dubbed “2-D reverberation mapping”) requires observational resources of a level only practical for a dedicated instrument (e.g., Kronos; Peterson et al. 2004b).

Reverberation mapping campaigns can be rather demanding of observational resources. To accurately resolve the timing of the BLR response to continuum

fluctuations, the sampling interval between observations must be much shorter than the delay time and the duration should exceed the time delay by more than a factor of two. Beyond these requirements, AGN variability is irregular, often entering relatively stable states of significant length. A strong event (either rising or falling) in the continuum light curve is essential for measuring an accurate time delay—without an extremely long campaign, a bit of luck is crucial.

In practice, almost all reverberation mapping data has been insufficient to constrain the transfer function, and reverberation analysis has instead relied on cross-correlation techniques. The light curves are measured for the emission line and the continuum region (measured over some line-free wavelength range) and then cross-correlated with one of two methods: the interpolated cross-correlation function

(ICCF), or the discrete correlation function (DCF).

5 The ICCF method uses linear interpolation between points in the light curve to allow calculation of the cross-correlation coefficient at any amount of time delay between the continuum light curve and the emission line response (Gaskell

& Peterson 1987; White & Peterson 1994). The typical implementation of this technique (1) shifts one dataset; (2a) interpolates between points of the shifted light curve; (3a) calculates the correlation coefficient; (2b) interpolates between the points in the other light curve; (3b) calculates the correlation coefficient; (4) takes the mean of the two cross-correlation coefficients.

Alternatively, the DCF (Edelson & Krolik 1988) uses an unmodified version of one light curve and the mean flux value of data within a bin centered on whatever time delay is being considered. The bin width can be modified for a given dataset to balance between resolving the time delay and providing enough data points per bin. A variant of this technique, the Z-Transformed DCF (ZDCF; Alexander 1997), adapts the bin size to produce the desired number of points in each bin. However,

Gaskell (1994) notes that the DCF also relies on interpolation, but does so in the correlation function, rather than the original time series.

Both methods operate by calculating the correlation coefficient, r:

(x x¯)(y y¯) r = i − i − (1.2) (xP x¯)2 (y y¯)2 i − i − ³qP ´ ³qP ´ 6 where (xi, yi) is the ith flux value pair (i.e., a flux point in one light curve and the corresponding interpolated or binned flux value), and x¯ and y¯ are the mean fluxes in the x and y light curves.

Once the ICCF or DCF has been computed for a range of time delay values, the task remains of identifying the best estimate of the “true” time delay. One approach is to simply take the peak, τp, the time delay value with the highest value of the cross-correlation coefficient. While the determination of this quantity from the data is unambiguous, it can be extremely sensitive to small changes in the data. A preferable method is to measure the centroid of the correlation function,

τc. One virtue of using the centroid is that it is more clearly related to the transfer function: in the limit of infinite-duration light curves, the centroid of the correlation function is equal to the centroid of the transfer function. In practice, the centroid is calculated based only on values above some threshold, typically 80% of the peak r value. Koratkar & Gaskell (1991a) show how the choice of this threshold value can impact the calculated centroid.

To assess the uncertainties in the time lags, we used the Monte Carlo method described by Peterson et al. (1998). This technique for model-independent error estimation consists of two components, each testing for a separate contribution to the cross-correlation uncertainty. To account for the uncertainty in an individual

flux measurement, each data point in the light curve is altered by a random Gaussian deviation that corresponds to the flux error. The result of many such realizations,

7 referred to as “flux randomization” (FR), should yield average values equal to the original data with standard deviations given by the original uncertainties.

Secondly, the effects of non-uniform temporal sampling of the AGN fluctuations are investigated with “random subset selection” (RSS). Given a sample of N observations, N data points are randomly chosen from the set (ignoring whether they have been chosen previously). While DCF (and ZDCF) analyses can weight multiply-selected data, the ICCF does not consider the flux uncertainties and simply excludes the redundant data points. Ignoring these data reduces the set by N/e ∼ on average and so should yield a wider range of peak lags from the ICCF. Repeated

Monte Carlo realizations (typically at least 103, combining the FR/RSS methods for each calculation) are used to create a cross-correlation peak distribution (CCPD;

Maoz & Netzer 1989), which provides an empirical measurement of the uncertainties for both τc and τp, based on the τ values which encompass 68% of the realizations.

For well-sampled light curves, the ICCF is generally in good agreement with the results from the DCF and usually has much smaller uncertainties in the time lag, although it is more susceptible to producing spurious correlations (White &

Peterson 1994; Litchfield et al. 1995).

8 1.2.2. Velocity Width

The other quantity used in reverberation mapping is the width of the variable emission lines, V . Many early reverberation mapping efforts calculated V from a mean spectrum created from the entire dataset (typically having a very high signal-to-noise ratio, S/N). However, a better way to isolate the BLR gas that is giving rise to the time delay measurement is to construct the rms spectrum. The rms spectrum will remove any constant spectral features, e.g., narrow emission lines arising from much larger distances from the BH (and so varying over much slower time scales), and host galaxy contribution. The advantage of only measuring the variable part of the spectrum comes at the cost of a lower S/N spectrum from which to measure V . Previous work examining the difference between using the mean and rms spectra have not produced significantly different results (e.g., Kaspi et al. 2000), but in principle the rms spectrum should better trace the gas with which we are concerned.

The most common measurement of V is the full-width-at-half-maximum

(FWHM), which, for a Gaussian emission line shape, is equal to 2.35σG, where σG is the dispersion of the Gaussian. Recently, Peterson et al. (2004a) advocated the use of σline, the line dispersion, defined as:

2 2 2 λ F (λ)dλ λF (λ)dλ σline = , (1.3) R F (λ)dλ − "R F (λ)dλ # R R 9 where λ is the wavelength, F (λ) is the line flux at λ, and the integrals extend over the wavelength range of the line. Measurements of σline are preferable to the FWHM because they are well-defined for any line profile and are less sensitive to features near the line center (e.g., residual narrow-line components).

1.2.3. BH Mass

The virial theorem predicts a relationship between time delay and velocity width of a form, τ V −2. For objects in which multiple emission lines have been ∝ reverberation mapped, the data are in excellent agreement with the virial expectation

(Fig. 1.3). The time delay and the velocity width can be used to estimate the BH mass via the following relationship:

c τV 2 M = f , (1.4) BH G where c is the speed of light, G is Newton’s constant, and f is a factor that accounts for the geometry and detailed kinematics of the BLR gas. The quantity (c τV 2/G) is called the virial product, and is equal to the BH mass to within the factor f. The close fit to a virial relationship among the different emission lines of a given object in Figure 1.3 means that the lines give a consistent value of the virial product.

The value of the scale factor f has not been well constrained. The most commonly used value has been 0.75 (Netzer 1990), which comes from assuming that

10 the velocity distribution in the BLR is isotropic and that the circular velocity at the radius given by τ is equal to half of the FWHM. A few authors have pursued other estimates of f (e.g., McLure & Dunlop 2001; Wu & Han 2001; McLure & Dunlop

2002), but these rely on modeling with some severe assumptions. Chapter 4 provides the first empirical calibration of the average value of f that is independent of any

BLR modeling.

1.3. Stellar Dynamical BH Masses

In a few dozen quiescent galaxies, the technique of stellar dynamics has been used to estimate BH masses. The general principles of this method will be described here, leaving a more detailed treatment for Chapter 5.

In contrast to the approach used to measure the mass of the BH at the center of the , the motions of individual stars cannot be observed in galaxies beyond the . Instead, the gravitational influence of the BH must be discerned from the collective distribution of orbital motions for all stars within the observed region of the galaxy (where the observed region is commonly defined by the slit width of a long-slit spectroscopic observation and the spatial extraction window of the spectrum). Moreover, because the velocity information is coming from the

Fizeau shift (the photon equivalent of a sound wave’s Doppler shift; Fizeau 1848), only the component of the velocity along the line of sight is measurable. These shifts

11 are determined with stellar absorption lines arising in the atmospheres of red giant stars (spectral classes G and K). The lines are intrinsically much narrower than the observed profiles in the centers of galaxies, so the broadening is taken to be a measure of the distribution of velocities among the different stars. The line-of-sight velocity distribution (LOSVD) is a composite, weighted by luminosity and line strength, of all stars within the observed window. Studies of stellar dynamics use the detailed shape of the LOSVD as a constraint for modeling the mass of the central

BH.

The first step in the modeling process is to build a picture of the mass distribution of the stars in the galaxy. Images of the galaxy are used to estimate the total light being emitted by the stars. One then assumes (often based on the observed colors of the stars) a value for the mass-to-light ratio (M/L), and this provides a 2-D map of the stellar mass distribution. The 2-D mass distribution is then deprojected into a 3-D mass profile based on assumptions about the intrinsic shape and orientation of the galaxy.

An integral of motion is any function of a particle’s 3-D position and 3-D velocity that is constant along a given orbit. Orbits in galactic gravitational potentials are often defined by three integrals of motion: the energy of the orbit, the vertical angular momentum of the orbit, and a third quantity that has no simple form but can implemented numerically (see, e.g., Ollongren 1962; Richstone 1982;

Famaey et al. 2002). Applying the technique of orbit superposition (Schwarzschild

12 1979; Cretton et al. 1999), within the 3-D mass distribution for the galaxy, one selects a series of regularly spaced positions and values for the parameters defining the integrals of motion. Test particles are placed at these positions in phase space and followed for several hundred orbits through the potential. This is done for thousands of different orbits. A spatial grid is overlaid on the galaxy and the amount of time an orbit spends inside each box of the grid and its line-of-sight velocity within the box are tabulated. The linear combination of orbits that provides the best fit to the observed LOSVD is then determined.

To the original mass model, one can add the gravity of a central BH with a given mass and then recalculate the orbits and the best fit to the observations. The relative quality of the fits from one BH mass to the next provides an estimate of the most likely value of the BH mass and its uncertainty.

Recently, Silge et al. (2005) made an estimate for the BH mass in the Seyfert 2 galaxy, Centaurus A (NGC 5128)—the first stellar dynamic BH measurement in an

AGN. This measurement was made possible by the combination of a large telescope

(the Gemini-South 8 m), sensitive near-infrared instruments (to penetrate the broad swath of obscuring material in front of the nucleus of Cen A), and the modern computing resources required for the timely integration of 10,000 orbits for each of many tens of parameter sets. This measurement is important because it provides a needed (although strongly model-dependent) point of comparison with estimates of the BH mass from gas dynamics (Marconi et al. 2001; H¨aring-Neumayer et al. 2005;

13 Marconi et al. 2005). What is still missing, however, is a direct point of comparison between stellar dynamics and reverberation mapping. In Chapter 5, we seek to address this missing link.

1.4. Scaling Relations

The remarkable success of the reverberation mapping approach (e.g., Chap. 2) implies that most of the flux in each emission line is coming from a narrow spatial extent, which is able to reverberate quite coherently. The locally optimally emitting cloud model (Baldwin et al. 1995) provides an explanation for this fact, even if cloud-based models have fallen out of favor. Baldwin et al. (1995) show that the maximum flux from a given emission line is produced from a relatively small range in the combination of gas density and ionizing photon density. While AGN conditions may span a large range in, say, ionizing luminosity (both on an object-by-object basis, and as a function of distance), the region of parameter space that can give rise to a strong emission line at any particular luminosity remains tightly confined.

Hence, most of the flux of a given line arises from a small radial span in the BLR, and reverberates in a coherent manner.

That the emission lines are dominated by the small regions where they are most favorably formed explains the incredible similarity among AGN spectra across

14 a wide range of luminosity and redshift, and also offers the possibility of exploiting this fact in the vast number of AGNs which have not been reverberation mapped.

In addition to the tight relationship between time delay and line width described above and illustrated in Chapter 2, reverberation mapping discovered a strong correlation between the radius (i.e. time delay) and the continuum luminosity.

The radius-luminosity (R L) relationship has been established using the optical − continuum at 5100 A˚ and the Hβ emission line (Kaspi et al. 1996, 2000, 2005), and deviates in a slight, but statistically significant way from the na¨ıve prediction of photoionization modeling that R L0.5. This difference may be caused by the ∝ non-linear relationship between the ionizing luminosity in the UV and the optical continuum luminosity. In the local AGN, NGC 5548, Peterson et al. (2002) find that the optical luminosity is proportional to L0.56, which then leads to an R L UV − UV relationship that agrees with the expectations from simple photoionization. Another possibility is that host galaxies have contaminated the optical luminosity estimates of these AGNs and have skewed the observed R L relation away from its true − nature (M. Bentz et al., in preparation).

The existence of the R L relationship means that a measurement of the − continuum luminosity from a single spectrum allows one to estimate the characteristic radius of the emission line gas. Coupled with a single-epoch measurement of the line width, this yields an estimate of the BH mass. To allow BH mass measurements at higher redshifts, these scaling relations have been extended to emission lines at

15 shorter rest wavelengths: Mg II λ2800 (McLure & Jarvis 2002) and C IV λ1549

(Vestergaard 2002). However, it should be noted that the new relationships rely exclusively on intercalibration with the Hβ relation, and are not established independently by reverberation mapping of Mg II or C IV. Several C IV reverberation studies have been conducted, which typically give virial products consistent with those from Hβ (see Peterson et al. 2004a). However, reverberation results for

Mg II have only been published for two objects (NGC 3783: Reichert et al. 1994;

NGC 5548: Clavel et al. 1991, Dietrich & Kollatschny 1995), and these give rather uncertain and somewhat conflicting indications as to where Mg II arises in the BLR.

Although subject to greater uncertainties than full reverberation mapping studies, the BH masses produced by these scaling relationships provide an opportunity to estimate the distribution of BH masses for AGNs observed in large spectroscopic surveys (i.e., the Sloan Digital Sky Survey (SDSS; McLure & Dunlop

2004; M. Vestergaard, in preparation), the Two-Degree Field QSO Redshift survey

(2QZ; Corbett et al. 2003), and the AGN and Galaxy Evolution Survey (AGES;

Chap. 6)).

1.5. Focus of This Study

This dissertation examines the key steps in the process of measuring BH masses in AGNs via reverberation mapping and tries to calibrate those masses through two

16 independent means. Reverberation-derived scaling relations are then used to study the population of AGNs at z = 0.5 4. −

In Chapter 2, we present our reverberation mapping analysis of a nearby AGN,

NGC 3783, in which we revise the estimate of its BH mass and dramatically reduce the uncertainty in the mass.

Following similar procedures for three AGNs with lower quality data, we show in Chapter 3 that these objects also benefit from the improved reverberation mapping techniques developed since the data were originally obtained.

Chapter 4 exploits the tight correlation between the stellar velocity dispersions of galaxies and their BH masses to provide the first empirical calibration of reverberation-based masses.

An independent test of this calibration is studied in Chapter 5, in which we attempt to measure the BH mass in NGC 4151, a reverberation-mapped AGN, via modeling of stellar dynamics near the BH.

In Chapter 6, scaling relationships established from reverberation mapping results are applied to over 400 AGNs from the AGN and Galaxy Evolution Survey

(AGES), and point to a very narrow distribution of Eddington ratios among AGNs.

This result suggests that BHs tend to be “on” as AGNs, or “off” as quiescent galaxies with a very sharp (i.e., quick) transition between the two states.

17 Chapter 7 addresses the implications of all of these results and notes a few avenues of future work that would build upon these studies and further test the conclusions we have drawn.

1.6. Published Work

Chapter 2 has been published as Christopher A. Onken, and Bradley M.

Peterson, “The Mass of the Central Black Hole in the Seyfert Galaxy NGC 3783”,

2002, The Astrophysical Journal, 572, 746.

Chapter 3 has been published as C. A. Onken, B. M. Peterson, M. Dietrich,

A. Robinson, and I. M. Salamanca, “Black Hole Masses in Three Seyfert Galaxies”,

2003, The Astrophysical Journal, 585, 121.

Chapter 4 has been published as Christopher A. Onken, Laura Ferrarese, David

Merritt, Bradley M. Peterson, Richard W. Pogge, Marianne Vestergaard, and Amri

Wandel, “Supermassive Black Holes in Active Galactic Nuclei. II. Calibration of the

Black Hole Mass-Velocity Dispersion Relationship for Active Galactic Nuclei”, 2004,

The Astrophysical Journal, 615, 645.

Chapter 6 will be published as Juna A. Kollmeier, Christopher A. Onken,

Christopher S. Kochanek, Andrew Gould, David H. Weinberg, Matthias Dietrich,

Richard Cool, Arjun Dey, Daniel J. Eisenstein, Buell T. Jannuzi, Emeric Le Floc’h,

18 and Daniel Stern, “Black Hole Masses and Eddington Ratios at 0.3 < z < 4”, 2005,

The Astrophysical Journal, submitted.

The previously published works have been formatted to better conform to

Graduate School rules, and edited to improve continuity and clarity.

19 20

Fig. 1.1.— Cartoon of AGN structure from Elvis (2000). To observer

0.01r/c

0.5r/c

2r/c r/c 1.5r/c

Fig. 1.2.— Cross-sections of reverberating light echo. As time advances, the signal reaching the observer from radiation emitted near the BH is defined by the surface of an expanding ellipsoid with foci at the BH and the observer, and a semi-major axis equal to (0.5d + 2ct), where d is the distance between the BH and the observer, c is the speed of light, and t is the time elapsed since the radiation reached the observer directly. From Peterson (2001).

21 1016 cm 1017 cm 30,000

3C 390.3

10,000 ) 1 − s m k ( M H W F V 3000 NGC 5548 NGC 7469

1000 1 10 100 Lag (days)

Fig. 1.3.— Log V as a function of log τ for three AGNs in which multiple emission lines have been reverberation mapped. The virial theorem predicts a relationship with a slope of 2. The solid lines show the best fits to the data from each object with fixed − slope equal to the virial value. Dotted lines show the best fits with unconstrained slopes (which are consistent with the virial expectation in all three cases). From

Peterson (2001). 22 Chapter 2

Reverberation Mapping of NGC 3783

2.1. Introduction

As the AGN community’s experience with RM campaigns and datasets increased, the techniques of RM analysis were revised and improved. While the treatment of new data made use of these advances, there were a number of existing datasets to which the updated approaches had not yet been applied. We returned to some of these archival data and refined the BH mass measurements for four AGNs

(described here and in Chapter 3).

This chapter details our analysis of the Seyfert 1 galaxy, NGC 3783, and the extensive optical and UV datasets that were obtained from a mixture of space- and ground-based observatories. We describe these data in 2.2, the details of our § analysis techniques in 2.3, our results in 2.4, and their implications in 2.5. § § § 23 2.2. Data

Our study of NGC 3783 combined optical and UV observations from a monitoring campaign carried out in 1991-1992 by the International AGN Watch consortium. The initial results of the seven-month program were published by

Reichert et al. (1994), Stirpe et al. (1994), and Alloin et al. (1995).

2.2.1. UV Data

The IUE observations of NGC 3783 were conducted in 69 separate epochs, with two sampling rates. The first interval (of 45 epochs) had an average spacing of 4.0 days, while the final 24 epochs observed the AGN with an average spacing of 2.0 days. A more complete description of the UV observing program is provided by

Reichert et al. (1994).

In addition to the original IUE Spectral Image Processing System (IUESIPS),

Reichert et al. (1994) used a Gaussian extraction method (GEX; see Clavel et al. 1991) to obtain the spectra of NGC 3783. After the original data had been taken, a new standard processing pipeline was introduced. The main advantages of the New Spectral Image Processing System (NEWSIPS; Nichols et al. 1993) with respect to the older IUESIPS are the improved photometric accuracy and higher S/N of the spectra; these characteristics have been achieved by introducing a new method of raw data science registration (which both reduces the fixed

24 pattern noise in the images and improves the photometric corrections), a weighted slit extraction method, and re-derived absolute flux calibrations. NEWSIPS also includes corrections for non-linearity that might have affected previous studies.

Overall, NEWSIPS-processed spectra show average S/N increases of 10–50% over

IUESIPS data (Nichols & Linsky 1996).

We retrieved the NEWSIPS-extracted short wavelength prime camera (SWP;

Harris & Sonneborn 1987) spectra from the IUE Final Archive1. While Reichert et al. (1994) analyzed data from both the SWP and long wavelength prime cameras, we have limited our study to observations made with the SWP instrument, which has a wavelength range of 1150–1975 A˚ in the low-dispersion mode (Newmark et al. 1992).

Each spectrum was examined and several types of problems led to spectra being removed from further consideration: (1) low S/N (determined by inspection, but corresponding roughly to a continuum S/N limit of 10); (2) unusual spectral features (possibly due to grazing cosmic-ray impacts); (3) short exposure times

(when longer-exposure data were available from the same epoch and the line flux data were discrepant). Some anomalous features were checked against the GEX frames, from which impacts were carefully removed. Problems with the spectra were ignored in cases where they occurred in spectral regions outside those used in computing line and continuum fluxes. Continuum and emission line flux values were measured using the wavelengths limits listed in Table 3.

1http://ines.laeff.esa.es/ines/

25 The emission line fluxes were measured after the continuum was defined by a linear fit through four spectral regions (1340–1370 A,˚ 1440–1480 A,˚ 1710–1730 A,˚ and 1840–1860 A).˚ An alternate fit through the first three of these regions produced consistent results. Wavelength-specific problems in two cases (SWP 45150, SWP

45206) led us to substitute the alternate continuum fit for these spectra.

We have estimated the flux uncertainties by considering instances in which multiple independent exposures were obtained at the same epoch (i.e., a single pointing toward the target). Flux ratios between pairs of points within each epoch were calculated and the standard deviation of the flux ratios was taken as the fractional uncertainty for all observations. This analysis was conducted independently for each emission line and continuum band. As noted above, however, highly discrepant data were removed prior to this analysis. In spite of our use of an edited dataset, the large number of data pairs contributing to our error estimate

(about 35) justifies our continued use of these values in the analysis. The final UV dataset is given in Table 2.2 for the continuum measurements and in Table 2.3 for the emission lines.

2.2.2. Optical Data

Ground-based optical spectroscopy was conducted over the same time period as the IUE observations. The optical data analyzed here were retrieved from

26 the AGN Watch website2, and details of the observations are described by Stirpe et al. (1994). We have limited our investigation to the data gathered at the

Cerro Tololo Inter-American Observatory (CTIO) 1.0 m telescope to ensure the most homogeneous dataset possible for the cross-correlation analysis and for the construction of the mean and rms spectra. Spectra that were excessively noisy or contained other anomalies were discarded from consideration, leaving 37 CTIO observations for further analysis.

Narrow spectral lines are assumed not to vary over the timescales these data are probing. Thus the individual spectra were scaled to a constant flux by using the spectral scaling technique of van Groningen & Wanders (1992). This method computes a smooth scaling function between the input spectrum and a reference (the mean spectrum in this case, shown in Figure 2.2) over a specified wavelength range.

We scaled over the spectral region 4972–5150 A˚ in order to span the redshifted

[O III] λλ4959, 5007 emission lines and a suitable amount of continuum. We found that two iterations were required for full convergence. This reduced the fractional rms scatter in the [O III] λ5007 light curve (measured between 5028 and 5090 A)˚ to less than 2.5%. Additional iterations failed to produce any light curves with smaller scatter. The mean [O III] λ5007 flux was normalized to 8.44 10−13 erg s−1 cm−2, × the value derived by the careful analysis of Stirpe et al. (1994).

2http://www.astronomy.ohio-state.edu/~agnwatch/

27 Following the calibration of the spectra, the Hβ line was measured between

4830 and 4985 A˚ (with the continuum set by a linear fit between 4800–4820 A˚ and

5130–5150 A).˚ The flux uncertainties were measured in the same way as for the UV data ( 2.2.1), and the results are given in Table 2.5. Due to the smaller optical § dataset, the flux errors for Hβ and the 5150 A˚ continuum rely on only seven data pairs. To be cautious, we have been more conservative in our estimation of the optical flux uncertainties.

2.3. Methods and Techniques

2.3.1. Spectral Analysis

For the reverberation masses calculated here, we measure the line width V as the full width at half maximum (FWHM) of each emission line, and assume f = 0.75 to maintain consistency with previous work (e.g., Wandel et al. 1999; Kaspi et al.

2000).

An rms spectrum was created from the data to isolate the varying parts of the emission lines and it was from this spectrum that the primary FWHM values for the emission lines were measured. The FWHM data were constructed by considering the extreme flux values within the continuum regions, fitting two continuum slopes (to the highest flux levels and lowest flux levels), and averaging the measures of FWHM

28 derived from the two continuum determinations. Finally, the data were converted to their rest-frame widths using z = 0.009730 0.000007 (Theureau et al. 1998). §

2.3.2. Light Curve Analysis

Applying the techniques described by Peterson et al. (1998) and in Chapter 1, we generated cross-correlation functions (CCFs) relating the various emission line light curves to the 1355 A˚ continuum flux. We report both peak (τp) and centroid

(τc) cross-correlation lags. However, the reader should be warned that “lags” in the text will hereafter refer to centroids, unless otherwise noted, and that such lags do not represent a simple phase shift between the light curves.

As Koratkar & Gaskell (1991a) noted for NGC 3783 (and other reverberation- mapped AGNs), the choice of what threshold to use for the centroid calculation can significantly affect the resulting time lag. Figure 2.4 shows that some lines tend toward larger lags and others toward smaller values as the centroid becomes increasingly dominated by the peak value. For the ICCF, we experimented with different interpolation lengths and different thresholds for the calculation of the lag centroid. Our subsequent analysis uses an interpolation unit of 0.1 days in both light curves (interpolating one dataset at a time, with the resulting lags averaged) and a centroid threshold of 80% of the peak correlation coefficient. In addition to the ICCF, we calculated the DCF for each continuum band and emission line. The

29 uncertainties in the time lag calculations were calculated with the FR/RSS methods described in Chapter 1.

2.4. Results

2.4.1. Line widths

We have measured Vrest from both the mean and rms UV spectra (Figure 2.1) and report the results in Table 3. Geocoronal Lyα emission blended into the Lyα spectral region precludes a FWHM measurement for this line and thus also prevents

Lyα contribution to the mass determination, but cross-correlation analysis is still feasible by excluding the contaminated portion of the spectrum (see Table 3).

The same method for measuring Vrest was also applied to the optical data and yielded values of (2.91 0.19) 103 km s−1 for the rms spectrum and (2.65 0.02) 103 § × § × km s−1 for the mean optical spectrum.

2.4.2. Time delays

Figure 2.3 shows the light curves for each of the UV and optical emission lines and continuum bands.

30 In Table 3 we compare the sampling characteristics of our data with the previously published light curves. When we bin the data in each epoch, the variability parameters of the old and new UV datasets appear nearly identical. The

“excess variance”, Fvar, represents the mean fractional variation of each dataset

(see Rodr´igues-Pascual et al. 1997); Rmax is the ratio of maximum to minimum

flux levels. The updated optical dataset is much more sparse than the previously published data because of our desire for the most homogeneous dataset possible.

Emission Lines

The light curves for each emission line (He II λ1640 + O III] λ1663, Si IV λ1400

+ O IV] λ1402, Lyα, C IV λ1549, Si III] λ1892 + C III] λ1909, and Hβ) were run through the ICCF, DCF, and FR/RSS programs, using the 1355 A˚ continuum data as the “driving” light curve. The CCFs and CCPDs are shown for each emission line in the panels of Figure 2.5. The CCPDs are shown to give a graphical indication of the empirical uncertainties and are scaled to the maximum value in each panel.

Each of the emission lines was very well correlated with the continuum flux.

The poorest correlation with the continuum was found for Si III] λ1892 + C III]

λ1909, which was found to have a peak ICCF value of rmax=0.354 (i.e., a probability of arising from an uncorrelated parent population of roughly < 0.001) and we limited the range of computation for this line to 16 days to avoid aliasing. Table 3 § summarizes the previous data and our new results.

31 Our results for the UV emission lines are generally in agreement with those of Reichert et al. (1994). It should be noted, however, that the peak and centroid lags calculated by Reichert et al. (1994) and Stirpe et al. (1994) used the 1460 A˚ continuum as the driving light curve. The results quoted here are consistent with those derived from the recalibrated data with the continuum centered at 1460 A˚ rather than at 1355 A.˚ Because of the large uncertainties assigned to previous lag values, most of our NEWSIPS lags are within 1-σ of the old data. The exceptions are Si III] λ1892 + C III] λ1909, for which the IUESIPS-based data failed to produce any lag at all, and Hβ. The GEX extraction method yielded a peak lag for Si III]

λ1892 + C III] λ1909 similar to what we found, but a centroid lag approximately

2-σ larger than the current result. Our centroid lag for Hβ was only slightly more than 1-σ greater than the previous value.

The significant discrepancy between our Hβ results and those of Wandel et al.

+3.6 (1999, 4.5−3.1 days) arises from the double-peaked nature of the CCF. The centroid lag calculated for the 5150 A˚ continuum-Hβ CCF is based on fewer points than the

1355 A˚ continuum-Hβ lag, and gives precedence to the peak at smaller lags. We have greater confidence in the results that use the UV continuum data, and those results closely match the UV-Hβ correlation found by Stirpe et al. (1994).

32 Continuum

Strong evidence for wavelength-dependent continuum lags has been found for only two AGNs (NGC 7469 and Akn 564), but appears to be consistent with simple accretion disk models that predict τ λ4/3 (see Wanders et al. 1997; Collier et ∝ al. 1998, 2001). However, Korista & Goad (2001) note that diffuse emission from broad-line clouds can produce a similar wavelength dependence, so the origin of this phenomenon is not clear.

The large uncertainties still present in the NEWSIPS continuum lags prevent us from reasonably testing the τ–λ relationship because the continuum-continuum time lags we find are not statistically significant (see Table 3).

2.5. Discussion

As the tabular data indicate, the expected pattern of more highly ionized lines having smaller time lags (i.e., originating closer to the ionization source) is reconfirmed by our analysis.

rest Figure 2.6 plots Vrest(rms) versus τc for the five emission lines we measured.

The virial assumption predicts a slope of 0.5 (in log-log space). Deviation from − this relationship would contradict our model, but agreement with the predicted

33 slope cannot rule out other dynamical possibilities (see Krolik 2001, and references therein).

The statistical problem of fitting to intrinsically scattered data with heteroscedastic errors has been addressed with computational methods by Akritas

& Bershady (1996). However, our data has the additional difficulty of asymmetric errors in the lags. To account for the asymmetric time lag uncertainties we first used the larger of the two lag errors and then assessed in which direction the data points differed from the regression. We recalculated the fit using the errors toward the previous regression and confirmed that those were the appropriate choices in the final

fit. The slope of the V (rms) τ rest relation derived by the regression software3 was rest − c 0.450 0.070, consistent with our expectations for a virial relationship (irrespective − § of the specific multiplicative factor relating the line widths and FWHM values).

Hence, we fixed the slope at 0.5 and calculated the mass independently for each − emission line, applying our previously stated assumption of isotropic BLR gas motion and inserting the appropriate rest-frame values into the following equation:

3 c τ V 2 M = . (2.1) 4 G

Weighting the data by the uncertainty in the direction of the mean (since the lag

6 errors are still asymmetric) yields an average BH mass of (8.7 1.1) 10 M¯. § ×

3available at http://www.astro.wisc.edu/ mab/archive/stats/stats.html ∼ 34 Previous work with IUE archival data having much poorer temporal resolution measured a much larger C IV λ1549 time lag and derived a mass of 7.3+3.5 107 −3.6 ×

M¯ (Koratkar & Gaskell 1991a,b). Wandel et al. (1999) calculated the Hβ lag with respect to the 5100 A˚ continuum from the light curves of Stirpe et al. (1994) and

+1.1 7 then used the rms velocity width to estimate a mass of 1.1 10 M¯. Applying −1.0 × +4.7 6 this method to our version of the optical data yields a BH mass of 6.2 10 M¯, −6.1 × within the 1-σ error bars for our mass measurement with the full dataset. Fromerth

& Melia (2000) employed a different means of measuring the velocity dispersion

+0.8 7 from the data of Reichert et al. (1994) and derived masses of 1.6 10 M¯ and −0.4 × +0.8 7 1.3 10 M¯ from Lyα and C IV λ1549, respectively. Various disk accretion −0.5 × 7 models predicting a BH mass in the range of 2.0–7.0 10 M¯ were cited by Alloin × et al. (1995). However, they note the simple nature of these spatially thin, optically models and the potential for a large discrepancy from the true BH mass.

The dramatic reduction in the relative mass error for NGC 3783 demonstrates the value of both a homogeneous dataset and the improved analysis techniques. The revised results for NGC 3783 comprised one of just four AGNs at the time in which several lines yielded consistent answers for the BH mass, providing increased support for the idea that reverberation mapping is indeed probing the gravity of the BH and is not subject to excessive statistical uncertainties.

35 Fig. 2.1.— Top: rms UV spectrum. Bottom: Mean UV spectrum. Wavelengths delineated above the spectrum indicate emission-line ranges; those below the spectrum mark ranges of continuum flux measurement.

36 Fig. 2.2.— Top: rms optical spectrum. Bottom: Mean optical spectrum. Wavelength indications as in Fig. 1.

37 Fig. 2.3.— UV and optical light curves: (a) 1355 A˚ continuum, (b) 1460 A˚ continuum,

(c) 1835 A˚ continuum, (d) 5150 A˚ continuum, (e) He II λ1640 + O III] λ1663, (f)

Si IV λ1400 + O IV] λ1402, (g) Lyα, (h) C IV λ1549, (i) Si III] λ1892 + C III] λ1909, and (j) Hβ. Continuum fluxes are in units of 10−15 ergs cm−2 s−1 A˚−1. Emission line

fluxes are given in units of 10−13 ergs cm−2 s−1. 38 Fig. 2.4.— Rest-frame centroid time lag versus threshold level for centroid determination (as a fraction of the peak correlation coefficient). The data plotted are for He II λ1640 + O III] λ1663 (solid), Si IV λ1400 + O IV] λ1402 (short dashed),

Lyα (dotted), C IV λ1549 (long dashed), Si III] λ1892 + C III] λ1909 (dot-short dashed), and Hβ (dot-long dashed).

39 Fig. 2.5.— Results of cross-correlation of the 1355 A˚ continuum with (a) itself; (b)

1460 A˚ continuum; (c) 1835 A˚ continuum; (d) 5150 A˚ continuum; (e) He II λ1640

+ O III] λ1663; (f) Si IV λ1400 + O IV] λ1402; (g) Lyα; (h) C IV λ1549; (i) Si III]

λ1892 + C III] λ1909; (j) Hβ. The solid lines show the ICCFs, the data points are the DCFs, and the dashed lines represent the CCPDs. Note that the y-axis scale for the CCPDs is the fraction of MC realizations producing a centroid of that lag value and is scaled to the maximum in each panel.

40 Fig. 2.6.— Rest-frame velocity FWHM versus rest-frame centroid time lag for the

five emission lines we have measured. The dashed line is the best fit to the data; the solid line is the best fit with fixed slope of -0.5.

41 Table 2.1. NGC 3783 Continuum and Emission Line Wavelength Limits

Line/Band Wavelength Range (A)˚

1355 A˚ continuum 1340–1370 1460 A˚ continuum 1445–1475 1835 A˚ continuum 1820–1850 5150 A˚ continuum 5140–5160 Lyα 1225–1280 Si IV λ1400 + O IV] λ1402 1355–1460 C IV λ1549 1460–1624 He II λ1640 + O III] λ1663 1624–1710 Si III] λ1892 + C III] λ1909 1890–1948 Hβ 4830–4985

42 Table 2.2:: NGC 3783 UV Continuum Flux Data

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 43438 8611.948 60.811 2.432 57.899 2.142 52.244 1.776 § § § SWP 43439 8612.031 57.353 2.294 57.395 2.124 50.582 1.720 § § § SWP 43472 8615.949 51.984 2.079 50.204 1.858 47.259 1.607 § § § SWP 43473 8616.029 51.979 2.079 45.407 1.680 45.024 1.531 § § § SWP 43485 8618.118 47.391 1.896 47.043 1.741 44.713 1.520 § § § SWP 43539 8624.202 60.846 2.434 58.123 2.151 48.605 1.653 § § § SWP 43540 8624.294 61.371 2.455 58.173 2.152 50.031 1.701 § § § SWP 43541 8624.385 58.033 2.321 59.779 2.212 48.735 1.657 § § § SWP 43557 8628.595 46.814 1.873 47.146 1.744 40.850 1.389 § § § SWP 43587 8631.680 36.975 1.479 43.277 1.601 37.136 1.263 § § § SWP 43588 8631.765 41.661 1.666 42.163 1.560 35.594 1.210 § § § SWP 43636 8635.688 45.680 1.827 49.405 1.828 39.628 1.347 § § § SWP 43676 8639.862 53.819 2.153 52.490 1.942 45.264 1.539 § § § Continued on next page...

43 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 43716 8643.948 53.593 2.144 56.866 2.104 47.434 1.613 § § § SWP 43871 8649.367 56.758 2.270 55.327 2.047 48.963 1.665 § § § SWP 43872 8649.454 55.680 2.227 52.634 1.947 46.580 1.584 § § § SWP 43894 8651.756 49.667 1.987 48.990 1.813 43.264 1.471 § § § SWP 43895 8651.855 48.352 1.934 52.337 1.936 41.435 1.409 § § § SWP 43921 8656.130 61.961 2.478 62.436 2.310 53.664 1.825 § § § SWP 43945 8660.040 64.188 2.568 67.311 2.491 51.531 1.752 § § § SWP 43946 8660.121 63.656 2.546 63.607 2.353 52.300 1.778 § § § SWP 43962 8664.048 43.983 1.759 43.716 1.617 40.467 1.376 § § § SWP 43995 8668.022 32.648 1.306 30.813 1.140 31.508 1.071 § § § SWP 43996 8668.115 31.491 1.260 32.760 1.212 32.328 1.099 § § § SWP 44020 8672.106 40.887 1.635 44.671 1.653 38.755 1.318 § § § SWP 44048 8676.272 44.897 1.796 49.399 1.828 44.148 1.501 § § § SWP 44072 8680.310 50.035 2.001 49.829 1.844 42.224 1.436 § § § Continued on next page...

44 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 44099 8684.199 49.072 1.963 38.191 1.413 33.057 1.124 § § § SWP 44100 8684.275 36.331 1.453 39.336 1.455 32.631 1.109 § § § SWP 44126 8688.228 33.912 1.356 27.773 1.028 30.118 1.024 § § § SWP 44149 8692.216 28.542 1.142 30.410 1.125 28.579 0.972 § § § SWP 44176 8695.910 27.667 1.107 29.197 1.080 26.109 0.888 § § § SWP 44189 8699.713 28.465 1.139 27.397 1.014 29.131 0.990 § § § SWP 44208 8703.723 27.708 1.108 24.374 0.902 26.402 0.898 § § § SWP 44237 8707.871 25.508 1.020 33.736 1.248 31.970 1.087 § § § SWP 44267 8711.731 32.335 1.293 45.001 1.665 39.240 1.334 § § § SWP 44307 8715.612 44.933 1.797 49.245 1.822 45.781 1.557 § § § SWP 44349 8719.932 46.150 1.846 55.077 2.038 47.230 1.606 § § § SWP 44350 8720.020 56.100 2.244 54.303 2.009 46.392 1.577 § § § SWP 44381 8724.004 50.768 2.031 44.383 1.642 42.641 1.450 § § § SWP 44408 8727.967 49.801 1.992 44.960 1.664 41.495 1.411 § § § Continued on next page...

45 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 44409 8728.068 42.096 1.684 43.414 1.606 38.242 1.300 § § § SWP 44410 8728.168 45.123 1.805 43.196 1.598 40.839 1.389 § § § SWP 44434 8731.946 43.086 1.723 49.410 1.828 46.087 1.567 § § § SWP 44435 8732.205 49.680 1.987 48.115 1.780 43.598 1.482 § § § SWP 44461 8735.946 48.231 1.929 39.557 1.464 38.632 1.313 § § § SWP 44486 8739.990 41.475 1.659 35.484 1.313 35.752 1.216 § § § SWP 44492 8744.129 37.318 1.493 47.640 1.763 40.569 1.379 § § § SWP 44581 8747.874 44.276 1.771 45.235 1.674 41.850 1.423 § § § SWP 44627 8751.845 45.411 1.816 55.741 2.062 47.994 1.632 § § § SWP 44628 8751.950 51.002 2.040 53.867 1.993 48.992 1.666 § § § SWP 44629 8752.056 53.204 2.128 54.219 2.006 47.375 1.611 § § § SWP 44659 8755.872 52.803 2.112 46.318 1.714 45.237 1.538 § § § SWP 44660 8755.949 47.175 1.887 44.828 1.659 40.003 1.360 § § § SWP 44682 8759.702 47.448 1.898 51.934 1.922 48.939 1.664 § § § Continued on next page...

46 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 44731 8763.559 52.863 2.115 44.915 1.662 42.596 1.448 § § § SWP 44760 8767.535 47.646 1.906 45.992 1.702 41.306 1.404 § § § SWP 44803 8771.689 47.353 1.894 47.380 1.753 41.833 1.422 § § § SWP 44804 8771.772 52.010 2.080 46.543 1.722 42.305 1.438 § § § SWP 44830 8775.633 46.666 1.867 44.004 1.628 38.184 1.298 § § § SWP 44873 8779.607 54.715 2.189 54.094 2.001 53.352 1.814 § § § SWP 44907 8783.948 56.424 2.257 54.764 2.026 49.191 1.672 § § § SWP 44918 8785.949 57.362 2.294 55.285 2.046 50.312 1.711 § § § SWP 44921 8787.786 57.131 2.285 57.471 2.126 49.797 1.693 § § § SWP 44922 8787.864 56.597 2.264 59.267 2.193 46.567 1.583 § § § SWP 44935 8790.113 54.404 2.176 60.190 2.227 50.947 1.732 § § § SWP 44949 8791.769 60.256 2.410 55.655 2.059 49.797 1.693 § § § SWP 44950 8791.857 70.247 2.810 55.465 2.052 55.578 1.890 § § § SWP 44964 8793.963 57.656 2.306 56.226 2.080 48.742 1.657 § § § Continued on next page...

47 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 44974 8795.768 56.318 2.253 53.796 1.990 48.098 1.635 § § § SWP 44992 8797.448 54.836 2.193 53.977 1.997 46.297 1.574 § § § SWP 44993 8797.538 52.793 2.112 54.654 2.022 46.258 1.573 § § § SWP 45010 8799.460 54.291 2.172 47.346 1.752 50.123 1.704 § § § SWP 45024 8801.764 55.805 2.232 50.718 1.877 45.454 1.545 § § § SWP 45025 8801.887 49.579 1.983 54.021 1.999 43.688 1.485 § § § SWP 45026 8801.996 44.498 1.780 52.914 1.958 44.756 1.522 § § § SWP 45038 8803.458 52.139 2.086 58.489 2.164 49.668 1.689 § § § SWP 45052 8805.543 50.742 2.030 61.563 2.278 52.663 1.791 § § § SWP 45063 8807.520 51.641 2.066 59.254 2.192 50.333 1.711 § § § SWP 45064 8807.603 51.587 2.063 62.612 2.317 54.971 1.869 § § § SWP 45081 8809.509 59.759 2.390 57.843 2.140 50.158 1.705 § § § SWP 45082 8809.601 60.228 2.409 53.138 1.966 51.993 1.768 § § § SWP 45096 8811.493 63.860 2.554 55.395 2.050 49.903 1.697 § § § Continued on next page...

48 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 45097 8811.595 56.369 2.255 52.657 1.948 50.072 1.702 § § § SWP 45106 8813.384 51.571 2.063 54.998 2.035 47.810 1.626 § § § SWP 45118 8816.028 57.179 2.287 51.851 1.918 45.563 1.549 § § § SWP 45133 8818.024 57.305 2.292 42.796 1.583 39.882 1.356 § § § SWP 45150 8819.700 54.163 2.167 37.296 1.380 § § · · · SWP 45151 8819.800 52.810 2.112 38.618 1.429 38.881 1.322 § § § SWP 45152 8819.904 41.373 1.655 37.208 1.377 38.237 1.300 § § § SWP 45167 8821.689 39.568 1.583 45.306 1.676 44.545 1.515 § § § SWP 45168 8821.791 48.325 1.933 47.669 1.764 46.635 1.586 § § § SWP 45169 8821.892 51.088 2.044 53.395 1.976 43.804 1.489 § § § SWP 45194 8824.353 50.394 2.016 63.151 2.337 52.316 1.779 § § § SWP 45195 8824.440 61.519 2.461 60.085 2.223 51.566 1.753 § § § SWP 45206 8825.701 60.748 2.430 63.864 2.363 § § · · · SWP 45207 8825.798 66.760 2.670 62.632 2.317 53.268 1.811 § § § Continued on next page...

49 Table 2.2 – Continued

Image Julian Date

Name (2,440,000+) F(λ1355) F(λ1460) F(λ1835)

SWP 45219 8827.904 67.384 2.695 67.817 2.509 59.430 2.021 § § § SWP 45227 8829.302 69.254 2.770 59.651 2.207 56.059 1.906 § § § SWP 45237 8831.317 70.910 2.836 67.452 2.496 58.335 1.983 § § § SWP 45246 8833.326 72.855 2.914 76.033 2.813 59.754 2.032 § § §

Continuum fluxes are given in units of 10−15 ergs cm−2 s−1 A˚−1.

50 Table 2.3:: NGC 3783 UV Emission Line Flux Data

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 43438 8611.948 17.051 1.705 12.769 1.430 41.751 1.837 76.614 2.988 9.707 1.175 § § § § §

51 SWP 43439 8612.031 14.790 1.479 10.856 1.216 41.387 1.821 74.662 2.912 9.180 1.111 § § § § § SWP 43472 8615.949 15.837 1.584 10.362 1.161 38.720 1.704 69.633 2.716 8.968 1.085 § § § § § SWP 43473 8616.029 15.365 1.536 11.732 1.314 39.614 1.743 71.235 2.778 13.764 1.665 § § § § § SWP 43485 8618.118 13.188 1.319 8.559 0.959 37.369 1.644 73.775 2.877 10.270 1.243 § § § § § SWP 43539 8624.202 14.384 1.438 10.623 1.190 39.431 1.735 66.575 2.596 10.175 1.231 § § § § § SWP 43540 8624.294 17.992 1.799 9.900 1.109 36.819 1.620 69.660 2.717 13.212 1.599 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 43541 8624.385 18.288 1.829 12.007 1.345 39.643 1.744 74.851 2.919 10.466 1.266 § § § § § SWP 43557 8628.595 18.169 1.817 9.672 1.083 36.259 1.595 65.172 2.542 11.489 1.390

52 § § § § § SWP 43587 8631.680 12.129 1.213 9.326 1.045 37.353 1.644 68.481 2.671 11.690 1.414 § § § § § SWP 43588 8631.765 11.233 1.123 9.692 1.086 35.715 1.571 71.312 2.781 13.907 1.683 § § § § § SWP 43636 8635.688 13.382 1.338 9.124 1.022 33.320 1.466 64.623 2.520 9.733 1.178 § § § § § SWP 43676 8639.862 17.282 1.728 10.188 1.141 36.059 1.587 68.877 2.686 10.664 1.290 § § § § § SWP 43716 8643.948 18.714 1.871 11.037 1.236 40.307 1.774 73.296 2.859 10.240 1.239 § § § § § SWP 43871 8649.367 13.185 1.319 11.574 1.296 42.142 1.854 72.814 2.840 10.015 1.212 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 43872 8649.454 15.884 1.588 12.645 1.416 45.933 2.021 76.959 3.001 10.892 1.318 § § § § § SWP 43894 8651.756 12.264 1.226 10.932 1.224 38.313 1.686 67.904 2.648 12.443 1.506

53 § § § § § SWP 43895 8651.855 14.863 1.486 9.827 1.101 38.525 1.695 67.918 2.649 10.770 1.303 § § § § § SWP 43921 8656.130 15.008 1.501 9.425 1.056 40.596 1.786 67.940 2.650 8.425 1.019 § § § § § SWP 43945 8660.040 14.856 1.486 11.559 1.295 39.347 1.731 70.191 2.737 10.595 1.282 § § § § § SWP 43946 8660.121 13.709 1.371 11.282 1.264 42.857 1.886 68.775 2.682 8.694 1.052 § § § § § SWP 43962 8664.048 13.022 1.302 11.275 1.263 40.307 1.774 71.699 2.796 10.706 1.295 § § § § § SWP 43995 8668.022 7.140 0.800 35.489 1.562 66.840 2.607 11.294 1.367 · · · § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 43996 8668.115 9.901 1.109 35.812 1.576 69.053 2.693 10.296 1.246 · · · § § § § SWP 44020 8672.106 11.764 1.176 10.083 1.129 36.446 1.604 64.753 2.525 9.706 1.174

54 § § § § § SWP 44048 8676.272 14.299 1.430 10.509 1.177 35.112 1.545 60.287 2.351 10.137 1.227 § § § § § SWP 44072 8680.310 11.451 1.145 9.171 1.027 33.214 1.461 65.161 2.541 9.986 1.208 § § § § § SWP 44099 8684.199 13.925 1.393 10.682 1.196 33.220 1.462 67.461 2.631 11.558 1.399 § § § § § SWP 44100 8684.275 11.177 1.118 12.329 1.381 31.722 1.396 64.203 2.504 14.211 1.720 § § § § § SWP 44126 8688.228 8.477 0.848 7.211 0.808 29.611 1.303 57.594 2.246 9.462 1.145 § § § § § SWP 44149 8692.216 8.912 0.891 7.930 0.888 29.519 1.299 63.911 2.493 9.846 1.191 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44176 8695.910 10.943 1.094 8.069 0.904 26.618 1.171 58.785 2.293 10.981 1.329 § § § § § SWP 44189 8699.713 8.945 0.894 8.916 0.999 28.493 1.254 57.753 2.252 7.723 0.934

55 § § § § § SWP 44208 8703.723 7.663 0.766 8.471 0.949 24.982 1.099 53.504 2.087 8.793 1.064 § § § § § SWP 44237 8707.871 9.909 0.991 8.974 1.005 27.480 1.209 56.975 2.222 7.860 0.951 § § § § § SWP 44267 8711.731 13.391 1.339 9.154 1.025 31.722 1.396 60.820 2.372 8.547 1.034 § § § § § SWP 44307 8715.612 11.593 1.159 12.355 1.384 33.015 1.453 65.884 2.569 9.412 1.139 § § § § § SWP 44349 8719.932 11.976 1.198 9.588 1.074 33.504 1.474 61.410 2.395 9.790 1.185 § § § § § SWP 44350 8720.020 15.663 1.566 13.091 1.466 36.571 1.609 67.330 2.626 10.934 1.323 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44381 8724.004 12.488 1.249 8.475 0.949 38.718 1.704 64.045 2.498 9.589 1.160 § § § § § SWP 44408 8727.967 11.592 1.159 11.942 1.338 39.422 1.735 68.315 2.664 7.924 0.959

56 § § § § § SWP 44409 8728.068 11.346 1.271 37.210 1.637 67.746 2.642 11.943 1.445 · · · § § § § SWP 44410 8728.168 15.773 1.577 11.716 1.312 37.375 1.645 72.573 2.830 11.679 1.413 § § § § § SWP 44434 8731.946 16.459 1.646 11.614 1.301 41.213 1.813 72.594 2.831 8.378 1.014 § § § § § SWP 44435 8732.205 13.598 1.523 40.652 1.789 68.736 2.681 9.980 1.208 · · · § § § § SWP 44461 8735.946 14.037 1.404 10.790 1.208 37.320 1.642 69.492 2.710 11.347 1.373 § § § § § SWP 44486 8739.990 10.913 1.091 11.056 1.238 33.716 1.484 63.079 2.460 10.973 1.328 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44492 8744.129 10.972 1.097 8.646 0.968 31.276 1.376 63.631 2.482 9.638 1.166 § § § § § SWP 44581 8747.874 13.851 1.385 10.644 1.192 37.659 1.657 71.354 2.783 9.569 1.158

57 § § § § § SWP 44627 8751.845 15.350 1.535 10.311 1.155 34.631 1.524 72.491 2.827 8.062 0.976 § § § § § SWP 44628 8751.950 14.578 1.458 10.779 1.207 39.274 1.728 71.496 2.788 9.285 1.123 § § § § § SWP 44629 8752.056 12.920 1.292 8.708 0.975 36.988 1.627 67.524 2.633 9.705 1.174 § § § § § SWP 44659 8755.872 13.868 1.387 10.749 1.204 37.794 1.663 68.993 2.691 9.236 1.118 § § § § § SWP 44660 8755.949 9.701 1.087 37.884 1.667 71.405 2.785 13.026 1.576 · · · § § § § SWP 44682 8759.702 13.730 1.373 11.811 1.323 37.249 1.639 68.033 2.653 7.400 0.895 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44731 8763.559 15.880 1.588 10.974 1.229 36.521 1.607 65.598 2.558 9.866 1.194 § § § § § SWP 44760 8767.535 10.554 1.055 9.223 1.033 31.337 1.379 54.612 2.130 10.060 1.217

58 § § § § § SWP 44803 8771.689 12.725 1.273 11.678 1.308 30.261 1.331 60.436 2.357 9.198 1.113 § § § § § SWP 44804 8771.772 12.169 1.217 8.260 0.925 30.620 1.347 63.469 2.475 12.367 1.496 § § § § § SWP 44830 8775.633 15.957 1.596 10.018 1.122 34.556 1.520 61.264 2.389 11.703 1.416 § § § § § SWP 44873 8779.607 12.230 1.223 13.309 1.491 33.902 1.492 62.857 2.451 7.113 0.861 § § § § § SWP 44907 8783.948 15.930 1.593 9.829 1.101 35.165 1.547 65.009 2.535 8.760 1.060 § § § § § SWP 44918 8785.949 10.634 1.063 11.991 1.343 36.745 1.617 62.983 2.456 6.630 0.802 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44921 8787.786 12.702 1.270 10.055 1.126 34.865 1.534 64.779 2.526 10.548 1.276 § § § § § SWP 44922 8787.864 11.840 1.184 8.687 0.973 34.194 1.505 66.079 2.577 11.881 1.438

59 § § § § § SWP 44935 8790.113 14.784 1.478 9.681 1.084 33.982 1.495 67.992 2.652 11.232 1.359 § § § § § SWP 44949 8791.769 15.495 1.549 11.760 1.317 37.262 1.640 70.532 2.751 9.822 1.188 § § § § § SWP 44950 8791.857 15.004 1.500 10.425 1.168 36.716 1.616 67.412 2.629 14.068 1.702 § § § § § SWP 44964 8793.963 14.658 1.466 9.572 1.072 37.016 1.629 67.780 2.643 10.045 1.215 § § § § § SWP 44974 8795.768 14.577 1.458 10.878 1.218 37.405 1.646 69.919 2.727 8.464 1.024 § § § § § SWP 44992 8797.448 16.789 1.679 11.013 1.233 37.606 1.655 71.690 2.796 11.005 1.332 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 44993 8797.538 15.144 1.514 10.938 1.225 37.047 1.630 71.722 2.797 10.318 1.248 § § § § § SWP 45010 8799.460 15.066 1.507 10.861 1.216 38.602 1.698 61.098 2.383 9.052 1.095

60 § § § § § SWP 45024 8801.764 11.581 1.158 11.610 1.300 63.585 2.480 11.248 1.361 § § · · · § § SWP 45025 8801.887 13.310 1.331 9.372 1.050 34.032 1.497 64.283 2.507 10.079 1.220 § § § § § SWP 45026 8801.996 11.679 1.168 9.878 1.106 35.559 1.565 64.598 2.519 11.732 1.420 § § § § § SWP 45038 8803.458 11.433 1.143 11.278 1.263 36.813 1.620 66.819 2.606 8.374 1.013 § § § § § SWP 45052 8805.543 16.844 1.684 12.541 1.405 41.348 1.819 73.452 2.865 10.795 1.306 § § § § § SWP 45063 8807.520 16.108 1.611 13.618 1.525 40.783 1.794 72.034 2.809 11.587 1.402 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 45064 8807.603 15.292 1.529 12.634 1.415 36.449 1.604 72.370 2.822 12.714 1.538 § § § § § SWP 45081 8809.509 14.375 1.438 12.532 1.404 35.314 1.554 70.550 2.751 11.009 1.332

61 § § § § § SWP 45082 8809.601 14.549 1.455 13.052 1.462 38.033 1.673 74.732 2.915 9.878 1.195 § § § § § SWP 45096 8811.493 15.791 1.579 13.392 1.500 35.964 1.582 69.099 2.695 11.158 1.350 § § § § § SWP 45097 8811.595 15.899 1.590 13.384 1.499 38.451 1.692 72.306 2.820 10.461 1.266 § § § § § SWP 45106 8813.384 14.546 1.455 10.726 1.201 36.440 1.603 64.005 2.496 12.925 1.564 § § § § § SWP 45118 8816.028 13.700 1.370 8.430 0.944 34.493 1.518 67.487 2.632 11.906 1.441 § § § § § SWP 45133 8818.024 10.933 1.093 11.187 1.253 32.175 1.416 66.611 2.598 11.225 1.358 § § § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 45150 8819.700 8.954 0.895 8.490 0.951 27.355 1.204 55.596 2.168 § § § § · · · SWP 45151 8819.800 7.794 0.779 6.574 0.736 31.274 1.376 52.838 2.061 12.267 1.484

62 § § § § § SWP 45152 8819.904 7.073 0.707 8.813 0.987 30.107 1.325 59.321 2.314 9.929 1.201 § § § § § SWP 45167 8821.689 11.037 1.104 8.706 0.975 30.555 1.344 62.589 2.441 10.167 1.230 § § § § § SWP 45168 8821.791 12.697 1.270 7.656 0.857 29.693 1.306 57.353 2.237 9.218 1.115 § § § § § SWP 45169 8821.892 11.109 1.111 8.925 1.000 29.510 1.298 56.358 2.198 11.512 1.393 § § § § § SWP 45194 8824.353 14.707 1.471 9.936 1.113 31.339 1.379 59.284 2.312 8.923 1.080 § § § § § SWP 45195 8824.440 13.914 1.391 29.242 1.287 61.297 2.391 9.275 1.122 § · · · § § § Continued on next page... Table 2.3 – Continued

Image Julian Date He II λ1640 + Si IV λ1400 + Si III λ1892 +

Name (2,440,000+) O III] λ1663 O IV] λ1402 Lyα C IV λ1549 C III] λ1909

SWP 45206 8825.701 14.222 1.422 10.173 1.139 33.615 1.479 64.641 2.521 § § § § · · · SWP 45207 8825.798 16.205 1.620 9.322 1.044 30.904 1.360 60.589 2.363 10.583 1.281

63 § § § § § SWP 45219 8827.904 14.897 1.490 12.322 1.380 35.942 1.581 63.532 2.478 8.180 0.990 § § § § § SWP 45227 8829.302 14.370 1.437 14.745 1.651 36.139 1.590 72.627 2.832 9.507 1.150 § § § § § SWP 45237 8831.317 14.427 1.443 13.534 1.516 35.966 1.583 66.867 2.608 9.803 1.186 § § § § § SWP 45246 8833.326 18.615 1.862 11.108 1.244 36.969 1.627 69.398 2.707 11.312 1.369 § § § § §

Note. — Emission line fluxes are in units of 10−13 ergs cm−2 s−1. Table 2.4. NGC 3783 UV Velocity Data

Emission Line Vrest(rms) Vrest(mean)

He II λ1640 + O III] λ1663 6.34 0.90 4.74 0.69 § § Si IV λ1400 + O IV] λ1402 5.73 2.71 4.81 0.50 § § Lyα · · · · · · C IV λ1549 3.55 0.59 3.03 0.07 § § Si III] λ1892 + C III] λ1909 2.61 0.16 2.82 0.31 § §

Note. — Velocity data are in units of 103 km s−1.

64 Table 2.5:: NGC 3783 Optical Flux Data

Image Julian Date

Name (2,440,000+) F(λ5150) F(Hβ)

n38607a 8607.830 11.406 0.605 9.829 0.403 § § n38623a 8623.830 11.707 0.620 10.100 0.414 § § n38627a 8627.830 11.094 0.588 10.272 0.421 § § n38631a 8631.840 10.865 0.576 10.494 0.430 § § n38635a 8635.830 10.781 0.571 10.183 0.418 § § n38639a 8639.840 11.303 0.599 10.327 0.423 § § n38643a 8643.720 12.788 0.678 10.423 0.427 § § n38647a 8647.760 12.668 0.671 10.459 0.429 § § n38651a 8651.670 11.190 0.593 10.603 0.435 § § n38656a 8656.770 12.163 0.645 10.669 0.437 § § n38660a 8660.770 11.629 0.616 10.109 0.414 § § n38664a 8664.770 11.005 0.583 10.894 0.447 § § n38668a 8668.810 10.204 0.541 10.969 0.450 § § Continued on next page...

65 Table 2.5 – Continued

Image Julian Date

Name (2,440,000+) F(λ5150) F(Hβ)

n38676a 8676.710 10.817 0.573 9.760 0.400 § § n38677a 8677.800 10.786 0.572 9.483 0.389 § § n38678a 8678.680 10.703 0.567 9.513 0.390 § § n38704a 8704.580 8.763 0.464 8.469 0.347 § § n38712a 8712.590 10.350 0.549 8.410 0.345 § § n38716a 8716.600 10.055 0.533 8.840 0.362 § § n38720a 8720.590 11.842 0.628 9.395 0.385 § § n38724a 8724.560 11.257 0.597 9.640 0.395 § § n38732a 8732.560 11.212 0.594 9.357 0.384 § § n38736a 8736.550 10.149 0.538 9.410 0.386 § § n38744a 8744.610 10.812 0.573 10.537 0.432 § § n38752a 8752.570 11.174 0.592 9.606 0.394 § § n38763a 8763.580 11.880 0.630 11.678 0.479 § § n38764a 8764.560 10.754 0.570 10.131 0.415 § § Continued on next page...

66 Table 2.5 – Continued

Image Julian Date

Name (2,440,000+) F(λ5150) F(Hβ)

n38772a 8772.630 11.799 0.625 9.627 0.395 § § n38776a 8776.560 10.636 0.564 9.272 0.380 § § n38794a 8794.570 11.444 0.607 10.803 0.443 § § n38804a 8804.470 12.281 0.651 11.780 0.483 § § n38808a 8808.480 12.773 0.677 10.531 0.432 § § n38822a 8822.480 12.917 0.685 11.060 0.453 § § n38824a 8824.470 11.765 0.624 10.108 0.414 § § n38826a 8826.470 12.717 0.674 9.912 0.406 § § n38830a 8830.480 12.108 0.642 11.026 0.452 § § n38832a 8832.460 13.281 0.704 11.086 0.455 § §

Note. — Continuum fluxes are in units of 10−15 ergs cm−2 s−1 A˚−1.

Emission line fluxes are in units of 10−13 ergs cm−2 s−1.

67 Table 2.6. NGC 3783 Sampling Statistics

Sampling Interval (days)

Subset Number Average Median Fvar Rmax Reference

Previous 1460 A˚ continuum dataset 69 3.3 3.9 0.201 3.027 0.380 1 § New UV dataset; binned by epoch 69 3.3 3.9 0.203 3.119 0.163 § New UV dataset; complete sample 101 2.2 2.0 0.192 2.856 0.162

68 § New UV dataset; 4-day sampling period 62 2.8 3.9 0.193 2.762 0.145 § New UV dataset; 2-day sampling period 40 1.3 1.7 0.140 2.043 0.107 § Previous 5150 A˚ continuum dataset 72 3.2 2.0 0.078 1.517 0.042 2 § New optical dataset; binned by epoch 35 6.6 4.0 0.065 1.516 0.114 § New optical dataset; complete sample 37 6.2 4.0 0.064 1.516 0.114 §

References. — (1) Reichert et al. 1994; (2) Stirpe et al. 1994.

Note. — Previously published light curves were collected from the AGN Watch website. Table 2.7. NGC 3783 Cross-Correlation Results

Previous Resultsa Current Results

rest rest rest rest Line/Band τc τp τc τp

+0.3 +0.2 F(λ1460) 0.1−0.2 0.0−0.4 · ·+3· · ·+2· − +0.3 +0.6 F(λ1835) 0.1−3 0−2 0.0−0.3 0.0−0.5 +2 +2 +3.1 +1.9 F(λ5150) 1.6−2 1−2 0.4−1.6 0.7−1.6 +4 +2 +0.9 +0.6 He II λ1640 + O III] λ1663 0.5−4 1−2 1.3−0.5 1.4−1.1 +4 +2 +0.9 +0.8 Si IV λ1400 + O IV] λ1402 3.9−4 5−2 2.1−1.5 2.3−2.4 +3 +2 +1.1 +2.5 Lyα 3.8−3 4−2 3.6−0.7 2.2−0.1 +3 +2 +1.0 +0.4 C IV λ1549 5.4−3 5−2 3.8−0.9 4.5−2.2 b +4 +2 +1.3 +0.2 Si III] λ1892 + C III] λ1909 15.6−4 9−2 8.5−2.6 10.2−5.3 +2 +2 +4.1 +5.1 Hβ 7.1−2 8−2 10.4−2.3 9.0−2.4

Note. — All time lag data are in units of days. aThe previous results listed here are adapted from the GEX- extracted UV data of Reichert et al. (1994) and from the optical results of Stirpe et al. (1994), both of which used the 1460 A˚ continuum as the driving light curve. bThe range of time lags we included in our analysis of Si III] λ1892 + C III] λ1909 was limited to 16 days to avoid aliasing. §

69 Chapter 3

Additional Reverberation Mapping Studies

3.1. Introduction

Most of the statistical studies of reverberation-based black-hole masses used the homogeneous compilations of Wandel et al. (1999, hereafter, WPM), and Kaspi et al. (2000). Missing from these compilations were AGNs observed by the “Lovers of Active Galaxies (LAG)” collaboration that monitored several AGNs in early

1990 (Robinson 1994). Three of the LAG sources, NGC 3227 (Salamanca et al.

1994), NGC 3516 (Wanders et al. 1993), and NGC 4593 (Dietrich et al. 1994), have well-determined emission-line lags and certainly meet the quality criteria for inclusion in these studies, and indeed were included in the compilation of Ho (1999, based on Hβ only). These galaxies were omitted from the WPM compilation only because the data had not been analyzed in the same fashion as the other sources; specifically, (a) WPM use the model-independent Monte Carlo method of Peterson et al. (1998) to assess uncertainties in emission-line lags, and (b) WPM use the

FWHM of the emission-line in the rms spectrum, rather than the mean spectrum, to characterize the BLR line-of-sight velocity width, V , that is used to form the

70 virial product ( V 2 τ). Because we have converted both our time lags and velocity ∝ widths to the AGN rest frame, our results supplement the compilation of Kaspi et al. (2000). (WPM did not correct for the time dilation of their τ values.)

In this Chapter, we reanalyze the LAG data on these three sources in the same fashion as Kaspi et al. (2000), with the intent of enlarging that homogeneous database. The total number of AGNs which have been reverberation-mapped (a few dozen) is small enough that additional objects will assist in on-going statistical investigations of (what are hoped to be) fundamental physical relationships. The main aspect on which we have worked is the careful construction the rms spectra over the spans of the respective observing campaigns. With the rms spectra for these

AGNs and updated time lag values and uncertainties, we are able to derive masses for the SMBHs in these three galaxies.

In 3.2, we briefly describe the original observing campaigns and our data § reduction. We explain the details of our time series and velocity width analyses in § 3.3. Our estimates of the SMBH mass in each galaxy are presented in 3.4. In 3.5, § § we address the MBH-σ∗ relationship, and we then summarize our conclusions ( 3.6). §

3.2. Observations and Data Reduction

The LAG consortium examined the Hα and Hβ wavelength regimes with long slit spectroscopy of moderate resolution and with the temporal sampling rates

71 listed in Table 3.1. The line marked “cont.” in Table 3.1 for NGC 3516 indicates the sampling statistics derived for the combination of imaging and spectroscopic measurements of the 4000-5000A˚ continuum. The continuum statistics for ∼

NGC 3227 and NGC 4593 are identical to the Hβ and Hα entries, respectively. Fvar indicates the “excess variance,” the mean fractional variation of each dataset (see

Rodr´igues-Pascual et al. 1997). The ratio of the maximum to minimum flux value is given in Table 3.1 as Rmax.

In general, no alterations were applied to the published light curve data.

However, in order to investigate the velocity width characteristics of the broad emission lines, we used the wavelength region around the narrow lines ([O III] and

[S II], which should only vary on timescales much longer than the campaigns) to scale the spectra (van Groningen & Wanders 1992). The absolute scaling was done so as to yield mean narrow-line fluxes equal to the values derived in the original papers. The slit width selected was an attempt to balance spectral resolution and total detected flux. However, the small slit width introduced a non-trivial degree of difficulty in the removal of seeing effects, as noted below.

3.2.1. NGC 3227

The LAG campaign on NGC 3227 was described by Salamanca et al. (1994).

The analysis was complicated by the fact that the two methods they employed for

72 removing the contribution of the stellar host gave significantly different answers: they derived a 24% contribution to the continuum by examining the magnesium triplet of a template galaxy of the same Hubble type, and a 40% continuum contribution from comparisons with the magnesium triplet absorption of the bulge of NGC 3227. We analyzed the data with each of the potential stellar levels removed and found the differences to be insignificant.

A detailed investigation of various origins of uncertainty in their flux calibration was conducted by a portion of the LAG group (Baribaud et al. 1994). In addition to the effects of slit mis-centering and seeing, they estimate the uncertainties incurred by their scaling of the spectra by the narrow lines, and conclude an overall flux error of 5–8%. We have assumed the more conservative 8% errors for all of the flux data.

Our analysis made use of the full LAG dataset, consisting of 23 Hα observations and

14 Hβ observations. The light curve cross-correlations were made with respect to the continuum at 5020A.˚

The systemic redshift of NGC 3227 was taken to be z = 0.00380 0.00002 § (Keel 1996).

3.2.2. NGC 3516

Spectra of NGC 3516 were taken at 22 separate epochs for Hα and 21 epochs for

Hβ. The continuum light curve for this object was constructed from a combination

73 of B-band imaging and spectral fits to the power-law continuum at 4905A.˚ The results of the LAG study have been published by Wanders et al. (1993) and Wanders

& Horne (1994). The spectra we obtained for this object had already been corrected for seeing, the stellar flux from the host galaxy, and contributions from narrow lines (see Wanders et al. 1992). In addition, the [S II] and [O III] narrow lines had been subtracted from the Hα and Hβ spectra, respectively, preventing any rescaling of the flux levels. Several observations were removed from the calculation of the rms spectrum for a variety of reasons (extremely poor seeing, object mis-centering, and other anomalies). Among the Hα dataset, we have excised observations from the following dates: 31 January, 16 February, 15 April, and 13 May. From the Hβ analysis we have removed 24 March, 15 April, and 13 May.

Conversion of time lags and line-of-sight velocity widths to the rest frame was made with z = 0.00884 0.00002 (Keel 1996). §

3.2.3. NGC 4593

Spectra of NGC 4593 were obtained at 22 epochs around Hα and at 11 epochs around Hβ, and the continuum used was at 6310A.˚ The Hα line was contaminated by strong [N II] λλ6548, 6584 emission, as well as a narrow Hα contribution. Using the shape of the narrow [O III]λ5007 line as a template for all of the narrow lines, and the flux ratios between the [N II] lines and narrow Hα that Dietrich et al.

74 (1994) calculated, we have bracketed the amount of narrow line removal by limiting cases in which too much or too little narrow line flux was subtracted. Conducting this for each Hα spectrum, we then separately analyzed our over-subtracted and under-subtracted datasets. The V difference between the over-subtraction and under-subtraction was on the order of 1%, much smaller than the V uncertainties as well as the range of values derived by doing no rescaling of the spectra or no removal of the narrow-line contribution.

We converted our results to the galaxy’s rest frame using z = 0.00900 0.00013 § (Strauss et al. 1992).

3.3. Analysis

Using the published light curves, we generated the cross-correlation functions

(CCFs) between the continua and the emission lines with the ICCF. We calculated both the peak of each CCF (τp) and its centroid (τc; above a threshold of 80% of the peak value correlation coefficient) with an interpolation unit of 0.1 days.

Uncertainties in the time lag values were estimated with the FR/RSS method described by Peterson et al. (1998) and in Chapter 1. The FR/RSS technique was implemented 1000 times for each continuum-line combination to build up the cross-correlation peak distribution (CCPD), and to estimate our time lag uncertainties. Table 3.2 and Figure 3.3 show our cross-correlation results.

75 The procedure we applied to the mean and rms spectra of each AGN was to make two estimates of the continuum, at the extrema of where the level could potentially be set. The two resulting measurements of the velocity full-width at half-maximum, V , were then averaged and the difference taken to be twice the one-sigma errors. We give the results of our V measurements from both the mean and rms spectra in Table 3.3.

3.4. Mass Calculations

For consistency with previously published compilations, we calculate the reverberation mass as

3 c τ V 2 M = (3.1) BH 4 G where V is taken from the rms spectrum, c is the speed of light, and G is the gravitational constant. The asymmetric errors in the time lags are propagated through to the masses derived from the individual lines. Then, the masses for each line are combined in a weighted average. We have measured the reverberation masses for our three targets and present the results in Table 3.4. Object-specific comments and comparisons with previous results follow.

76 3.4.1. NGC 3227

8 The original LAG estimate of the SMBH mass in NGC 3227 was 2 10 M¯ ∼ × 7 (Salamanca et al. 1994). Our value of (3.6 1.4) 10 M¯ is consistent with the § × +2.7 7 mass of 4.9 10 M¯ tabulated by WPM, which was based on an independent −5.0× monitoring campaign (Winge et al. 1995). The value derived by Ho (1999) from Hβ

7 was 3.8 10 M¯, also consistent with our new SMBH mass in this galaxy. Schinnerer × 7 et al. (2000) used CO gas kinematics to calculate a lower mass limit of 1.5 10 M¯ ∼ × enclosed within the central 25 pc. Again, this value is fully consistent with our results.

3.4.2. NGC 3516

Kazanas & Nayakshin (2001) have noted the difficulty in reconciling the

SMBH mass estimates for NGC 3516 arising from a variety of optical, UV, and

X-ray constraints within certain models of AGN structure. Although our final

7 estimate of MBH = (1.68 0.33) 10 M¯ lies slightly above a recent mass upper limit § × 7 of (1.12 0.05) 10 M¯ derived from models of the X-ray warm absorber in this § × galaxy (Morales & Fabian 2002), other models of X-ray properties point toward a

8 mass closer to 10 M¯ (Kazanas & Nayakshin 2001). Our results for this AGN are

7 consistent with the value of 2 10 M¯ from the earlier analysis of the LAG data by ∼ × 7 Wanders & Horne (1994), and also matches well the mass of 2.3 10 M¯ Ho (1999) ×

77 calculated using Hβ alone. Recent attempts at accretion disk modeling from X-ray

7 7 data have given values of 3.1 10 M¯ (Czerny et al. 2001) and 2 10 M¯ (Chiang × × 2002), both in rough agreement with the present work.

3.4.3. NGC 4593

6 For NGC 4593, we calculate that MBH = (6.6 5.2) 10 M¯. An earlier § × +1.4 6 reverberation mapping campaign derived a mass of 2.2 10 M¯ from Lyman α −1.1×

(Santos-Lle´o et al. 1995), and Kollatschny & Dietrich (1997) calculated MBH =

6 7 10 M¯ from the LAG Hα data (when corrected for our conversion between V and × the velocity dispersion). Ho (1999) estimated a slightly larger Hβ time delay and

6 found a SMBH mass of 8.1 10 M¯. With such large uncertainties, these differences × cannot be considered statistically significant.

3.5. The BH Mass—Stellar Velocity Dispersion

Relationship

Since the discovery of the exceptionally tight correlation between a galaxy’s

SMBH mass and the central value of the velocity dispersion of the galaxy’s spheroid

(i.e. MBH σ∗; Ferrarese & Merritt 2000; Gebhardt et al. 2000a), a number − of attempts have been made to reconcile differences in the derived slope of this relationship (e.g., Tremaine et al. 2002) as well as to develop a physical framework

78 that will naturally produce such a correlation. In addition, Gebhardt et al. (2000b) and Ferrarese et al. (2001) have examined whether SMBH masses derived by reverberation mapping are consistent with the MBH σ∗ relationship. Because bulge − velocity dispersion measurements exist for all three of our targets, we are able to enlarge the AGN sample used to investigate MBH σ∗. −

Nelson & Whittle (1995) measured bulge velocity dispersions for 85 galaxies,

−1 including two of our targets. NGC 3227 was found to have σ∗ = 128 13 km s , § −1 and NGC 4593 has σ∗ = 124 29 km s . Di Nella et al. (1995) published a value § −1 of σ∗ = 124 5 km s for NGC 3516. These values for σ∗, in conjunction with our § derived MBH data, are consistent with the current fits to MBH σ∗ (Figure 3.4). This − excellent agreement between the AGN and quiescent galaxy MBH σ∗ relationships − is additional evidence that reverberation masses are not subject to large systematic errors.

In addition, we note the relative concentrations of the AGN and quiescent galaxy data in Figure 3.4 toward the two ends of the relationship. This is a reflection of two facts: (1) in order to measure σ∗ in AGNs, the continuum emission must be weak enough so as not to wash out the stellar absorption lines, implying the accessibility of only the dimmer AGNs, and therefore only the smaller SMBH masses;

(2) the quiescent galaxies need a relatively large SMBH mass to allow the probing of the gravitational sphere of influence, whether by stellar dynamics or gas kinematics.

The combination of these requirements gives rise to the observed distribution, in

79 spite of the probable span of both galaxy types over the entire range of SMBH masses.

3.6. Conclusion

As many studies of AGNs are examining correlations between SMBH masses and properties of the AGN host galaxies, a uniform analysis methodology is important for minimizing systematic offsets between datasets. Time lags, calculated from the published light curves, were used in conjunction with rms velocity widths for optical hydrogen emission lines to estimate the reverberation masses for NGC 3227,

NGC 3516, and NGC 4593. These masses, listed in Table 3.4, can now be added to the compilation of Kaspi et al. (2000), as they were calculated in the same manner.

In addition, these masses, when combined with published data on the host galaxy spheroid velocity dispersions, strengthen the claim that reverberation masses are accurate measures of the true SMBH mass.

80 Fig. 3.1.— Mean and rms Hα spectra for NGC 3227 (a and b), NGC 3516 (c and d), and NGC 4593 (e and f). The y-axis scales of Fλ are arbitrary and the continua have been subtracted.

81 Fig. 3.2.— Mean and rms Hβ spectra for NGC 3227 (a and b), NGC 3516 (c and d), and NGC 4593 (e and f). The y-axis scales of Fλ are arbitrary and the continua have been subtracted.

82 Fig. 3.3.— Cross-correlation results for Hα (left panels) and Hβ (right panels) with continuum auto-correlation functions (dotted lines; cross-correlation of the continuum with itself) for NGC 3227 (a and b), NGC 3516 (c and d), and NGC 4593 (e and f).

83 Fig. 3.4.— The relationship between black-hole mass and the central value of host- galaxy bulge velocity dispersion for quiescent and active galaxies. The filled circles represent quiescent galaxies (data kindly provided by L. Ferrarese) and the filled triangles are AGNs from Ferrarese et al. (2001). The open triangles are the three

AGNs discussed in this paper. The dashed line is the best fit to the quiescent-galaxy

4.58 data, MBH σ (Ferrarese 2002). ∝

84 Table 3.1. Sampling Statistics

Sampling Interval (days)

Galaxy Dataset Number Average Median Fvar Rmax

85 NGC 3227 Hα 23 6.9 5.1 0.017 1.367 0.155 § Hβ 14 10.6 10.9 0.133 1.597 0.181 · · · § NGC 3516 Hα 18 8.9 8.8 0.129 1.541 0.065 § Hβ 18 9.0 8.0 0.112 1.474 0.063 · · · § cont. 35 4.5 3.0 0.280 4.027 0.364 · · · § NGC 4593 Hα 22 7.4 5.0 0.136 1.653 0.069 § Hβ 11 15.8 11.7 0.185 1.996 0.102 · · · § Table 3.2. Cross-Correlation Results

rest rest Galaxy Emission Line Continuum τcent τcent τpeak τpeak (A)˚ (days) (days) (days) (days)

86 +14.9 +14.8 +18.8 +18.7 NGC 3227 Hα (24% stellar) 5020 25.9−10.0 25.8−10.0 25.0−9.6 24.9−9.6 +26.8 +26.7 +26.6 +26.5 Hβ (24% stellar) 12.0−9.1 12.0−9.1 13.0−13.1 13.0−13.1 · · · · · · +5.9 +5.8 +5.3 +5.2 NGC 3516 Hα composite 13.2−2.6 13.1−2.6 13.8−3.8 13.7−3.8 +5.4 +5.4 +0.9 +0.9 Hβ 7.4−2.6 7.3−2.5 10.2−6.7 10.1−6.6 · · · · · · +2.5 +2.5 +7.8 +7.7 NGC 4593 Hα 6310 4.6−5.0 4.6−5.0 1.7−1.7 1.7−1.7 Hβ 3.1+7.6 3.1+7.5 9.6+1.2 9.5+1.2 · · · · · · −5.1 −5.1 −12.6 −12.5 Table 3.3. Velocity Width Results

Hα Hβ

Galaxy Dataset V V rest V V rest (103 km s−1) (103 km s−1) (103 km s−1) (103 km s−1) 87

NGC 3227 rms 3.10 0.15 3.09 0.15 4.36 1.32 4.34 1.31 § § § § NGC 3516 3.11 0.04 3.08 0.04 3.14 0.15 3.11 0.15 · · · § § § § NGC 4593 3.10 1.09 3.07 1.08 4.42 0.95 4.38 0.94 · · · § § § § NGC 3227 mean 2.62 0.22 2.61 0.22 1.90 0.57 1.89 0.57 § § § § NGC 3516 3.51 0.10 3.48 0.10 4.09 0.77 4.05 0.76 · · · § § § § NGC 4593 3.27 0.66 3.24 0.65 3.31 0.93 3.28 0.92 · · · § § § § Table 3.4. Reverberation Masses

Galaxy Hα-based Hβ-based Combined 6 6 6 (10 M¯) (10 M¯) (10 M¯)

NGC 3227 36.1+21.0 33.1+76.3 36 14 −14.4 −32.1 § NGC 3516 18.2+8.1 10.3+7.7 16.8 3.3 −3.6 −3.7 § NGC 4593 6.3+5.6 8.7+21.3 6.6 5.2 −8.2 −14.8 §

88 Chapter 4

The Black Hole Mass Stellar Velocity − Dispersion Relation for AGNs

4.1. Introduction

The advent of techniques for measuring masses of supermassive black holes

(BHs) has led to the identification of correlations between the BH mass (MBH) and various properties of the host galaxies. One of the tightest of these relationships is with the velocity dispersion of the bulge or spheroid (σ∗; Ferrarese & Merritt 2000;

Gebhardt et al. 2000a). The objects defining the initial MBH–σ∗ relationship were primarily quiescent galaxies, with MBH determined from or gas dynamics. However, galaxies hosting an (AGN), in which the BH mass was measured via reverberation mapping (Blandford & McKee 1982;

Peterson 1993), have been found to be consistent with following the same correlation

(Gebhardt et al. 2000b; Ferrarese et al. 2001; Chap. 3).

89 The values of MBH derived from reverberation mapping are subject to certain systematic uncertainties in that the kinematics and geometry of the BLR introduces a scaling factor, f, into the reverberation mass equation:

r V 2 M = f . (4.1) BH G

Simple models of BLR morphologies yield f parameters on the order of unity, and the application of such estimates, as in the above references, places the previous

AGN MBH–σ∗ data among the locus of quiescent galaxies.

A uniform relationship between the BH mass and properties of the host galaxy on size scales beyond the strong gravitational influence of the BH implies a causal connection between the formation of the galaxy and the central black hole. Many investigators have explored possible mechanisms for how the evolution of the BH and galaxy could be linked (e.g., Silk & Rees 1998; Haehnelt & Kauffmann 2000;

Adams, Graff, & Richstone 2001; Umemura 2001; Miralda-Escud´e & Kollmeier 2003;

Merritt & Poon 2004), while others have looked for outliers from these relationships as probes of the physical drivers of the correlations (e.g. Mathur, Kuraszkiewicz, &

Czerny 2001; Wandel 2002; Bian & Zhao 2004; Grupe & Mathur 2004).

We present measurements of stellar velocity dispersions for 6 reverberation- mapped AGNs, significantly enlarging the sample of objects which can be used to observationally investigate the MBH–σ∗ relationship for active galaxies. In 4.2, we § 90 describe our observations and analysis method. We present our results and discuss the implications for the ensemble average value of the reverberation mapping scaling factor, f , in 4.3. Our conclusions are summarized in 4.4. h i § §

4.2. Observations and Data Reduction

The velocity dispersions for our sample of AGNs were measured using the near-infrared Ca II triplet (CaT). These stellar absorption lines, at rest wavelengths of λ8498, 8542, and 8662 A,˚ occur in a region of relatively low AGN contribution

(Nelson & Whittle 1995), and are accessible to ground-based observations for sources with redshift z < 0.068 (water vapor bands begin to reduce atmospheric ∼ transparency at longer wavelengths).

Observations were conducted at Kitt Peak National Observatory (KPNO),

Cerro Tololo Inter-American Observatory (CTIO), and MDM Observatory. The general observing strategy at each of the telescopes was similar. Long-slit spectra of each target were bracketed by quartz-lamp flat fields and wavelength calibration exposures. The total exposure time for each AGN typically exceeded 10,000 seconds.

In addition to our targets, we obtained spectra of late-type giant stars (G8 III–K6

III) to use as spectral templates. Details of the observing runs are given below and in Table 4.1.

91 KPNO Observations. We observed at the Mayall 4 m telescope with the

Ritchey-Chr´etien (R-C) Spectrograph. The BL380 grating (1200 lines mm−1, blazed at 9000 A)˚ was used with an RG695 blocking filter and the LB1A CCD. The slit width was 200, with a dispersion of 0.45 A˚ pixel−1 and a wavelength range of

8250–9130 A.˚

CTIO Observations. We used the V.M. Blanco 4 m telescope with the R-C

Spectrograph, equipped with the KPGLD-1 grating (790 lines mm−1, blazed at

8500 A)˚ and an RG665 filter. We used the lower right amplifier for the Loral 3K 1

CCD detector within a window of 3072 585 pixels. The reduced spectra have a × dispersion of 0.83 A˚ pixel−1 over a wavelength range of 7500–10060 A.˚ The slit dimensions were 2 34400. ×

MDM Observations. Our first observations at the MDM 2.4 m telescope were made with the MDM Observatory Modular Spectrograph (ModSpec), using a grating of 830.8 lines mm−1, blazed at 8465 A.˚ The OG515 order-separating filter was used.

The detector for these observations was “Charlotte”, a thinned, backside-illuminated,

SITe 1024 1024 CCD. The spectra covered a range of 1400 A,˚ with a dispersion × ∼ of about 1.43 A˚ pixel−1, and a slit width of approximately 200. The second MDM observing run used the ModSpec setup described above, but utilized a different

92 detector. “Echelle”, a thinned, backside-illuminated, SITe 2048 2048 CCD was used × with a windowing of 330 2048 pixels. ×

4.2.1. Data Reduction and Analysis

We used IRAF1 and XVista2 for the reduction of different subsets of the data. However, the basic strategy was the same: the spectra were flat-fielded, sky-subtracted, and placed on a linear wavelength scale.

The AGN spectra in our sample were often contaminated by the broad O I

λ8446 emission line that appears blueward of the CaT, and these lines were removed with a high order polynomial in order to isolate the CaT absorption lines. None of our objects show evidence of emission from high-order Paschen emission lines, which can also contaminate the CaT lines.

1IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation 2See http://ganymede.nmsu.edu/holtz/xvista

93 4.3. Results and Discussion

4.3.1. Measuring Velocity Dispersions

We measured the velocity dispersions from the reduced spectra with the Fourier correlation quotient method (FCQ; Bender 1990). Cross-correlating the galaxy spectra with those of the template stars produces values for the central velocity dispersions.

Our galaxy spectra and the best template-star fits from the FCQ routine are shown in Figure 4.1. The results of our velocity dispersion analysis are listed in

Table 4.2. As noted in Ferrarese et al. (2001), the formal errors reported by the FCQ

fitting routine tend to underestimate the total uncertainties for the σ∗ measurements.

Thus, in cases where the FCQ output errors were smaller than 15%, the error bars have been brought up to this threshold.

Based on visual inspection, the least satisfactory fit from Figure 4.1 is that of Akn 120. However, consistent (within 1 σ) results are found from three other methods: (1) a simple Gaussian fit to the strongest absorption line; (2) the second moment of the profile for that line; and (3) an IDL routine for penalized pixel

fitting3.

3See Cappellari & Emsellem (2004), and Chap. 5.

94 We examined whether the AGN continuum could dilute the stellar absorption lines in a manner that would affect the σ∗ measurements. This was tested by artificially including an additional continuum contribution and running the FCQ routine again. Over a large range of continuum levels, no significant change was seen in the resulting σ∗ values.

Additional stellar velocity dispersion data for galaxies hosting AGNs have been taken from the literature.

4.3.2. Measuring Virial Products

New analysis of reverberation mapping data has produced updated measurements of the virial products (r V 2/G) for a sample of 35 AGNs, including the 16 which now have velocity dispersions (Peterson et al. 2004). These virial products, which, from Equation 4.1, can also be written as MBH/f, differ from the majority of previous reverberation mapping analyses in the calculation of V .

Whereas earlier work typically characterized the velocity width via the full-width at half-maximum (FWHM) of the emission line, Peterson et al. (2004) have found more consistent virial products when using the line dispersion (i.e., the second moment of the line profile; σline) instead. This will have important implications for comparing our results with previous publications.

95 4.3.3. The BH Mass–Velocity Dispersion Relation

We assume that the MBH–σ∗ correlation found in quiescent galaxies also holds for AGNs and their host galaxies. We can then determine the factor required to scale the AGN virial products (MBH/f) to the quiescent galaxy relation (holding σ∗

fixed at the measured value); that is, we calculate the average scaling factor, f , h i that, when multiplied by the virial products, brings the AGN MBH–σ∗ relation into agreement with the quiescent galaxy relationship.

Because published determinations of the slope of the quiescent galaxy relation have yet to converge, we selected the most prominent slope values from either end of the quoted range. Thus, we consider both the Tremaine et al. (2002; hereafter T02) slope of 4.02, and the Ferrarese (2002; F02, henceforth) slope of 4.58, and determine separate scaling factors, fT and fF , respectively. h i h i

While questions remain as to the reliability of the current fits to the quiescent galaxy MBH–σ∗ relation, especially with respect to the accuracy of the black hole masses from stellar kinematics (see Valluri, Merritt, & Emsellem 2004; Cretton &

Emsellem 2004; Richstone et al. 2004), we hope our use of the two slope values will give a more reliable estimate of the uncertainties involved while also avoiding entanglement in the continuing controversy.

96 The best fit to the AGN MBH–σ∗ values was accomplished with the orthogonal regression program, GaussFit4 (version 3.53; Jefferys, Fitzpatrick, & McArthur

1988), which accounts for errors in both (MBH/f) and σ∗. For the purposes of our

GaussFit analysis, the asymmetric variances in the virial products were symmetrized as the mean of the upper and lower variances. A fit was made to the equation

M σ log BH = α + β log ∗ , (4.2) Ã f ! µ200¶ where α is the normalization of the relationship, β is the (fixed) value of the slope,

−1 and where (MBH/f) is in solar masses and σ∗ is in km s . For each slope, the best-fit value of α was calculated. The χ2 which was minimized to determine the best fit is given by

2 N ( MBH ) α β( σ∗ ) 2 f i − − 200 i χ h 2 2 2 i , (4.3) ≡ σMi + β σσi Xi=1

where σMi and σσi are the uncertainties in the virial product and the velocity dispersion, respectively, for the ith data pair (Press et al. 1992). The virial products and velocity dispersions used in our fits are given in Table 4.3.

The published values of the quiescent galaxy intercept correspond to αF =

8.22 0.08 and α = 8.13 0.06. Our fits with fixed slopes are shown in Figure 4.2, § T § and yield αAGN = 7.48 0.13 and αAGN = 7.39 0.12. Remarkably, both slopes give F § T § 4GaussFit is available at ftp://clyde.as.utexas.edu/pub/gaussfit/

97 the same best fit value, log f = 0.74. Hence, fF = 5.5 1.9 and fT = 5.5 1.7. h i h i § h i § 2 2 The χ per degree of freedom, χν, for these two fits are 2.90 and 2.87, respectively.

Table 4.4 shows the resulting α and β values for a variety of fitting constraints.

When we allow β to vary in the fit to the AGNs, the result is consistent with both the T02 and F02 slopes, albeit with a large uncertainty. The distribution of points with previous σ∗ measurements in Figure 4.2 differs from that of Ferrarese et al. (2001) because of the improved virial product data from Peterson et al. (2004).

Because the reverberation mapping results for IC 4329A and NGC 4593 provide only upper limits on the black hole masses, these data were omitted from the above analysis. The formal error bars on MBH can extend below zero because the uncertainty of the time delay is not restricted to positive values. If we take the uncertainty in log MBH to be σM/(M ln 10), we can avoid taking the logarithms of negative numbers, and these two data points can be included in the

fit. The uncertainties are large enough, however, that the fit is not changed in a statistically significant way. Additionally, the fit is not strongly affected by the exclusion of NGC 4051 (see Table 4.4), which seems to be an outlier in the AGN radius-luminosity relationship (see the Appendix of Vestergaard 2002). When we do not include NGC 4051, we find fF =5.1 and fT =5.2, differences of less than 8% h i h i from our best fit value of 5.5.

98 The most common assumption for converting a virial product to MBH is that of

Netzer (1990)

3 r FWHM2 M = , (4.4) BH 4 G

which implicitly assumes an isotropic velocity distribution and V = FWHM/2.

Peterson et al. (2004) measure the second moment of the line profile, σline, rather than the FWHM, so an isotropic velocity field simply gives f = 3 with no other assumptions. Denoting the scale factor as ² when using the FWHM and f when using σline, we find that whereas Netzer’s assumptions give ² = 0.75 (as in Eq. 4.4), the value of f we derive above implies ² = 1.4, i.e., black hole masses 1.8 times h i h i ∼ 5 higher . Our masses are only 62% as large as those expected from the σline FWHM − relation found in a model-dependent fit to the distribution of FWHMs in a sample of and Seyfert galaxies (McLure & Dunlop 2001).

Applying our derived value of f = 5.5 (or ² = 1.4) will remove the systematic h i h i bias between the virial product and mass of the black hole (modulo the accuracy of the particular MBH σ∗ fit chosen). Yet, estimates for individual AGN black − hole masses may fall substantially off the MBH σ∗ relation, as there are still rms −

5 Our value of ² is calculated assuming σline = FWHM/2; for typical line profiles, changing the shape of the line alters the σline FWHM conversion by an amount smaller than the observed scatter − in the MBH σ∗ data. − 99 deviations from the fits in the mass direction of factors of 2.9 (F02 slope) and 2.6

(T02 slope).

Whether the scatter in MBH is due to differences in the BLR properties in each

AGN (corresponding to slightly different f values for each object) or a problem in our assumption that all AGNs fall on the quiescent galaxy relationship remains unclear. However, models of the BLR may shed light on this question.

4.3.4. BLR Models

Given our determination of f , we can try to learn something about the h i structure and kinematics of the BLR. Modeling of the BLR, taking into account different radial profiles and discrete time sampling of example light curves, will be undertaken in a later paper. Here we describe only a thin, rotating ring. In this case, f can be thought of as relating the Keplerian velocity (Vrot) at the radius of the ring

(which is traced by the appropriate time delay) to the observed line dispersion:

V 2 f = rot , (4.5) µσline ¶

For a ring of inclination i (where i=0◦ for face-on), this model gives f = 2 ln 2 / sin2 i.

Working within a thin-ring model context, Wu & Han (2001) calculated the inclinations necessary to scale published reverberation mapping results for 11 AGNs to the quiescent galaxy MBH σ∗ relation. With our revised virial product results, − 100 we find that this very simple model fails for two of our 16 AGNs. NGC 3783 and

Mrk 110 would require sin i values greater than unity to scale them to the quiescent galaxy MBH σ∗ relation. For our other sources, we do not find any evidence to − support the inclination trends with line width or radio loudness that were claimed by Wu & Han. Further constraints on the BLR geometry are not feasible given the sizes of the current uncertainties.

4.3.5. Gravitational Redshifts

In principle, an independent measure of the AGN black hole mass can be obtained by detection of gravitational redshifting of the emission lines (e.g., Netzer

1977; Peterson et al. 1984). Kollatschny (2003b) finds that the variable parts of the strong broad lines in the optical spectrum of Mrk 110 are redshifted relative to the systemic velocity. The reverberation time lag and redshift of each line relative to systemic are anticorrelated and appear to be consistent with a gravitational

8 redshift caused by a central mass Mgrav 1.4 10 M¯, rather higher than our ≈ × 6 7 reverberation-based mass (Table 3) of MBH = 2.5 10 M¯. ×

6 7 Kollatschny (2003a) obtains a reverberation-based mass estimate of 1.8 10 M¯, but using × FWHM as the line-width measure and ² = 1.5 (see 3.1). The FWHM data of Peterson et al. (2004) § gives a result consistent with that of Kollatschny, however our preferred method of measuring σline and using the empirical calibration of f yields a black hole mass approximately 40% larger. h i 101 Regardless of the particular value of f appropriate to Mrk 110, the offset between the reverberation-based virial product and the gravitational redshift mass is in the direction expected. However, the value of Mgrav places Mrk 110 even further

6 above the MBH σ∗ relation, which predicts MBH 4 10 M¯. − ≈ ×

If the BLR is modeled as a simple disk, then

(M /f) sin2 i BH = , (4.6) Mgrav 2 ln 2

and the difference between these two measurements allows us to infer that i 30◦, ≈ slightly larger than the 21 5 degrees found by Kollatschny (2003b). §

Measuring gravitational redshifts appears to be a promising technique for independently estimating central masses. However, applying the method is non-trivial because the gravitational redshifts are small, typically only several tens to a few hundred kilometers per second. Moreover, it is not a shortcut around reverberation mapping because (a) the redshift must be measured in the variable part of the emission line (i.e., the root-mean-square spectrum formed from the monitoring data; see Peterson et al. 2004) and (b) the time lag, which can vary with time, needs to be measured simultaneously. The spectral resolution required to measure such small redshifts reliably is higher than usually employed in reverberation mapping campaigns, with the exception of Kollatschny’s (2003a) program on Mrk 110 and the monitoring program on NGC 5548 (Korista et al.

102 1995). A re-examination of the NGC 5548 spectra from the latter campaign shows some evidence for redshifts at the expected levels, although the errors are quite large. Additional data will thus be required to examine this method of black hole mass measurement more thoroughly and to determine whether Mrk 110 is truly a significant outlier from the MBH σ∗ relation. −

4.3.6. Velocity Dispersion Versus FWHM([O III]λ5007 A)˚

Several recent studies have suggested the use of the FWHM of the [O III]λ5007 A˚ emission line as a proxy for σ∗. Nelson & Whittle (1996) examined a sample of 66

Seyfert galaxies with both σ∗ and FWHM([O III]) measurements and found scatter of 0.20 dex around a 1:1 correspondence (see their Figures 5 and 7a). Based on these trends, Nelson (2000) and Shields et al. (2003) used [O III] data as a substitute for σ∗ in AGNs out to z 3. Boroson (2003) applied the same arguments to data from the ∼ Sloan Digital Sky Survey Early Data Release and found that using FWHM([O III]) could reproduce an MBH σ∗-like relationship, although with a larger scatter than − has been found for quiescent galaxies (i.e., a factor of 5). Statistically identical ∼ distributions in FWHM([O III]) were found for the broad-line and narrow-line

Seyfert 1s in an X-ray selected sample of AGNs, which placed the narrow-line Seyfert

1s preferentially below the MBH σ∗ relation (Grupe & Mathur 2004). −

103 With the FWHM([O III]) data tabulated by Nelson (2000) for our 16 AGNs, we looked at the relationship with σ∗. Figure 4.3 shows that 25% of the sources have FWHM([O III]) data that deviate by >0.2 dex from the values expected based on their velocity dispersions. There is no evidence of the correlation between the discrepant objects and their radio power that was found for a larger sample of

Seyferts (Nelson & Whittle 1996), indicating that these differences may not be due to a systematic acceleration of the [O III]-emitting gas by the radio source. Overall, the weighted mean difference from equality is 0.03 dex (in the sense of larger FWHM relative to σ∗) with an rms scatter of 0.15 dex. In the way of commentary, we simply echo sentiments expressed by others that while use of such a proxy may be valid for a large sample of AGNs, it can also fail dramatically for individual objects.

4.4. Conclusion

With the addition of 6 new velocity dispersion measurements for reverberation- mapped AGNs and making use of improved reverberation mapping results, we tie together the MBH σ∗ relationships for quiescent galaxies and AGNs. This allows − us to calculate the average scaling factor, f , which removes the statistical bias h i between the virial product generated by reverberation mapping and the black hole mass. For the F02 and T02 fits to the quiescent galaxy MBH σ∗ relationship, − we find f = 5.5 1.9 and 5.5 1.7, respectively. These values of f apply h i § § h i

104 specifically to virial products using the dispersion of the emission lines, rather than measurements of the FWHM. While modeling of the BLR and studies of emission line gravitational redshifts may eventually lead to a better understanding of the structure and kinematics of the BLR, further work is needed.

105 Fig. 4.1.— Normalized spectra of the CaT region for the 6 AGNs in this study. Each spectrum has been offset in flux for clarity. The dashed line indicates the best fit obtained with the FCQ method. The spectrum of Akn 120 is truncated where the

FCQ fit becomes unstable. The spectra shown for Mrk 590 and NGC 7469 are from

KPNO; IC 4329A and NGC 3783 are from MDM. 106 Fig. 4.2.— Virial product, MBH/f, versus host galaxy velocity dispersion, σ∗. Solid points indicate objects with σ∗ measurements presented here. Open points represent

AGNs with previously published σ∗ data. The solid line indicates the F02 slope of

4.58, with the y-intercept shifted downward by fF . The dotted line denotes the T02 h i slope of 4.02, with the y-intercept scaled down by fT . The vertical scale on the h i right uses our derived offset: fF = fT =5.5. h i h i

107 Fig. 4.3.— σ∗ versus FWHM([O III]) for our 16 AGNs. The solid line shows a 1:1 correspondence between σ∗ and the equivalent for a Gaussian line profile,

FWHM([O III]) / 2.35.

108 Table 4.1. Observing Log

Resolution Targets Run Telescope Instrument UT Dates (km s−1) Observed 109

MDM-a MDM 2.4m ModSpec 2001 Oct 20–24 95 Mrk 590, NGC 3227, NGC 7469 KPNO KPNO 4m R-C Spec 2001 Oct 29–Nov 1 60 Akn 120, Mrk 590, NGC 7469 MDM-b MDM 2.4m ModSpec 2003 Mar 12–13 95 IC 4329A, NGC 3783 CTIO CTIO 4m R-C Spec 2003 Apr 18 90 IC 4329A, NGC 3783 Table 4.2. Velocity Dispersion Measurements

Redshift σ∗ Galaxy Redshift References Run (km s−1)

Akn 120 0.032296 0.000143 1 KPNO 239 36 § § IC 4329A 0.016054 0.000050 2 CTIO 131+20 § −60 MDM-b 121 18 · · · § Mrk 590 0.026385 0.000040 3 KPNO 201 30 § § MDM-a 188 28 · · · § NGC 3227 0.003859 0.000010 3 MDM-a 139 21 § § NGC 3783 0.009730 0.000007 4 CTIO 87 13 § § MDM-b 108 16 · · · § NGC 7469 0.016317 0.000007 5 KPNO 149 22 § § MDM-a 157 24 · · · §

References. — (1) Falco et al. 1999; (2) Willmer et al. 1991; (3) de Vaucouleurs et al. 1991; (4) Theureau et al. 1998; (5) Keel 1996.

110 Table 4.3. MBH σ∗ Data −

a b Virial Product Black Hole Mass σ∗ (avg) σ∗ 6 6 −1 Galaxy (10 M¯) (10 M¯) (km s ) References

3C 120 10.1+5.7 55.6+31.4 162 24 2 −4.1 −22.3 § 3C 390.3 52.2 11.7 289 64 240 36 3 § § § Akn 120 27.2 3.5 150 19 239 36 1 § § § IC 4329Ac 1.80+3.25 9.9+17.9 122 13 1 −2.16 −11.9 § Mrk 79 9.52 2.61 52.4 14.4 130 20 4 § § § Mrk 110 4.57 1.10 25.1 6.1 86 13 4 § § § Mrk 590 8.64 1.34 47.5 7.4 194 20 1 § § § Mrk 817 8.98 1.40 49.4 7.7 142 21 4 § § § NGC 3227 7.67 3.90 42.2 21.5 131 11 1, 2 § § § NGC 3516 7.76 2.65 42.7 14.6 164 35 5 § § § NGC 3783 5.42 0.99 29.8 5.4 95 10 1 § § § NGC 4051 0.348 0.142 1.91 0.78 84 9 2, 4 § § § NGC 4151 2.42 0.83 13.3 4.6 93 14 4 § § § NGC 4593c 0.98+1.70 5.36+9.37 124 29 2 −1.26 −6.95 § NGC 5548 12.20 0.47 67.1 2.6 183 27 4 § § § NGC 7469 2.21 0.25 12.2 1.4 152 16 1 § § §

aFrom Peterson et al. (2004).

bScaled using f = 5.5.

cExcluded from fits.

References. — (1) This work; (2) Nelson & Whittle 1995; (3) Green et al. 2003; (4) Ferrarese et al. 2001; (5) Arribas et al. 1997.

111 Table 4.4. MBH σ∗ Fitting Results −

Constraint Slope (β) Intercept (α) χ2 f ν h i F02 slope 4.58 7.48 0.13 2.90 5.5 1.9 § § T02 slope 4.02 7.39 0.12 2.87 5.5 1.7 § § none 4.11 1.07 7.40 0.21 3.11 N/A § § F02 slope, no NGC 4051 4.58 7.50 0.14 3.09 5.2 1.9 § § T02 slope, no NGC 4051 4.02 7.42 0.12 2.96 5.1 1.6 § §

112 Chapter 5

Stellar Dynamics of NGC 4151

5.1. Introduction

Another way to calibrate the BH masses from reverberation mapping is to apply to reverberation-mapped AGNs one of the techniques commonly used for estimating

BH masses in quiescent galaxies. The advantage of this approach over that described in Chapter 4 is that it does not assume anything about the relationship between the host galaxy and the BH. One disadvantage is that the number of AGNs accessible to these types of studies with existing facilities is quite small.

Based on the revised reverberation-based BH masses of Peterson et al. (2004a), we identified two candidates for our stellar dynamics program: NGC 3227 and

NGC 4151, both local Seyfert galaxies for which the sphere of influence for the BH was expected to be large enough to probe with Hubble Space Telescope (HST; see

5.2 for more details). In order to determine the dynamics of the stars, we needed § to be able to accurately measure the absorption lines arising in the atmospheres of stars near the BH. The bulge velocity dispersions of both of our candidate AGNs had

113 been previously measured with the Ca II triplet (CaT; for NGC 3227, see Chap. 4, and for NGC 4151, see Ferrarese et al. 2001), allowing us to estimate the strength of these absorption lines for spectra from smaller angular/spatial scales. However, the continuum flux of the AGN acts like a source of noise over the spectral region of the CaT lines, leading us to propose a Target-of-Opportunity (ToO) observation for the Space Telescope Imaging Spectrograph (STIS) on HST, which could execute our program when the AGN went into a faint flux state.

Monitoring of our two targets was carried out with ground-based spectroscopy at the Crimean Astrophysical Observatory (CrAO) and with space-based X-ray

flux measurements from the Rossi X-ray Timing Experiment (RXTE). In late

November 2003, NGC 4151 was found to have fallen below our critical threshold for observation (confirmed with spectra from the MDM Observatory), and the ToO was triggered. Observations began approximately 3 weeks later and the spectroscopy was completed by the end of December 2003. The goal of the project was to combine the HST spectra with a mixture of archival and new ground-based spectroscopy, along with imaging from both space and the ground, and then to construct stellar dynamical models of NGC 4151 to test whether a BH mass consistent with the reverberation-based value was required to match the observations.

114 In 5.3 we describe our imaging data, and our spectroscopic data in 5.4.1 We § § explain the process of the stellar dynamical modeling and show the results of our fits in 5.5. We discuss the outcome of the project and future steps in 5.6. § §

5.2. Stellar Dynamics Overview

Here we briefly recap the procedure involved in stellar dynamics studies and note some of the data and modeling requirements. Our approach makes use of the orbit superposition algorithm first described by Schwarzschild (1979).

Imaging data is used to construct a mass model of the galaxy, and many thousands of possible stellar orbits are followed through the mass distribution for several hundred rotations to build up an orbit library. The contribution of each orbit to the observed distribution of stellar velocities is determined within the constraints of the mass model. Comparisons of the quality of the fit are made between models having different values of the BH mass to determine the best estimate of the BH mass and its uncertainty.

In order to obtain a reliable estimate of the BH mass, it is important to spatially resolve the BH sphere of influence (Ferrarese & Merritt 2000), which is defined as the region where the circular velocity of stars in orbit around the BH

1Observations reported here were obtained at the MMT Observatory, a joint facility of the

University of Arizona and the Smithsonian Institution.

115 exceeds the velocity dispersion, σ∗, of the spheroidal component of the host galaxy

(meaning either σ∗ of a ’s bulge or of an itself). For

NGC 4151, the bulge velocity dispersion has been measured to be 93 14 km s−1 § (Ferrarese et al. 2001). If we assume the value of the BH mass from reverberation

7 mapping ((1.33 0.46) 10 M¯; Peterson et al. 2004a), then the sphere of influence § × has a radius of 6.9 pc. The published heliocentric of the galaxy is

998 km s−1 (Pedlar et al. 1992), implying a distance of 13.9 Mpc if in pure Hubble

−1 −1 expansion with H0 = 72 km s Mpc . However, other estimates of the distance to NGC 4151 have ranged from 10 to 30 Mpc based on various estimates of the

Virgocentric correction (see discussion in 1 of Mundell et al. 1999). We adopt a § distance of 13.9 Mpc, and note later the impact of this assumption. At our inferred distance, 100 corresponds to 67 pc, so the HST resolution of 000.10 corresponds to

6.7 pc, just inside the BH sphere of influence.

Concerns have been raised regarding the robustness and reliability of some stellar dynamic BH mass measurements. Two-integral models (2-I; in which the distribution of orbits is constrained only by their energies, E, and vertical angular momenta, Lz) were used by Magorrian et al. (1998) to estimate BH masses in 32 galaxies. The application of three-integral (3-I) models (see 1.3 or 5.5.3) have §§ caused many of these measurements to be revised downward (see, e.g., Tremaine et al. 2002). Some degeneracies found in 2-I models are thought to be similarly important in 3-I models (Cretton & Emsellem 2004). Degeneracies in 3-I models

116 have also been discussed by Valluri et al. (2004; although see the counter-claims by

Richstone et al. 2004). Among the main groups developing stellar dynamics codes, no consensus has been reached regarding the importance of these potential degeneracies or how to best deal with them. One way to minimize the role of degeneracies is via the process of regularization, imposing a level of smoothness upon the distributions of the various orbit parameters (e.g., E and Lz in 2-I models). Although one must be careful about biasing the derived solution by the use of regularization (see Cretton

& Emsellem 2004), we adopt this tool in the analysis presented below.

5.3. Imaging

Along with the STIS spectroscopy, our ToO proposal included 320 s of imaging with the Near-Infrared Camera and Multi-Object Spectrometer (NICMOS) on HST in the F160W filter. It was hoped that these images would be able to penetrate the gas and dust around the nucleus of the target galaxy and provide a clear view of the stellar distribution. The brightness of the AGN relative to the stellar surface brightness meant that the short exposure times required to keep the nucleus from saturating prevented any significant stellar signal from rising above the wings of the nuclear point spread function (PSF). As an alternative, we made use of a set of ground-based and HST imaging data obtained with the High Resolution Camera

(HRC) of the Advanced Camera for Surveys (ACS) for an independent project

117 designed to study host galaxy contamination in AGN spectra (Program ID 9851, PI:

B. Peterson). These data used an optical filter (F550M) rather than the infrared of

NICMOS, and so are more susceptible to obscuration, causing complications which we describe below ( 5.5.1). §

5.3.1. ACS/HRC

Imaging data with ACS/HRC were obtained in the F550M filter, totaling

1020 s of exposure time. These images were processed through the standard Space

Telescope Science Data Analysis System (STSDAS) pipeline for IRAF2 for bias removal, flat-fielding, and cosmic-ray removal. Additional bad pixels and residual cosmic-rays were removed by hand. Then the individual images were drizzled and combined. The final ACS image is shown in Figure 5.1. The image has a plate scale of 000.027 pixel−1 and a PSF full width at half maximum (FWHM) of 000.06. ≈

5.3.2. MDM

The MDM Observatory 1.3 m telescope was used with the “Templeton” CCD to take direct images of NGC 4151 in the B, V , and R filters. The images in each

filter were bias-subtracted, flat-fielded, registered, and then median-combined. The

R-band image was used to determine the surface brightness profile, and it is shown

2See http://www.stsci.edu/resources/software_hardware/stsdas

118 in Figure 5.2. The image has a plate scale of 000.508 pixel−1 and a PSF FWHM of

200.3. Flux calibration was achieved through aperture photometry of a nearby star ≈ (Eyermann et al. 2005). The B and V images were used with the R-band data to estimate the stellar mass-to-light ratio (M/L) in the galaxy. The images were aligned, convolved with Gaussians of appropriate size to bring the PSFs to the same size as the band with the worst seeing, and a pixel-by-pixel color map was created

(Figure 5.3). The (B R) and (V R) colors were determined as a function of − − radius and from the colors, we estimate the R-band M/L to be 3 (Bell & de Jong ≈ 2001; Bell et al. 2003).

5.4. Spectroscopy

For the purposes of stellar dynamical modeling, it is useful to have data on both small and large spatial scales, and also at multiple spectroscopic slit position angles (PAs). As the STIS PA was left unconstrained to facilitate the execution of our ToO, the supplementary spectroscopy at the same PA could not be obtained in advance. Once the ToO was completed, we were able to follow up with ground-based spectra at larger radii.

The threshold for triggering the ToO had been derived from ground-based

R-band observations and bulge-disk-nucleus decompositions of NGC 3227 and

NGC 4151 (Virani et al. 2000). These data allowed us to estimate the stellar flux

119 we could expect to fall within the STIS slit and also to independently examine the effects of varying the nuclear brightness on the signal-to-noise ratio, S/N. We assumed a stellar contribution near the center of each galaxy to be roughly like an elliptical galaxy (flat in Fλ), with a spatial distribution given by the bulge fit of

Virani et al. (2000). Thus, we were able to parameterize the exposure time as a function of the nuclear non-stellar flux. We calculated the thresholds of nuclear flux for each target that allowed us to reach our desired S/N of 50 A˚−1. For NGC 4151, this threshold was equal to 4 10−14 erg s−1 cm−2A˚−1, which the AGN had been × near in the recent past.

On 19 November 2003, the CrAO observations of NGC 4151 indicated that the continuum flux level had fallen below the threshold required for the triggering of the ToO. Figure 5.4 shows a historical light curve for NGC 4151 and indicates with arrows the date of a previous STIS observation (one employing the STIS occulting bar to block the nuclear light, but which also blocked the stellar signature near the

BH), the time when we were able to execute our ToO, and the date at which the

STIS instrument stopped functioning.

5.4.1. HST/STIS ToO

The instrumental setup for our STIS observations made use of the 5200 000.1 slit × and the G750M grating. A total exposure time of 29,730 s was achieved. In addition

120 to the observations of NGC 4151, we obtained spectra of a K-giant (HR 4521) with the same instrumental setup to act as our comparison for the CaT line profiles.

While the continuum flux stayed close to the threshold value between the time of triggering on 20 November 2003 and the time of observation on 22-29 December

2003, the extracted STIS spectrum (Figure 5.5) does not show the presence of the CaT lines, and so we are unable to use the STIS data to constrain the stellar kinematics near the BH in NGC 4151. It remains unclear as to whether the failure to detect the stellar lines originates from a miscalculation of the required integration time (either because of an error in extrapolating from the Virani et al. (2000) photometric parameters or because of a flawed bulge-disk-nucleus decomposition) or from a change in the emission characteristics of the AGN at low flux levels (e.g, at low accretion rates, the BLR gas goes into a state with strong CaT lines in emission, which fill in the signal expected to be seen from the stars near the nucleus) or from a change in the stellar population near the center of the galaxy (i.e., there are fewer

G- and K-giants to produce strong CaT absorption lines than at larger bulge radii).

The lack of absorption lines can be used to place limits on the number of giant stars in the central region of NGC 4151. The final spectrum of the comparison star

HR 4521 was used to estimate the maximum number of such stars that could exist in the center of NGC 4151 without producing measurable lines. HR 4521 has a

Hipparcos parallax measurement of 15.80 0.59 mas, implying a distance of 63 3 pc. § § We then scaled the luminosity of HR 4521 to the flux appropriate for our assumed

121 distance of NGC 4151 (13.9 Mpc), and found the multiplicative factors required to produce visible spectral features when subtracting the scaled stellar spectra from

NGC 4151 spectra of different aperture sizes. For the inner 000.1 (6.7 pc), we derive an upper limit of 7.5 105 giant stars, while the limit moves to 106 stars for an × extraction window of 000.2 (13.4 pc). A wide-aperture extraction ( 300 000.1, or ∼ × 200 pc 6.7 pc) only increases the upper limit to 1.25 106 stars. ∼ × ×

If the AGN itself is filling in the CaT lines by producing those lines in emission, then we may be underestimating the allowable stellar density. On the other hand, if we assume MV = 0.2 (based on an interpolation of Table 3.10 in Binney & Merrifield

1998, and the spectral classification of HR 4521 of K2.5 IIIb by Keenan & McNeil

1989) rather than using the Hipparcos parallax, this suggests a much larger distance to HR 4521 of 103 pc, and the corresponding limits on the stellar density are reduced by a factor of 3. ∼

5.4.2. Ground-based

Following the execution of the ToO observations of NGC 4151 (although before it became clear that the STIS data would not produce a useful stellar signature), we arranged to obtain follow-up spectra of the galaxy at the PA at which the STIS observations had been conducted. The timing of the ToO meant that the first opportunity to propose for ground-based telescope time was through the MMT

122 Observatory. In addition to the MMT observations, we revisited the Kitt Peak

National Observatory (KPNO) spectra of NGC 4151 that had been used by Ferrarese et al. (2001) to determine σ∗ and we attempted to extract more detailed information on the stellar kinematics along the (different) PA of the KPNO data.

MMT

We observed NGC 4151 and comparison stars with the Blue Channel

Spectrograph at the 6.5 m MMT on UT 29 May 2004. We used the 1200 lines mm−1 grating with the LP-530 order-blocking filter, a 100 slit width, and the “ccd35” detector. The spectra were centered at 8600 A˚ for measurement of the CaT with a dispersion of 0.50 A˚ pixel−1 and a resolution of 1.4 A.˚ The plate scale was 000.3 pixel−1, and the PA was set to match the STIS slit angle of 69◦. While the photometric conditions were quite good, strong and gusty winds caused both poor seeing and problematic guiding. Hence, our total exposure time of 11700 s was not able to take full advantage of the MMT’s large collecting area.

The spectra were rectified and wavelength-calibrated using the bright sky lines that permeate the observed spectral region (the same wavelength solutions were applied to the spectra of HR 4521, as the exposures for the stellar targets were too short for sky lines to appear and the lines present in the HeNeAr calibration lamp spectra were much more sparse). Series of flat fields taken among the observations of

123 NGC 4151 were employed to correct the strong fringing (20% peak-to-peak) present in the spectra.

Next, 1-D spectra were extracted in 1-pixel apertures out to 1200, tracing § along the peak of the spatial profile. While the seeing disk (modeled as a Gaussian with σ = 100.27) was much larger than our individual extraction windows, the dynamical modeling routine takes the seeing into account when fitting to the observed kinematics. The extracted spectra were then normalized with a spline fit to both the continuum and any extant AGN emission features to isolate the CaT lines for profile fitting.

KPNO

The work of Ferrarese et al. (2001), on which Chapter 4 expanded, used the

Mayall 4 m telescope at KPNO on UT 9 April 2001 to measure the bulge velocity dispersion of NGC 4151 with the Ritchey-Chr´etien Spectrograph. Details of the observations were fully described in Ferrarese et al. (2001), but we restate the key features here. The observations were taken with a 200 slit at PA=135◦, the detector plate scale was 000.69 pixel−1, and the spectral resolution was 1.7 A.˚ The total ≈ exposure time was 3600 s.

The observations were rectified and wavelength-calibrated with the sky lines present in the spectra, and then combined. Extractions were made with a variable

124 window size, ranging from 1 pixel at the peak of the light to a 7-pixel width in the furthest bins (extending to radii of 1400). The seeing for these observations was ≈ modeled as a Gaussian with σ = 000.76. Observations of HR 4521 were obtained for the absorption line profile measurements and were reduced in the same way as the

AGN spectra.

5.5. Analysis & Results

We extracted the required information from the data in two distinct steps.

The first involved the analysis of the imaging data and the construction of the the mass model for NGC 4151 from the surface brightness distribution. The second task was to determine the stellar dynamics from the absorption line profiles in the spectroscopic data. These components were then combined in the orbital modeling to establish limits on the mass of the BH in the center of NGC 4151. We describe each of these portions of the analysis in turn.

5.5.1. Surface Brightness Profile Modeling

The two pieces of data from which we derive the surface brightness profile are the MDM R-band image and the ACS/HRC F550M image. As the goal of this procedure is to develop a model for the mass distribution in the galaxy, we need to remove the AGN contribution to the light profile and extract the stellar

125 surface brightness. The process of removing the AGN light is complicated by the very different spatial characteristics and filter bandpasses of the two images. We use GALFIT3, the publicly available 2-D galaxy fitting software (Peng et al. 2002), to remove the AGN contamination and also to provide a smoothed version of the underlying surface brightness for comparison in later modeling steps. In order to most accurately subtract the AGN flux, we perform a multi-component fit to the major surface brightness features of the galaxy and jointly determine the galaxy properties and the AGN flux to be removed.

For the MDM R-band image, we subtract a sky level equal to the median flux beyond the visible extent of NGC 4151, fit two S´ersic profiles for the disk and the bulge, and use a PSF for the nuclear emission that is based on one of the stellar profiles in the image. The fit is done in several steps, initially masking out the bulge and nucleus in order to fit the disk. Next, we remove the mask from the bulge and use the earlier results from the disk-only fit as the starting values for that component.

Then we remove the mask from the nucleus and allow the PSF magnitude to vary.

We left the mask in place over several stars in the field and over a number of bright knots near the ends of the weak bar in NGC 4151. Because of slight differences in the profile shapes of the PSF and the AGN light, manual adjustment of the PSF magnitude and position was required to produce the final fit. The parameters of our ultimate fit are shown in Table 5.6, and the residual image is shown in Figure 5.6.

3http://zwicky.as.arizona.edu/~cyp/work/galfit/galfit.html

126 The small field of view of the ACS/HRC image meant there were no galaxy-free regions from which to measure the sky flux. Instead, the background count rate for the HRC in the F550M filter was retrieved from the ACS Instrument Handbook

(Pavlovsky et al. 2004) and combined with the total exposure time to estimate the background flux. The effective radii of the two S´ersic profiles from the MDM image fit were converted to the appropriate pixel scale for the ACS image and held

fixed. The total magnitudes of the components were allowed to vary to account for differences arising from the different bandpasses between the images. We attempted to model the HRC PSF with Tiny Tim4, which produces PSFs for the various HST instruments at any desired position in the field. Tiny Tim also allows one to input a user-defined spectrum to better match any chromatic distortions in the optical path.

We joined two archival STIS spectra of NGC 4151 taken with the G430L and G750L low-resolution gratings and a 5200 000.1 slit (Program ID 7569, PI: J. Hutchings; × Kaiser et al. 2000), and fed the resulting spectrum to Tiny Tim. However, the resulting PSF proved to be a poor match to the nuclear emission. Thus, we searched for all archival stellar images taken with the HRC in the F550M filter. Unfortunately, this search turned up only a handful of white dwarf observations, all with relatively short exposure times. The best of these was a 2 s image of BD+17◦4708 (Program

ID 9664, PI: R. Gilliland; Bohlin & Gilliland 2004), and this image was then used as

4http://www.stsci.edu/software/tinytim/tinytim.html

127 our HRC PSF. The residual of our fit to the ACS image is shown in Figure 5.7 and the fitting parameters are given in Table 5.6.

It is unclear whether the strong residuals near the center of the HRC image are primarily due to dust obscuration or to [O III] λ5007 emission leaking into the

F550M bandpass. To check for the second case, we retrieved an [O III] image of

NGC 4151 taken with the linear ramp filter on the Wide Field Planetary Camera 2

(WFPC2) instrument of HST (Program ID 8019, PI: J. Hutchings; Hutchings et al.

1999). After registering the images and resampling the HRC image to the slightly larger pixel scale of the WFPC2 image, we were unable to find a scaling factor for the [O III] image that produced a satisfactory subtraction of the remaining residuals.

While some of the positive-flux regions in the residual image may be due to [O III] emission, it is likely that obscuration also plays an important role in shaping the

F550M image.

With the nuclear emission subtracted to the best of our ability, we proceeded with the surface brightness fitting. Rather than simply using the results of our

GALFIT exercise, we employed the Multi-Gaussian Expansion (MGE) routine described by Cappellari (2002), as it provides for a more trivial deprojection of the

2-D distribution into a 3-D model. The MGE routine fits a series of 10 concentric ≈ elliptical Gaussian profiles to the surface brightness distribution. (The program has

128 been made publicly available5 for use with IDL6.) In addition to the MGE fits to the nuclear-subtracted images, we also ran the program on the best-fit GALFIT model to the ACS image. This allowed us to test the impact of the ACS residuals on the

MGE results.

The MGE routine performs a simultaneous fit to the MDM and ACS images, and this allowed us to rescale the ACS flux level to match that of the MDM image where the two overlapped in radius. Thus, we fixed all of the flux levels to that determined by the calibrated stellar photometry in the R-band, for which straight-forward M/L calculations can be made. The results of our MGE fits are shown in Figures 5.8 and 5.9 for the two cases of using the ACS image and using the

GALFIT fit to the ACS image. Using the distance to NGC 4151, our photometric calibration, and the R-band extinction for the direction of NGC 4151 given by NED7, we converted the MGE output of counts in each Gaussian to solar luminosities. Our calculated Gaussian parameters are given in Table 5.6.

5.5.2. Gauss-Hermite Parameter Estimation

The spectroscopic data were analyzed with the Penalized Pixel-Fitting (pPXF) method of Cappellari & Emsellem (2004)8. The pPXF program finds the best fit

5http://www.strw.leidenuniv.nl/~mcappell/idl/ 6http://www.rsinc.com/idl/ 7http://nedwww.ipac.caltech.edu/ 8The IDL routine is publicly available from http://www.strw.leidenuniv.nl/~mcappell/idl/.

129 to the line-of-sight velocity distribution (LOSVD) based on a maximum penalized likelihood approach to the broadening of a template spectrum. The LOSVD is most often described by a Gauss-Hermite (GH) parameterization (even in cases where the

LOSVD is fit non-parametrically), which is given by

2 e−y /2 M (v) = 1 + hmHm(y) , (5.1) L σ√2π " =3 # mX

where (v) is the LOSVD as a function of velocity v, V is the velocity difference L between the target and template spectra, σ is the velocity dispersion, y (v V )/σ, ≡ −

Hm are the Hermite polynomials, and hm are the weights of the respective Hms.

The GH expansion is typically taken to the 4th or 6th components in stellar dynamics applications. The pPXF routine uses a stellar spectrum as its template and simultaneously fits to all desired GH terms over a user-defined wavelength region. The ability to select multiple fitting regions allows us to exclude regions of noise or AGN emission that might lie between the lines we are trying to measure.

The pPXF program also allows for multiplicative or additive scaling factors between the template and target spectra. Legendre polynomials of a user-specified order can also be fit by the routine and we take advantage of this feature to remove the flux level and broad shape differences between the spectra of HR 4521 and

NGC 4151.

130 Each of the extracted spectra from the MMT and KPNO data were run through the pPXF algorithm with the corresponding template of HR 4521. Examples of the resulting fits are shown in Figures 5.10 and 5.11, and the components of the fits are shown as functions of slit position in Figures 5.12 and 5.13.

5.5.3. Orbit Modeling

The results of our MGE fit were deprojected into a 3-D luminosity distribution and, with the R-band M/L determination from the (B R) and (V R) colors (which − − we assume to be constant with radius), were converted into a mass distribution. This provided the first input to our Schwarzschild orbit superposition routine. Details of the 3-I modeling technique can be found in Cretton et al. (1999); Valluri et al.

(2004, 2005). While the inclination of the galaxy disk is closer to face-on (i 20◦ ≈ Simkin 1975), our model assumes an edge-on axisymmetric mass distribution. For the bulge and for orbits near the BH, the assumption of edge-on axisymmetry should be reasonable, however this is an area which deserves further consideration.

Also, the weak, large-scale bar in NGC 4151 may perturb the stellar orbits in an important way, but such effects cannot be tested with the simple models employed here. Instead, we simply try to establish a best-case upper limit on the BH mass in

NGC 4151.

131 Our orbit models use the light distribution between 000.01 and 30000, and a spatial grid with 20 logarithmically spaced radial bins and 16 tangential bins spaced linearly in azimuth. We computed 5040 total orbits with 35 values of E, 8 values of Lz, and 9 values of the third isolating integral. These orbits are then fit to our

56 velocity constraints and 251 spatial (luminosity) constraints. To the basic mass model we add the gravitational influence of a BH of different masses and compare the quality of the fit from each case. We show the results of our orbit model fitting to the spatial and kinematic data in Figures 5.14–5.18 for a potential that includes a BH of the mass predicted by reverberation mapping. The fits with no BH are indistinguishable from those with a BH.

Unfortunately, the lack of STIS kinematic constraints means we are unable to provide a meaningful limit on the BH mass. The ground-based data simply do not probe the dynamics on a small enough spatial scale to reveal the presence of the

BH. Moreover, the basic nature of our modeling leaves open the possibility that large degeneracies might have remained even if the STIS observations had been successful. As a consequence of our inability to provide a measurement of the BH mass, the systematic uncertainty arising from the wide range of possible distances to NGC 4151 becomes moot.

132 5.6. Discussion

The current program failed to adequately constrain the BH mass in NGC 4151.

However, this galaxy remains one of the best targets for future stellar dynamical measurements in broad-line AGNs. While the failure of STIS precludes deeper studies of the CaT at the spatial resolution of HST, ground-based observations with adaptive optics (AO) on large-aperture telescopes still hold promise for unraveling the stellar dynamics of NGC 4151. A further improvement over our approach would be to expand from a few slit positions to many, or even to the use of integral

field units (IFUs), which can take simultaneous spectra of 50 small, densely ∼ packed spatial regions. A wider array of instrumentation becomes available if one moves from the CaT to near-infrared observations of the CO bandhead at 2.29 µm.

There are several large telescopes with AO and IFUs capable of observing the CO bandhead. The observations of Ivanov et al. (2000) showed very weak CO features in NGC 4151, but those observations were taken in 1996, during the period of peak

flux seen in Figure 5.4.

The other broad-line AGN which is likely close enough to have its BH sphere of influence resolved is our other ToO target, NGC 3227. The spectra of its CO bandhead also look more promising (see Ivanov et al. 2000). How much the CO observations would benefit from the AGN being in a faint flux state is not yet

133 clear, but the best way forward might in fact be another ToO program on a large ground-based telescope.

Ultimately, if the remaining questions regarding the reliability of the stellar dynamical modeling are removed, the insensitivity of stellar orbits to the outflows and bulk motions that affect gas dynamics means that stellar dynamical BH masses will act as the linchpin in determining the calibration of masses produced by reverberation mapping. But, sadly, that time has not yet arrived.

134 Fig. 5.1.— ACS image of NGC 4151 with the F550M filter. North is up and East is to the left. The indicated scale denotes 1 arcsec.

135 Fig. 5.2.— MDM image of NGC 4151 with the R-band filter. North is up and East is to the left. The indicated scale denotes 1 arcmin.

136 1

0.8

0.6

0.4

0.2

0

-0.2 -1 0 1 2 log R (arcsec)

2

1.5

1

0.5 -1 0 1 2 log R (arcsec)

Fig. 5.3.— Top: (V R) color as a function of galactocentric radius (in arcseconds). − Bottom: (B R) color as a function of radius. −

137 Fig. 5.4.— Historical light curve of optical continuum flux. Arrows indicate date of an earlier STIS program that used the occulting bar (blue; GO-7350), the date of our

ToO observations (green; GO-9849), and the date when STIS stopped functioning

(red).

138 Fig. 5.5.— Extracted spectrum from the combined STIS data. The arrows indicate the positions of the CaT lines for the redshift of NGC 4151. The feature seen near the central arrow is not the CaT line, but a residual fringe not removed by the flat

fields. The y-axis scale is arbitrary.

139 Fig. 5.6.— Residuals from the GALFIT fit to the R-band image. The orientation is as in Figure 5.2 and the indicated scale denotes 1 arcmin.

140 Fig. 5.7.— Residuals from the GALFIT fit to the F550M image. The orientation is as in Figure 5.1 and the indicated scale denotes 1 arcsec.

141 Fig. 5.8.— Contours of the MGE fit to the F550M image surface brightness. Smooth lines show the model fit, while the noisy lines indicate the contours of the data.

142 Fig. 5.9.— Contours of the MGE fit to the R-band image surface brightness. Smooth lines show the model fit, while the noisy lines indicate the contours of the data.

143 144

Fig. 5.10.— Example pPXF fit to MMT spectroscopic data. Smooth lines show the stellar template convolved with the

GH parameters inferred from the fit. 145

Fig. 5.11.— Example pPXF fit to KPNO data. Smooth lines show the stellar template convolved with the GH parameters

inferred from the fit. Fig. 5.12.— Profile of GH parameter values as a function of slit position in arcseconds for MMT extractions.

146 Fig. 5.13.— Profile of GH parameter values versus slit position in arcseconds for

KPNO extractions.

147 148

Fig. 5.14.— Mass profile as a function of radius in arcseconds. Left-hand panels show values inferred from the MMT

7 data (crosses) and the results of the orbit model with a 10 M¯ BH (solid lines). Right-hand panels show the same for

the KPNO data. 149

Fig. 5.15.— Recessional velocity as a function of radius. Left- and right-hand panels as in Figure 5.14. 150

Fig. 5.16.— Radial profiles of velocity dispersion. Left- and right-hand panels as in Figure 5.14. 151

Fig. 5.17.— Radial profiles of the third GH term. Left- and right-hand panels as in Figure 5.14. 152

Fig. 5.18.— Radial profile of the fourth GH term. Left- and right-hand panels as in Figure 5.14. Table 5.1. GALFIT Results for MDM Image of NGC 4151

Component Profile Index Profile Scale Radius Total Magnitude

Bulge S´ersic 2.74 22.73 11.09 Disk S´ersic 0.69 128.8 10.87 Nuclear PSF 11.77 · · · · · ·

Note. — Radius given in pixels, with a plate scale of 000.508 pixel−1.

153 Table 5.2. GALFIT Results for ACS Image of NGC 4151

Component S´ersic Profile Index S´ersic Profile Scale Radius Total Magnitude 154 Bulge S´ersic 2.65 314.7 12.39 Nuclear PSF 15.51 · · · · · ·

Note. — Radius given in pixels, with a plate scale of 000.027 pixel−1. Table 5.3. MGE Results for Combined Imaging Data of NGC 4151

MDM Data + ACS Data MDM Data + ACS Model Surface Brightness Gaussian σ Gaussian Surface Brightness Gaussian σ Gaussian 2 2 (L¯/arcsec ) (arcsec) Axis Ratio (L¯/arcsec ) (arcsec) Axis Ratio

2.41 1010 0.135 1.00 2.90 107 0.017 0.81 × × 4.45 109 0.352 1.00 2.76 108 0.091 1.00 × × 1.93 109 1.350 1.00 7.22 108 0.223 1.00 × × 4.72 108 3.176 1.00 1.30 109 0.469 1.00 × × 155 2.37 108 8.121 1.00 1.78 109 0.899 1.00 × × 4.23 106 41.075 0.51 1.35 109 1.692 1.00 × × 3.08 107 44.872 0.68 6.32 108 4.058 1.00 × × 1.07 106 229.418 1.00 7.03 107 7.354 1.00 × × 9.29 107 10.521 1.00 · · · · · · · · · × 3.15 107 44.422 0.67 · · · · · · · · · × 9.06 105 127.003 1.00 · · · · · · · · · × 8.99 105 127.003 0.53 · · · · · · · · · ×

Note. — All non-circular Gaussians are oriented with their major axis aligned with the major axis of the large-scale isophotes. Chapter 6

Black Hole Masses and Eddington Ratios at 0.3 < z < 4

6.1. Introduction

For well over 30 years, the basic theory of active galactic nuclei (AGNs) has been that they are luminous because of the accretion of matter onto black holes (BHs;

Salpeter 1964; Zel’dovich & Novikov 1964; Lynden-Bell 1969). In this picture, the luminosity produced by a BH of mass MBH has a natural maximum, the Eddington limit (LEdd) at which the radiation pressure due to the accretion of the infalling matter balances the gravitational attraction of the BH. Most models for AGNs assume that they are BHs radiating near the Eddington limit, and as techniques for independently estimating BH masses in AGNs have been developed, it has become possible to test this supposition. In particular, large, modern spectroscopic surveys can provide estimates of the Eddington ratios (the ratio of the AGN bolometric luminosity to the Eddington limit) for thousands of AGNs (see, e.g., analyses of the

Sloan Digital Sky Survey [SDSS] by McLure & Dunlop [2004] and M. Vestergaard et al., in preparation). Unfortunately, the shallowness of these large, wide-area surveys

156 imposes severe restrictions on the combinations of Eddington ratio and BH mass

9 that are observable, especially at z > 1. For MBH < 10 M¯, the SDSS is sensitive ∼ only to near-Eddington radiators above this redshift, and even at z < 1 the SDSS analyses to date have not clearly established whether there is a lower cutoff to the

Lbol/LEdd distribution at fixed BH mass. The Warner et al. (2004) analysis of over

500 BH masses with (0 z 5) AGNs included data from a heterogeneous sample ≤ ≤ of smaller, deeper studies. While they found a broad range of Eddington ratios, it is difficult to draw any conclusions about the underlying distribution of Lbol/LEdd because their dataset is composed of multiple samples with different selection criteria.

The AGN and Galaxy Evolution Survey1 (AGES; Kochanek et al. 2004) probes nearly a decade further down the AGN luminosity function than the SDSS. For the first time, this permits a relatively unbiased measurement of the distribution of

Eddington ratios at fixed BH mass at z 1. The distribution at fixed luminosity ≥ can be measured down to z = 0.5. The AGES survey uses the 300-fiber Hectospec robotic spectrograph on the MMT (Fabricant et al. 1998) to survey galaxies and

AGNs in the Bo¨otes field of the NOAO Deep Wide-Field Survey2 (NDWFS; Jannuzi

& Dey 1999). Both a population of high mass BHs radiating significantly below

Eddington and a population of low mass BHs radiating near or above Eddington would be observable with AGES. In this paper we show: (1) that the distribution

1http://cmb.as.arizona.edu/~{}eisenste/AGES/ 2http://www.noao.edu/noao/noaodeep/

157 of Eddington ratios at fixed BH mass or at fixed luminosity is narrowly peaked and well-described by a single log-normal distribution independent of redshift and luminosity, (2) that this peak occurs at roughly 1/3 of the Eddington limit, and

(3) that the rms error in BH mass estimates from emission-line scaling relations is less than 0.3 dex at fixed luminosity. The first two conclusions imply that the luminous growth of BHs over cosmic time is dominated by objects radiating near the

Eddington limit.

We present a brief overview of our data from the AGES survey in 6.2 and § describe our method of analysis in 6.3. We present our BH mass estimates and § Eddington ratios as functions of redshift and luminosity in 6.4. Finally, we § discuss the implications of these results in 6.5. In our analysis we assume an § −1 −1 H0 = 72 km s Mpc , Ωm = 0.3, ΩΛ = 0.7, flat .

6.2. Data

AGES is a redshift survey in the roughly 9 deg2 Bo¨otes Field of the NOAO

Deep Wide-Field Survey (NDWFS; Jannuzi & Dey 1999). Subsequent surveys have imaged the field at many wavelengths, providing a rich multi-wavelength data set for the detected sources. In this paper, we make particular use of data from the

Chandra X-ray Observatory (XBo¨otes; Murray et al. 2005; Brand et al. 2005; Kenter

158 et al. 2005), and the Spitzer Space Telescope (MIPS3: Soifer et al. 2004; IRAC:

Eisenhardt et al. 2004).

The AGES-I Survey (C. Kochanek et al., in preparation) selected AGN candidates as either X-ray or 24µm sources with R 21.5 mag (Vega) optical ≤ point source counterparts. Objects were considered to be point sources if they were point-like in any one of (R, I, BW). Since the luminosities we consider here are higher than the canonical Seyfert luminosity, there should be little contribution from host galaxies to the fluxes measured for these AGNs. The primary sample of z > 1

AGNs consists of either XBo¨otes sources with 4 counts (in exposures averaging ≥ 5 ks) or 24µm sources brighter than 1 mJy that are off the stellar locus [2MASS

J > 12 + log(F24µm)]. These are supplemented by X-ray sources with 2 or 3 counts and 24µm sources with 0.5 mJy < F24µm < 1 mJy. Redshifts were obtained with

Hectospec, the multi-object fiber spectrograph at the MMT Observatory (Fabricant et al. 1998). The spectra were analyzed by two independent pipelines and verified by eye. This led to a sample of 733 broad-line AGNs with z > 0.1. We analyze a subset of this sample for which we can reliably measure emission line widths, as described in 6.3.1. §

3The Spitzer MIPS survey of the Bootes region was obtained using Guaranteed Time Observations provided by the Spitzer Infrared Spectrograph Team (James Houck, P.I.) and by M. Rieke.

159 6.2.1. Completeness

There are three issues for understanding the completeness of our sample: the literal completeness of the spectroscopy, the effects of the optical flux limits, and the effects of the 24µm/X-ray flux limits. The first issue, the completeness of the spectroscopy, is not an issue. While we did not obtain spectra of every candidate, we did obtain redshifts for 97% of the candidates for which we obtained spectra.

These represented only 66% of the candidates, but the candidates with spectra can be regarded as a “random” sub-sample of the candidates dictated by the fraction of the NDWFS region covered by spectroscopy and whether it was possible to assign

fibers to the candidates. Presumably neither of these issues are correlated with either Eddington factors or black hole masses. Thus, the only issue is whether the spectrum allowed the measurement of the line FWHM, and we discuss this in 6.3.1. § The second issue, the optical flux limit imposed for the spectrographic targets we will include explicitly.

It is the third issue, the consequences of the 24µm and X-ray flux limits that we must consider in more detail. If we examine, for example, the distribution of

AGNs in the plane of the R-band and 24µm fluxes it is clear that both flux limits matter. Our spectroscopic limit is nominally defined by R = 21.5 mag. However, the

24µm+X-ray selected sample is only complete to R = 19.1 mag. We evaluate the problem and estimate a correction by using the deeper samples from the AGES-II

160 catalogs (C. Kochanek et al., in preparation). AGES-II includes AGNs selected using mid-infrared colors from the IRAC Shallow Survey (Eisenhardt et al. 2004) based on the color selection method outlined in Stern et al. (2005), as well as the

24µm and X-ray criterion used for AGES-I. This leads to an AGN sample with an

L-band (3.6µm) flux limit of L = 18 mag and an optical limit of I = 21.5 mag that fills in most of the missing quasars to the optical flux limit of AGES-I. We determine the completeness of the sample as a function of the R-band flux and then weight each detected AGN by the inverse of the completeness. The AGES-II

L = 18 mag limit implies that it is complete to R = 20.9 mag. Brighter than this magnitude, we simply compare the AGES-I and AGES-II samples to determine the required completeness correction. For fainter AGNs, we must apply an additional correction that incorporates the fraction of AGNs that are lost in AGES-II owing to their relatively blue colors with respect to the L-band limit. This correction factor is known from the brighter magnitudes where AGES-II is complete. We show the functional form for our completeness correction in Figure 6.1.

These corrections for completeness are important for the interpretation of the results in 6.4.2, and we return to the completeness corrections and how it affects § our conclusions there.

161 6.3. Analysis

BH masses have been estimated from Hβ emission line widths and luminosities for some time (e.g., Dibai 1980). However, it was only with the application of reverberation mapping (for an overview, see Peterson 2001) that this relationship became firmly established, and then was revised with improved data and techniques

(Kaspi et al. 1996; Wandel et al. 1999; Kaspi et al. 2000; McLure & Jarvis 2002;

Vestergaard 2002; Kaspi et al. 2005). The general form of the relation is:

log MBH = a + b log(λL44) + 2 log V, (6.1)

where MBH is the estimated BH mass in units of M¯, V is the full width at half

−1 maximum (FWHM) of the emission line in km s , and λL44 is the continuum luminosity near the line (at 5100 A˚ for Hβ) in units of 1044 erg s−1. Because optical spectra can only probe Hβ to a maximum redshift of z 0.75, studies have been ∼ undertaken to find scaling relationships for UV lines allowing the estimation of black hole masses at high redshift. For intermediate redshifts (0.4 < z < 2), McLure

& Jarvis (2002) determined a scaling relationship for Mg II λ2800 by comparing single-epoch measurements of Mg II FWHM and 3000 A˚ continuum luminosity with results from Hβ reverberation mapping (and assuming that the similar ionization potentials for the two lines cause them to be emitted at the same distance from the BH). A relation useful at higher redshifts (1.6 < z < 5) was established with

C IV λ1549 by Vestergaard (2002), who found that the assumption of an identical

162 luminosity scaling yielded consistency between Hβ reverberation mapping results and masses derived from single-epoch FWHM(C IV) and 1350 A˚ continuum luminosity measurements. While the response of Mg II in reverberation mapping campaigns has been rather weak (see Clavel et al. 1991 and Dietrich & Kollatschny 1995 for the best result, that of NGC 5548), studies of C IV have yielded BH masses in good agreement with those from Hβ (e.g., Chap. 2). There has been a long struggle to measure black hole masses from these emission lines and each line has associated peculiarities. We refer the reader to the original works, which address these issues in greater detail.

We adopt (a, b)=(0.68, 0.61) in Equation 6.1 for the Hβ relation (from McLure

& Jarvis 2002, although other versions of the relation are little different); we use

(a, b)=(0.20, 0.7) for C IV (from Vestergaard 2002); and we estimate in 6.3.3 that § (a, b)=(0.24, 0.91) for Mg II. We measure the line widths from the AGES-I spectra

( 6.3.1) and the continuum luminosity from the NDWFS photometry ( 6.3.2). § §

6.3.1. Line Width

The AGES-I spectra have a pixel scale of 1.2 A˚ and a resolution of 6 A˚ ≈ FWHM. For our analysis, we boxcar-smooth the spectra over 11 pixels, then subtract a locally-defined linear continuum from the region around the emission line of interest. The wavelength regions used to set the continuum around each line are

163 (4740-4765 A,˚ 5075-5100 A)˚ for Hβ, (2670-2682 A,˚ 2940-2970 A)˚ for Mg II, and

(1455-1465 A,˚ 1700-1705 A)˚ for C IV. We subtract narrow-line contributions to Hβ using the [O III] λ5007 line as a model, with the [O III] flux scaled by a factor of

0.15. This fiducial value for the (narrow Hβ)-to-[O III] flux ratio lies within the range defined by local AGNs that have only narrow emission lines (following Baldwin et al. 1981 and Veilleux & Osterbrock 1987). The subtraction of the narrow-line component based on this value proved satisfactory. We determine the peak flux in the line region and measure the FWHM of the line as follows (adapted from Peterson et al. 2004a). First we determine two wavelengths on either side of the line: (1) the first crossing of the half-max flux level moving downward from the line peak

(indicated as Blue1 and Red1 in Fig. 6.2); (2) the first half-max crossing moving upward from the line edge (Blue2 and Red2). The mean of these two wavelengths is taken as the half-max point for that side of the line (Blue and Red). The FWHM is defined as the difference between the Blue and Red points. The boxcar smoothing we apply serves to minimize the sensitivity of our automated procedure to noise in the spectra. This smoothing results in a final resolution of 11 A˚ corresponding to 500 km s−1. This adds a negligible systematic contribution to our measured ∼ FWHM at the 2% level, for which we do not correct. We estimate the error in ∼ the FWHM with the technique described by Corbett et al. (2003), in which line gradients at the half-max points are used to convert the Poisson noise of the spectral

164 counts at half-max into wavelength uncertainties. A minimum FWHM error of 10% is imposed on all measurements.

To remove spurious measurements from the dataset, every spectrum is examined for a series of problems: significant absorption in the line profile, strong Fe II emission around Mg II4, low signal-to-noise ratio (S/N), and other anomalous features. We remove 169 of the 733 AGNs in this process: 75 for absorption features, 18 with problematic Fe II emission, 42 with low spectroscopic S/N, and 34 other anomalies

(e.g., CCD defects). Representative spectra for the first three classes (and one good spectrum) are shown in Figure 6.3. Some spectra fall into multiple categories, and for our accounting they are assigned to the first class in this ordering of problems.

As a quality-control measure, we also estimate the FWHM by manual analysis for the best line in each spectrum. Objects for which the automated routine yields a FWHM differing by more than 1-σ from the manual estimate are deleted. This removes an additional 138 objects, leaving 426 AGNs in our final dataset. We argue below ( 6.4.2) that the removal of all of these objects is justified and that the § inclusion of these objects does not substantially alter our conclusions.

4Because of the shape of the Fe II complex near Mg II, the effects of weaker Fe II in objects that remained in our sample are not immediately clear. A careful removal of the Fe II in several test cases revealed no obvious trends between pre- and post-removal measurements (M. Dietrich 2005, private communication).

165 6.3.2. Continuum Luminosity

Our Hectospec observations were made during the inaugural season of operation of the instrument. During our runs, the atmospheric dispersion corrector did not function consistently, making it difficult to accurately flux calibrate our spectra.

We therefore estimate the continuum luminosities required for the MBH estimates from the broad-band magnitudes. Using the 600-aperture NDWFS R magnitude and its estimated bandpass, we calculate the flux at the band’s effective wavelength

(6515 A),˚ then compute the rest-frame 5100, 3000, or 1350 A˚ fluxes assuming a power-law continuum with F λ−1.7. This spectral index is the average spectral λ ∼ slope measured for the 700 AGNs of comparable luminosity in the sample of ∼ Dietrich et al. (2002). We attempted to estimate an individual spectral slope for each AGN using the broad-band colors, but this procedure failed to provide satisfactory fits to reliable portions of the spectra, leading us to impose the average spectral slope on all of our AGNs. The range around the median slope is less than

0.15 (M. Dietrich & F. Hamann, in preparation), so our adoption of a fixed slope contributes little, on average, to our total observational uncertainties. While Francis et al. (1991) found a larger scatter among spectral slopes in the Large Bright Quasar

Survey (a dispersion of 0.6 for 718 AGNs), for reasons described in 6.4.2, we ∼ § do not believe the spectral slopes for our AGNs can typically be that discrepant.

We also note that the Dietrich & Hamann measurements are consistent with the results from the SDSS composite spectra (Vanden Berk et al. 2001). Contributions

166 of emission lines to the R flux should be less than 0.2 mag, or 20% in luminosity, over our redshift range (M. Dietrich 2005, private communication).

6.3.3. Calibrating the Mg II Relation

Initially, we used the Mg II relation of McLure & Jarvis (2002), for which

(a, b)=(0.53, 0.47) in Equation (6.1). However, for AGNs in the redshift ranges in which we could estimate MBH using both Mg II and a second line

(0.4 z 0.75 for Hβ and 1.6 z 2.0 for C IV), we found systematic ≤ ≤ ≤ ≤ differences when using this relation. As shown in Figure 6.4, there is a linear trend in [log MBH(Mg II) log MBH(Hβ or C IV)] with log Lbol that has a similar slope for − both redshift ranges/lines. Fitting for the luminosity trend, we find a consistent relationship for both Hβ and C IV, showing that all three lines will be on a common scale. We redefine the Mg II relation from this fit and find that (a, b)=(0.24, 0.91), while the McLure & Jarvis (2002) slope is ruled out at the 7-σ level. Although we use this modified relation in the current work, we are not at this time advocating our revised relation for other studies because of the differences between our analysis and that presented in the McLure & Jarvis (2002) paper. After adjusting the slopes, the residual scatter between Mg II and Hβ or C IV mass estimates indicates an intrinsic dispersion beyond the measurement errors of 0.14 dex. This is significantly less than the 0.4 dex of intrinsic scatter found in earlier studies of Mg II (McLure &

Dunlop 2004) and suggests that Mg II MBH estimates are of comparable accuracy to

167 those from Hβ or C IV, although in principle all three line determinations could be affected by the same zero-point error.

In addition to the scaling with luminosity, we examined the mass difference as a function of Mg II FWHM, using the redefined luminosity-dependence of the Mg II relation (Fig. 6.5). We see that the masses are generally consistent, with a possible indication of a trend above 3700 km s−1. The potential correlation of mass difference with Mg II velocity above this threshold is an issue to which we will return in 6.4.2. §

6.3.4. Bolometric Luminosity Calculation

Following Kaspi et al. (2000), we estimate the bolometric luminosity as

Lbol 9 λL (5100 A),˚ a value that assumes an AGN spectral energy distribution ' × λ (SED) typical of optically selected quasars with little dust obscuration. We calculate the rest-frame flux at 5100 A˚ employing the same method that we used to estimate continuum luminosities in 6.3.2. While this extrapolation to 5100 A˚ is quite far § for the highest redshift AGNs in our sample, published conversions to Lbol using alternative continuum regions affect the resulting luminosities by less than 30% (for our assumed spectral slope). Of course, departures from the standard AGN SED could change the relation between Lbol and Lλ(5100 A),˚ so it is important to keep in mind that the “bolometric” luminosities referred to throughout this paper are really

168 just multiples of the optical luminosity. However, the fact that Elvis et al. (1994)

find a 5400 A˚ bolometric correction with a dispersion of just 0.16 dex is reassuring.

A primary goal of this study is to understand the mass distribution of active

BHs at a fixed luminosity and redshift. The homogeneous nature of our sample and the depth of the AGES-I survey combine to allow us to address the underlying distribution over a wider range of Lbol and z than other studies. Figure 6.6 shows our sample of AGN luminosities as a function of redshift along with curves at fixed magnitudes to illustrate how the flux limits restrict the sample. We also show how the AGES-I data are distributed relative to the knee of the luminosity function as determined by the 2dF-SDSS LRG and QSO Survey (2SLAQ; Richards et al. 2005).

In a bright survey like the SDSS, for which the i0 = 19.1 mag limit (for z < 3 AGNs) corresponds to R 19.3 mag and the i0 = 20.2 mag limit (for z > 3) corresponds ≈ to R 20.4 mag, one cannot compare AGNs of fixed luminosity over as broad a ≈ redshift range as with AGES-I.

6.4. Results

Figure 6.7 shows the distribution of AGNs in inferred BH mass, MBH, and bolometric luminosity, Lbol. The points are color-coded by redshift, and we show only a typical error bar to avoid clutter. These uncertainties reflect only the statistical errors in the line width and luminosity estimates, and we will discuss the

169 possible consequences of systematic errors in 6.5. Figure 6.7 has several striking § features. First, it is characterized by a fairly narrow ridge that extends diagonally, with unit slope, across the entire diagram. Second, this ridge is separated from the solid line representing the Eddington limit by approximately 0.5 dex. Third, at a given luminosity, the density of points falls rapidly toward higher MBH, i.e., at small Eddington ratio. Most of the conclusions of this paper are derived by quantifying these features. At fixed Lbol, the ridge is broader than the statistical uncertainties, indicating that it is dominated either by intrinsic scatter in the underlying distribution or by scatter between the inferred values of MBH and

Lbol and their true values. While the systematic errors in Lbol should be modest, determining the systematic errors in MBH is a notoriously difficult problem to which we will return. However, from the ridge’s relatively narrow width it is already clear that this scatter cannot be too large. The fact that the ridge is displaced from the Eddington limit shows that the great majority of observed AGNs radiate at modestly sub-Eddington rates, assuming that our MBH calibration is correct in the mean. The fact that there are few AGNs above the ridge implies that the observed

(in practice, luminosity-selected) AGNs are dominated by BHs radiating close to

Eddington rather than more massive BHs radiating well below Eddington. The data presented in Figure 6.7 also permit one to pose an orthogonal question: what is the distribution of AGN luminosities at fixed MBH? However, this question cannot be addressed by mere inspection of this figure and will require additional analysis.

170 In the next two subsections we examine the distributions of Eddington ratios at fixed Lbol ( 6.4.1) and fixed MBH ( 6.4.2) as functions of redshift. § §

6.4.1. Luminosity-Redshift Bins

In order to characterize the true distributions of masses and luminosities for the BHs in our sample we perform a very simple maximum likelihood analysis,

fitting the data assuming a model in which the true distribution of log(Lbol/LEdd) is

Gaussian (i.e., Lbol/LEdd is log-normal). We divide the sample into 2 bins in redshift and 3 bins in luminosity. The redshift division is at z = 1.2, and the luminosity cuts are at 1045.5 erg s−1 and 1046.0 erg s−1 in both redshift bins. Due to the different redshift distributions of sources at high and low luminosity, it is impossible to make cuts such that there are similar numbers of objects in each bin. We therefore choose cuts such that we have at least 10 objects in each bin. In each (L, z) bin we estimate the mean and dispersion of log(Lbol/LEdd), corrected for the statistical uncertainties.

The histograms in Figure 6.8 show the distributions of estimated log(Lbol/LEdd) in each (L, z) bin. For each panel, we calculate the unweighted mean and standard deviation of the data points, and we plot a Gaussian with those parameters as a dashed curve in the panel. Solid curves show the Gaussian with our maximum likelihood fit parameters, which accounts for the statistical errors in the linewidth and luminosity of each data point. Overall, these statistical errors are small

171 compared to the widths of the histograms, so there is little difference between the dashed and solid curves. Table 6.1 lists the measured mean, dispersion, skewness, and kurtosis of the distributions in each (L, z) bin, and the mean and dispersion of the maximum likelihood Gaussian fits.

The first point to note from Figure 6.8 is that the distribution of log(Lbol/LEdd) measurements in each luminosity bin is nearly Gaussian, with a center and width that is approximately independent of redshift and luminosity. This result quantifies the impression from Figure 6.7 that the observed AGNs in our sample are predominantly

BHs radiating fairly near Eddington (Lbol/LEdd 1/3), rather than more massive ∼

BHs at Lbol/LEdd 1. The skewness and kurtosis of the distributions are in good ¿ agreement with the values Sk = 0 and Ak = 3 expected for a Gaussian distribution.

The second point to note is that these Gaussians are rather narrow:

σ 0.3 dex after accounting for the statistical errors in our linewidth and luminosity ≈ measurements. These widths reflect the width of the intrinsic distribution of

Eddington ratios at fixed luminosity and the scatters between the inferred and true values of MBH and between the inferred and true values of Lbol. One might expect the scatter in bolometric corrections to be small, but the conclusion that the quadrature sum of measurement errors in MBH and variations in Lbol/LEdd is only 0.3 dex rms is rather remarkable. The data give no clear indication of how to partition the scatter between these two contributions. A rms scatter of 0.3 dex in inferred

MBH seems plausible on geometrical grounds, since the relation between observed

172 linewidth and BH mass may depend on viewing angle (Krolik 2001; C. Kuehn & B.

Peterson, in preparation). If the observational errors or the distribution of viewing angles do dominate the width of the histograms, then the intrinsic distribution of

Lbol/LEdd is very narrow indeed.

Figure 6.9 shows the log Lbol vs. log MBH positions of the AGNs eliminated due to either low S/N spectra or 1-σ discrepancies between automated and manual

FWHM measurements, for comparison with Figure 4. We have not included in

Figure 6.9 objects rejected due to absorption, Fe II emission, or other anomalous features. These objects (for which we show the Lbol distribution in the inset of

Fig. 6.9) have unmeasurable FWHMs, and it therefore is not sensible to assign any mass value, in contrast to AGNs for which we simply have low confidence in the measured value. Objects were rejected without reference to their position in

Figure 6.9, and there is no strong clustering of their position in this diagram. The simplest interpretation of this broad, flat distribution is that the measurements are simply bad, and the calculated MBH values bear only a casual relation to the true values. While the inclusion of these uniformly distributed bad measurements would slightly increase the dispersion in our Eddington ratio distribution, we believe that their rejection is justified. In any case, including them does not qualitatively alter our results. Hence, we adopt the conclusions based on their exclusion.

We stress that from an observational standpoint, the procedure followed in this section is extremely clean. Since the sample completeness is well-characterized

173 at fixed optical luminosity, and independent of Eddington ratio, there are almost no selection effects coming into play in any of the luminosity bins. Application of our completeness corrections have nearly zero effect on these histograms. On the other hand, within each luminosity bin there is a contribution to the dispersion in the histogram from uncertainties in estimating MBH. The robust nature of binning by luminosity also applies to AGN surveys with other flux limits. For example, in

Figure 3 of McLure & Dunlop (2004), who analyzed more than 12,000 AGNs from

SDSS, one can make similar cuts at constant M˙ BH ( Lbol). McLure & Dunlop ∝ (2004) calculated the mean Eddington ratio as a function of redshift, finding the mean to increase slightly with z (from 0.15 at z 0.2 to 0.5 at z 2.0), but ' ' they do not address the shape of the distribution at fixed luminosity, nor, for the redshifts that they can probe, the distribution at fixed mass. A much broader distribution of Eddington ratios is found in a sample of 234 AGNs compiled from the literature (Woo & Urry 2002). Their sample consists of low redshift (z < 1.0) and local (z < 0.1) systems, and extends to bolometric luminosities as low as

1043 erg s−1, so the difference from our results could plausibly be explained by the very different properties of the samples. The Warner et al. (2004) and Vestergaard

(2004) studies of heterogeneous samples of AGNs are most akin to what we have presented here. In each of those works, Eddington ratios are presented over a wide range of redshifts and luminosities. However, because each study relies on a diverse mixture of samples, the selection effects are not easily understood, and therefore,

174 the underlying distribution of Eddington ratios cannot be determined. Two of the main advantages of the AGES-I survey are the homogeneity of the sample and the relatively straightforward completeness corrections (see 6.2.1), which allow us to § address the underlying distribution of Eddington ratios.

6.4.2. Mass-Redshift Bins

The distribution of Eddington ratios at fixed luminosity is a convolution of the luminosity distribution at fixed BH mass with the BH mass function (see Steed &

Weinberg 2003). Thus, the falloff at low Eddington ratios in Figure 6.8 could in principle be attributed primarily to a rapid falloff in the MBH distribution toward higher masses. Since we know from our previous discussion that the errors in BH mass assignments must be fairly small (0.3 dex rms at most), we can bin the sample by estimated BH mass and look directly at the distribution of Eddington ratios at

fixed MBH. Figure 6.10 shows these distributions for mass bins log(MBH/M¯) = 7 8, − 8 9, and 9 10, and redshift bins z = 1 2, 2 3, and 3 4. As with the − − − − − luminosity-bin histograms, the distributions tend to peak at about 1/3 Eddington and fall off sharply toward both lower and higher Eddington ratios.

In contrast to Figure 6.8, however, the distribution of Eddington ratios at fixed mass is affected by the magnitude limit of the survey. We have included the effects of incompleteness at the faint end as described in 6.2.1. The solid histograms in §

175 each panel are completeness-corrected. This alters the histograms, but does not substantially alter our conclusions. The vertical arrow in each panel of Figure 6.10 indicates the point in that mass-redshift bin at which AGNs are first lost to optical selection effects (which begin at the low-MBH, high-z corner of the bin), and the shaded region indicates the point at which all AGNs (extending to the highest

MBH and lowest z) are lost. For the lower-mass bins (particularly at high redshift), the optical magnitude limits truncate the distributions, and we cannot determine whether there is a true cutoff at low Lbol/LEdd. However, in five mass-redshift bins

(three shown in Fig. 6.10 and two additional, smaller bins in Fig. 6.11), the optical magnitude limits only affect the regime log(Lbol/LEdd) < 0.1, and it is possible to ∼ test for a real cutoff. In one of these, the highest-mass highest-redshift bin, there are too few objects to make a robust determination. However, Figure 6.11 shows that in the remaining four bins — log(MBH/M¯) = 9 10 at z = 1 2 and 2 3, and − − − log(MBH/M¯) = 8.5 9 at z = 1 1.5 and 1.5 2 — the statistics are relatively − − − good and the falloff appears to be real. We have calculated the Poisson probabilities that the peak in each case is a statistical fluctuation, and determined this to be low. The peak is therefore robust and not merely a product of counting statistics.

The filled circles at the bottom of each panel in Figures 6.10 and 6.11 show the relevant spectroscopic limit of the SDSS for each redshift bin. The SDSS data are not deep enough to probe the cutoff in any of these bins, i.e., the SDSS cannot study

8.8 the distribution of Eddington ratios for BHs with MBH < 10 M¯ above z 1, or ∼

176 9.4 MBH < 10 M¯ above z 2. Vestergaard (2004) plots Eddington ratio values for ∼

150 SDSS AGNs above z = 3 with different point types for bins of MBH (her ∼ Fig. 6), but does not examine the effects of the SDSS flux limit on the underlying distribution. Because of its large area, the SDSS can obtain better statistics for high mass BHs at low redshift, or for very massive BHs at higher redshift, and it will be interesting to see whether these have a peaked Eddington ratio distribution similar to that found here.

Caveats

There are two possible ways to circumvent the conclusions we draw from the

AGES-I data. We consider these two possibilities in turn.

First, AGNs rejected from the survey due to poor MBH measurement could induce the illusion of a falloff if these lay preferentially near the selection boundary.

In Figure 6.9, however, we showed that the AGNs eliminated because of low S/N spectra or discrepant manual versus automated FWHM measurements did not depend on their location in the log Lbol vs. log MBH plane. Even if we include all of these points, as shown by the dashed histograms in Figure 6.12, it generally does not affect the appearance of the peaks nor our conclusions.

Second, the scatter in inferred MBH might move AGNs from one mass bin to another, which could potentially enhance the appearance of a peak either on

177 the rising or the falling sides. Recall from 6.4.1 that for high luminosities (and § correspondingly high masses) the rms scatter in the inferred MBH must be less than

0.3 dex. At least for the two high-mass bins in Figure 6.11, which have a width of 1 dex, scatter between mass bins must be a small effect. For the more narrow

(0.5 dex) mass bins, this argument does not apply, but the smooth distribution of Eddington ratios seen in Figure 6.7 implies that it would be difficult to import structure into or out of a particular mass bin without having significant structure in another part of the distribution from which objects can scatter. No such structure is seen in our data.

Mg II Line Width Validity

Also of concern is the nearly linear dependence of MBH on luminosity for the

Mg II relationship ( 6.3.3). With such a scaling relation, it would be possible § for a random distribution of Mg II FWHMs to reproduce a tight Eddington ratio distribution without actually being related to the BH mass. However, as

Figure 6.5 shows, the correlation between Mg II FWHM and the BH mass estimated from other lines is quite good. There may be an indication that the FWHM of

Mg II over-estimates the BH mass above 3700 km s−1. Approximately half of our

Mg II-measured AGNs have FWHM values above this value. Overestimation of the Mg II masses at high FWHM would only influence the upper left and lower right panels of Figure 6.11, and would do so in a way that tightens even further

178 the distribution of Eddington ratios. Because of the limited number of objects

−1 in Figure 6.5 with FWHM above 3700 km s and the already narrow Lbol/LEdd distribution, we refrain from adjusting further our Mg II-derived masses.

6.5. Discussion

Our survey of R 21.5 mag, X-ray and 24µm-selected AGNs in the Bo¨otes ≤ NDWFS field yields four basic results. First, the rms scatter in BH masses inferred from linewidth-luminosity scaling relations is less than 0.3 dex, at least for

8 MBH > 10 M¯ and the luminosity range studied here. Second, luminous AGNs at z > 0.5 are powered by BHs radiating at roughly 1/3 of the Eddington rate.

There are few cases of higher mass BHs radiating well below Eddington or of lower

8.5 mass BHs radiating well above Eddington. Third, at fixed mass above 10 M¯

(where selection effects are inconsequential), the distribution of Eddington ratios at z > 1 peaks strongly at Lbol/LEdd 0.3. Fourth, the distribution around this ∼ peak is confined to within 1 dex in Lbol/LEdd, with many fewer AGNs at either ∼

Lbol/LEdd = 1 or Lbol/LEdd = 0.1 than at the peak.

Our analysis does not test the zero-point or luminosity scaling of the MBH relations. To aid comparisons with previous work, we have used the scaling relations of McLure & Jarvis (2002) for Hβ and Vestergaard (2004) for C IV. The results of

Chapter 4, in which reverberation-based BH masses were calibrated by assuming that

179 AGNs follow the same correlation between BH mass and stellar velocity dispersion as quiescent galaxies, suggest that the zero points of these relations should be adjusted upward by factors of 1.4 and 1.8, respectively. We did adjust the McLure &

Jarvis (2002) Mg II relation based on the evidence for a different luminosity scaling shown in Figure 6.4. Changing the zero-points of the MBH relations, or changing the bolometric correction, would change our conclusions about the location of peaks in the Lbol/LEdd distributions, but they would not change our conclusions about the existence of these peaks. For example, the Chapter 4 revision would shift the means of our Lbol/LEdd distributions downward by 0.2 dex but would not broaden the ∼ distributions. Given the numerous complexities of BH mass estimates from emission line widths, our empirical evidence that these estimates have less than 0.3 dex rms scatter at fixed luminosity is remarkable. One might expect scatter nearly this large from geometrical effects alone, since the relation between observed linewidth and the

BH mass may depend on viewing angle.

The dominance of the z > 0.5 AGN population by near-Eddington accretors is very different from the local-universe sample of AGNs, for which the distribution of inferred Eddington ratios is broader (e.g., Woo & Urry 2002; Ho 2004; Heckman et al. 2004)). The luminosities of these local AGNs are typically much lower than those in our sample, so it is unclear whether luminosity or redshift dependence is responsible for most of the difference. The AGES-I sample does not probe down to

Seyfert luminosities (below 1044 ergs s−1), and we therefore cannot comment on ∼

180 the contribution of broad-lined Seyfert galaxies to the overall black hole growth. In addition, while narrow-line objects cannot be analyzed using the methods employed here, they are a very small fraction of the overall AGES-I AGN sample, with only 29 objects with 0.5 z 1 matching a narrow-line template and no such objects above ≤ ≤ z = 1. High Eddington ratios for the most luminous quasars at high redshift are not surprising, since the steep high-mass falloff of the BH mass function (see, e.g., Aller

& Richstone 2002) means that the BHs required to power them at sub-Eddington luminosities would be very rare. However, the dominance of near-Eddington accretors at the knee of the luminosity function and at redshifts at which the quasar luminosity function is declining in amplitude imposes strong constraints on the distribution of BH fueling rates (Steed & Weinberg 2003). Of course, there must also be many BHs with much lower Lbol/LEdd (perhaps approaching zero), otherwise it would be hard to reconcile the local BH mass function with the observed number of AGNs. However, these low Eddington-ratio objects do not contribute significantly to the observed AGN population in the luminosity and redshift range probed by

AGES-I.

For high mass BHs at z > 1, we can make these constraints more direct by computing the Lbol/LEdd distributions at fixed BH mass. The depth of the

AGES-I survey is crucial in allowing us to compute histograms for Lbol/LEdd > 0.1 that are unaffected by the survey magnitude limit (see Fig. 6.11). Accounting for completeness corrections, these histograms clearly decline toward both high and low

181 Lbol/LEdd from the peak at Lbol/LEdd 1/3. This peak could be further studied by ∼ the expansion of the AGES survey presently underway (AGES-II) and the 2SLAQ survey, which probes to approximately the same depth as AGES-I over a much wider area (Richards et al. 2005). Because 2SLAQ AGNs are selected by optical colors, the effects of dust extinction may be more severe for this survey than AGES-II.

Our results strongly suggest that BHs gain most of their mass while accreting at near-Eddington rates. Since the height of the Eddington ratio distribution falls as

Lbol/LEdd drops from 0.3 to 0.1, any rise at lower values would have to be extremely steep to contribute more mass growth than the observed peak. The main loopholes are the possibilities that much of the growth occurs in objects that are obscured at optical wavelengths or are accreting with very low efficiency. Such objects could possibly have a different Lbol/LEdd distribution. However, the reasonable agreement between the integrated emissivity of the optical quasar population and the local BH mass density (Soltan 1982; see Shankar et al. 2004 for a recent analysis) suggests that any such population cannot dominate by a large fraction. In fact, the Shankar et al. (2004) study concluded that Eddington ratios of 1/3 at z = 3 were required ∼ in order to obtain a good match between the emissivity from optical quasars and the local BH mass density. We also note that our AGN sample is not as adversely affected by obscuration effects as most existing surveys. While the X-ray selected

AGNs are affected by soft X-ray absorption, the 24µm selected AGNs are not.

Furthermore, the bluest optical band used in this survey is the R band instead of the

182 more commonly used optical B-band. This relative immunity to obscuration will be enhanced further in the AGES-II AGN sample, which will have an optical flux limit of I 21.5 mag that is both at a longer wavelength and corresponds to a deeper ≤ flux limit.

For a BH to become active, some event (merger, tidal interaction, dynamical instability, etc.) must drive material into the central few pc of its host galaxy, and this inner material must then form an accretion flow onto the BH itself, on the much smaller scale of hundreds to thousands of AU. The galactic-scale fueling events are likely to have a broad mass distribution, and there is no reason for them to know about the precise mass of the central BH. The sharp peak of the observed Lbol/LEdd distribution suggests that these events are often sufficient to provide super-Eddington fuel supplies and that the actual BH accretion rates are determined by the BH’s self-regulation of the inner accretion flow. The narrow width ( 0.3 dex) and central ≤ value (Lbol/LEdd 1/3) are important targets for theoretical models of accretion ∼

flows, though further investigation of the zero-point calibration of MBH indicators is desirable to firm up the latter constraint. Overall, the population of active BHs in the AGES-I survey is simpler than one might have imagined beforehand, and explaining this simplicity is a new challenge for theories of AGN evolution.

183 1

0.8

0.6

0.4

0.2

0

18 19 20 21

Fig. 6.1.— Completeness correction as a function of R-band magnitude. The solid line shows our overall completeness correction. The dashed line shows the AGES-I completeness relative to AGES-II. The departure between the two lines shows where we make an additional correction for loss of blue objects due to the AGES-II L-band

flux limit.

184 Max Flux

Blue = Blue1 = Blue2 Red

Red1 Red2

Half-Max

Fig. 6.2.— Illustration of the method for measuring the FWHMs. Blue1 and Red1 indicate the wavelengths of the first crossing of the half-max point descending from the line peak, and Blue2 and Red2 show the first half-max crossing ascending from the line limits. Red1 and Red2 are averaged to produce the wavelength Red. In this cartoon, Blue1 and Blue2 are identical and so are equal to Blue. The FWHM is determined by the difference between Red and Blue.

185 Fig. 6.3.— Examples of typical smoothed spectra. Panels from top to bottom show examples of an acceptable spectrum, a spectrum rejected for low S/N, a spectrum with strong Fe II emission, and a spectrum for which it was not possible to make a measurement of the line width.

186 Fig. 6.4.— Comparison of BH masses derived from the McLure & Jarvis (2002)

Mg II scaling relation and our adopted Hβ (filled stars) or C IV (open circles) scaling relations in redshift regimes of overlap, as a function of bolometric luminosity, Lbol.

The line shows the best fit to the combined dataset and forms the basis for the modified calibration of the Mg II relation.

187 Fig. 6.5.— Comparison between BH masses derived from our adopted Mg II relation and BH masses from the relations for our measured Hβ (filled stars) or C IV (open circles) FWHMs as a function of measured Mg II FWHM. Below 3700 km s−1, the masses are in good agreement, although there may be an indication of some deviation at higher FWHM.

188 Fig. 6.6.— AGN luminosity as a function of redshift. Points are coded according to the emission line used for the mass measurement, with filled stars, open squares, and asterisks corresponding to Hβ, Mg II, and C IV, respectively. Also shown are solid curves at R = (17, 19, 21) mag, the SDSS spectroscopic limit as a function of redshift (dashed line; z < 3: R 19.3 mag, z > 3: R 20.4 mag), and the ≈ ≈ AGES-I spectroscopic limit (dotted line; R=21.5 mag). The dot-dashed line shows the evolution of the knee in the luminosity function with redshift, as determined by the 2SLAQ survey (Richards et al. 2005). 189 Fig. 6.7.— Estimated BH masses as a function of AGN bolometric luminosity.

Objects are color-coded by redshift range as indicated in the legend. The solid line denotes the Eddington limit, LEdd: objects to the right of the line are radiating above Eddington for the measured mass. The dotted line denotes one-tenth of the

Eddington limit. Point types denote the emission line used for the mass measurement, with open stars, filled squares, and asterisks corresponding to Hβ, Mg II, and C IV, respectively.

190 Fig. 6.8.— Distributions of Eddington ratios in bins of luminosity and redshift.

The panels are divided between z < 1.2 (left) and z > 1.2 (right) and in increasing luminosity from bottom to top. The histograms show the fraction, f, of data points in each bin. Dashed curves are Gaussians with the same mean and dispersion as the data.

Solid curves are the best-fit Gaussians accounting for measurement uncertainties in luminosity and linewidth. These curves are similar because the measurement errors are small compared to the distribution widths. At all redshifts and luminosities, we

find that most AGNs are radiating close to Eddington, with a dispersion of only

0.3 dex. ∼ 191 30 20 10 0

Fig. 6.9.— Estimated BH masses versus AGN bolometric luminosity for objects eliminated from our sample. Open squares show the cases for which the S/N of our data was insufficient to make a reliable FWHM measurement. Crosses show objects removed from the sample for having manual mass measurements that were more than 1-σ discrepant with the automated measurements (with the automated measurements shown here). The inset histogram shows the luminosity distribution of objects for which no sensible FWHM determination could be made (independent of the quality of the spectrum), using the same x-axis scale.

192 Fig. 6.10.— Distribution of Eddington ratios in bins of BH mass and redshift.

Parameters for the columns and rows are given at the top and the right of the

figure. The solid histograms in each panel show the distributions for the “clean” dataset of 426 objects corrected for completeness. The dotted histograms show our non completeness-corrected values. The arrow shows where AGNs within the bin are

first hitting the survey optical flux limit. The shaded region marks where AGNs are completely lost to optical selection. The position of the solid dot along the bottom of each panel shows the equivalent of the arrow for the SDSS spectroscopic limit for that bin. 193 Fig. 6.11.— Peak in Eddington ratio distribution at fixed mass in bins for which distribution shape is not determined by optical selection. The mass and redshift range for each bin is designated within each panel. Line types and symbols are the same as in Figure 6.10.

194 Fig. 6.12.— Effect of object removal on the peak in Eddington ratio distribution at

fixed mass in bins for which distribution shape is not determined by optical selection.

The mass and redshift range for each bin is designated within each panel. Solid lines show our completeness-corrected measurements as in Fig. 6.11 and dashed histograms include objects removed for low S/N or > 1-σ discrepancy in FWHM (also corrected for completeness). Symbols are the same as in Figure 6.10

195 Table 6.1. Gaussian Parameters of Data and Fits to Data

a b c d e zbin Lbin N µdata σdata Sk Ak µmodel σmodel

Low Low 118 0.56 0.34 0.07 2.81 0.56 0.34 − − − High Low 12 0.70 0.38 0.24 1.82 0.70 0.36 − − Low Med 60 0.61 0.31 0.18 2.50 0.61 0.31 − − − High Med 73 0.61 0.31 0.09 3.10 0.60 0.29 − − − Low High 23 0.59 0.34 0.31 2.11 0.59 0.33 − − − High High 140 0.51 0.30 0.19 3.42 0.51 0.29 − − −

aLow: z < 1.2; High: z > 1.2 b Low: log Lbol < 45.5; Med: 45.5 < log Lbol < 46; High: log Lbol > 46

c σµ = σdata/ (N) q d σSk = (6/N) q e σAk = (24/N) q Note. — For each bin in redshift and luminosity, we list the number of objects, N, the mean Eddington ratio, µdata, the dispersion in Eddington ratios, σdata, the skew (Sk) and kurtosis (Ak) of the distribution, the maximum likelihood fits to the mean Eddington ratio, µmodel, and the dispersion in Eddington ratios, σmodel. Also listed are formulas for the errors in the mean, skew, and kurtosis.

196 Chapter 7

Conclusion

Our understanding of the interplay between supermassive BHs and their host galaxies has developed significantly in the last few years. As a result, simulations and semi-analytic modeling of galaxy-BH co-evolution have become more sophisticated and are making more detailed predictions about observable properties of the universe. Thus, measuring BH masses in galaxies can provide a crucial constraint on models of BH growth.

This dissertation has studied reverberation mapping, the primary method of determining BH masses in relatively nearby broad-line AGNs, and has provided an empirical calibration for the reverberation mass scale while also attempting to produce an independent calibration to directly tie the AGN BH masses to those measured in quiescent galaxies. We then take advantage of scaling relationships established from reverberation mapping to investigate the distribution of Eddington ratios from a sample of AGNs identified in a deep multi-wavelength survey.

197 The main results are summarized as follows:

Virial relationship in NGC 3783

1. A reanalysis of archival IUE and ground-based reverberation mapping data for

the Seyfert 1 galaxy, NGC 3783, finds that the time delays, τ, of different UV

and optical emission lines are correlated with the velocity widths, V , of the

lines as τ V −2, consistent with expectations from the virial theorem for the ∝ gas near the BH.

6 2. The mass of the BH is determined to be (8.7 1.1) 10 M¯, a significant § × reduction in the relative uncertainty of the mass compared to previously

published values.

Revised BH mass determinations in three other AGNs

1. We use archival ground-based reverberation mapping data to determine the

BH mass in three Seyfert 1 galaxies.

7 (a) NGC 3227 has a BH mass of (3.6 1.4) 10 M¯. § ×

7 (b) NGC 3516 has a BH mass of (1.68 0.33) 10 M¯. § ×

6 (c) NGC 4593 has a BH mass of (6.6 5.2) 10 M¯. § × 198 2. Published velocity dispersion measurements for these galaxies place them in

agreement with the BH mass stellar velocity dispersion relationship found in − quiescent galaxies.

Empirical calibration of reverberation mapping masses

1. Stellar velocity dispersion measurements in 16 AGNs having virial products

measured from reverberation mapping are used to calibrate the reverberation-

based BH mass scale on the assumption that all of the AGNs fall on the BH

mass velocity dispersion. −

2. For a virial product determined with the dispersion (rather than FWHM) of

the emission line, the scaling factor between the virial product and the BH

mass is 5.5 1.8. §

3. The FWHM of the [O III]λ 5007 A˚ emission line is correlated with the stellar

velocity dispersion of these 16 AGNs, although with large scatter.

Direct calibration of reverberation mapping masses

1. We attempted to measure the BH mass in the reverberation-mapped Seyfert

galaxy, NGC 4151, with the technique of stellar dynamics. Imaging and

spectroscopy from both ground-based telescopes and HST were obtained in

this effort. Unfortunately, the resulting data were unable to constrain the BH

mass.

199 Measurement of Eddington ratios

1. We estimate BH masses and bolometric luminosities for 400 AGNs in the ∼ AGES-I survey.

2. The distribution of Eddington ratios at fixed optical luminosity is strongly

peaked at L/LEdd 1/3, with a dispersion of 0.3 dex. '

3. We establish that the Eddington ratio distribution of AGNs at z = 1 3 is − also strongly peaked at fixed BH mass. For the combination of redshift and

BH mass probed by AGES-I, the shape of the Eddington ratio distribution at

fixed mass had not been previously measurable.

4. The shape of the Eddington ratio distribution suggests that BHs transition

quickly between being “on” in AGNs and being “off” in quiescent galaxies.

7.1. Future Work

There are several research projects which are natural outgrowths of the work presented here. Revisiting the work of Chapter 5 with an alternative approach

(e.g., measuring the stellar dynamics with the CO bandhead), or just with better data, could allow us to determine the precise relationship between the virial product and the BH mass. Successful execution of stellar dynamics programs for multiple reverberation-mapped AGNs could reveal differences in the BLR structure in

200 different objects, which would help inform the physical picture of AGNs that has been developing.

As described in Chapter 6, the expansion of the AGES dataset to even fainter

flux limits (AGES-II) will allow us to test the conclusions drawn from the AGES-I data and probe the Eddington ratio distributions to even higher redshifts and lower

BH masses.

A similar exercise with the publicly available data from the Two-Degree Field

QSO Redshift Survey (2QZ) for over 25,000 AGNs may provide significantly better statistics for some redshift and mass bins, and could serve as a compliment to the analysis of the large SDSS AGN sample that is currently in progress by other authors.

The ultimate goal of these studies is to provide the observational constraints needed to test the emerging theoretical framework for the growth and evolution of

BHs. The work presented here represents my first step on that path.

201 BIBLIOGRAPHY

Adams, F. C., Graff, D. S., & Richstone, D. O. 2001, ApJ, 551, L31

Akritas, M. G. & Bershady, M. A. 1996, ApJ, 470, 706

Alexander, T. 1997, in Astronomical Time Series, ed. D. Maoz, A. Sternberg, & E. M. Leibowitz (Dordrecht: Kluwer), 163

Aller, M. C., & Richstone, D. 2002, AJ, 124, 3035

Alloin, D., et al. 1995, A&A, 293, 293

Arribas, S., Mediavilla, E., Garcia-Lorenzo, B., & del Burgo, C. 1997, ApJ, 490, 227

Baldwin, J., Ferland, G., Korista, K., & Verner, D. 1995, ApJ, 455, L119

Baribaud, T., Salamanca, I., Alloin, D., & Wagner, S. 1994, A&AS, 103, 121

Bell, E. F., & de Jong, R. S. 2001, ApJ, 550, 212

Bell, E. F., McIntosh, D. H., Katz, N., & Weinberg, M. D. 2003, ApJS, 149, 289

Bender, R. 1990, A&A, 229, 441

Bian, W., & Zhao, Y. 2004, MNRAS, 347, 607

Binney, J., & Merrifield, M. 1998, Galactic Astronomy (Princeton, NJ : Princeton University Press)

Blandford, R. D., & McKee, C. F. 1982, ApJ, 255, 419

Bohlin, R. C., & Gilliland, R. L. 2004, AJ, 128, 3053

Boroson, T. 2003, ApJ, 585, 647

Brand, K., et al. 2005, ApJ, submitted

202 Cappellari, M. 2002, MNRAS, 333, 400

Cappellari, M., & Emsellem, E. 2004, PASP, 116, 138

Chiang, J. 2002, ApJ, 572, 79

Clavel, J., et al. 1991, ApJ, 366, 64

Collier, S. J., et al. 1998, ApJ, 500, 162

Collier, S., et al. 2001, ApJ, 561, 146

Corbett, E. A., et al. 2003, MNRAS, 343, 705

Cretton, N., & Emsellem, E. 2004, MNRAS, 347, L31

Cretton, N., de Zeeuw, P. T., van der Marel, R. P., & Rix, H. 1999, ApJS, 124, 383

Czerny, B., NikoÃlajuk, M., Piasecki, M., & Kuraszkiewicz, J. 2001, MNRAS, 325, 865 de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Jr., Buta, R., Paturel, G., & Fouque, P. 1991, Third Reference Catalogue of Bright Galaxies (New York: Springer)

Di Nella, H., Garcia, A. M., Garnier, R., & Paturel, G. 1995, A&AS, 113, 151

Dibai, E.´ A. 1980, Astron. Zh., 57, 677 (English translation: Sov. Astron., 24, 389)

Dietrich, M., & Hamann, F. 2004, ApJ, 611, 761

Dietrich, M., & Kollatschny, W. 1995, A&A, 303, 405

Dietrich, M., et al. 1993, ApJ, 408, 416

Dietrich, M., et al. 1994, A&A, 284, 33

Dietrich, M., Hamann, F., Shields, J. C., Constantin, A., Vestergaard, M., Chaffee, F., Foltz, C. B., & Junkkarinen, V. T. 2002, ApJ, 581, 912

Edelson, R. A., & Krolik, J. H. 1988, ApJ, 333, 646

Eisenhardt, P. R., et al. 2004, ApJS, 154, 48

Elvis, M. et al., 1994, ApJS, 95, 1

Elvis, M. 2000, ApJ, 545, 63

203 Eyermann, S. E., Rumstay, K. S., Boone, R. H., & Cortes, S. R. 2005, Publications I.A.P.P.P. (in press)

Fabricant, D. G., Hertz, E. N., Szentgyorgyi, A. H., Fata, R. G., Roll, J. B., & Zajac, J. M. 1998, Proc. SPIE, 3355, 285

Falco, E. E., et al. 1999, PASP, 111, 438

Famaey, B., Van Caelenberg, K., Dejonghe, H. 2002, MNRAS, 335, 201

Ferrarese, L. 2002, in Current High-Energy Emission around Black Holes, ed. C.-H. Lee (Singapore: World Scientific), in press (astro-ph/0203047)

Ferrarese, L., & Merritt, D. 2000, ApJ, 539, L9

Ferrarese, L., Pogge, R. W., Peterson, B. M., Merritt, D., Wandel, A., & Joseph, C. L. 2001, ApJ, 555, L79

Fizeau, A. H. 1848, lecture before the Soci´et´e Philomathique, 1848 December 23

Francis, P. J., Hewett, P. C., Foltz, C. B., Chaffee, F. H., Weymann, R. J., & Morris, S. L. 1991, ApJ, 373, 465

Fromerth, M. J., & Melia, F. 2000, ApJ, 533, 172

Fromerth, M. J., & Melia, F. 2001, in ASP Conf. Ser. 224, Probing The Physics Of Active Galactic Nuclei By Multiwavelength Monitoring, ed. B. M. Peterson, R. S. Polidan, & R. W. Pogge (San Francisco: ASP), 377

Gaskell, C. M. 1994, in ASP Conf. Ser. 69, Reverberation Mapping Of The Broad-Line Region In Active Galactic Nuclei, ed. P. M. Gondhalekar, K. Horne, & B. M. Peterson (San Francisco: ASP), 111

Gaskell, C. M., & Peterson, B. M. 1987, ApJS, 65, 1

Gebhardt, K., et al. 2000a, ApJ, 539, L13

Gebhardt, K., et al. 2000b, ApJ, 543, L5

Green, R. F., Nelson, C. H., & Boroson, T. 2003, in Carnegie Observatories Astrophysics Series, Vol. 1: Coevolution of Black Holes and Galaxies, ed. L. C. Ho (Pasadena: Carnegie Observatories, http://www.ociw.edu/ociw/symposia/series/symposium1/proceedings.html)

Grupe, D., & Mathur, S. 2004, ApJ, 606, L41

204 Haehnelt, M. G., & Kauffmann, G. 2000, MNRAS, 318, L35

H¨aring-Neumayer, N., Cappellari, M., Rix, H.-W., Hartung, M., Prieto, M. A., Meisenheimer, K., & Lenzen, R. 2005, ApJ, submitted (astro-ph/0507094)

Harris, A. W., & Sonneborn, G. 1987, in Exploring the Universe with the IUE Satellite, ed. Y. Kondo (Dordrecht: Reidel), 729

Heckman, T. M., Kauffmann, G., Brinchmann, J., Charlot, S., Tremonti, C., & White, S. D. M. 2004, ApJ, 613, 109

Ho, L. C. 1999, in Observational Evidence for Black Holes in the Universe, ed. S. K. Chakrabarti (Dordrecht: Kluwer), 157

Ho, L. C. 2004, in Carnegie Observatories Astrophysics Series, Vol. 1: Coevolution of Black Holes and Galaxies, ed. L. C. Ho (Cambridge: Cambridge Univ. Press), in press (astro-ph/0401527)

Hutchings, J. B., et al. 1999, AJ, 118, 2101

Ivanov, V. D., Rieke, G. H., Groppi, C. E., Alonso-Herrero, A., Rieke, M. J., & Engelbracht, C. W. 2000, ApJ, 545, 190

Jannuzi, B. T. & Dey, A. 1999, in ASP Conf. Ser. 191: Photometric Redshifts and the Detection of High Redshift Galaxies, ed. L. Weymann, L. Storrie-Lombardi, M. Sawicki, & R. Brunner, 111

Jefferys, W. H., Fitzpatrick, M. J., & McArthur, B. E. 1988, Celes. Mech., 41, 39

Kaiser, M. E., et al. 2000, ApJ, 528, 260

Kaspi, S., Maoz, D., Netzer, H., Peterson, B. M., Vestergaard, M., & Jannuzi, B. T. 2005, ApJ, in press (astro-ph/0504484)

Kaspi, S., Smith, P. S., Maoz, D., Netzer, H., & Jannuzi, B. T. 1996, ApJ, 471, L75

Kaspi, S., Smith, P. S., Netzer, H., Maoz, D., Jannuzi, B. T., & Giveon, U. 2000, ApJ, 533, 631

Kazanas, D., & Nayakshin, S. 2001, ApJ, 550, 655

Keel, W. C. 1996, AJ, 111, 696

Keenan, P. C., & McNeil, R. C. 1989, ApJS, 71, 245

205 Kenter, A., et al. 2005, ApJ, submitted (astro-ph/0507615)

Kochanek, C. S., Eisenstein, D., Caldwell, N., Cool, R., Green, P., & AGES 2004, American Astronomical Society Meeting Abstracts, 205

Kollatschny, W. 2003a, A&A, 407, 461

Kollatschny, W. 2003b, A&A, 412, L61

Kollatschny, W., & Dietrich, M. 1997, A&A, 323, 5

Koratkar, A. P., & Gaskell, C. M. 1991a, ApJS, 75, 719

Koratkar, A. P., & Gaskell, C. M. 1991b, ApJ, 370, L61

Korista, K. T., et al. 1995, ApJS, 97, 285

Korista, K. T., & Goad, M. R. 2001, in ASP Conf. Ser. 224, Probing The Physics Of Active Galactic Nuclei By Multiwavelength Monitoring, ed. B. M. Peterson, R. S. Polidan, & R. W. Pogge (San Francisco: ASP), 411

Krolik, J. H. 2001, ApJ, 551, 72

Litchfield, S. J., Robson, E. I., & Hughes, D. H. 1995, A&A, 300, 385

Lynden-Bell, D. 1969, Nature, 223, 690

Magorrian, J., et al. 1998, AJ, 115, 2285

Maoz, D., & Netzer, H. 1989, MNRAS, 236, 21

Marconi, A., Capetti, A., Axon, D. J., Koekemoer, A., Macchetto, D., & Schreier, E. J. 2001, ApJ, 549, 915

Marconi, A., Pastorini, G., Pacini, F., Axon, D. J., Capetti, A., Macchetto, D., Koekemoer, A. M., & Schreier, E. J. 2005, A&A, submitted (astro-ph/0507435)

Mathur, S., Kuraszkiewicz, J., & Czerny, B. 2001, NewA, 6, 321

McLure, R. J., & Dunlop, J. S. 2001, MNRAS, 327, 199

McLure, R. J., & Dunlop, J. S. 2002, MNRAS, 331, 795

McLure, R. J., & Dunlop, J. S. 2004, MNRAS, 352, 1390

McLure, R. J., & Jarvis, M. J. 2002, MNRAS, 337, 109

206 Merritt, D., & Poon, M. Y. 2004, ApJ, 606, 788

Miralda-Escud´e, J., & Kollmeier, J. A. 2003, ApJ, 619, 30

Morales, R., & Fabian, A. C. 2002, MNRAS, 329, 209

Mundell, C. G., Pedlar, A., Shone, D. L., & Robinson, A. 1999, MNRAS, 304, 481

Murray, S. S., et al., 2005, astro-ph/0504084

Nelson, C. H. 2000, ApJ, 544, L91

Nelson, C. H., & Whittle, M. 1995, ApJS, 99, 67

Nelson, C. H., & Whittle, M. 1996, ApJ, 465, 96

Netzer, H. 1977, MNRAS, 181, 89P

Netzer, H. 1990, in Active Galactic Nuclei, ed. R. D. Blandford, H. Netzer, & L. Woltjer (Berlin: Springer), 137

Newmark, J. S., Holm, A. V., Imhoff, C. L., Oliverson, N. A., Pitts, R. E., & Sonneborn, G. 1992, IUE Observing Guide, NASA IUE Newsletter, 47, 1

Nichols, J. S., Garhart, M. P., De La Pena,˜ M. D., & Levay, K. L. 1993, International Ultraviolet Explorer New Spectral Image Processing System Information Manual: Low-Dispersion Data, Version 1.0, CSC/SD-93/6062

Nichols, J. S., & Linsky, J. L. 1996, AJ, 111, 1

Ollongren, A. 1962, Bull. Astron. Inst. Netherlands, 16, 241

Pavlovsky, C., et al. 2004, ”ACS Instrument Handbook”, Version 5.0, (Baltimore: STScI)

Pedlar, A., Howley, P., Axon, D. J., & Unger, S. W. 1992, MNRAS, 259, 369

Peng, C. Y., Ho, L. C., Impey, C. D., & Rix, H.-W. 2002, AJ, 124, 266

Peterson, B. M. 1993, PASP, 105, 247

Peterson, B. M. 2001, in Advanced Lectures on the Starburst-AGN Connection, ed. I. Aretxaga, D. Kunth, & R. Mujica´ (Singapore: World Scientific), 3

Peterson, B. M., Meyers, K. A., Capriotti, E. R., Foltz, C. B., Wilkes, B. J., & Miller, H. R. 1984, ApJ, 292, 164

207 Peterson, B. M., Polidan, R. S., & Horne, K. 2004b, Astron. Nachr., 325, 257

Peterson, B. M., & Wandel, A. 1999, ApJ, 521, L95

Peterson, B. M., & Wandel, A. 2000, ApJ, 540, L13

Peterson, B. M., Wanders, I., Horne, K., Collier, S., Alexander, T., Kaspi, S., & Maoz, D. 1998, PASP, 110, 660

Peterson, B. M., et al. 1991, ApJ, 368, 119

Peterson, B. M., et al. 2002, ApJ, 581, 197

Peterson, B. M., et al. 2004a, ApJ, 613, 682

Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, Numerical Recipes (2nd ed.; Cambridge: Cambridge Univ. Press)

Reichert, G. A., et al. 1994, ApJ, 425, 582

Richards, G. et al. 2005, MNRAS, 360, 825

Richstone, D. 1982, ApJ, 252, 496

Richstone, D., et al. 2004, ApJL, submitted (astro-ph/0403257)

Robinson, A. 1994, in ASP Conf. Ser. 69, Reverberation Mapping of the Broad-Line Region in Active Galactic Nuclei, ed. P. M. Gondhalekar, K. Horne, & B. M. Peterson (San Francisco: ASP), 147

Rodr´iguez-Pascual, P. M., et al. 1997, ApJS, 110, 9

Salamanca, I., et al. 1994, A&A, 282, 742

Salpeter, E., 1964, ApJ, 140, 796

Santos-Lle´o, M., Clavel, J., Barr, P., Glass, I. S., Pelat, D., Peterson, B. M., & Reichert, G. 1995, MNRAS, 274, 1

Schinnerer, E., Eckart, A., & Tacconi, L. J. 2000, ApJ, 533, 826

Schwarzschild, M. 1979, ApJ, 232, 236

Shankar, F., Salucci, P., Granato, G.L., DeZotti, G., Danese, L., 2004, MNRAS, 354, 1020

208 Shields, G. A., Gebhardt, K., Salviander, S., Wills, B. J., Xie, B., Brotherton, M. S., Yuan, J., & Dietrich, M. 2003, ApJ, 583, 124

Silge, J. D., Gebhardt, K., Bergmann, M., & Richstone, D. 2005, AJ, 130, 406

Silk, J., & Rees, M. J. 1998, A&A, 331, L1

Simkin, S. M. 1975, ApJ, 200, 567

Soifer, B. T. & Spitzer/NOAO Team 2004, American Astronomical Society Meeting Abstracts, 204

Soltan, A. 1982, MNRAS, 200, 115

Steed, A., & Weinberg, D. H. 2003, ApJ, submitted, astro-ph/0311312

Stern, D., et al. 2005, ApJ, submitted (astro-ph/0410523)

Stirpe, G. M., et al. 1994, ApJ, 425, 609

Strauss, M. A., Huchra, J. P., Davis, M., Yahil, A., Fisher, K. B., & Tonry, J. 1992, ApJS, 83, 29

Theureau, G., Bottinelli, L., Coudreau-Durand, N., Gouguenheim, L., Hallet, N., Loulergue, M., Paturel, G., & Teerikorpi, P. 1998, A&AS, 130, 333

Tremaine, S., et al. 2002, ApJ, 574, 740

Umemura, M. 2001, ApJ, 560, L29

Valluri, M., Merritt, D., & Emsellem, E. 2004, ApJ, 602, 66

Valluri, M., Ferrarese, L., Merritt, D., & Joseph, C. L. 2005, ApJ, 628, 137 van Groningen, E., & Wanders, I. 1992, PASP, 104, 700

Vanden Berk, D., et al. 2001, AJ, 122 549

Vestergaard, M. 2002, ApJ, 571, 733

Vestergaard, M. 2004, ApJ, 601, 676

Virani, S. N., De Robertis, M. M., & VanDalfsen, M. L. 2000, AJ, 120, 1739

Wandel, A. 2002, ApJ, 565, 762

Wandel, A., Peterson, B. M., & Malkan, M. A. 1999, ApJ, 526, 579

209 Wanders, I., et al. 1993, A&A, 269, 39

Wanders, I., et al. 1997, ApJS, 113, 69

Wanders, I., & Horne, K. 1994, A&A, 289, 76

Wanders, I., Peterson, B. M., Pogge, R. W., DeRobertis, M. M., & van Groningen, E. 1992, A&A, 266, 72

Warner, C., Hamann, F., & Dietrich, M. 2004, ApJ, 606, 136

White, R. J., & Peterson, B. M. 1994, PASP, 106, 879

Willmer, C. N. A., Focardi, P., Chan, R., Pilligrini, P. S., & da Costa, N. L. 1991, AJ, 101, 57

Winge, C., Peterson, B. M., Horne, K., Pogge, R. W., Pastoriza, M. G., & Storchi-Bergmann, T. 1995, ApJ, 445, 680

Woo, J., & Urry, C. M. 2002, ApJ, 579, 530

Wu, X.-B., & Han, J. L. 2001, ApJ, 561, L59

Yu, Q. & Tremaine, S. 2002, MNRAS, 335, 965

Zel’dovich, Y. B., & Novikov, I. D. 1964, Dokl. Akad. Nauk SSSR, 158, 811 (English translation: 1965, Sov. Phys. Dokl., 9, 834)

210