ABSTRACT Title of dissertation: THE COOL SIDE OF GALACTIC WINDS: EXPLORATION WITH HERSCHEL-PACS AND SPITZER-IRS Myra Jean Stone, Doctor of Philosophy, 2020 Dissertation directed by: Professor Sylvain Veilleux Department of Astronomy Galactic-scale outflows driven by starbursts and/or active galactic nuclei (AGN) are key ingredients to theoretical models and numerical simulations of galaxy as- sembly and evolution. The feedback induced by the presence of these outflows (or winds) may affect the evolution and formation of a galaxy by regulating the amount of cold, dense gas responsible for star formation and black hole accretion. We present the results from a systematic search for galactic-scale, molecu- lar (OH 119 ?m) outflows in a sample of 52 Local Volume (d < 50 Mpc) Burst Alert Telescope detected active galactic nuclei (BAT AGN) with Herschel -PACS. We combine the results from our analysis of the BAT AGN with the published Herschel/PACS data of 43 nearby (z < 0.3) galaxy mergers, mostly ultraluminous infrared galaxies (ULIRGs) and QSOs. Our data show that both the starburst and AGN contribute to driving OH outflows, but the fastest OH winds require AGN with quasar-like luminosities. We also analyze Spitzer InfraRed Spectrograph (IRS) observations of the OH 35 ?m feature in 15 nearby (z . 0.06) (ultra-)luminous infrared galaxies (U/LIRGs). The measured OH 35 ?m equivalent widths are used to compute an average OH column density which is then compared to the hydrogen column density for a typical optical depth at 35 ?m of ?0.5 and gas-to-dust ratio of 125 to derive an OH?to?H abundance ratio of XOH = 1.01 ? 0.15 ? 10?6. The OH 35 ?m line profiles predicted from published radiative transfer models constrained by observa- tions of OH 65, 79, 84, and 119 ?m in five objects are found to be consistent with the IRS OH 35 ?m spectra. Finally, we analyze Herschel -PACS observations of five atomic fine-structure transition lines ([O I] 63 ?m, [O III] 88 ?m, [N II] 122 ?m, [O I] 145 ?m, and [C II] 158 ?m) in seven nearby (d < 16 Mpc) galaxies with well-known galactic- scale outflows (Cen A, Circinus, M 82, NGC 253, NGC 1068, NGC 3079, and NGC 4945). With this suite of atomic emission lines, we investigate the cool neutral atomic (T ? 103 K) and warm ionized (T ? 104 K) gas phases within each outflow. The outflows in the Herschel data are spatially isolated from the galactic disk based on the kinematic signatures of the outflows. The spatial distribution and physical properties of the outflows detected in the Herschel data are compared with published results at other wavelengths. For completeness, an analysis of the molecular gas traced by OH 119 ?m is also presented. THE COOL SIDE OF GALACTIC WINDS: EXPLORATION WITH HERSCHEL-PACS AND SPITZER-IRS by Myra Jean Stone Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2020 Advisory Committee: Professor Sylvain Veilleux, Chair/Advisor Professor Alberto Bolatto Professor Bill Dorland Dr. Marcio Mele?ndez Professor Stuart Vogel ?c Copyright by Myra Jean Stone 2020 Preface Much of the work in this thesis has been published and is presented here with minimal modification and the rest is in preparation for submission. Chapter 2 has been published in The Astrophysical Journal under the title ?The Search for Molecular Outflows in Local Volume AGNs with Herschel-PACS? (Stone, M., Veilleux, S., Mele?ndez, M., Sturm, E., Gracia?-Carpio, J., & Gonza?lez- Alfonso, E., 2016, ApJ, 826,111). Chapter 3 has been published in The Astrophysical Journal under the title ?Constraints on the OH-to-H Abundance Ratio in Infrared-bright Galaxies Derived from the Strength of the OH 35 ?m Absorption Feature? (Stone, M., Veilleux, S., Gonza?lez-Alfonso, E., Spoon, H., & Sturm, E., 2018, ApJ, 853, 132). Chapter 4 is currently in preparation for submission to The Astrophysical Journal (Stone, M. & Veilleux, S.). ii Acknowledgments I would not have made it this far and completed this thesis without the im- mense support from the people around me. I am very grateful to my advisor, Sylvain Veilleux, for his advice and guidance in my research and his enduring ability to keep me from getting too far down into many a rabbit hole. I am indebted to Marcio Mele?ndez for always being available to chat about life, the universe, and everything (mostly universe related though). His invaluable assistance with my research and coding throughout my entire graduate career was one of the main propellers that kept me moving forward. His obvious excitement about astronomy and his cheerful eagerness to share his knowledge about it has always been a reminder to me as to why I too loved the field. I thank Blake and Zeeve for our many take-out dinners and TV watching after long days of work. These outings always kept me sane ... and looking forward to food. Their support both personal and professional has meant a great deal to me. I would like to thank Alex who was a constant source of advice, laughter, and unwavering support which kept me afloat early on in these uncharted waters. I?d like to thank Qian and Tiara for our food adventures while seeking out new and interesting restaurants. I thank Drew for our many chats which helped me navigate the ins and outs of the department and for the many chats which just helped brighten my mood. I would like to thank my old roommates Tanya and Paulo who endured with me a very interesting 2.5 year stint in an apartment whose description would cause iii your brow to furrow and your eyes to squint. I?m grateful for our little traditions such as RenFest and BBQ, Blues, and Brews which were always great ways to relax and reset. To my more recent roommates, Aaron, Lauren, Silvy and Chris, I am grateful for the welcoming home we made which was such a comfort during hard times. I?m grateful for our dinners and cocktail hours which kept me fueled for another day of research. Last but not least, I?d like to thank my boyfriend Chris who has been amazing during this last leg of my graduate school journey. You?ve pushed me to strive for the best and set me back on course if ever I faltered. Your belief in me has helped me maintain belief in myself, something for which I will always be grateful for. iv Table of Contents Preface ii Acknowledgements iii Table of Contents v List of Tables viii List of Figures ix List of Abbreviations xvi 1 Introduction 1 1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Starburst and Active Galaxies . . . . . . . . . . . . . . . . . . . . . . 5 1.2.1 Starburst-Driven Winds . . . . . . . . . . . . . . . . . . . . . 5 1.2.2 AGN-Driven Winds . . . . . . . . . . . . . . . . . . . . . . . . 8 1.3 The Multiphase ISM in Outflows . . . . . . . . . . . . . . . . . . . . 11 1.3.1 General Considerations . . . . . . . . . . . . . . . . . . . . . . 11 1.3.2 The OH Molecule . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.3.3 Atomic Fine-Structure Lines . . . . . . . . . . . . . . . . . . . 15 1.4 Observatories and Instruments . . . . . . . . . . . . . . . . . . . . . . 17 1.4.1 Herschel -PACS . . . . . . . . . . . . . . . . . . . . . . . . . . 18 1.4.2 Spitzer -IRS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 1.5 Outline of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2 Search for Molecular Outflows in Local Volume AGN with Herschel -PACS 22 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.2 BAT AGN Sample Selection . . . . . . . . . . . . . . . . . . . . . . . 29 2.3 ULIRG and PG QSO Sample . . . . . . . . . . . . . . . . . . . . . . 36 2.4 Properties of the Sample Galaxies . . . . . . . . . . . . . . . . . . . . 36 2.5 Observations, Data Reduction, and Spectral Analysis . . . . . . . . . 44 2.5.1 OH 119 ?m Feature . . . . . . . . . . . . . . . . . . . . . . . . 44 2.5.1.1 OH Observations . . . . . . . . . . . . . . . . . . . . 44 2.5.1.2 OH Data Reduction . . . . . . . . . . . . . . . . . . 45 2.5.1.3 Spectral Analysis of the OH Doublet . . . . . . . . . 46 2.5.2 The 9.7 ?m Silicate Feature . . . . . . . . . . . . . . . . . . . 49 v 2.5.2.1 Data Reduction of the 9.7 ?m Silicate Feature . . . . 49 2.5.2.2 Spectral Analysis of the 9.7 ?m Silicate Feature . . . 50 2.6 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2.6.1 The OH 119 ?m Feature . . . . . . . . . . . . . . . . . . . . . 52 2.6.2 The S9.7?m Feature . . . . . . . . . . . . . . . . . . . . . . . . 58 2.6.3 OH Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . 59 2.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 2.7.1 Outflows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 2.7.1.1 BAT AGN with Molecular Outflows . . . . . . . . . 70 2.7.1.2 Driving Mechanisms of Molecular Outflows . . . . . 74 2.7.2 Inflows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 2.7.3 The 9.7 ?m Silicate Feature . . . . . . . . . . . . . . . . . . . 78 2.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 3 Constraints on the OH-to-H Abundance Ratio in Infrared-Bright Galaxies Derived from the Strength of the OH 35 ?m Absorption Feature 85 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 3.2 Background: The OH Molecule . . . . . . . . . . . . . . . . . . . . . 90 3.3 Sample Selection, Data Reduction, and Spectral Analysis . . . . . . . 91 3.3.1 The Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 3.3.2 Data Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . 92 3.3.3 Spectral Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 96 3.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 3.4.1 OH Column Density . . . . . . . . . . . . . . . . . . . . . . . 102 3.4.2 XOH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 3.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 3.5.1 XOH: Comparison with the Literature . . . . . . . . . . . . . 106 3.5.2 XOH: A Check on Radiative Transfer Models . . . . . . . . . . 107 3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 4 Far-Infrared Integral-Field Spectroscopy of Nearby Galactic Winds with Herschel-PACS 110 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4.2 Atomic Fine-Structure Emission Lines . . . . . . . . . . . . . . . . . 113 4.3 The Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.4 Observations, Archival Data, and Spectral Analysis . . . . . . . . . . 126 4.5 Velocity-Integrated Emission Line Flux and Ratio Maps . . . . . . . . 128 4.5.1 Constraints from the Emission Line Ratios . . . . . . . . . . . 129 4.5.2 Results from the Emission Line Ratios . . . . . . . . . . . . . 130 4.6 Gas Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 4.7 PDR Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 4.7.1 [C II] 158 Emission . . . . . . . . . . . . . . . . . . . . . . . . 142 4.7.2 [O I] 63 Emission . . . . . . . . . . . . . . . . . . . . . . . . . 142 4.7.3 Total Infrared Emission . . . . . . . . . . . . . . . . . . . . . 143 4.7.4 Hydrogen Density and UV Radiation Field . . . . . . . . . . . 144 vi 4.7.5 Caveats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 4.8 Methods to Derive the Outflow Properties . . . . . . . . . . . . . . . 146 4.8.1 Beam Smearing . . . . . . . . . . . . . . . . . . . . . . . . . . 146 4.8.2 Tilted Ring Model . . . . . . . . . . . . . . . . . . . . . . . . 147 4.8.3 Defining the Spatial Location of the Outflow . . . . . . . . . . 148 4.8.4 Line Luminosities . . . . . . . . . . . . . . . . . . . . . . . . . 150 4.8.5 Mass of the Neutral Atomic Wind . . . . . . . . . . . . . . . . 151 4.8.6 Mass of the Ionized Wind . . . . . . . . . . . . . . . . . . . . 152 4.8.7 Kinetic Energy of the Wind . . . . . . . . . . . . . . . . . . . 153 4.9 Properties of the Outflows and Multi-Phase Comparisons . . . . . . . 153 4.9.1 M 82 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 4.9.2 Cen A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 4.9.3 Circinus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 4.9.4 NGC 253 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 4.9.5 NGC 1068 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 4.9.6 NGC 3079 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176 4.9.7 NGC 4945 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 4.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 5 Summary and Future Work 187 5.1 Main Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187 5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 A Spitzer MIR SPECTRA 192 B Results on OH 119 ?m 199 Bibliography 222 vii List of Tables 2.1 Herschel Observations of BAT AGN . . . . . . . . . . . . . . . . . . 23 2.1 Herschel Observations of BAT AGN . . . . . . . . . . . . . . . . . . 24 2.2 Galaxy Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 2.2 Galaxy Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.2 Galaxy Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.3 Properties of the OH 119 ?m Profiles . . . . . . . . . . . . . . . . . 39 2.3 Properties of the OH 119 ?m Profiles . . . . . . . . . . . . . . . . . 40 2.3 Properties of the OH 119 ?m Profiles . . . . . . . . . . . . . . . . . 41 2.4 S9.7?m Continuum Parameters and Measured Values of the BAT AGN Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 2.4 S9.7?m Continuum Parameters and Measured Values of the BAT AGN Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 2.4 S9.7?m Continuum Parameters and Measured Values of the BAT AGN Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 2.5 S9.7?m Continuum Parameters and Measured Values of the ULIRG/PG QSO Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 2.5 S9.7?m Continuum Parameters and Measured Values of the ULIRG/PG QSO Sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 2.6 Results from Statistical Analyses of Host Galaxy Properties . . . . . 64 2.7 Results from Statistical Analyses of the Kinematics . . . . . . . . . . 65 2.7 Results from Statistical Analyses of the Kinematics . . . . . . . . . . 66 3.1 Spitzer -IRS Spectra: Observations . . . . . . . . . . . . . . . . . . . 88 3.2 Galaxy Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 3.3 OH 35 ?m Profile Properties . . . . . . . . . . . . . . . . . . . . . . 101 4.1 Fine-structure Lines . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 4.2 Galaxy Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 4.3 PACS Fine-Structure Line Observations . . . . . . . . . . . . . . . . 121 4.3 PACS Fine-Structure Line Observations . . . . . . . . . . . . . . . . 122 4.4 Wind Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 B.1 PACS OH 119 ?m Observations . . . . . . . . . . . . . . . . . . . . 200 viii List of Figures 1.1 SMBH vs galaxy properties . . . . . . . . . . . . . . . . . . . . . . . 5 1.2 Schematic of two evolutionary phases of a wind-blown bubble . . . . 8 1.3 Schematic of a shock driven into the ISM by an AGN . . . . . . . . . 10 1.4 Multiphase AGN Outflow . . . . . . . . . . . . . . . . . . . . . . . . 12 1.5 Multiphase SB Outflow in M 82 . . . . . . . . . . . . . . . . . . . . . 14 1.6 Grotrian diagram of OH ground state transitions . . . . . . . . . . . 16 1.7 Grotrian diagram of the electronic transitions in atomic fine-structure lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 1.8 Herschel, Spitzer, model SEDs . . . . . . . . . . . . . . . . . . . . . . 18 2.1 BAT/AGN and ULIRG/PG QSO property distributions . . . . . . . 24 2.2 Fits to the continuum subtracted OH 119 line profiles . . . . . . . . . 27 2.2 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 2.2 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.2 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.2 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.3 OH 119 EW vs M? vs ?AGN vs LAGN . . . . . . . . . . . . . . . . . . 45 2.4 The apparent strength of the 9.7 ?m silicate feature relative to the local mid-infrared continuum as a function of the AGN fractions. Note that S9.7?m is a logarithmic quantity and can be interpreted as the apparent silicate optical depth. The strength of silicate absorption increases upward. Sign conventions and meanings of the symbols are the same as those in Section 2.5.1.1. Crosses represent objects with a null OH detection. Vertical lines indicate objects with a null 9.7 ?m silicate feature detection. . . . . . . . . . . . . . . . . . . . . . . . . . 49 2.5 Title . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 ix 2.6 Histograms showing the distributions of the 50% (median; left panels) and 84% (right panels) velocities derived from the multi-Gaussian fits to the OH profiles of the BAT AGN. Top panels show pure absorption components (blue), pure emission components (red dashed), P-cygni emission components (red, left diagonals), and inverse P-cygni emis- sion components (red, right diagonals). Bottom panels show pure absorption components (filled grey), P-Cygni absorption components (blue, left diagonals), inverse P-cygni absorption components (blue, right diagonals), and total absorption components (pure + P-Cygni + inverse P-Cygni). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 2.7 v50 and v84 as a function of the stellar masses. The meanings of the symbols are the same as those in Section 2.5.1.1. The data points joined by a segment correspond to F14394+5332 W and E. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) . . . . . . . . . . . . . . . . . . . . . . . . 67 2.8 v50 and v84 as a function of the star formation rates. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 2.9 v50 and v84 as a function of the specific star formation rates. The meanings of the symbols are the same as those in Section 2.5.1.1. The data points joined by a segment correspond to F14394+5332 E and W. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) The black vertical line indicates the approximate location of the Main Sequence of star-forming galaxies (Shimizu et al., 2015). . . . . . . . . . . . . . . . . . . . . . . . . . . 69 2.10 v50 and v84 as a function of the AGN fractions. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) 70 2.11 v50 and v84 as a function of the AGN luminosities. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 x 2.12 Total (absorption + emission) equivalent widths of OH 119 ?m as a function of the apparent strength of the 9.7 ?m silicate feature relative to the local mid-infrared continuum. Note that S9.7?m is a logarithmic quantity and can be interpreted as the apparent silicate optical depth. The strength of this absorption feature increases to the right. Also note that objects classified as LINERs have been excluded from this plot. Filled markers refer to Type 1 and open markers refer to Type 2. Blue squares and red circles represent BAT AGN and ULIRGs/PG QSOs, respectively. Vertical lines represent objects in which OH was not detected. Horizontal lines represent objects with a null S9.7?m detection. Dotted lines and dash-dotted lines refer to Type 1 and Type 2, respectively. Blue and red lines indicate BAT AGN and ULIRGs/PG QSOs, respectively. . . . . . . . . . . . . . . . 80 2.13 Ratio of the semi-minor axis to the semi-major axis (a proxy for the inclination of the host galaxy disk) as a function of S9.7?m for the BAT AGN sample. Squares, triangles, and circles represent BAT AGN in which OH is observed purely in absorption, purely in emission, com- posite absorption/emission, respectively. Diamonds represent objects in which OH was undetected. Filled points and dash-dotted lines in- dicate Type 1 while open points and dotted lines indicate Type 2 AGN. Horizontal lines represent objects with a null 9.7 ?m silicate feature detection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 3.1 Example of twelve different continuum fits to the mid-infrared spec- trum of IRAS F17208-0014. The black solid histograms are the data separated from each other by an arbitrary offset. Spline pivot points are anchored at wavelengths of 33.0 ?m, 33.8 ?m, 34.5 ?m, and 35.0 ?m. The blue, green, and red dots indicate the set of pivot points which are respectively the lower limits (v1), upper limits (v2), and best-fit-by-eye values (v3) used to fit a continuum spline. The top most set of fits is the result of using a third-order (k = 3) spline fit to each pivot point set. Below are the results for second-order (k = 2) and first-order (k = 1) spline fits. In each case, the blue, green, and red dotted lines show the results of the spline fits to the lower-limit, upper-limit, and ?best-fit-by-eye? continuum flux pivot points. The last spectrum at the bottom is fit with a polynomial. The magenta regions in this spectrum indicate the regions used to fit the first (solid blue line), second (solid green line), and third (solid red line) order polynomials to the continuum. The two dotted vertical lines in grey mark the rest wavelengths of the OH 35 doublet at systemic velocity. The dotted vertical lines in red mark the locations of the emission fea- tures [S III] 33.48 ?m and [Si II] 34.82 ?m. The vertical light-green band shows the region where a 20% dip in the detection response in the NOD 1 position occurs. . . . . . . . . . . . . . . . . . . . . . . . 94 xi 3.2 Fits to the 33? 35?m continua for all 15 objects in the sample. The grey shaded area shows the full range of the twelve different contin- uum fits described in Figure 3.1. Black dots are median values of the six fluxes at each pivot point, and the dotted line is the resultant third-order spline fit to those pivot points. The grey dotted vertical lines mark the rest wavelengths of the OH 35 doublet at systemic velocity. The red dotted vertical lines mark the locations of the emis- sion features [S III] at 33.48 ?m and [Si II] at 34.815 ?m. The vertical light-green band shows the region where a 20% dip in the detection response in the NOD 1 position occurs. . . . . . . . . . . . . . . . . . 95 3.3 Two-Gaussian fits to the continuum-subtracted OH 35 line profiles of the 15 objects in our sample; see Section 3.3.3. In each figure, the solid black histogram is the data. Blue dashed lines indicate the two Gaussian components which best fit the line profile, and the magenta line is the sum of those two components. The grey dotted vertical lines mark the rest wavelengths of the OH 35 doublet at systemic velocity. The red dotted vertical line marks the location of the [Si II] emission line at 34.815 ?m. . . . . . . . . . . . . . . . . . . . . . . . 98 3.4 Examples of OH 35 line profile fits for each of the twelve different continuum subtracted spectra in IRAS F17208-0014. Line colors and styles are the same as that for Figure 3.3. ?p1, p2, p3? indicate first-, second-, and third-order polynomial fits to the continuum, respec- tively. ?v1? refers to the ?lower-limit? pivot points in Figure 3.1. ?v2? refers to the ?upper-limit? pivot points, and ?v3? to the ?best- fit-by-eye? pivot points. ?k1, k2, k3? are the orders of the spline fitted to the continuum. For example, ?v2k3? is the third-order spline fit to the continuum using the ?upper-limit? pivot points. . . . . . . . . 99 3.5 Distributions of the measured OH 35 (a) integrated fluxes and (b) equivalent widths. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 3.6 Distributions of the (a) OH column densities and (b) OH-to-H abun- dance ratios, XOH, inferred from the fits of the OH 35 feature. . . . . 102 3.7 OH 35 spectra normalized to the continuum. Red and blue lines are the two Gaussian components of the fitted line and the magenta line is the sum of the two components. The green line is the profile pre- dicted by the radiative transfer model described in Gonza?lez-Alfonso & Cernicharo (1999); Gonza?lez-Alfonso et al. (2017). The modeled line profiles are constrained by observations of four OH lines at 65, 79, 84, and 119 ?m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.1 Signal-to-noise ratio maps. Contours are Contours are 0.1, 0.3, 0.5, 0.7, 0.8, and 0.9 of the peak value in each image. Black crosses mark the adopted galaxy center. . . . . . . . . . . . . . . . . . . . . . . . . 124 4.2 Total integrated emission line fluxes in 10?17 W m?2. Contours are 0.1, 0.3, 0.5, 0.7, and 0.9 of the peak flux in the image. Black crosses mark the adopted galaxy center. . . . . . . . . . . . . . . . . . . . . . 125 xii 4.3 Maps of the emission line ratios. Contours are 0.1, 0.3, 0.5, 0.7, 0.8, and 0.9 of the peak value in each image. . . . . . . . . . . . . . . . . 131 4.4 Maps of the median velocities, v in units of km s?150 . Contours are in eight equal steps between the minimum velocity and the maximum velocity in each image. . . . . . . . . . . . . . . . . . . . . . . . . . . 132 4.5 Maps of the 1?? line widths, W1? , in units of km s?1 . Contours are 0.4, 0.5, 0.6, 0.7, 0.8, and 0.9 of the peak width in each image. . . . . 133 4.6 Maps of the hydrogen density, nH in each object. Units are in cm ?3. The ?low? density range is defined as 0 < nH< 10 4 cm?3 and the ?high? density range is defined as 104 < nH< 10 7 cm?3. See Sec- tion 4.7.4 for details about the determination of these density ranges and definitions of Solutions A?D. Solution A is the low density solution with only the [C II] 158 flux correction (see Section 4.7.1 and Section 4.7.4), Solution B is the low density solution with both the fluxes of [C II] 158 and [O I] 63 corrected (see Section 4.7.1, Section 4.7.2 and Section 4.7.4), Solution C is the high density solution with only the [C II] 158 flux correction, and Solution D is the high density solution with both the fluxes of [C II] 158 and [O I] 63 corrected. As discussed in Section 4.7.4, the high density PDR solutions ?C? and ?D? lead to unphysical ISRFs. Therefore, we have excluded them from the analysis of the outflow. . . . . . . . . . 139 4.7 Maps of the strength of the UV radiation field, G0, for the low density and high density limits PDR solutions. Definitions of the Solutions are the same as those in Figure 4.6. . . . . . . . . . . . . . . . . . . . 140 4.8 M 82: Results from modeling the disk velocity field with 3DBarolo (which accounts for both the instrumental spectral and spatial resolutions) and the location of the outflow based on excess line broadening. Color bar values are in units of km s?1 . For each line, left to right: observed data, model result, data - model residual, spatial location of the wind in regions where ?W > 25 km s?11? , and spatial location of the wind in regions where ?W1? > 50 km s ?1 . Contours in (a) are in five equal steps between the minimum and maximum velocities in each image. Contours in (b) are 0.3, 0.5, 0.7, and 0.9 of the peak value in each image. The solid blue line marks the galaxy major axis. The magenta cross marks the adopted galaxy center. . . . . . . . . . . . . . . . . . 156 4.9 M 82: PACS contours (magenta) overlaid on the total integrated intensity contours of SiO(2-1) (black, Garc??a-Burillo et al., 2001) and the radio continuum at 4.8 GHz (grey scale, Wills et al., 1999). . 157 4.10 Mass and KE in the wind derived from the PDR solutions (see Sec- tion 4.7.4 for definitions of Solutions A, B, C, and D) and the [O III] 88 flux. Units are in 104M and 1051 erg for the masses and KEs, respectively.158 4.11 Cen A: Results from modeling the disk velocity field with 3DBarolo. Symbols and plots are the same as those in Figure 4.8. . . . . . . . . 161 4.12 Cen A: PACS contours overlaid on 8.4 GHz image from Hardcastle et al. (2003). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 xiii 4.13 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 4.14 Circinus: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. . . . . 165 4.15 Circinus does not have a discernible outflow in either the neutral or ionized gas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 4.16 NGC 253: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. . . . . 168 4.17 NGC 253: PACS contours (yellow), contours of receding and ap- proaching CO outflow (magenta and blue, respectively) from Bolatto et al. (2013). Composite image from Heesen et al. (2011) which shows H? in red from Westmoquette et al. (2011), ?20 cm contin- uum (green), and soft X-ray in blue from Hardcastle et al. (2010). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 4.18 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 4.19 NGC 1068: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. . . . . 172 4.20 NGC 1068: PACS contours overlaid on 349 GHz continuum and CO(3-2) residual mean-velocity field from Garc??a-Burillo et al. (2014). 173 4.21 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 4.22 NGC 3079: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. . . . . 177 4.23 Top row: PACS contours of the [C II] 158 v50 and W1? residuals in the wind in NGC 3079 overlaid on 1.4GHz observations from Sebastian et al. (2019). Middle row: PACS contours of the [C II] 158 v50 and W1? residuals in the wind in NGC 3079 overlaid on H?+ [N II] image from Cecil et al. (2001). Bottom row: Chandra image (blue) + HST (red and green). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 4.24 NGC 4945: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. . . . . 180 4.25 Top row: PACS contours overlaid on the [NII] residual velocity field of NGC 4945 from Venturi et al. (2017). Bottom row: PACS contours overlaid on the [NII] W70 of NGC 4945 from Venturi et al. (2017). . . 181 4.26 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 xiv A.1 Mid-infrared (5-37?m) spectra used to measure S9.7?m . The dashed line is the continuum calculated from the cubic spline interpolation fit- ted to the pivot points shown as black dots. Red dots show fcont(9.7?m) (located on dashed continuum line) and fobs(9.7?m) (located on the solid black line or the observed flux density). The blue line shows the integration range used to calculate the flux and total equivalent width of the 9.7 ?m silicate feature (see Table 2.4 and Table 2.5). . . 192 A.2 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193 A.3 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194 A.4 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 A.5 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196 A.6 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 A.7 (Continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198 B.1 Top left: PACS IFU footprint of OH 119 (black squares) overlaid on the 22 ?m WISE image of M 82. The white contour marks the spatial extent of the [O III] 88 wind. The black contour outlines the PACS footprint of the [O III] 88 observation. The black cross marks the adopted galaxy center and the blue line marks the galaxy major axis. Top right: Spline fits to the OH 119 continuum (blue dashed lines). Black lines are the observed data. Magenta areas indicate the regions used to fit the continuum. Blue dots mark the pivot points used to fit the spline. Bottom left: Line profile fitting results of the continuum-subtracted spectra. Solid blue lines indicate gaussian absorption components. Solid red lines indicate gaussian emission components. Vertical dashed blue (red) lines mark the v16 , v50 , and v84 velocities in absorption (emission). Bottom right: Top row, from left to right shows the total velocity-integrated flux of the fitted OH line profiles, the total flux in the absorption components only, and the total flux in the emission components only. Middle row shows v50 for the absorption (left) and emission components (right). Bottom row shows the 1?? line widths of the absorption (left) and emission (right) components. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 B.2 Cen A. Lines and symbols are the same as those in Figure B.1. . . . . 203 B.3 Circinus. Lines and symbols are the same as those in Figure B.1. . . . 204 B.4 NGC 253. Lines and symbols are the same as those in Figure B.1. . . 205 B.5 NGC 1068. Lines and symbols are the same as those in Figure B.1. . 206 B.6 NGC 3079. Top left: The white contour marks the spatial extent of the [C II] 158 wind. Lines and symbols are the same as those in Figure B.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 B.7 NGC 4945. Lines and symbols are the same as those in Figure B.1. . 208 xv List of Abbreviations AGN Active Galactic Nucleus ALMA Atacama Large Millimeter Array BAT Burst Alert Telescope CASSIS Cornell AtlaS of Spiter/IRS Sources EW Equivalent Width FIR Far-Infrared HST Hubble Space Telescope IR Infrared ICM Intracluster Medium IGM Intergalactic Medium IRS Infrared Spectrograph ISM Interstellar Medium IFS Integral Field Spectroscopy IFU Integral Field Unit LF Luminosity Function LIRG Luminous Infrared Galaxy MIR Mid-Infrared NED NASA/IPAC Extragalactic Database PACS Photodetector Array Camera and Spectrometer PAH Polycyclic Aromatic Hydrocarbon PG QSO Palomar Green Quasi-Stellar Object RSRF Relative Spectral Response Function SB Starburst SED Spectral Energy Distribution SHA Spitzer Heritage Archive SMBH Supermassive Black Hole SF Star Formation SNe Supernovae sSFR Specific Star Formation Rate SFR Star Formation Rate ULIRG Ultra-Luminous Infrared Galaxy UV Ultraviolet xvi Chapter 1: Introduction 1.1 Background Galactic-scale outflows (or winds) are powered by stellar and/or supermassive black-hole (SMBH) accretion processes which inject energy and momentum into the interstellar medium (ISM). On a grand scale, winds may significantly affect the assembly and evolution of galaxies (Veilleux et al., 2005, 2020). Outflows are ubiq- uitous at any redshift and they are invoked to help explain a number of observations in the Universe. For example, they help explain some of the observable properties of galaxies (e.g. color and gas abundance; for example, red, gas-poor ellipticals) and of the intergalactic medium (IGM). Outflows can account for the correlations that exist between the physical properties of a galaxy (e.g. the tight correlation between the SMBH mass and the mass of the spheroidal bulge). Moreover, outflows may help reconcile (or mitigate) the discrepancies between predictions from cosmological simulations and models with actual observations. In the following list, I expound upon some specific scenarios where winds have been incorporated to explain a num- ber of observations. This list is by no means exhaustive, but it does highlight how winds may provide a significant contribution to our understanding of the formation and evolution of galaxies. 1 ? Star formation regulation - There are a number of ways an outflow can affect the star formation activity in a galaxy. If the outflow has enough energy to escape the gravitational potential well of its host galaxy, then the outflow may be able to empty the reservoir of gas within the system and halt star formation completely (Fabian, 2012; Veilleux et al., 2005). Winds may heat the cold ISM via shocks or from the deposition of mechanical energy, thereby preventing the gas from condensing and forming stars (e.g. Cecil et al., 2001; Heckman & Thompson, 2017; Martin, 2006). On the other hand, if the material in the outflow lacks the energy to escape its host system, it will fall back onto the disk in the form of a ?galactic fountain? and may thus be recycled again for star formation. ? Scaling relations with the SMBH mass and the host galaxy properties - Winds may be responsible for the fueling or quenching of accretion onto the SMBH in a galaxy (Di Matteo et al., 2005). The brightness of an active galactic nucleus (AGN) is regulated by the rate of infall from surrounding material (e.g. gas and dust) while the infall rate of material is regulated by the brightness of the AGN. This feedback loop may explain the remarkable correlations observed between the mass of the SMBH (MBH) and the properties of the host galaxy, such as luminosity (MBH?L relation), velocity dispersions in (MBH?? relation) and masses (MBH?Mbulge relation) of the spheroidal bulge (see the review and references therein of Kormendy & Ho, 2013). These relationships indicate that SMBHs and the bulges coevolve by regulating each other?s growth. Figure 1.1 2 shows the three relationships. ? Correlations with galaxy type - Winds that can empty their host galaxies of gas may explain the bimodal color distribution between early (E, S0, and Sa) and late (Sb, Sc, and Irr) morphological types observed in large galaxy surveys (e.g. Strateva et al., 2001). These surveys indicate that disk-dominated galaxies are predominantly blue and star forming, whereas spheroid-dominated galaxies are largely red and quiescent (e.g. Blanton & Moustakas, 2009). This bimodality indicated that some sort of quenching mechanism was required to prevent gas from cooling and/or forming stars in observed gas-poor, quiescent galaxies (Peng et al., 2010). The inclusion of winds into semi-analytic models has qualitatively reproduced this observed bimodal color distribution (e.g. Kimm et al., 2009). ? Metal enrichment of the IGM and CGM - Outflows may not only impact in- dividual galaxies, but they may also play a significant role in the chemical and thermal evolution of the universe. They may even influence the metal en- richment of the intergalactic medium (IGM) and the circumgalactic medium (CGM) via the regulation of gas and dust exchange into and out of galaxies (Bertone et al., 2007; Heckman & Thompson, 2017; Zhang, 2018). Observa- tions of absorption systems in quasar sightlines through the IGM/CGM show that the IGM/CGM is enriched with metals to a small fraction of solar metal- licity and simulations have shown how outflows may be responsible for lifting gas and metal up and out of the galaxy disk into the IGM/CGM (Aguirre 3 et al., 2001; Madau et al., 2001). ? Paucity of very high and low mass galaxies - Early attempts to predict the statistical distributions of global properties of galaxy populations have shown a mismatch in shape between the predicted distribution of dark halo masses and the observed distribution of galaxy luminosities (or equivalently, masses). Galaxy abundances at higher luminosities (e.g. those brighter than galaxies with Milky Way-like luminosities, L? ? 3 ? 1010 L ) are observed to decline exponentially compared to CDM simulations, while the shape of galaxy abun- dances are observed to flatten at the fainter end. Inclusion of feedback from the SMBH appears to explain the abundance deficit at the high luminosity end (e.g. Croton et al., 2006; Somerville et al., 2008), while the inclusion of supernovae (SNe) and stellar processes is able to reproduce the shape of the luminosity function (LF) at the fainter end (e.g. Bower et al., 2006). Essen- tially, feedback due to stellar processes plays a crucial role in shaping the LF for dark matter halos with MH . 1012M while feedback from AGN processes plays a more dominant role at halo masses M & 1012H M (Somerville & Dave?, 2015). Although we now have a better understanding of how winds may significantly impact the formation and evolution of galaxies there still exists a plethora of unan- swered questions concerning the detailed physics and properties of galactic outflows. What is the primary mechanism of energy and momentum injection that powers an outflow? How can we constrain outflow properties such as energetics, morphologies, 4 b Figure 1.1 Left and Middle: MBH?L and MBH?? relations, respectively (from Gu?ltekin et al., 2009). Right: MBH?Mbulge relation (from Kormendy & Ho, 2013). lifetimes, spatial extent, radial-dependent velocities, mass outflow rates? What is the efficiency of entraining material from the ISM and ejecting it out of the disk? The answers to these questions can fundamentally measure the impact of outflows on galaxy evolution. In this chapter, I describe the basic physics behind outflows. In particular, the sources of energy and momentum (either starburst (SB) or AGN) and the driving mechanisms of the outflows. I also discuss the multiphase nature of outflows and the tools we used to probe different ISM phases in outflows. 1.2 Starburst and Active Galaxies 1.2.1 Starburst-Driven Winds SB galaxies contain regions (typically nuclear or circumnuclear) that are un- dergoing intense and spatially concentrated star formation activity; much higher than the rate of star formation in galaxies similar to the Milky Way. Such a SB can 5 be triggered when gas rapidly accumulates into a small volume, say from a merger of two gas-rich galaxies Barnes & Hernquist (1996), or from gas funneled into the nuclear region of a galaxy by dynamical processes associated with bar instabilities Combes & Charmandaris (2000). The SB regions host large numbers of young, hot, massive stars that ionize their environment and evolve quickly to form supernovae. Outflows can be driven by the powerful energy reservoirs in these SB regions. Chevalier & Clegg (1985, hereafter, C85) was first to establish a conceptual picture of how a SB-driven wind would work. The evolution of a bubble in the in- terstellar medium (ISM) evolves in several stages. First, kinetic energy is launched from a collection of stellar processes such as winds from high-mass, high-luminosity stars and explosions from supernovae (SNe). The ejecta from the winds and explo- sions intersect and collide with each other creating shocks that rapidly thermalize the kinetic energy of the ejecta, thus producing hot (T ? 107 K), high pressure (P/k ? 107 K cm?3), low density gas (Strickland & Stevens, 2000). The wind- shocked material creates an over-pressurized cavity (or bubble) that adiabatically expands as a free wind with vwind ? several thousand km s?1 (Garc??a-Burillo et al., 2001). During this ?energy-conserving? phase, little energy is lost to radiation due to the high post-shock temperatures and low densities. In the next phase, the free wind passes through an internal wind shock where the ejecta is decelerated and re- heated to T ? 107 K. The pressure of the shocked wind continues to inflate a bubble which is collimated by the density gradient between the galactic disk and its halo. The bubble preferentially propagates along the direction of the steepest pressure gradient. This forms a bipolar flow which sweeps up cooler, denser, ambient disk or 6 halo gas that is accelerated to vgas ? few hundreds km s?1 . A shell of accumulated ISM gas then forms at the edge of the bubble. Once the velocity of the shell has decelerated enough so that the cooling time is shorter than the expansion time, the shell will begin to cool radiatively. The bubble then enters the pressure-driven snow- plow phase in which the internal thermal energy of the bubble cannot be neglected. The bubble leaves the snowplow phase once the thermal energy has been radiated away. The total momentum of the inner wind is transferred to the gas and the bubble enters a ?momentum-conserving? phase. Assuming a stratified halo, once the bubble diameter is comparable with the scale height of the disk, the ambient gas density will be low enough that the bubble can accelerate and fragment into clumps and filaments through Rayleigh-Taylor instabilities (Schiano, 1985). During this ?break-out? phase, the freely flowing and shocked wind is injected into the galaxy halo. In summary, the structure of a bubble expanding into an ISM due to a SB includes (a) the energy injection zone, (b) hot, freely-expanding wind, (c) wind ma- terial that has been shocked, (d) a thin, dense shell of post-shock gas that has been accumulated by the wind and is now radiatively cooling, and finally (e) undisturbed ISM material. A schematic displaying the essential features of a wind-blown bubble is shown in Figure 1.2. 7 Figure 1.2 A schematic showing two evolutionary phases of a wind-blown bubble. Panel (a) shows the radiative phase where the energy injection zone from the SB is indicated by open stars, the free wind (FW, long arrows) expands outward at several thousand km s?1 where it passes through an internal wind shock (WS) creating shocked wind material (SW). The expanding bubble will sweep up ambient gas and accelerate it to a few hundred km s?1 (short arrows). The shocked gas will then radiatively cool into a thin shell (S). Eventually, when the diameter of the bubble is comparable to the scale height of the disk, it will fragment, panel (b), where the freely expanding wind fluid will flow into the halo. Bow shocks (BS) can form around shocked clouds (SC) which may be accelerated outward (from Heckman et al., 1990). 1.2.2 AGN-Driven Winds An active galaxy contains a core of bright emission that is embedded in the nuclear region of an otherwise normal galaxy. This core is typically brighter than the rest of the galaxy, shows variations in luminosity, and is known as an AGN. Spectral energy distributions (SEDs) of this higher-than-normal luminosity core indicate that the excess brightness is non-stellar in origin. It is generally accepted that the source of energy from an AGN is the accretion of matter (e.g. gas or dust) onto the SMBH, where the accreted material loses its gravitational potential energy via conversion into heat and is partly radiated away. This accretion disk is mostly seen in the 8 ultra-violet (UV) and in X-rays. In AGN, the nature of the energy outflow near the SMBH differentiates AGN feedback into two modes:?radiative? or ?quasar? mode which is powered by radiation and ?radio? or ?kinetic? mode which is powered by mechanical jets (Cielo et al., 2018). Radiative mode feedback is typically observed in high-luminosity AGN that are radiatively efficient and close to the Eddington limit (within about one or two orders of magnitude). It is most concerned with pushing cold gas about (Fabian et al., 2013). This mode is the most likely explanation for the observed black hole mass?stellar velocity dispersion relation (the so-called ?MBH??? relation). This mode was probably most effective at z ? 2 ? 3 when quasar activity was more common and galaxies were more rich in gas. The radio mode AGN are radiatively inefficient and typically host the most massive SMBH. These low-luminosity AGN (LAGN < 0.01LEdd; Combes, 2014) typically harbor light relativistic radio jets and are surrounded by hot halos. They typically correspond to massive early type galaxies (e.g. radio ellipticals). If feedback successfully empties a galaxy of its gas, eventually stellar mass loss and/or intracluster plasma will inevitably refill the gas reservoir. Maintaining an empty reservoir, or at least heating the gas enough to prevent star formation, appears to be a consequence of the radio mode feedback. In general, an AGN-driven wind will collide with the surrounding ISM, driving a reverse shock into the wind and a forward shock into the ISM. If the shocked wind cools efficiently, the outflow is ?momentum-conserving? (or ?momentum-driven?) and transfers only its ram pressure (and, hence, momentum flux) to the ISM (Zubo- vas & Nayakshin, 2014). If, on the other hand, the shocked wind cools inefficiently, 9 Figure 1.3 A schematic view of the shock driven into the ISM by a wind launched from an AGN is displayed in the left image where (a) is the wind region, delimited by the reverse wind shock (RSW ), (b) is the shocked wind that ends at the discontinuity surface (RC), and (c) is the shocked ISM, bounded by the forward shock (RS). The top right image shows a schematic of the ?energy-conserving? outflow where the hot and thick region (b) cools inefficiently and expands adiabatically. The bottom right image shows the ?momentum-conserving? outflow where the shocked wind (light blue region (b)) cools efficiently, experiences a drop in pressure, and becomes thin (from Costa et al., 2014). the outflow is ?energy-conserving? (or ?energy-driven?) and transfers most of its kinetic luminosity (Lkin ' 0.05LAGN; Zubovas & Nayakshin, 2014) to the ambient medium. This hot outflow may create a bubble in the galaxy disk that thermally expands after the shock. This phenomenon is analogous to the adiabatic phase in a SB-driven bubble described in Section 1.2.1. This kind of wind can drive outflows to velocities greater than 1000 km s?1 (King et al., 2011). A schematic view of the shock driven into the ISM by a wind launched from an AGN is displayed in the left image of Figure 1.3. The two images on the right of Figure 1.3 show the two 10 different driving modes. The bubble scenario described above is valid for an AGN driving a wide- angle outflow from the accretion disk or for a low energy AGN jet where all of the kinetic energy of the jet is deposited to the ISM. If the jet is powerful and highly collimated, the jet will simply ?drill? through the ISM without imparting much energy or momentum to the surrounding gas (e.g. Scheuer, 1974). 1.3 The Multiphase ISM in Outflows 1.3.1 General Considerations It is well established that outflows are of a multiphase nature (as revealed in observations and expected from simulations), spanning from the cold and dense molecular clouds to the very hot highly ionized medium. This multiphase structure complicates the calculations for the masses and energetics in outflows because we must concurrently analyze the outflow in multiple wavelength regimes, otherwise, measurements based on a single-phase will most likely result in misleading and incomplete conclusions. However, if we wish to assess the degree to which galactic winds impact the evolution and formation of galaxies, a fundamental understanding of the dynamic, energetic, and physical properties of the outflow must be established. Therefore, it is essential to consider all phases of the ISM in outflows in order to gain an understanding of the cosmological significance of outflows. 11 Figure 1.4 Artistic and observational views of AGN-driven winds in different ISM phases and at different physical scales. The cartoon in panels a, b, and c depicts the multiphase outflow based on observations and AGN feedback models. The outflow is launched from the galaxy nucleus (< 1 pc; panel a), propagates through the surrounding host galaxy ISM (1 pc ? 1 kpc; panel b), and extends out into the galaxy halo (> 10 kpc; panel c). Observations of different outflow phases at different physical scales in three well known AGN are shown in panels d, e, and f. Panel d shows the highly ionized accretion disk wind observed in X-ray from PDS456 (Nardini et al., 2015). Panel e shows the neutral atomic and molecular outflow phases traced by the mm and radio in Mrk 231 on a scale of ? few hundreds pc (Cicone et al., 2012; Morganti et al., 2016). Finally, panel f shows the ionized optical outflow of NGC 1365 extending out a few kpc (Venturi et al., 2017). (Figure is reproduced from (Cicone et al., 2018)). 12 Multiwavelength observations from the radio to the ??ray have shown that the multiphase structure of an outflow are generally divided into four gas phases: hot (highly ionized, Tgas? 106?107 K; n 6 8 ?3gas? 10 ?10 cm ), warm (ionized, T ? 103gas ? 104 K; n 2 4 ?3gas? 10 ?10 cm ), cool (neutral atomic, T 2 3gas? 10 ?10 K; ngas? 1?102 cm?3), and cold (molecular, Tgas? 10?102 K; n ? 103 cm?3gas ). These phases have been studied in several objects through various techniques which span the length of the electromagnetic spectrum. I will elucidate how these ISM phases have been traced in previous studies and I note that this list is by no means exhaustive or complete. Chandra and XMM-Newton have been used to observe the highly ionized X- ray emitting gas (e.g. Bravo-Guerrero & Stevens, 2017; Strickland & Heckman, 2007; Strickland & Stevens, 2000). Optical emission (e.g. H? , [N II], [O III]) and absorp- tion (e.g. Na I D and Mg II) lines have been used to trace the warm ionized (Cecil et al., 2002; Heckman et al., 2015; Martin et al., 2013; Rubin et al., 2014; Shop- bell & Bland-Hawthorn, 1998; Weiner et al., 2009) and neutral atomic (Heckman et al., 2000; Kornei et al., 2013; Martin, 2005; Rupke et al., 2002, 2005a,b,c; Rupke & Veilleux, 2015) gas phases in outflows. The warm molecular phase in outflows has been traced with H2 (e.g. Veilleux et al., 2009). Cold molecular gas has been traced using OH, CO, HCN, and HCO+ (e.g. Bolatto et al., 2013; Leroy et al., 2015; Veilleux et al., 2013; Walter et al., 2017; Westmoquette et al., 2013). Dust has also been traced in outflows (e.g. Hutton et al., 2014; McCormick et al., 2018; Mele?ndez et al., 2015). Figure 1.4 shows artistic representations and observational data of dif- ferent phases of AGN-driven outflows at different physical scales. Figure 1.5 shows 13 Figure 1.5 A schematic of the multiphase outflow in the SB galaxy M 82. The red statements are supported by the observational evidence described in blue (from Leroy et al., 2015). the multiphase SB-driven outflow in M 82. 1.3.2 The OH Molecule We use the FIR lines of the hydroxyl molecule (OH) as a tracer of the molecular gas component in an outflow. Furthermore, these lines can be used as powerful diagnostics of molecular winds because of their operative excitation mechanisms and their range in optical depths. These properties (e.g. OH is mainly excited by the absorption of FIR) also allow OH to probe nuclear regions and star forming complexes within galaxies where the FIR radiation density is strong. Within the 34 14 ?m to 163 ?m FIR wavelength range, a total of 14 lines (exhibiting both optically thick and optically thin features) arise from the eight lowest rotational levels of OH. In this thesis, we focus on two of the OH transitions. The first is the ground- state OH 2?3/2 J = 5/2? 3/2 rotational ?-doublet at 119.233 ?m and 119.441 ?m (hereafter, OH 119). It is the strongest of these FIR OH transitions (therefore it is more likely to be detected) and its wavelength is conveniently positioned at the peak sensitivity range of Herschel -PACS. Previous studies have already demonstrated the powerful diagnostic capability of OH 119 to determine wind characteristics (e.g. Fischer et al., 2010; Spoon et al., 2013; Sturm et al., 2011; Veilleux et al., 2013). Moreover, these studies have shown that the molecular gas can dominate the mass and energy budget of the galactic outflow. We also look at the cross ladder OH 2? ?23/2 ?5/2 ??doublet transition at 34.60 and 34.63?m (hereafter, OH 35). OH 35 is optically thin compared to the other FIR OH transitions and because only 4% of all 35?m absorption events result in re-emission at 35?m, OH 35 absorption line can be used to provide a meaningful constraint on the true column density. See Figure 1.6 for a schematic of the OH energy levels. Transitions analyzed in this thesis are outlined in red. 1.3.3 Atomic Fine-Structure Lines Fine-structure line emission from neutral and ionized carbon atoms, as well as other species such as oxygen and nitrogen, is an important key to understanding the 15 Figure 1.6 OH ?-doublet transitions. Transitions analyzed in this thesis are outlined in red. properties of the ISM in the region from which they originate, as it provides almost all of the gas cooling. Warm, ionized, and relatively tenuous gas can be traced by [O III] 88, [N II] 122, and (to a smaller extent) [C II] 158. Warm, neutral, dense gas is exclusively traced by [O I] 63 and [O I] 145, but [C II] 158 emission also predominately occurs in the atomic medium. The observed lines cover a range in density and temperature be- havior, which probe different phases of the ISM and are expected to probe different regions of their host galaxies. Figure 1.7 shows a schematic of the energy levels for the five atomic fine structure lines. 16 Figure 1.7 Energy level diagrams of the transitions in the electronic ground states of the atomic fine-structure lines. Top left: C+ (Draine, 2011). Top right: Neutral O (Draine, 2011). Bottom left: N+ ion (Goldsmith et al., 2015). Bottom right: O++ ion. The optical transitions are on the left and the fine-structure transitions are on an expanded scale on the right (Dinerstein et al., 1985). 1.4 Observatories and Instruments The online archival data presented in this thesis were acquired with the Her- schel Space Observatory and the Spitzer Space Telescope. Both telescopes have run through their cryogenic fuel and are no longer in operation. In this section, I detail these observatories and the instruments used to obtain the data analyzed in this thesis. Chapter 2 utilizes PACS and Spitzer. Chapter 3 utilizes only Spitzer data. Chapter 4 utilizes only PACS data. 17 Figure 1.8 Top left: Herschel Space Observatory that carried PACS. Top right: Spitzer Space Telescope carried IRS. Bottom: Image shows a series of model spec- tral energy distributions (SEDs) for a star-forming galaxy. Wavelength coverage is indicated for other IR instruments. PACS covers the principal FIR atomic fine structure cooling lines of the ISM: [O I] 63 , [O III] 88 , [N II] 122 , [O I] 145 , and [C II] 158 . Notice that OH 119 falls at the peak sensitivity of PACS. 1.4.1 Herschel -PACS The bulk of the data in this thesis are FIR observations from the Herschel Space Observatory (Pilbratt et al., 2010) which carries a 3.5 m diameter Cassegrain 18 telescope. It is the largest infrared space telescope to have ever been launched. Spectroscopic data was taken with the Photodetector Array Camera and Spectrom- eter (PACS; Poglitsch et al., 2010) on board the Herschel. PACS covers a field of view of 47??? 47??with a 5 ? 5 squared spaxel (spatial pixel) integral field unit spectrograph. It has a wavelength dependent spectral resolving power that ranges from R = 1000 ? 4000 (?v ? 75 ? 300 km s?1 ). PACS sampled the ? 50 ? 210 ?m FIR wavelength range with an unprecedented combination of sensitivity and spatial resolution (? 6 ? 11??; this is about 4? higher than previous infrared space missions). PACS covers the principal FIR fine-structure cooling lines (see bottom panel of Figure 1.8). Archival Herschel -PACS data were retrieved via the Herschel Science Archive (HSA). 1.4.2 Spitzer -IRS Launched in 2003 The Spitzer Space Telescope (Werner et al., 2004), operated by the Jet Propulsion Laboratory and the Spitzer Science Center (SSC), incorporates a 0.85 m diameter primary mirror and has a wavelength coverage from 3.6 to 160?m. The mid-infrared (5 ? 37?m), ?high-resolution? (R ? 600) data presented in this thesis were obtained with the Long High (LH) module of the Infrared Spectrograph (IRS; Houck et al., 2004). The data were retrieved from The Cornell AtlaS of Spitzer/IRS Sources (CASSIS) or from the Spitzer Heritage Archive (SHA). 19 1.5 Outline of Thesis This thesis focuses on the study of cool galactic winds in SBs and AGNs. Chapters 2 and 3 focus on the molecular phase in outflows. Chapter 4 is a multiphase study of outflows that looks at the neutral atomic, ionized, and molecular phases of the ISM. In Chapter 2, we study molecular outflows and inflows in a combined sample of local Burst Alert Telescope-detected AGN (BAT AGN)+ULIRGs+QSOs. We measure the wind detection rate in the BAT AGN and compare it with the rate of detection in the higher luminosity ULIRGs+QSOs. We also explore what rela- tionships exist between the properties of the outflows and the properties of their host galaxies. We then perform statistical tests to measure the strength of the correlations. In Chapter 3, we analyze Spitzer -IRS data of 15 U/LIRGs and constrain the OH-to-H abundance ratio, XOH, by taking advantage of the optically thin OH 35 ?m transition in the MIR. We compare our value with those in the literature and evaluate the accuracy of radiative transfer models to predict absorption line profile properties. Chapter 4 explores the multiphase nature of outflows. We analyze Herschel?PACS data of seven nearby galaxies with well-known winds. We mainly study the neutral atomic and ionized gas phases, but also include the molecular phase. After delin- eating the winds from their host galaxy disks, we estimate the spatial (e.g. location, extent, morphology) and kinematic (e.g. radial velocities, 1-? line widths) proper- 20 ties of the neutral atomic and ionized gas phases in the outflows. We compare our results with data in other wavelengths. We also present the results of the analysis of the molecular gas traced by OH 119 ?m. In Chapter 5, we summarize the main results of each chapter in this thesis and discuss future work. 21 Chapter 2: Search for Molecular Outflows in Local Volume AGN with Herschel -PACS 2.1 Introduction Massive, galactic-scale outflows driven by star formation and/or active galactic nuclei (AGNs) may be the dominant form of feedback in galaxies (Fabian, 2012; Veilleux et al., 2005). These outflows (or winds) likely affect the evolution of galaxies by regulating star formation and black hole (BH) accretion activity. These winds may shut off the gas feeding process and stop the growth of both the BH and the spheroidal component (Di Matteo et al., 2005), thereby explaining the tight ?bulge?BH mass relation? (e.g. Fabian, 2012; Kormendy & Ho, 2013; Kormendy & Richstone, 1995; Marconi & Hunt, 2003). They may also quench star formation altogether and help explain the presence of ?red-and-dead,? gas-poor ellipticals, and the bimodal color distribution observed in large galaxy surveys (e.g. Baldry et al., 2004; Strateva et al., 2001). Winds may also be the primary mechanism by which metals are transferred from galaxies to their surrounding halos and, to a lesser extent, to the intergalactic medium. 22 Table 2.1. Herschel Observations of BAT AGN Name OBSID texp [sec] Program (1) (2) (3) (4) CenA 1342225989 976 OT1 shaileyd 1 Circinus 1342225147 958 OT1 shaileyd 1 ESO 005?G004 1342245457 746 OT2 sveilleu 6 ESO 137?34 1342252089 1452 OT2 sveilleu 6 IC 5063 1342241848 2897 OT2 sveilleu 6 IRAS 04410+2807 1342249997 2897 OT2 sveilleu 6 IRAS 19348?0619 1342241495 1452 OT2 sveilleu 6 MCG?05.23.16 1342245454 14251 OT2 sveilleu 6 MCG?06.30.15 1342247813 14251 OT2 sveilleu 6 Mrk18 1342253541 4309 OT2 sveilleu 6 NGC 1052 1342247734 14251 OT2 sveilleu 6 NGC 1068 1342191154 3944 SHINING target NGC 1125 1342247722 4309 OT2 sveilleu 6 NGC 1365 1342247546 746 OT2 sveilleu 6 NGC 1566 1342244440 746 OT2 sveilleu 6 NGC 2110 1342250314 2158 OT2 sveilleu 6 NGC 2655 1342246552 1452 OT2 sveilleu 6 NGC 2992 1342246246 746 OT2 sveilleu 6 NGC 3079 1342221391 8045 DDT esturm 4 NGC 3081 1342245955 2897 OT2 sveilleu 6 NGC 3227 1342197796 2923 GT1 lspinogl 4 NGC 3281 1342248310 746 OT2 sveilleu 6 NGC 3516 1342245980 3128 GT1 lspinogl 6 NGC 3718 1342253721 7129 OT2 sveilleu 6 NGC 3783 1342247816 2897 OT2 sveilleu 6 NGC 4051 1342247512 1452 OT2 sveilleu 6 NGC 4102 1342247002 746 OT2 sveilleu 6 NGC 4138 1342256950 1452 OT2 sveilleu 6 NGC 4151 1342247511 746 OT2 sveilleu 6 NGC 4258 1342257242 746 OT2 sveilleu 6 NGC 4388 1342197911 3453 GT1 lspinogl 4 NGC 4395 1342247533 1452 OT2 sveilleu 6 NGC 4579 1342248535 746 OT2 sveilleu 6 NGC 4593 1342248372 1452 OT2 sveilleu 6 NGC 4939 1342248509 1452 OT2 sveilleu 6 NGC 4941 1342248508 2862 OT2 sveilleu 6 NGC 4945 1342247792 958 OT1 shaileyd 1 NGC 5033 1342247011 746 OT2 sveilleu 6 NGC 5273 1342246800 14251 OT2 sveilleu 6 NGC 5290 1342247535 2158 OT2 sveilleu 6 NGC 5506 1342247811 746 OT2 sveilleu 6 NGC 5728 1342249309 2508 GT1 lspinogl 6 NGC 5899 1342247007 746 OT2 sveilleu 6 23 Table 2.1 (cont?d) Name OBSID texp [sec] Program (1) (2) (3) (4) NGC 6221 1342252088 746 OT2 sveilleu 6 NGC 6300 1342253357 746 OT2 sveilleu 6 NGC 6814 1342241849 746 OT2 sveilleu 6 NGC 7172 1342218490 2933 GT1 lspinogl 4 NGC 7213 1342245961 1452 OT2 sveilleu 6 NGC 7314 1342245225 746 OT2 sveilleu 6 NGC 7465 1342245963 1452 OT2 sveilleu 6 NGC 7479 1342258846 746 OT2 sveilleu 6 NGC 7582 1342257273 746 OT2 sveilleu 6 Note. ? Column 1: Galaxy name. Column 2: Obser- vation ID. Column 3: Exposure time. Column 4: Pro- gram. 10 10 8 BAT AGN ULIRG/PG QSO 8 8 6 6 6 4 4 4 2 2 2 0 0 0 0.0 0.005 0.01 0.0 0.05 0.1 0.15 0.2 8 9 10 11 12 13 z z log(M?/M ) 12 24 12 10 20 10 8 16 8 6 12 6 4 8 4 2 4 2 ?0 01 0 1 2 3 0 ? 0 20 40 60 80 100 8 9 10 11 12 13 log(SFR/[M yr 1]) ?AGN log(LAGN/L ) Figure 2.1 Histograms showing the distributions of the BAT AGN and ULIRG/PG QSO properties: redshifts, stellar masses, star formation rates (SFR), AGN frac- tions, and AGN luminosities. 24 Number of Sources Number of Sources Until recently, searches for galactic-scale outflows have focused on the brightest sources in bands often affected by obscuration and/or contamination from the host galaxy light (e.g. Lehnert & Heckman, 1996). These searches have generally been directed at either the ionized phase (e.g. Crenshaw et al., 1999; Dunn et al., 2008; Rubin et al., 2010) or the neutral phase (e.g. Heckman et al., 2000; Krug et al., 2010; Martin, 2005; Rupke et al., 2005a,b,c; Rupke & Veilleux, 2011, 2013; Schwartz & Martin, 2004) of the ISM. If winds are to inhibit star formation in the host galaxy, then the mass outflow must affect the phase of the ISM from which stars form (i.e. the cold molecular gas). Our knowledge of molecular outflows is quickly improving. Recent studies of galactic-scale winds have effectively demonstrated that far infrared (FIR) spectroscopy of the hydroxyl molecule (OH) with Herschel -PACS is well suited to identify molecular outflows in the nearby universe (Spoon et al., 2013; Sturm et al., 2011; Veilleux et al., 2013, hereafter, S11, V13, and S13 respectively). S11 reported preliminary evidence of a correlation between AGN luminosity (LAGN) and OH terminal velocity in six ultra-luminous infrared galaxies (ULIRGs; i.e. ULIRGs with higher terminal velocities hosted AGNs with higher luminosities). V13 and S13 later confirmed this correlation via the analyses of larger samples (43 and 24 ULIRGs, respectively). In particular, V13 reported a nonlinear relationship between log(LAGN/L ) and outflow velocity, but noted that better statistics were required at lower AGN luminosities in order to confirm this nonlinearity. These Herschel - based studies, supplemented with millimeter-wave, interferometric studies, have also shown that the molecular gas often dominates the mass and energy budget of these outflows (e.g. Alatalo et al., 2011; Cicone et al., 2014; Feruglio et al., 2010; Fischer 25 et al., 2010; Gonza?lez-Alfonso et al., 2014b; Morganti et al., 2013; Sturm et al., 2011; Veilleux et al., 2013). There is thus a clear need to extend this type of study to lower AGN luminosi- ties and star formation rates (SFRs). For this, we examine Herschel observations of a complete sample of local Swift-BAT selected AGNs. Since stellar processes con- tribute negligibly to the 14-195 keV emission, the Burst Alert Telescope detected active galactic nucleus (BAT AGN) survey is not sensitive to star formation activity within the host galaxy. Additionally, this survey is unbiased to column densities of N . 1024H cm?2. The characteristics and host galaxy properties of these ultra-hard X-ray detected BAT AGNs have been studied extensively across most wavelengths (e.g. Koss et al., 2013, 2011; Matsuta et al., 2012; Mele?ndez et al., 2014; Mushotzky et al., 2014; Shimizu et al., 2016, 2015; Vasudevan & Fabian, 2009; Vasudevan et al., 2010; Winter et al., 2012). By combining the results of our analysis on 52 BAT AGNs with those of V13 on 43 ULIRGs, we extend the range of AGN luminosities, SFRs, and stellar masses sampled in this study by 1-2 orders of magnitude, from which we can draw stronger statistical conclusions on the driving mechanisms of these outflows. We also examine mid-infrared (MIR; 5?37 ?m) spectra from the Infrared Spectrograph (IRS) on board the Spitzer Space Telescope of the combined BAT AGN + ULIRG + PG QSO sample in an attempt to constrain the distributions of the dust, as measured by the strength of the 9.7 ?m silicate feature, and OH gas within these systems. We note that although robust, statistical conclusions can only be drawn from a well-defined sample, valuable insight into molecular winds may still be obtained by combining these two samples which are distinct in their 26 0.5 1.5 0.2 CenA ESO 005-G004 1.0 Circinus 0.1 0.0 0.5 0.0 ?0.5 0.0 ?0.1 ?1.0 ?0.5 ?0.2 ?1.0 ?0.3 ?1.5 ?1.5 ?0.4 ?2?.0 ? ?2.0 ?0.53000 2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.15 0.2 0.20 IC 5063 0.16 0.10 ESO 137-34 IRAS 04410+28070.1 0.12 0.05 0.08 0.0 0.04 0.00 0.00 ?0.1 ?0.04 ?0.05 ?0.08 ?0.10 ?0.2 ?0.12 ?0.16 ?0.?15 ? ? ?0.33000 2000 1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.150 0.10 0.10 0.125 IRAS 19348-0619 MCG-05.23.160.100 MCG-06.30.15 0.05 0.075 0.05 0.050 0.025 0.00 0.00 0.000 ?0.025 ? ?0.05 ?0.050.050 ?0.075 ? ? ? ?0.?10 ? ? ?0.103000 2000 1000 0 1000 2000 3000 3000 2000 1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.3 7 0.04 0.2 Mrk18 NGC 1052 6 5 NGC 1068 0.1 0.02 4 0.0 0.00 3 ? 20.1 ?0.02 1 ?0.2 ?0.04 0 ?0?.3 ?13000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 2.2 Fits to the central spaxel, continuum subtracted OH 119 ?m profiles of the 52 objects in our sample; see Section 2.5.1.3. In each figure, the solid black line represents the data and the magenta line is the best fit to the data. The blue and red dashed lines represent the Gaussian components of the fits. The origin of the velocity scale corresponds to OH 119.233 ?m at the systemic velocity. The two vertical dashed and dotted lines mark the positions of 16OH and 18OH, respectively. selection methods and in their properties. The samples used in this analysis are described in Section 2. The observations 27 Jy Jy Jy Jy Jy Jy JyJy Jy Jy Jy Jy 0.20 0.6 0.4 0.15 NGC 1125 0.4 NGC 1365 0.3 0.10 NGC 1566 0.2 0.05 0.2 0.00 0.1 ?0.05 0.0 0.0 ?0.10 ?0.2 ? ?0.10.15 ?0.?20 ? ? ?0.4 ?0.23000 2000 1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.25 0.10 0.1 NGC 29920.20 NGC 2655 NGC 2110 0.15 0.05 0.0 0.10 ?0.1 0.05 0.00 0.00 ?0.2 ?0.05 ?0.05 ?0.3 ?0.10 ?0.?15 ? ? ?0.?10 ? ? ?0.43000 2000 1000 0 1000 2000 3000 3000 2000 1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 2 0.20 1.0 NGC 3079 1 0.15 0.8 0 NGC 3081 NGC 3227 0.6 ?1 0.10 ?2 0.4 0.05 ?3 0.2 ?4 0.00 0.0 ?5 ?0.05 ? ?0.26 ??73000 ? ?0.102000 ?1000 0 1000 2000 3000 ?3000 ? ?0.42000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.3 0.15 0.2 NGC 3281 0.2 NGC 3516 0.10 NGC 3718 0.0 0.05 ?0.2 0.1 0.00 ?0.4 0.0 ?0.05 ?0.6 ?0.1 ?0.10 ?0.8 ? ? ?0.2 ?0.153000 2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 2.2 (Continued) and data reduction techniques are outlined in Section 3. The results of the analysis are presented in Section 4, while the implications of these results are reported in Section 5. The conclusions are summarized in Section 6. 28 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy 0.10 0.3 0.5 NGC 4102 NGC 3783 0.2 NGC 4051 0.05 0.0 0.1 0.00 0.0 ?0.5 ?0.05 ?0.1 ?1.0 ?0.?10 ?0.23000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.10 0.25 0.15 NGC 4138 0.20 0.10 NGC 4258NGC 4151 0.05 0.15 0.05 0.10 0.00 0.05 0.00 0.00 ?0.05 ?0.05 ?0.05 ? ?0.100.10 ?0.?10 ?0.15 ?0.153000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.3 0.20 0.20 NGC 4388 0.15 0.15 0.2 NGC 4395 NGC 4579 0.10 0.10 0.1 0.05 0.05 0.0 0.00 0.00 ? ?0.05 ?0.050.1 ?0.10 ?0.10 ?0.2 ?0.15 ?0.15 ?0?.3 ?0.20 ?0.203000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.25 0.15 0.15 0.20 0.10 NGC 4939 NGC 4941NGC 4593 0.10 0.15 0.05 0.05 0.10 0.05 0.00 0.00 0.00 ?0.05 ?0.05 ?0.05 ? ?0.10 ?0.100.10 ?0.?153000 ?2000 ? ?0.15 ?0.151000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 2.2 (Continued) 2.2 BAT AGN Sample Selection The BAT AGN in our sample were selected using three criteria: (1) all targets are from the very hard X-ray selected (14-195 keV) 58-month Swift-BAT Survey (Baumgartner et al., 2011) of local AGN. Since the 14-195 keV flux is solely produced 29 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy 50 0.25 0.10 NGC 4945 0.20 NGC 5273 0 NGC 5033 0.15 0.05 ?50 0.10 0.05 0.00 ?100 0.00 ? ?0.05 ?0.05150 ?0.10 ?2?00 ? ?0.15 ?0.103000 2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.10 0.6 0.4 NGC 5290 0.4 NGC 5506 0.3 NGC 5728 0.05 0.2 0.2 0.1 0.00 0.0 0.0 ? ?0.10.2 ?0.05 ?0.2 ?0.4 ?0.3 ?0.?103000 ?2000 ? ?0.6 ?0.41000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.15 0.25 0.4 NGC 5899 0.20 NGC 6300 0.10 NGC 6221 0.15 0.2 0.05 0.10 0.0 0.05 0.00 ?0.2 0.00 ?0.4 ?0.05 ?0.05 ? ?0.60.10 ?0.10 ?0.15 ?0.8 ?0.?15 ? ? ?0.?20 ?1.03000 2000 1000 0 1000 2000 3000 3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.20 0.4 0.20 0.15 NGC 7172NGC 6814 0.150.2 NGC 7213 0.10 0.10 0.0 0.05 0.05 0.00 ?0.2 0.00 ?0.05 ? ?0.050.4 ?0.10 ?0.10 ? ?0.60.15 ?0.15 ?0.?20 ?0.8 ?0.203000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 2.2 (Continued) by the AGN and unaffected by the host galaxy light or obscuration along the line- of-sight (N . 1024H cm?2, this criterion removes any ambiguity as to the power of the AGN. Thus, the Swift-BAT survey is arguably superior to soft X-ray, UV, optical, IR, or radio surveys for understanding the role of AGN-driven winds in galaxies. (2) All targets have a total integrated flux at 120 ?m of Stot120 & 1 Jy so 30 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy 0.2 0.20 0.4 NGC 7314 NGC 74790.15 NGC 7465 0.2 0.1 0.10 0.0 0.0 0.05 ?0.2 0.00 ?0.1 ? ?0.40.05 ?0.6 ? ?0.100.2 ?0.15 ?0.8 ?0?.33000 ? ? ?0.?20 ? ? ?1.02000 1000 0 1000 2000 3000 3000 2000 1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.5 NGC 7582 0.0 ?0.5 ?1.0 ?1.5 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Figure 2.2 (Continued) that high S/N in the continuum can be reached in a reasonable amount of time with PACS. (3) Finally, all targets are located within 50 Mpc. The low redshifts in this sample provide the best possible scale (? 0.02-0.2 kpc arcsec?1) for spatially separating star-formation emission from the nuclear component. This distance is also large enough to properly sample the AGN luminosity function up to quasar-like values (log LBAT ? 43.5; log LBOL ? 45), without favoring IR-bright systems due to criterion #2. We find that 52 targets meet these three requirements. Of these targets, 42 objects are from the cycle 2 open-time program OT2 sveilleu 6 (PI: S. Veilleux), 3 objects are from the guaranteed time program GT1 lspinogl 4 (PI: L. Spinoglio), 2 objects are from the guaranteed time program GT1 lspinogl 6 (PI: L. Spinoglio), 3 objects are from the cycle 1 open-time program OT1 shaileyd 1 (PI: S. Hailey- Dunsheath), 1 object is from the Director?s Discretionary Time DDT esturm 4 (PI: 31 Jy Jy Jy Jy E. Sturm), and 1 object is from the guaranteed time key program Survey with Herschel of the ISM in Nearby Infrared Galaxies (SHINING; PI: E. Sturm) (see Table 2.1). 32 Table 2.2. Galaxy Properties (Name) z Distance ?AGN log LAGN log M? log SFR fcen/ftot f30?m/f15?m Type (Mpc) (%) (L ) (M ) (M yr ?1 ) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) CenA 0.001901a 3.7 93.2 9.78 ? ? ? ? ? ? 0.14 2.44 Type 2 Circinus 0.001573a 4.2 86.6 9.19 ? ? ? ? ? ? 0.22 3.32 Type 1 ESO 005?G004 0.006379b 22.4 89.4 9.73 ? ? ? 0.72 0.15 2.95 Type 2 ESO 137?34 0.009144 33 ? ? ? 9.99 ? ? ? 0.5 0.2 ? ? ? Type 2 IC 5063 0.011198c 49 93.3 10.75 ? ? ? 0.61 0.38 2.43 Type 2 IRAS 04410+2807 0.011268 48.7 ? ? ? 10.61 ? ? ? 0.43 0.37 ? ? ? Type 2 IRAS 19348?0619 0.010254 44.3 ? ? ? 10.17 ? ? ? 0.66 0.29 ? ? ? Type 1 MCG?05.23.16 0.008486 36.6 96.9 10.94 9.56 ?0.53 0.59 1.97 Type 2 MCG?06.30.15 0.007749 33.4 99.9 10.36 ? ? ? ?0.4 0.57 1.60 Type 1 Mrk18 0.011345d 47.9 79.0 9.96 9.57 0.41 0.48 4.40 Type 2 NGC 1052 0.005037 19.5 95.7 9.56 10.35 ? ? ? 0.46 2.12 Type 2 NGC 1068 0.003931a 12.7 100 9.26 ? ? ? ? ? ? 0.14 1.20 Type 2 NGC 1125 0.010931 47.2 60.8 10.1 ? ? ? 0.39 0.42 7.31 Type 2 NGC 1365 0.005349a 17.9 75.3 9.82 ? ? ? 1.51 0.18 4.96 Type 1 NGC 1566 0.005017 12.2 ? ? ? 9.01 ? ? ? ? ? ? 0.21 ? ? ? Type 1 NGC 2110 0.007789 35.6 94.0 11.11 10.63 0.49 0.43 2.34 Type 2 NGC 2655 0.00467 24.4 91.8 9.41 ? ? ? ?0.1 0.25 2.63 Type 2 NGC 2992 0.00771 31.6 92.1 9.94 10.31 0.75 0.38 2.59 Type 2 NGC 3079 0.003797d 19.1 68.7 9.59 9.98 1.22 0.29 5.99 Type 2 NGC 3081 0.007976 26.5 93.8 10.27 10.31 0.12 0.35 2.37 Type 2 NGC 3227 0.004001a 18.7 88.6 10.09 9.98 0.55 0.33 3.05 Type 1 NGC 3281 0.010674 46.1 92.7 10.77 10.24 0.84 0.46 2.51 Type 2 NGC 3516 0.008889e 52.5 95.5 11.02 10.46 0.19 0.44 2.15 Type 1 NGC 3718 0.003312 17 ? ? ? 9.05 9.98 ?0.3 0.23 ? ? ? LINER NGC 3783 0.009791f 47.8 95.2 11.12 ? ? ? 0.68 0.11 2.19 Type 1 33 Table 2.2 (cont?d) (Name) z Distance ?AGN log LAGN log M? log SFR fcen/ftot f30?m/f15?m Type (Mpc) (%) (L ) (M ) (M yr ?1) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) NGC 4051 0.002490a 14.3 95.2 9.41 9.44 0.57 0.34 2.19 Type 1 NGC 4102 0.002823 20.4 61.7 9.57 9.68 1.02 0.51 7.15 LINER NGC 4138 0.002962 20.7 99.3 9.62 9.61 0.04 0.05 1.67 Type 1 NGC 4151 0.003502a 9.9 100 10.23 ? ? ? ?0.38 0.24 1.48 Type 1 NGC 4258 0.001494 7.5 98.4 8.59 ? ? ? ? ? ? 0.07 1.79 Type 2 NGC 4388 0.008467a 21 84.3 10.6 10.53 0.58 0.14 3.64 Type 2 NGC 4395 0.001064 4.5 72.7 8.21 8.28 ? ? ? 0.1 5.34 Type 1 NGC 4579 0.00506 19.6 ? ? ? 9.11 ? ? ? ? ? ? 0.21 ? ? ? Type 2 NGC 4593 0.008455a 30.8 76.5 10.43 10.75 ? ? ? 0.4 4.78 Type 1 NGC 4939 0.010374 44.8 90.2 10.21 ? ? ? 0.82 0.06 2.84 Type 2 NGC 4941 0.003707d 18.7 89.5 9.36 ? ? ? ?0.19 0.22 2.93 Type 2 NGC 4945 0.001836g 4.1 10.4 9.18 ? ? ? ? ? ? 0.41 18.96 Type 2 NGC 5033 0.003211a 19.6 ? ? ? 8.84 ? ? ? 0.95 0.1 ? ? ? Type 1 NGC 5273 0.003619 16 77.4 9.06 9.64 ? ? ? 0.55 4.64 Type 1 NGC 5290 0.008583 35 ? ? ? 9.88 10.23 0.62 0.12 ? ? ? Type 2 NGC 5506 0.006228a 23.8 93.3 10.64 10.02 0.25 0.52 2.43 Type 1 NGC 5728 0.009475a 30.6 70.4 10.43 10.78 0.78 0.41 5.71 Type 2 NGC 5899 0.008546 38.1 91.0 9.97 10.28 0.95 0.13 2.73 Type 2 NGC 6221 0.004999 12.2 65.4 8.99 ? ? ? 0.76 0.24 6.53 Type 2 NGC 6300 0.003699 14.4 87.0 9.82 ? ? ? 0.63 0.31 3.27 Type 2 NGC 6814 0.005214 22.8 98.0 10.11 10.3 0.6 0.05 1.83 Type 1 NGC 7172 0.009180e 33.9 92.4 10.8 ? ? ? 0.83 0.24 2.54 Type 2 NGC 7213 0.005983d 22 100 9.82 ? ? ? 0.25 0.12 1.46 Type 1 NGC 7314 0.004839h 18.6 89.0 9.77 10.06 ? ? ? 0.13 3.00 Type 1 NGC 7465 0.006666i 27.2 ? ? ? 9.54 ? ? ? 0.21 0.35 ? ? ? Type 2 34 Table 2.2 (cont?d) (Name) z Distance ?AGN log LAGN log M? log SFR fcen/ftot f30?m/f15?m Type (Mpc) (%) (L ?1 ) (M ) (M yr ) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) NGC 7479 0.007942 33.9 83.7 9.88 ? ? ? 1.14 0.37 3.73 Type 2 NGC 7582 0.005541j 20.9 73.8 10.06 ? ? ? 1.2 0.49 5.18 Type 2 aRedshift is calculated using the [OI] 63.18 ?m , [OI] 145.53 ?m, [CII] 157.74 ?m, and [NII] 121.90 ?m emission lines. bRedshift is calculated using the [OI] 145.53 ?m emission line. cRedshift is calculated using the [OI] 63.18 ?m and [CII] 157.74 ?m emission lines. dRedshift is calculated using the [CII] 157.74 ?m emission line. eRedshift is calculated using the [CII] 157.74 ?m and [NII] 121.90 ?m emission lines. fRedshift is calculated using the [OI] 63.18 ?m and [OI] 145.53 ?m emission lines. gRedshift is calculated using the [OI] 63.18 ?m, [OI] 145.53 ?m, and [CII] 157.74 ?m emission lines. hRedshift is calculated using the [OI] 145.53 ?m and [CII] 157.74 ?m emission lines. iRedshift is calculated using the [OI] 63.18 ?m, [CII] 157.74 ?m, and [NII] 121.90 ?m emission lines. jRedshift is calculated using the [OI] 14.53 ?m , [CII] 157.74 ?m, and [NII] 121.90 ?memission lines. Note. ? Column 1: Galaxy name. Column 2: Redshift value (When available, redshift is calculated from emis- sion lines. Otherwise, redshifts are from NED). Column 3: When available mean, redshift-independent distance is obtained from NED. Otherwise, the luminosity distance is calculated via Wright (2006). Column 4: ?AGN, fractional contribution of the AGN to the bolometric luminosity; see Section 2.4. Column 5: AGN luminosity. Column 6: stellar masses are adopted from Koss et al. (2011). Column 7: star formation rate; see Section 2.4. Column 8: Continuum flux density ratio at 119 ?m of the central spaxel to all 25 spaxels. For reference, the average continuum ratio for a point source calculated from the five PG QSOs in our ULIRG sample is fcen/ftot = 0.56. Column 9: 30 ?m to 15 ?m continuum flux density ratio. Column 10: Spectral type. 35 2.3 ULIRG and PG QSO Sample We include in our analysis the Herschel/PACS spectra of the ULIRGs and PG QSOs from V13. Of the 43 objects (38 ULIRGs + 5 QSOs) from that sample, 23 targets are from the key program SHINING (PI: E. Sturm), 15 targets are from the cycle 1 open-time program OT1 sveilleu 1 (PI: S. Veilleux), and 5 are from the cycle 2 open-time program OT2 sveilleu 4 (PI: S. Veilleux). This sample spans a broad range of merger stages. 20 objects are cool (f25/f60 ? 0.2 ) pre-merger ULIRGs, 18 objects are warm (f25/f60 > 0.2) quasar-dominated, late-stage, fully coalesced ULIRGs, and 5 objects are ?classic? IR-faint QSOs. These QSOs are in a late merger phase in which the quasar has finally shed its natal ?co- coon? of dust and gas and feedback effects may be receding (Veilleux et al., 2009, V13) 2.4 Properties of the Sample Galaxies The properties of our BAT AGN sample are listed in Table 2.2. The notes to Table 2.2 briefly explain the meaning of each of these quantities. Some quan- tities, however, require further clarification. We apply the bolometric correction from Winter et al. (2012) to our BAT AGN luminosities, which are derived from the fluxes from the Swift BAT 70-month survey. Since the Swift-BAT bandpass is at high enough energies (14-195 keV) to be unaffected by all but the highest levels of obscuration, the luminosities in this bandpass should be the direct unobscured 36 signature from the AGN. Thus, it is assumed to be a good proxy for the bolometric luminosity of the AGN. The correction is a scale factor derived from the correlation between the bolometric luminosity, which is determined from simultaneous broad- band fitting of the spectral energy distribution of 33 sources in the optical, UV, and X-ray (Vasudevan & Fabian, 2007, 2009), and the Swift BAT band 14-195 keV luminosities of those sources. The ordinary least-squares line through these data of Winter et al. (2012) yields the following correction: LAGN = 10.5? L14?195keV, (2.1) where LAGN is the bolometric luminosity of the AGN and L14?195keV is the Swift BAT luminosity in the 14-195 keV band. For the ULIRGs, we adopt the starburst and AGN luminosities from V13, which were calculated as follows: the bolometric luminosities were estimated to be LBOL = 1.15 LIR, where LIR is the infrared luminosity over 8 ? 1000 ?m (Sanders & Mirabel, 1996), and LBOL = 7L(5100 A?)+LIR for the PG QSOs (Netzer et al., 2007). Here, L(5100 A?) corresponds to ?L? at 5100 A?. The starburst and AGN luminosities were next calculated from LBOL = LAGN + LSB (2.2) = ?AGN LBOL + LSB, (2.3) where ?AGN is the fractional contribution of the AGN to the bolometric luminosity, 37 hereafter called the ?AGN fraction? for short. For the BAT AGN sample, we derive the AGN fractions from the rest frame f30?m/f15?m continuum flux density ratio, which was found by Veilleux et al. (2009) to be more tightly correlated with the PAH-free, silicate-free MIR/FIR ratio and the AGN contribution to the bolometric luminosity than any other Spitzer-derived continuum ratio. The fraction of the 15 ?m flux produced by the AGN is defined as: AGN%(f15) (f /f? 30 15 )agn 100 (f30/f15)agn + (f30/f15)sb (2.4) (f30/f15)obs ? (f30/f15)= sb? ,(f30/f15)agn (f30/f15)sb where (f30/f15)obs is the observed flux density ratio. (f30/f15)agn and (f30/f15)sb are the flux density ratios due to the AGN and starburst, respectively. We adopt from Table 9 of Veilleux et al. (2009) the zero-point values of log (f30/f15)agn = 0.2 and log (f30/f15)sb = 1.35. 38 Table 2.3. Properties of the OH 119 ?m Profiles (Name) v50 (abs) v84 (abs) Fluxabs EQWabs v50 (emi) v84 (emi) Fluxemi EQWemi EQWTotal (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) CenA 63 195 521.1 27 ?201: ?327: ?107.9 ?6 21 Circinus 81 201 494.3 8 ?225 ?327: ?313.9: ?5: 3 ESO 005?G004 9 ?141 167.1 70 622 754 ?32 ?14 55 ESO 137?34 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?9.4 ?14 ?14 IC 5063 ?33:: ?309:: 53.7: 16 ? ? ? ? ? ? ? ? ? ? ? ? 16 IRAS 04410+2807 ? ? ? ? ? ? ? ? ? ? ? ? 21:: 135:: ?36.8:: ?37:: ?37:: IRAS 19348?0619 ? ? ? ? ? ? ? ? ? ? ? ? 57 207 ?28.5 ?25: ?25: MCG?05?23?016 ? ? ? ? ? ? ? ? ? ? ? ? 57:: 207:: ?8.6:: ?21:: ?21:: MCG?06?30?015 ? ? ? ? ? ? ? ? ? ? ? ? 129:: 255:: ?6.9:: ?8:: ?8:: Mrk 18 ? ? ? ? ? ? 12.4: 6: ? ? ? ? ? ? ? ? ? ? ? ? 6 NGC 1052 ? ? ? ? ? ? ? ? ? ? ? ? ?39:: 117:: ?9.9:: ?27:: ?27:: NGC 1068 ? ? ? ? ? ? ? ? ? ? ? ? 15 213 ?2479.1 ?70 ?70 NGC 1125 123: ?3: 24.8: 20 ? ? ? ? ? ? ? ? ? ? ? ? 20 NGC 1365 ? ? ? ? ? ? ? ? ? ? ? ? 177: 267: ?104.8: ?4: ?4: NGC 1566 ? ? ? ? ? ? ? ? ? ? ? ? 51 177 ?93.6 ?36 ?36 NGC 2110 ? ? ? ? ? ? ? ? ? ? ? ? 75 351 ?103.9 ?38 ?38 NGC 2655 ? ? ? ? ? ? ? ? ? ? ? ? 9:: 135:: ?8:: ?9:: ?9:: NGC 2992 45 ?129 146.9 33 ? ? ? ? ? ? ? ? ? ? ? ? 33 NGC 3079 87 ?93 2580.2 119 ? ? ? ? ? ? ? ? ? ? ? ? 119 NGC 3081 ? ? ? ? ? ? ? ? ? ? ? ? 93 219 ?38.7 ?33 ?33 NGC 3227 ? ? ? ? ? ? ? ? ? ? ? ? ?27 147 ?225.3 ?53 ?53 NGC 3281 273 429 313.9 99 ?321 ?453 ?32 ?10 ?86 NGC 3516 ? ? ? ? ? ? ? ? ? ? ? ? ?99 39 ?52.4 ?78 ?78 NGC 3718 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?9.3: ?15: ?15: NGC 3783 ? ? ? ? ? ? ? ? ? ? ? ? ?69:: 81:: ?16.8:: ?33:: ?33:: 39 Table 2.3 (cont?d) (Name) v50 (abs) v84 (abs) Fluxabs EQWabs v50 (emi) v84 (emi) Fluxemi EQWemi EQWTotal (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) NGC 4051 ? ? ? ? ? ? ? ? ? ? ? ? ?9: 141: ?102.2: ?57: ?57: NGC 4102 ?3 ?135 343.4 12: ? ? ? ? ? ? ? ? ? ? ? ? 12: NGC 4138 ? ? ? ? ? ? ? ? ? ? ? ? ?99:: 33:: ?19.2:: ?81:: ?81:: NGC 4151 ? ? ? ? ? ? ? ? ? ? ? ? 75 333: ?96.7 ?82 ?82 NGC 4258 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?19.2 ?18 ?18 NGC 4388 ? ? ? ? ? ? 32.7 18 ? ? ? ? ? ? ? ? ? ? ? ? 18 NGC 4395 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?14.2:: ?254:: ?254:: NGC 4579 ? ? ? ? ? ? ? ? ? ? ? ? ?15:: 111:: ?40.3:: ?36:: ?36:: NGC 4593 ? ? ? ? ? ? ? ? ? ? ? ? 15 141 ?44.4 ?26: ?26: NGC 4939 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?11.1 ?45 ?45 NGC 4941 ? ? ? ? ? ? ? ? ? ? ? ? 3:: 129:: ?22.4:: ?40:: ?40:: NGC 4945 87 ?51 66839.9 171 ? ? ? ? ? ? ? ? ? ? ? ? 171 NGC 5033 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?15.1 ?4 ?4 NGC 5273 ? ? ? ? ? ? ? ? ? ? ? ? 87: 231: ?13.6: ?30: ?30: NGC 5290 9:: ?117:: 14:: 25:: ? ? ? ? ? ? ? ? ? ? ? ? 25:: NGC 5506 45:: ?357:: 234: 59: ? ? ? ? ? ? ? ? ? ? ? ? 59: NGC 5728 ?15:: ?141:: 96:: 20:: ? ? ? ? ? ? ? ? ? ? ? ? 20:: NGC 5899 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?11: ?15: ?15: NGC 6221 ? ? ? ? ? ? ? ? ? ? ? ? ?99:: 33:: ?32:: ?4:: ?4:: NGC 6300 ?3 ?201 301.5 70 ? ? ? ? ? ? ? ? ? ? ? ? 70 NGC 6814 ? ? ? ? ? ? ? ? ? ? ? ? 105:: 231:: ?33.8:: ?82:: ?82:: NGC 7172 ?51: ?207:: 252.6 85 ? ? ? ? ? ? ? ? ? ? ? ? 85 NGC 7213 ? ? ? ? ? ? ? ? ? ? ? ? ?3:: 141:: ?28.3:: ?42:: ?42:: NGC 7314 ? ? ? ? ? ? 18 24 ? ? ? ? ? ? ? ? ? ? ? ? 24 NGC 7465 ? ? ? ? ? ? ? ? ? ? ? ? ?69:: 57:: ?25.6:: ?13:: ?13:: 40 Table 2.3 (cont?d) (Name) v50 (abs) v84 (abs) Fluxabs EQWabs v50 (emi) v84 (emi) Fluxemi EQWemi EQWTotal (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (km s?1) (Jy km s?1) (km s?1) (km s?1) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) NGC 7479 ?51 ?658:: 379.6 73 ? ? ? ? ? ? ? ? ? ? ? ? 73 NGC 7582 105 3 383.6 13 604: 700:: ?50.1:: ?2 11 Note. ? Column 1: galaxy name. Column 2: v50(abs) is the median velocity of the fitted absorption profile i.e. 50% of the absorption takes place at velocities above - more positive than - v50. Column 3: v84(abs) is the velocity above which 84% of the absorption takes place. Column 4: the total integrated flux for the absorption component(s). Column 5: the total equivalent width for the absoprtion component(s). Column 6: v50 (emi) is the median velocity of the fitted emission profile. Column 7: v84 (emi) is the velocity below which 84% of the emission takes place. Column 8: the total integrated flux for the emission component(s). Column 9: the total equivalent width for the emission component(s). Column 10: The total equivalent width for the sum of the two components for one line of the OH doublet. Fluxes followed by a colon indicate uncertainties between 20% and 50%. Velocities followed by a colon indicate uncertainties between 50? 150 km s?1. Fluxes followed by a double colon indicate uncertainties > 50%. Velocities followed by a double colon indicate uncertainties > 150 km s?1. 41 The fraction of the bolometric luminosity produced by the AGN is then cal- culated from: ? AGN(Lbol) L(bol) agn ?AGN = 100 L(bol)agn + L(bol)sb [ ][1( ) ] (2.5)= (L? (15?m)/L(bol)1 + 100 ? 1 )agn AGN% (f15) L? (15?m)/L(bol) sb where we adopt the bolometric corrections of log [L? (15?m)/L(bol)]agn = ?14.33 log [L? (15?m)/L(bol)]sb = ?14.56 from Table 10 of Veilleux et al. (2009). As a check, we compare the AGN fractions derived via this method with those found in Tommasin et al. (2010). For the few objects in common (i.e. NGC 1125, NGC 3516, NGC 4388, NGC 7172, and NGC 7582), we find good agreement between the two methods. For the ULIRGs and QSOs, we adopt the AGN fractions from V13 which are calculated using the same method applied to the BAT AGN. BAT AGN stellar masses are adopted from Koss et al. (2011). ULIRG stellar masses are calculated by adopting H-band absolute magnitudes from V13. We then assume M?H = ?23.7 (the H-band absolute magnitude of a L? galaxy in a Schechter function description of the local field galaxy luminosity function (Bell & de Jong, 2001; Cole et al., 2001)) and m? = 1.4? 1011 M (the mass of an early-type galaxy 42 at the knee of the Schechter distribution (Cole et al., 2001; Veilleux et al., 2002)). BAT AGN SFRs are derived from equation [25] of Calzetti et al. (2010): L(160) [erg s?1] SFR(160)(M yr ?1) = ? , (2.6)4.8 1042 where L(160) is ?L? at 160 ?m and the factor 4.8?1042 assumes a Kennicutt (1998) calibration for SFR. SFRs for ULIRGs and PG QSOs are calculated from equation [1] of Rupke et al. (2005b): LSB SFR = ? , (2.7)5.8 109 L where LSB = LBOL(1? ?AGN). Table 2.1 shows the distributions of redshifts, stellar masses, SFRs, AGN fractions, and AGN luminosities for all 52 BAT AGN in our sample (blue diagonals) and for all 43 objects (38 ULIRGs + 5 PG QSOs; red diagonals) from V13. The AGN fractions of the BAT AGN are typically higher than the fractions of the V13 sample of objects.The AGN luminosities and SFRs of the BAT AGN, however, are typically lower by two orders of magnitude than those of the ULIRGs and QSOs. 43 2.5 Observations, Data Reduction, and Spectral Analysis 2.5.1 OH 119 ?m Feature 2.5.1.1 OH Observations The OH observations were obtained with the PACS FIR spectrometer (Poglitsch et al., 2010) on board Herschel (Pilbratt et al., 2010). We focus our efforts on the ground-state OH 119 ?m 2?3/2 J = 5/2? 3/2 rotational ?-doublet. This feature is the strongest transition in ULIRGs and is positioned near the peak spectroscopic sensitivity of PACS. Fischer et al. (2010), S11, and V13 have demonstrated the efficacy of using the OH 119 ?m feature to determine wind characteristics. The data for OT2 sveilleu 6 were obtained in a similar fashion as those in SHINING (S11) and OT1 sveilleu 1 (V13). PACS was used in range scan spec- troscopy mode in high sampling centered on the redshifted OH 119 ?m + 18OH 120 ?m complex with a velocity range of ? ?4000 km s?1 (rest-frame 118-121 ?m) to provide enough coverage on both sides of the OH complex for reliable continuum placement. As a result, the PACS spectral resolution is ? 270 km s?1. The to- tal amount of observation time (including overheads) for OT2 sveilleux 6 was 35.3 hours. The program ID and observing time for each target, including all overheads, are listed in Table 2.2. A large chopper throw of 3? is used in all cases. 44 300 Abs (BAT AGN)(a) Emi (BAT AGN) (b) (c) Abs/Emi (BAT AGN) Abs (ULIRG) Emi (ULIRG) 200 Abs/Emi (ULIRG) Emi (PG QSO) 100 0 ?100 9 10 11 12 13 0 20 40 60 80 100 8 9 10 11 12 13 log(M?/M ) ?AGN [%] log(LAGN/L ) Figure 2.3 Total (absorption + emission) equivalent widths of OH 119 ?m (positive values indicate absorption and negative values indicate emission) as a function of the (a) stellar masses, (b) AGN fractions, (c) AGN luminosities. The colors blue and red refer to BAT AGN and ULIRGs/PG QSOs, respectively. Filled squares, triangles, and circles represent BAT AGN or ULIRGs in which OH 119 ?m is seen purely in absorption, purely in emission, or with composite absorption/emission, respectively. Open triangles represent PG QSOs in which OH 119 ?m is seen purely in emission. The blue, vertical, dotted lines refer to BAT AGN in which OH 119 ?m is not detected. Similarly, the red, vertical, dash-dotted lines represent ULIRGs/PG QSOs with undetected OH 119 ?m. The gray, horizontal line marks the null OH equivalent width. 2.5.1.2 OH Data Reduction The reduction of the OH data on the BAT AGN sample was carried out using the standard PACS reduction and calibration pipeline (ipipe) included in HIPE 6.0. For the final calibration, fainter sources were normalized to the telescope flux (which dominates the total signal) and recalibrated using a reference telescope spectrum obtained from dedicated Neptune observations during the Herschel performance verification phase. Sturm et al. (2011) demonstrated that this telescope background technique can reliably recover the continuum from faint sources. In the following, we use the spectrum of the central 9.4?? ? 9.4?? spatial pixel (spaxel) only, without the application of a point source flux correction. In a few objects we note that the 45 EW [km s?1OH ] OH emission is extended. These objects will be the focus of a future paper. The reduced spectra were next smoothed using a Gaussian kernel of width 0.05 ?m (i.e. about half a resolution element) to reduce the noise in the data before the spectral analysis. A spline was fit to the continuum emission and subtracted from the spectra. Subsequently, spectral fitting was carried out on these continuum- subtracted spectra. 2.5.1.3 Spectral Analysis of the OH Doublet Line profile fits of the OH 119.233, 119.441 ?m doublet were computed by using PySpecKit, a spectroscopic analysis and reduction toolkit for optical, infrared, and radio spectra (Ginsburg & Mirocha, 2011). The toolkit uses the Levenberg- Marquardt technique to solve the least-squares problem in order to find the best fit for the observations. Profile fitting of the OH doublet followed a similar procedure as that outlined in V13, in which the doublet profile was modeled using four Gaussian components (two components for each line of the doublet), each characterized by their amplitude (either negative or positive), peak position, and standard deviation (or, equivalently FWHM). However, many of the OH profiles observed here were fit with only two Gaussian components (one component for each line of the doublet), since a fit with four Gaussian components often led to spurious results (i.e. skinny components with FWHM ? 10 km s?1, values too small to be considered real). The separation between the two lines of the doublet was set to 0.208 ?m in the rest frame (? 520 km s?1) and the amplitude and standard deviation were fixed to be 46 the same for each component in the doublet. In cases where OH was not detected, two Gaussian components, characterized by an amplitude consistent with the 1? level of the noise and a FWHM (300 km s?1) approximately equal to the resolution of PACS, were fit and maximum values for the OH flux and equivalent width were derived. Four distinct scenarios apply to our data: (1) pure OH absorption, (2) pure OH emission, (3) P Cygni profiles, and (4) inverse P Cygni profiles. In scenario 1, there is no evidence of OH emission and each line of the OH doublet is fitted with 1-2 absorption components. In the case of the two-component fit, one component traces the stronger low-velocity component of the outflow, while the other component traces the fainter high-velocity component. In the case of the single-component fit, only the low-velocity component is captured. Scenario 2 is treated similarly. In this scenario, there is no evidence for any OH absorption and 1-2 Gaussian components are used to model each line of the OH doublet. In scenario 3, each line of the doublet is modeled with a single blueshifted absorption and a single redshifted emission component. In scenario 4, each line of the doublet is modeled with a single blueshifted emission component and a single redshifted absorption component. In scenarios 3 and 4, just as in the one-component fits of Scenarios 1 and 2, only the low-velocity component of the outflow is captured. These fits were first used to quantify the strength and nature (absorption versus emission) of the OH feature: (1) the total flux and equivalent width of the OH 119.441 ?m line, adding up all of the absorption and emission components, (2) the flux and equivalent width of the absorption component(s) used to fit this line, 47 and (3) the flux and equivalent width of the emission component(s) used to fit this line. Following the same method as that outlined in V13, we also characterize the OH profile by measuring velocities: (1) v50(abs) is the median velocity of the fitted absorption profile, i.e., 50% of the absorption takes place at velocities above (more positive than) v50(abs), (2) v84(abs) is the velocity above which 84% of the absorption takes place, (3) v50(emi) is the median velocity of the fitted emission profile, i.e., 50% of the emission takes place at velocities below (less positive than) v50(emi), and (4) v84(emi) is the velocity below which 84% of the emission takes place. We note that three objects are fitted with inverted P-Cygni profiles, suggesting inflow. Circinus is an unambiguous fit while Centaurus A and NGC 3281 are marginally fit with inverted P-Cygni profiles. The velocities from the inverted P-Cygni profile fits are measured as thus: v50(abs) and v84(abs) are the velocities below which 50% and 84% of the absorption takes place, respectively, and v50(emi) and v84(emi) are the velocities above which 50% and 84% of the emission takes place, respectively. We note that the OH emission is extended in some of our BAT AGN sources. While a full analysis of the 5 ? 5 spaxels spectrum is outside the scope of this paper, we provide the ratio of the central spaxel 119 ?m continuum flux density to that of the summed 25 spaxels (fcen/ftot) to quantify the degree to which some of our BAT AGN are spatially extended (see Table 2.2). For reference, the average value of fcen/ftot for a point source (derived from the five PG QSOs in our sample) is 0.56. Note that the continuum is often considerably more extended than the OH 119 ?m feature. 48 8 Abs (BAT AGN) 7 Emi (BAT AGN) Abs/Emi (BAT AGN) Null (BAT AGN) 6 Abs (ULIRG) Emi (ULIRG) Abs/Emi (ULIRG) 5 Null (ULIRG) Emi (PG QSO) 4 3 2 1 0 ?1 0 20 40 60 80 100 ?AGN [%] Figure 2.4 The apparent strength of the 9.7 ?m silicate feature relative to the local mid-infrared continuum as a function of the AGN fractions. Note that S9.7?m is a logarithmic quantity and can be interpreted as the apparent silicate optical depth. The strength of silicate absorption increases upward. Sign conventions and meanings of the symbols are the same as those in Section 2.5.1.1. Crosses represent objects with a null OH detection. Vertical lines indicate objects with a null 9.7 ?m silicate feature detection. 2.5.2 The 9.7 ?m Silicate Feature 2.5.2.1 Data Reduction of the 9.7 ?m Silicate Feature Mid-infrared (MIR; 5-37 ?m) low resolution spectra (R ? 60 ? 127) used to measure the 9.7 ?m silicate feature for the BAT AGN, ULIRG, and PG QSO samples were extracted from The Cornell AtlaS of Spitzer/IRS Sources (CASSIS1). The archival data were obtained with the Infrared Spectrograph (IRS; Houck et al. (2004)) on board the Spitzer Space Telescope (Werner et al., 2004). There were two objects however (NGC 7479 and NGC 4945), in which the MIR spectrum was 1http://www.cassis.sirtf.com/atlas 49 S9.7?m extracted from the Spitzer Heritage Archive (SHA2). The Spitzer observations were made with the Short-Low (SL, 5.2-14.5 ?m) and Long-Low (LL, 14.0-38.0 ?m) modules of the IRS. The orders were stitched to LL order 2, requiring order-to-order scaling adjustments of less than ? 15%. 2.5.2.2 Spectral Analysis of the 9.7 ?m Silicate Feature We have measured the strength of the 9.7 ?m silicate feature for sources in our BAT AGN sample and in the ULIRG/QSO sample. The calculation of the mid- infrared continuum loosely follows the method of Spoon et al. (2007). For a majority of the sources, the mid-infrared continuum (e.g. the continuum flux density fcont) is determined from a cubic spline interpolation to continuum pivots at 5.2, 5.6, 14.0, 27.0, and 31.5 ?m. For objects in which the 9.7 ?m silicate feature dominates the spectrum (i.e. there is very little PAH emission), an additional pivot point is added at 7.8 ?m. The pivot points are adopted from Spoon et al. (2007), however we have placed a pivot point at 27 ?m instead of at 25 ?m due to the proximity of the [OIV] emission line at 25.89 ?m which is common in BAT AGN. Due to the diversity of our BAT AGN sample, the wavelength range in which the 9.7 ?m silicate feature is determined differs slightly from source to source (see Table 2.4). Typically, however, the observed flux density (e.g. fobs) is determined from a cubic spline interpolation to the data between ? 8?m and ? 14?m. This interpolation skips the H2 line at 9.6 ?m, [S IV] line at 10.51 ?m, PAH feature at 11.25 ?m, H2 line at 12.28 ?m, and [NeII] at 12.68 ?m. 2http://sha.ipac.caltech.edu/applications/Spitzer/SHA/ 50 The apparent strength of the 9.7 ?m silicate feature is then defined as: ( ) fobs(9.7?m) S9.7?m = ? ln . (2.8) fcont(9.7?m) Here, we adopt the sign convention consistent with our OH 119 ?m analysis (e.g. positive S9.7?m values indicate absorption while negative S9.7?m values indicate emission). Note that this convention is different from previous studies (e.g. Spoon et al., 2007). For sources with a silicate absorption feature, S9.7?m can be interpreted as the the apparent silicate optical depth. Table 2.4 and Table 2.5 lists pivot points, integration ranges, and measured equivalent widths and fluxes of the 9.7 ?m silicate feature for our samples. MIR spectra showing these pivot points and integration ranges may be found in the Appendix. 2.6 Results Figure 2.1 shows the fits to the OH 119 ?m profiles for our BAT AGN targets. The OH 119 ?m parameters for all targets derived from these fits are listed in Table 3.3. The meaning of each parameter is discussed in Section 2.5.1.3 and in the notes to Table 3.3. Note that the fluxes and equivalent widths in Table 3.3 need to be multiplied by a factor of two when considering both lines of the doublet. In this section we compare the OH 119 ?m and S9.7?m results listed in Table 3.3, Table 2.4, and Table 2.5, with the galaxy properties listed in Table 2.2. 51 300 200 100 0 Abs (BAT AGN) Emi (BAT AGN) Abs/Emi (BAT AGN) ? Abs (ULIRG)100 Emi (ULIRG) Abs/Emi (ULIRG) Emi (PG QSO) ?1 0 1 2 3 4 5 S9.7?m Figure 2.5 Total (absorption + emission) equivalent widths of OH 119 ?m as a function of the apparent strength of the 9.7 ?m silicate feature relative to the local mid-infrared continuum. Note that S9.7?m is a logarithmic quantity and can be interpreted as the apparent silicate optical depth. The strength of this absorption feature increases to the right. Sign conventions and meanings of the symbols are the same as those in Section 2.5.1.1. Horizontal, blue, dotted lines represent BAT AGN in which the 9.7 ?m silicate feature is not detected. Similarly, horizontal, red, dash-dotted lines represent ULIRGs/PG QSOs with a null 9.7 ?m silicate feature detection. The black diagonal line shows the ordinary least squares bisector linear regression: EWOH = 56.81?S9.7?m?26.7 kms?1. The Pearson r null probability for the linear relationship between S9.7?m and the total OH equivalent width is P [null]= 1.2 ? 10?9 for the BAT AGN sample and P [null] = 0.003 for the ULIRG + QSO sample. When the samples are combined, we find P [null] = 3.9? 10?9. 2.6.1 The OH 119 ?m Feature The OH 119 ?m doublet is detected in 42 of the 52 objects in our BAT AGN sample. Of the 42 detections, 25 are seen purely in emission, 12 are seen purely in ab- sorption, and 5 are seen with absorption+emission composite profiles. For compari- son, in the ULIRG + QSO sample (V13), 37 objects showed OH 119 ?m detections, 52 EW ?1OH [km s ] 14 12 Pure Abs Pure Abs 12 Pure Emi Pure Emi PCyg 10 PCyg 10 Inv PCyg Inv PCyg8 8 6 6 4 4 2 2 ?0800 ? 0400 0 400 800 ?800 ?400 0 400 800 8 8 Pure Abs Pure Abs PCyg PCyg Inv PCyg Inv PCyg 6 Total 6 Total 4 4 2 2 ?0 0400 ?200 0 200 400 ?800 ?400 0 400 v ?1 ?150 [km s ] v84 [km s ] Figure 2.6 Histograms showing the distributions of the 50% (median; left panels) and 84% (right panels) velocities derived from the multi-Gaussian fits to the OH profiles of the BAT AGN. Top panels show pure absorption components (blue), pure emission components (red dashed), P-cygni emission components (red, left di- agonals), and inverse P-cygni emission components (red, right diagonals). Bottom panels show pure absorption components (filled grey), P-Cygni absorption com- ponents (blue, left diagonals), inverse P-cygni absorption components (blue, right diagonals), and total absorption components (pure + P-Cygni + inverse P-Cygni). 17 of which were seen purely in absorption, 15 showed absorption+emission com- posite profiles, and 5 were seen purely in emission. Section 2.5.1.1a, Section 2.5.1.1c plot the OH equivalent width with the stellar mass and AGN luminosity, respec- tively. For these properties, we see no discernible correlation in either the individual samples (i.e. BAT AGN only and ULIRG + PG QSO only) or in the combined sample. There is however a weak trend between EWOH and AGN fraction (Sec- tion 2.5.1.1b) when the samples are combined. A more formal statistical analysis of these parameters indeed indicates a statistically significant correlation between 53 Number of Sources these quantities when all objects are considered (see Table 2.6). For the BAT AGN in which OH 119 ?m is seen purely in emission, we find that the AGN is dominant (i.e. ?AGN > 50%) in these objects. In agreement with V13, the low luminosities (log(LAGN/L ) . 12) of these 20 AGN-dominated objects with pure OH emission suggest that the AGN fraction is more important than AGN luminosity in setting the character (i.e. strength of emission relative to absorption) of the OH feature. Therefore, we see that dominant AGN in our sample can excite the OH molecule (via radiative pumping or collisional excitation) and produce the 2?3/2 J = 5/2? 3/2 rotational emission line. This trend was noted in V13. 54 Table 2.4. S9.7?m Continuum Parameters and Measured Values of the BAT AGN Sample (Name) Pivot Points Integration Range FluxS9.7?m EQWS9.7?m S9.7?m (?m) (?m) (Jy ?m) (?m) (1) (2) (3) (4) (5) (6) CenA 5.2, 5.6, 14.0, 27.0, 31.5 8.3, 11.3 1.06 1.01 0.79 Circinus 5.2, 5.6, 14.0, 31.5 7.8, 14 23.38 2.06 1.24 ESO 005?G004 5.2, 5.6, 14.0, 27.0, 31.5 8.6, 11.19 0.34 1.55 1.32 ESO 137?34 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? IC 5063 5.2, 5.6, 14.0, 27.0, 31.5 8.15, 12.3 0.37 0.55 0.27 IRAS 00410+2807 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? IRAS 19348?0619 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? MGC?05.23.16 5.2, 5.6, 14.0, 27.0, 31.5 8, 12.45 0.32 0.64 0.31 MGC?06.30.15 5.2, 5.6, 14.0, 27.0, 31.5 8.09, 10.4 0.05 0.16 0.09 Mrk 18 5.2, 5.6, 14.0, 27.0, 31.5 9, 10.4 0.01 0.12 0.11 NGC 1052 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 1068 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 1125 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 1365 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 1566 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 2110 5.2, 5.6, 14.0, 27.0, 31.5 7.8, 14 0.17 0.74 ?0.07 NGC 2655 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 2992 5.2, 5.6, 14.0, 27.0, 31.5 8.75, 11.1 0.13 0.48 0.37 NGC 3079 5.2, 5.6, 14.0, 27.0, 31.5 8.8, 11.13 0.66 1.5 1.56 NGC 3081 5.2, 5.6, 14.0, 27.0, 31.5 8, 11.3 0.09 0.44 0.22 NGC 3227 5.2, 5.6, 14.0, 27.0, 31.5 8.75, 11.09 0.15 0.46 0.28 NGC 3281 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 14 1.98 2.27 1.36 NGC 3516 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 3718 ? ? ? ? ? ? ? ? ? ? ? ? NGC 3783 5.2, 5.6, 14.0, 27.0, 31.5 7.9, 11.3 0.17 0.33 0.13 55 Table 2.4 (cont?d) (Name) Pivot Points Integration Range FluxS EQW S9.7?m S9.7?m 9.7?m (?m) (?m) (Jy ?m) (?m) (1) (2) (3) (4) (5) (6) NGC 4051 5.2, 5.6, 14.0, 21.9, 24.9 8, 10.8 0.07 0.24 ?0.07 NGC 4102 5.2, 5.6, 14.0, 27.0, 31.5 9, 10.95 0.23 0.52 0.41 NGC 4138 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 4151 ? ? ? 7.8, 14 ? ? ? ? ? ? ? ? ? NGC 4258 5.2, 5.6, 14.0, 27.0, 31.5 7.8, 13.39 0.14 1.15 ?0.16 NGC 4388 5.2, 5.6, 14.0, 27.0, 31.5 8.2, 12.2 0.43 1.35 0.79 NGC 4395 5.2, 5.6, 14.0, 27.0, 31.5 7.8, 11.09 0.004 0.67 0.25 NGC 4579 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 4593 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 4939 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 4941 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 8, 11.13 0.02 0.33 0.15 NGC 4945 7.8, 14.0, 27.0, 31.5 7.8, 11.09 1.95 2.72 3.65 NGC 5033 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 5273 4.6, 14.0, 27.0, 31.5 7.7, 11.05 0.01 0.66 0.38 NGC 5290 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 5506 5.2, 5.6, 14.0, 27.0, 31.5 8.35, 12.15 1.2 1.16 0.77 NGC 5728 5.2, 5.6, 14.0, 27.0, 31.5 8.63, 11.16 0.15 1.17 0.94 NGC 5899 5.2, 5.6, 14.0, 27.0, 31.5 8.1, 11.15 0.08 1.44 1.02 NGC 6221 5.2, 5.6, 14.0, 27.0, 31.5 9.2, 10.7 0.03 0.13 0.11 NGC 6300 5.2, 5.6, 14.0, 27.0, 31.5 8.58, 12.2 0.67 1.71 1.16 NGC 6814 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 7172 5.2, 5.6, 14.0, 27.0, 31.5 8.15, 12.52 0.68 2.37 1.97 NGC 7213 5.2, 5.6, 14.0, 27.0, 31.5 8.76, 13.3 0.36 2.26 ?0.41 NGC 7314 5.2, 5.6, 7.8, 14.0 7.8, 14 0.09 1.38 0.6 NGC 7465 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? 56 Table 2.4 (cont?d) (Name) Pivot Points Integration Range FluxS9.7?m EQWS9.7?m S9.7?m (?m) (?m) (Jy ?m) (?m) (1) (2) (3) (4) (5) (6) NGC 7479 4.6, 7.8, 14.0, 27.0, 31.5 7.8, 12.9 1.98 3.09 2.3 NGC 7582 5.2, 5.6, 14.0, 27.0, 31.5 8.8, 11.05 0.81 0.95 0.78 Note. ? Column 1: galaxy name. Column 2: Pivot points in microns used for the continuum interpolation. Column 3: Integration range of the 9.7 ?m silicate feature. Column 4: Total integrated flux of the 9.7 ?m silicate feature. Column 5: total equivalent width for the 9.7 ?m silicate feature. Column 6: S9.7?m ; see Section 2.5.2.2. 57 2.6.2 The S9.7?m Feature Figure 2.4 plots the apparent strength of the 9.7 ?m silicate absorption feature versus the AGN fraction ?AGN. Here, larger, more positive values of S9.7?m indicate stronger silicate absorption features. We see no discernible trend between the strength of this silicate feature and AGN fraction. This is surprising since the AGN fractions should be inversely proportional to the equivalent widths of the PAH features (Veilleux et al., 2009). Thus, Figure 2.4 should look similar to Figure 1 of Spoon et al. (2007) which plots the 6.2 ?m PAH emission feature and S9.7?m . We find that our data do not show the bifurcation observed in Spoon et al. (2007) because only two of our ULIRGs (F08572+3915 and F15250+3608) lie on the upper branch of the bifurcation, and none of our BAT AGN populate the upper branch. The two objects in Figure 2.4 which display ?AGN ? 10% and high S9.7?m are NGC 4945 and F15327+2340. Both of these objects show weak PAH emission and are therefore not displayed in Figure 1 of Spoon et al. (2007). Figure 2.5 shows a positive correlation between the measured OH equivalent width and the 9.7 ?m silicate optical depth. The black diagonal line shows the ordinary least squares bisector linear regression for the entire sample (BAT AGN + ULIRGs + PG QSOs): EWOH = 56.81? S ?19.7?m ? 26.7 [km s ]. (2.9) The Pearson correlation coefficient of the linear relationship for S9.7?m and the 58 total OH equivalent width is ?S9.7?m,EW = 0.7 with a P [null] (the probabil-OH ity of an uncorrelated system producing datasets that have a Pearson correlation at least as extreme as the one computed from these datasets) of 3.9 ? 10?9 for the combined sample. If we restrict this analysis to the BAT AGN only, we find ?S9.7?m,EW (BAT AGN) = 0.89 with a P [null] of 1.2? 10?9. Restricting the analy-OH sis to ULIRGs and PG QSOs only yields ?S9.7?m,EW (ULIRG/PG QSO) = 0.5 withOH a P [null] of 0.003. In Figure 2.5 we also see that objects with OH in emission show either weak silicate absorption or silicate emission. Objects with OH P-Cygni profiles show moderate silicate absorption features, while the strongest silicate absorption features are seen in objects in which OH is observed purely in absorption. We return to this result in Section 3.5. 2.6.3 OH Kinematics We quantify the visual trends observed (or not observed) in our kinematic investigation of the individual samples (i.e. BAT AGN only or ULIRG/QSO only) and of the combined sample by calculating the correlational significances between the observed OH velocities (v50 and v84) and the host galaxy properties (stellar masses, SFRs, specific star formation rates (sSFRs; i.e. the rate at which stars are formed divided by the stellar mass of the galaxy), and AGN fractions and luminosities. Since the spatial location of the OH emission is unknown, physical interpretations of the observed OH velocities in these cases will be ambiguous (e.g., blueshifted 59 velocities may correspond to outflow or inflow if the OH emission region is located in front or behind the continuum source, respectively). We therefore exclude from our analysis objects in which OH is detected purely in emission. The results of our statistical analyses on all objects with either redshifted or blueshifted absorption profiles are listed in Table 2.7. Section 2.6 shows the distributions of velocities derived from both the OH absorption and emission line features (v50(abs), v84(abs), v50(emi), and v84(emi), as defined in Section 2.5.1.3 and listed in Table 3.3). In V13, we adopted Rupke et al. (2005a)?s conservative definition of an outflow as having an OH absorption feature with a median velocity (v50) more negative than ?50 km s?1. Similarly, we can define an inflow as having an OH absorption feature with a median velocity (v50) more positive than 50 km s ?1. This cutoff is used to avoid contamination due to systematic errors and measurement errors in wavelength calibration, line fitting (see Section 2.5.1.3), and redshift determination. We find that only two objects in our BAT AGN sample meet this outflow criterion, and just barely: NGC 7172 and NGC 7479 (both have v ?150 = ?51 km s ). The presence of a significant blue wing in the absorption profile of OH 119 ?m in NGC 7479 with v84 = ?658 km s?1 adds considerable support to this interpretation. An extended blue wing with v84 < ?300 km s?1 may also be present in the OH absorption profile of IC 5063, NGC 5506, and NGC 7172, although these detections are more tentative than in NGC 7479. In contrast, seven objects have OH absorption features with median velocities larger than +50 km s?1: Centaurus A, Circinus, NGC 1125, NGC 3079, NGC 3281, NGC 4945, and NGC 7582. The clear detection of an inverted P-Cygni profile in Circinus 60 provides unambiguous evidence for the presence of an inflow in this object. The inverted P-Cygni profiles in Centaurus A and NGC 3281 are much less secure. We return to these objects in Section 3.5. 61 Table 2.5. S9.7?m Continuum Parameters and Measured Values of the ULIRG/PG QSO Sample (Name) Pivot Points Integration Range FluxS EQW9.7?m S S9.7?m 9.7?m (?m) (?m) (Jy ?m) (?m) (1) (2) (3) (4) (5) (6) 07251?0248 5.2, 14.0, 27.0, 31.5 8.45, 12.7 0.21 3.4 2.37 09022?3615 5.2, 5.6, 14.0, 27.0, 31.5 8.3, 12.6 0.43 1.84 1.22 13120?5453 5.2, 5.6, 14.0, 27.0, 31.5 8.7, 12.0 0.5 1.72 1.49 19542+1110 5.2, 5.6, 14.0, 27.0, 31.5 8.74,11.14 0.05 1.12 0.86 F00509+1225 5.2, 5.6, 14.0, 27.0, 31.5 7.8, 14.0 0.31 0.82 ?0.24 F01572+0009 5.2, 5.6, 14.0, 27.0, 31.5 8.1, 14.0 0.06 0.75 ?0.17 F05024?1941 5.2, 5.6, 10.8, 18.5 7.8, 12.6 0.01 1.65 0.22 F05189?2524 5.2, 5.6, 14.0, 27.0, 31.5 8.37, 14.0 0.54 1.01 0.37 F07598+6508 5.2, 5.6, 14.0, 27.0, 31.5 7.2, 14.0 0.11 0.49 ?0.15 F08572+3915 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 13.0 2.55 3.39 4.00 F09320+6134 5.2, 5.6, 14.0, 27.0, 31.5 8.65, 12.1 0.33 1.91 1.58 F10565+2448 5.2, 5.6, 14.0, 27.0, 31.5 8.76, 11.17 0.16 1.21 1.01 F11119+3257 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 12.6 0.17 1.21 0.62 F12072?0444 5.2, 5.6, 14.0, 27.0, 31.5 8.1, 12.6 0.22 2.36 1.41 F12112+0305 5.2, 5.6, 14.0, 27.0, 31.5 8.6, 12.4 0.13 2.42 1.58 F12243?0036 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 8.15, 12.8 4.73 3.95 4.03 F12265+0219 5.2, 5.6, 14.0, 27.0, 31.5 8.8, 14.0 0.08 0.26 ?0.06 F12540+5708 5.2, 5.6, 14.0, 27.0, 31.5 8.4, 12.6 2.62 1.49 0.62 F13305?1739 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? F13428+5608 5.2, 5.6, 14.0, 27.0, 31.5 8.2, 12.6 0.73 2.77 2.02 F13451+1232 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 12.2 0.06 0.61 0.35 F14348?1447 5.2, 5.6, 14.0, 27.0, 31.5 8.3, 12.4 0.15 2.66 1.93 F14378?3651 5.2, 5.6, 14.0, 27.0, 31.5 8.7, 12.1 0.06 1.76 1.55 F14394+5332 5.2, 5.6, 14.0, 27.0, 31.5 8.5, 12.3 0.18 2.26 1.71 F15206+3342 5.2, 5.6, 14.0, 27.0, 31.5 9.0, 11.1 0.02 0.47 0.29 62 Table 2.5 (cont?d) (Name) Pivot Points Integration Range FluxS EQW9.7?m S S9.7?m 9.7?m (?m) (?m) (Jy ?m) (?m) (1) (2) (3) (4) (5) (6) F15250+3608 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 14.0 0.94 3.54 3.42 F15327+2340 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 12.6 2.87 3.57 3.49 F15462?0450 5.2, 5.6, 14.0, 27.0, 31.5 8.6, 12.4 0.06 0.78 0.29 F16504+0228 5.2, 5.6, 14.0, 27.0, 31.5 8.65, 12.2 0.72 2.06 1.54 F17208?0014 5.2, 5.6, 14.0, 27.0, 31.5 8.6, 12.4 0.37 2.38 1.96 F19297?0406 5.2, 5.6, 14.0, 27.0, 31.5 8.6, 12.5 0.12 2.48 1.60 F20551?4250 5.2, 5.6, 7.8, 14.0, 27.0, 31.5 7.8, 14.0 1.52 3.43 3.09 F22491?1808 5.2, 5.6, 14.0, 27.0, 31.5 8.65, 12.4 0.07 2.2 1.14 F23128?5919 5.2, 5.6, 14.0, 27.0, 31.5 8.65, 12.2 0.23 1.45 0.88 F23233+2817 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? F23365+3604 5.2, 5.6, 14.0, 27.0, 31.5 8.65, 12.4 0.17 2.49 1.78 F23389+0300 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? PG 1126?041 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? PG 1351+640 5.2, 5.6, 14.0, 27.0, 31.5 8.1, 14.0 0.24 2.31 ?0.58 PG 1440+356 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? PG 1613+658 5.2, 5.6, 14.0, 27.0, 31.5 7.2, 14.0 0.03 0.38 ?0.01 PG 2130+099 5.2, 5.6, 14.0, 27.0, 31.5 7.75, 14.0 0.05 0.29 0.10 Z11598?0114 ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? Note. ? Column 1: galaxy name. Column 2: Pivot points in microns used for the continuum inter- polation. Column 3: Integration range of the 9.7 ?m silicate feature. Column 4: Total integrated flux of the 9.7 ?m silicate feature. Column 5: total equivalent width for the 9.7 ?m silicate feature. Column 6: S9.7?m ; see Section 2.5.2.2. 63 Table 2.6. Results from Statistical Analyses of Host Galaxy Properties Parameter Number of Objects ?s P? ? P? r Pr (1) (2) (3) (4) (5) (6) (7) (8) log M? - EWOH (ULIRGs Only) 21 ?0.49 2.5e?02 ?0.34 3.3e?02 ?0.56 8.4e?03 log M? - EWOH (BAT AGN only) 18 0.04 8.8e?01 0.04 8.2e?01 ?0.00 9.9e?01 log M? - EWOH (Combined Sample) 39 0.13 4.3e?01 0.09 4.0e?01 0.13 4.2e?01 ?AGN - EWOH (ULIRGs Only) 37 ?0.42 1.1e?02 ?0.29 1.0e?02 ?0.47 3.7e?03 ?AGN - EWOH (BAT AGN only) 38 ?0.11 5.2e?01 ?0.08 5.0e?01 ?0.27 1.0e?01 ?AGN - EWOH (Combined Sample) 75 ?0.42 1.8e?04 ?0.29 2.3e?04 ?0.44 7.0e?05 log LAGN - EWOH (ULIRGs Only) 37 ?0.25 1.3e?01 ?0.19 9.7e?02 ?0.31 6.4e?02 log LAGN - EWOH (BAT AGN only) 42 0.03 8.5e?01 0.02 8.6e?01 ?0.03 8.7e?01 log LAGN - EWOH (Combined Sample) 79 0.26 2.0e?02 0.18 2.1e?02 0.25 2.4e?02 Note. ? Column 1: quantities considered for the statistical test. Column 2: number of objects in which OH 119 ?m is detected. Column 3: Spearman rank order correlation coefficient. Column 4: null probability of the Spearman rank order correlation coefficient. Column 5: Kendall?s correlation coefficient. Column 6: null probability of Kendall?s correlation. Column 7: Pearson?s linear correlation coefficient. Column 8: Two-tail area probability of Pearson?s linear correlation. Null probabilities . 10?3 (shown in bold-faced characters) indicate statistically significant trends. 64 Table 2.7. Results from Statistical Analyses of the Kinematics Parameter Number of Objects ?s P? ? P? r Pr (1) (2) (3) (4) (5) (6) (7) (8) ULIRGs Only logM? ? v50 18 -0.15 5.6e-01 -0.15 3.8e-01 -0.04 8.7e-01 logM? ? v84 18 -0.10 7.0e-01 -0.08 6.4e-01 0.01 9.6e-01 logSFR? v50 32 0.28 1.2e-01 0.22 8.2e-02 0.06 7.6e-01 logSFR? v84 32 0.19 2.9e-01 0.14 2.4e-01 0.07 7.1e-01 sSFR ? v50 18 0.37 1.3e-01 0.26 1.3e-01 0.54 2.2e-02 sSFR ? v84 18 0.26 3.0e-01 0.21 2.3e-01 0.57 1.4e-02 ?AGN ? v50 32 -0.52 2.4e-03 -0.37 3.2e-03 -0.48 5.5e-03 ?AGN ? v84 32 -0.46 7.4e-03 -0.31 1.2e-02 -0.46 7.6e-03 logLAGN ? v50 32 -0.50 3.3e-03 -0.37 2.8e-03 -0.56 8.0e-04 logLAGN ? v84 32 -0.44 1.1e-02 -0.32 9.8e-03 -0.54 1.6e-03 BAT AGN Only logM? ? v50 6 -0.23 6.6e-01 -0.14 7.0e-01 -0.34 5.1e-01 logM? ? v84 6 0.09 8.7e-01 -0.07 8.5e-01 0.27 6.1e-01 logSFR? v50 13 -0.11 7.3e-01 -0.11 6.2e-01 -0.04 8.9e-01 logSFR? v84 13 0.24 4.4e-01 0.17 4.2e-01 -0.02 9.5e-01 sSFR ? v50 6 0.32 5.4e-01 0.28 4.4e-01 0.17 7.4e-01 sSFR ? v84 6 0.54 2.7e-01 0.33 3.5e-01 0.34 5.1e-01 ?AGN ? v50 13 -0.50 8.4e-02 -0.32 1.3e-01 -0.50 8.1e-02 ?AGN ? v84 13 -0.69 9.4e-03 -0.50 1.7e-02 -0.44 1.3e-01 logLAGN ? v50 14 -0.27 3.6e-01 -0.18 3.7e-01 -0.39 1.7e-01 logLAGN ? v84 14 -0.36 2.1e-01 -0.27 1.9e-01 -0.29 3.2e-01 Combined Sample logM? ? v50 24 -0.58 2.8e-03 -0.44 2.5e-03 -0.13 5.5e-01 logM? ? v84 24 -0.53 7.9e-03 -0.38 9.9e-03 -0.08 7.0e-01 logSFR? v50 45 -0.34 2.3e-02 -0.21 3.9e-02 -0.44 2.3e-03 logSFR? v84 45 -0.30 4.3e-02 -0.18 7.8e-02 -0.40 6.8e-03 65 Table 2.7 (cont?d) Parameter Number of Objects ?s P? ? P? r Pr (1) (2) (3) (4) (5) (6) (7) (8) sSFR ? v50 24 0.13 5.6e-01 0.10 4.8e-01 0.34 1.1e-01 sSFR ? v84 24 0.10 6.3e-01 0.09 5.2e-01 0.38 6.7e-02 ?AGN ? v50 45 -0.04 8.1e-01 -0.03 7.8e-01 -0.08 6.1e-01 ?AGN ? v84 45 -0.09 5.6e-01 -0.06 5.7e-01 -0.12 4.5e-01 logLAGN ? v50 46 -0.70 6.8e-08 -0.51 5.7e-07 -0.67 3.3e-07 logLAGN ? v84 46 -0.61 8.2e-06 -0.45 8.9e-06 -0.63 2.9e-06 Note. ? Column 1: quantities considered for the statistical test. Column 2: number of objects in which OH 119 ?m is detected in either redshifted or blueshifted absorption. Column 3: Spearman rank order correlation coefficient. Column 4: null probability of the Spearman rank order correlation coefficient. Column 5: Kendall?s correlation coefficient. Column 6: null probability of Kendall?s correlation. Column 7: Pearson?s linear correlation coefficient. Column 8: Two-tail area probability of Pearson?s linear correlation. Null probabilities . 10?3 (shown in bold-faced characters) indicate statistically significant trends. Figure 2.7, Figure 2.8, and Figure 2.9 plot the OH velocities (v50 and v84) as a function of the stellar masses, star formation rates, and specific star formation rates. If we consider our two samples individually, a visual inspection of these plots does not show a correlation between these properties. However, once the samples are combined we see a negative correlation between the observed OH velocities and the stellar mass or star formation rate of the galaxy (i.e. galaxies with larger stellar masses or star formation rates exhibit more negative v50 and v84 values). The strengths of these correlations are quantified in Table 2.7. No obvious trend is seen with sSFR for the combined sample. These results remain quantitatively the same if instead of using the global star formation rates we use the star formation rates from the central spaxel (scaled by the ratio of the continuum flux from the central 66 400 (a) 0 ?400 Abs (BAT AGN) ?800 Abs (ULIRG) Abs/Emi (ULIRG) 400 (b) 0 ?400 NGC 5506 ?800 9 10 11 12 13 log(M?/M ) Figure 2.7 v50 and v84 as a function of the stellar masses. The meanings of the symbols are the same as those in Section 2.5.1.1. The data points joined by a segment correspond to F14394+5332 W and E. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) . spaxal to that of all 25 spaxels, as listed in column (8) of Table 2.2). Figure 2.10 plots the OH velocities (v50 and v84) as a function of the AGN fractions ?AGN, where ?AGN is derived from the f30/f15 continuum flux density ratios. As described in V13, ULIRGs/PG QSOs with dominant AGN (?AGN ? 50%) appear to have larger negative velocities than ULIRGs/PG QSOs with dominant starbursts (?AGN ? 50%), but a K-S test between velocity distributions of dominant AGN and dominant starburst systems indicated no statistically significant difference. A K-S 67 v84 [km s ?1] v ?150 [km s ] 400 (a) 0 ?400 Abs (BAT AGN) Abs/Emi (BAT AGN) Abs (ULIRG) ?800 Abs/Emi (ULIRG) 400 (b) 0 NGC 7172 ?400 IC 5063NGC 5506 NGC 7479 ?800 ?1 0 1 2 3 log(SFR/[M yr?1]) Figure 2.8 v50 and v84 as a function of the star formation rates. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) test on the combined BAT AGN and ULIRG + QSO sample also does not show a significant trend between the measured OH velocity and AGN fraction. Figure 2.11 shows the OH velocities (v50 and v84) versus the AGN luminosi- ties, LAGN = 10.5 ? L14?195keV for our BAT AGN sample and LAGN = ?AGNLBOL for the ULIRG/QSO sample; see Section 2.4. A correlation between these quan- tities is not observed in the BAT AGN sample, but clear trends are seen in the ULIRG/QSO sample and in the combined sample (see Table 2.7). Objects with log (LAGN/L ) .11.5 show no evidence for fast outflows. This suggests that AGN 68 v84 [km s ?1] v ?150 [km s ] 400 (a) NGC 3079 F22491-1808 0 NGC 4102 ?400 Abs (BAT AGN) Abs (ULIRG) ?800 Abs/Emi (ULIRG) 400 (b) F22491-1808 0 NGC 3079 NGC 4102 ?400 NGC 5506 ?800 ?1200 0 2 4 6 sSFR [Gyr?1] Figure 2.9 v50 and v84 as a function of the specific star formation rates. The meanings of the symbols are the same as those in Section 2.5.1.1. The data points joined by a segment correspond to F14394+5332 E and W. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) The black vertical line indicates the approximate location of the Main Sequence of star-forming galaxies (Shimizu et al., 2015). of lower luminosities are not able to drive significant molecular winds. 69 v84 [km s ?1] v50 [km s?1] 400 (a) 0 ?400 ?800 400 (b) 0 ?400 ?800 0 20 40 60 80 100 ?AGN[%] Figure 2.10 v50 and v84 as a function of the AGN fractions. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) 2.7 Discussion 2.7.1 Outflows 2.7.1.1 BAT AGN with Molecular Outflows As mentioned above, only (one) four objects in our BAT AGN sample shows (unambiguous) evidence of a molecular outflow (NGC 7479 is the unambiguous case, while NGC 5506, NGC 7172, and IC 5063 are more uncertain; see Figure 2.8 and 70 v84 [km s ?1] v50 [km s?1] 400 (a) 0 ?400 Abs (BAT AGN) Abs/Emi (BAT AGN) Abs (ULIRG) ?800 Abs/Emi (ULIRG) 400 (b) 0 NGC 7172 IC 5063 ?400 NGC 5506 NGC 7479 ?800 9 10 11 12 13 log(LAGN/L ) Figure 2.11 v50 and v84 as a function of the AGN luminosities. The meanings of the symbols are the same as those in Section 2.5.1.1. The smaller symbols have larger uncertainties (values followed by double colons in Table 3.3) Figure 2.11). This corresponds to an outflow detection rate of (6%) 24%, if we take into account that this search for outflows was possible only in the 17 sources with OH 119 ?m in absorption. This outflow detection rate is significantly smaller than that found in ULIRGs (?70%). We note, however, that outflow detection may be easier in the ULIRGs due to their higher gas fractions. NGC 7479 is the best case for an outflow in our sample of BAT AGN. NGC 7479 is a barred galaxy with the faint broad line emission at H? but not at H? of a Seyfert 1.9 galaxy (Ve?ron-Cetty & Ve?ron, 2006). To quantify the uncertainty 71 v84 [km s ?1] v ?150 [km s ] of the OH blue-wing detection in NGC 7479, we have refit the continuum for this object with a third order polynomial and then refit the OH doublet via the method outlined in Section 2.5.1.3. Although we find a shift in the velocities measured with this different continuum (v50(abs) = ?45 km s?1 and v84(abs) = ?513 km s?1), this shift (a measure of the error in our velocity estimates) is not large enough to account for the location of NGC 7479 in Figure 2.11b. The origin for the high outflow velocity in NGC 7479 may lie in its unusually high far-infrared surface density (Lutz et al., 2015). The high wind velocity observed in NGC 7479 may thus be due to its unusually high SFR surface density (e.g. Diamond-Stanic et al., 2012). However, we cannot exclude the possibility that the OH outflow in NGC 7479 is driven by the radio jet detected by Laine & Beck (2008). NGC 5506 is an edge-on disk galaxy optically classified as a Seyfert 1.9 galaxy, just like NGC 7479. In the very hard X-ray regime, NGC 5506 is one of the most luminous and brightest Seyferts in the local universe (L14? 43195 keV ? 10 erg s?1; Baumgartner et al., 2011), and its obscuring column (N = 3.4?1022H cm?2; Bassani et al., 1999) is intermediate between typical values for Seyfert 1s and 2s. Mid- infrared 8 ? 13?m observations of NGC 5506 (Roche et al., 1984) suggest that the nuclear region of NGC 5506 within 5 arcsec is powered by highly obscured AGN activity with little starburst activity. The OH 119 ?m line profile of NGC 5506 shows a very broad, secondary gaussian component, suggestive of an outflow (see Figure 2.1), but the velocities of this component are uncertain (see Table 3.3). Interestingly, the optical forbidden emission lines of NGC 5506 also exhibit distinct blue wings extending to ?1000 km s?1 (Veilleux, 1991a,b,c), giving credence to the 72 idea that a fast outflow is indeed present in this object. IC 5063 is a massive (M? ? 1011 M ) early-type galaxy which hosts a pow- erful double-lobed radio source (P 23 ?11.4GHz = 3 ? 10 W Hz ). The presence of this strong radio source is contrary to the weakly collimated jets found in the BAT AGN of Middelberg et al. (2004). Observations of high velocity (? 600 km s?1) warm molecular hydrogen gas in the western lobe of IC 5063 suggests that molecules may be shock-accelerated by the expanding radio jets (Tadhunter et al., 2014). Thus, the presence of this powerful jet may explain the additional boost observed in the (uncertain) outflow velocity v84. NGC 7172 is an obscured (N 23 ?2H ? 10 cm ; Turner & Pounds, 1989), almost edge-on Type 2 Seyfert galaxy which belongs to the Hickson compact group HCG 90. Smajic? et al. (2012) derive the two-dimensional velocity field and velocity dis- tribution of the central 4??? 4?? region of NGC 7172 using several emission lines (i.e. Pa?, H2(1-0)S(1), and [Si VI]) and two CO stellar absorption features. All mea- surements indicate disk rotation from east to west with amplitude of at least ?100 km s?1. Comparison of HST F606W imaging (Malkan et al., 1998) with 2MASS images in J?, H?, and K?bands (Jarrett et al., 2003) shows a shift of the nucleus by more than 1?? going from the visible to the J?, H?, and K?band. The central spaxel aperture (9.4??? 9.4??) of Herschel does not have the resolution to detect such a shift, but if the central spaxel is not centered on the galaxy nucleus, and if the velocity structure for OH 119 ?m is comparable to that seen in the near-infrared, then it is possible that the blueshifted absorption we detect is due to incomplete sampling of the rotation curve rather than a molecular outflow. 73 This last object emphasizes the fact that the poor spatial resolution of Herschel makes the interpretation of the OH spectra challenging. We have tried to err on the side of being cautious by utilizing a centroid velocity threshold of ?50 km s?1 and/or the detection of blue wings with v ?184 < ?300 km s . The Herschel data are not always sensitive to small-scale winds. Indeed, small-scale molecular winds are known to exist in some of our sources. For instance, Garc??a-Burillo et al. (2014) detect AGN-driven CO(3-2), CO(6-5), HCN(4-3), HCO+(4-3), and CS(7-6) outflows in NGC 1068 on spatial scales of ? 20? 35 pc (? 0.3??? 0.5??), while OH is observed purely in emission in the Herschel data and is therefore ambiguous regarding the existence of a large-scale outflow until a full analysis of the velocity field is carried out. Additionally, molecular outflows (or inflows) may be present, but not centered on the central spaxel. For example, Herschel observations of OH 79, 84, and 65 ?m, as well as HCN suggest the existence of a spatially resolved outflow outside of the central spaxel in NGC 3079 even though the central spaxel apparently displays an OH inflow. A detailed discussion of these objects is outside the scope of this work, but these issues will be addressed in a future paper. 2.7.1.2 Driving Mechanisms of Molecular Outflows Correlations between the observed OH outflow velocities and the stellar masses, star formation rates, and AGN luminosities of the host galaxies can shed light upon the physical mechanisms responsible for driving molecular winds. There are good reasons to believe that energy and momentum injection from star formation ac- 74 tivity plays a role in driving massive, galactic-scale outflows. For example, the observed spatial coincidence between H-? filaments and the X-ray emission from galactic winds suggests a close physical connection between the origins of these fea- tures (Cooper et al., 2008). Additionally, previous studies (e.g. Cicone et al., 2014; Martin, 2005; Rupke et al., 2005a; Schwartz & Martin, 2004; Weiner et al., 2009) have shown the existence of a positive trend between SFRs and outflow velocities. Curiously, V13 found no such trend in their ULIRGs + QSO sample. It appears, as V13 surmised, that the lack of a trend between these quantities was due to the limited range in SFR of the V13 sample, which spans only ? 2 dex. As seen in Figure 2.8, when the BAT AGN are combined with the ULIRGs of V13, the sample spans a range ? 3 dex in SFR, and a positive correlation between outflow velocities and SFRs becomes apparent. Positive correlations between stellar masses and outflow velocities in the neu- tral and ionized phases of the ISM have also been reported in the literature (e.g Martin, 2005; Rupke et al., 2005a; Weiner et al., 2009). V13 did not see this trend in the molecular data of their ULIRG + QSO sample, but again attributed this negative result to the small range in stellar masses of their sample (? 1 dex). Our BAT AGN + ULIRG + QSO sample now extends over a stellar mass range of ? 2 dex and reveals a significant trend (see Figure 2.7). However, note that this trend may be of secondary origin since a positive correlation is well known to exist be- tween stellar mass and SFR in galaxies, both locally (e.g. 0.015 ? z ? 0.1; Elbaz et al., 2007; Renzini & Peng, 2015; Shimizu et al., 2015) and at high redshift (e.g. 0.7 < z < 1.5; Rubin et al., 2010; Whitaker et al., 2014) ? the so-called Main 75 Sequence of star-forming galaxies. V13 reported a weak trend between wind velocities (v50 and v84) and AGN fractions in their ULIRG sample. They cautioned that the correlation could be merely an obscuration effect where both the AGN and central high-velocity out- flowing material are more easily detectable when the dusty material has been swept away or is seen more nearly face-on. Indeed, by adding our BAT AGN sample to the ULIRG sample, we find that this weak correlation disappears. AGN frac- tions are thus not a good predictor of molecular outflows. On the other hand, a convincing causal connection was presented in V13 between AGN luminosities and molecular outflow velocities with a possible steepening of the relation above log(LbreakAGN /L ) = 11.8 ? 0.3. Our results do indeed strongly suggest that at higher AGN luminosities (log(LAGN/L ) & 11.5) the AGN dominates over star formation in driving the outflow. At lower luminosities (log(LAGN/L ) . 11.5) the AGN may not have the energetics required to drive fast molecular winds. Statistically, the correlation between wind velocity and AGN luminosity is stronger than the correlation between wind velocity and SFR (see Table 2.7). How- ever, while the AGN may be the dominant source for driving the winds observed in our sample, it is clear that SFR also plays a role in driving these winds. The presence of an AGN seems to boost the observed velocity over that which would be observed in purely star forming systems (Cicone et al., 2014), although large scatter is observed (see Figure 2.11). This scatter may have multiple origins: (1) If the outflow is not spherically symmetric, projection effects will produce scatter in the observed velocities. (2) The efficiency to entrain material in the outflow depends on 76 several complex factors associated with the acceleration mechanisms and the multi- phase nature of these processes. For instance, if radiation pressure is the dominant mechanism driving the outflow, radiative transfer effects are probably important so that not all of the OH-detected gas is ?seeing? the same radiation field and there- fore not experiencing the same radiation force. Similarly, if the dominant driving mechanism is ram pressure of a fast diffuse medium on dense cloudlets, one might expect to observe a range of velocities depending on the characteristics (e.g., sizes and densities) of the cloudlets entrained in the flow and their survival timescales (e.g. Cooper et al., 2008, 2009). (3) The AGN luminosity is measured from the 14 ? 195 keV flux and therefore represent the current value of the AGN luminosity. In contrast, the observed outflow was likely produced ? 1 to several ?106 years ago based on the measured OH velocities and inferred outflow sizes (? 1 kpc; Gonza?lez- Alfonso et al. (2014b, 2015); Sturm et al. (2011)). AGN variability may therefore cause considerable scatter in Figure 2.11. 2.7.2 Inflows Seven objects (Centaurus A, Circinus, NGC 1125, NGC 3079, NGC 3281, NGC 4945, and NGC 7582) have OH absorption features with median velocities larger than 50 km s?1, corresponding to an inflow detection rate of ?40%. By far the best case for inflow in our sample is Circinus where OH 119 ?m shows an inverted P-Cygni profile. Inverted P-Cygni profiles are also tentatively detected in Centaurus A and NGC 3281. 77 Interestingly, previous searches for neutral-gas (Na I D) outflows/inflows in IR-faint Seyfert galaxies have also shown distinctly higher detection rates of inflows than outflows. Specifically, in an analysis of 35 IR-faint Seyferts (109.9 < LIR/L < 1011.2), Krug et al. (2010) reported an inflow detection rate of ? 37% and an outflow detection rate of only ? 11%. These numbers are similar to those derived here, and considerably different from those measured in (U)LIRGs using OH (? 10%; V13) and Na I D (e.g. ? 15% inflow detection rate for 78 starbursting galaxies with log(LIR/L ) = 10.2 ? 12.0; Rupke et al., 2005a,b). The origin of this difference is unknown. The fast winds in (U)LIRGs may disturb the neutral-molecular gas and prevent it from infalling to the center. On average, IR-bright sources are also richer in gas and dust than IR-faint galaxies. This material may be masking the central regions where inflow is taking place. 2.7.3 The 9.7 ?m Silicate Feature The analysis of the strength of the S9.7?m feature can provide insight into the mechanism responsible for the excitation of OH 119 ?m observed in our sample. As seen in Figure 2.5, the comparison of the OH 119 ?m equivalent width, EWOH, and the strength of the silicate feature, S9.7?m , implies a rather tight connection between OH gas and mid-infrared obscuration (a correlation is also found between OH 65 ?m and S9.7?m ; see also Gonza?lez-Alfonso et al. (2015)). The clear trend of deeper OH 119 ?m absorption and fainter OH 119 ?m emission with increasing silicate obscuration (more positive values of S9.7?m ) suggests that OH 119 ?m emission is 78 strongly affected by the obscuring geometry. Our results expand on those of V13 and S13 who found a similar trend with OH equivalent width and the strength of the 9.7 ?m feature among ULIRGs. S13 argue that the OH 119 ?m emission region often lies within the buried nucleus and that radiative excitation is the dominant source of OH 119 ?m emission. In reality, the geometry of the silicate obscuration and the source of the OH 119 ?m emission may be more complex. Figure 2.12 plots the total equivalent width of OH 119 ?m as a function of S9.7?m and distinguishes between AGN spectral type for each object. Note that objects classified as LINERs have been excluded from this plot due to the ambiguous energy source in these objects (Sturm et al., 2006). None of the Type 1 galaxies (BAT AGN or ULIRGs) show a strong silicate absorption (S9.7?m & 1.5). Interestingly, we see that OH 119 ?m for some Type 2 BAT AGN and ULIRGs is observed in emission (EWOH < 0 km s ?1) while the silicate feature, S9.7?m , is seen in absorption. It is possible that the OH 119 ?m emission region is not nuclear, but is instead distributed throughout a circumnuclear starburst where the number density (n(H2) ? a few times 105 cm?3), temperature (? 100 K), and OH abundance (X(OH) ? 2?10?6) are sufficient for collisional rather than radiative excitation of the upper level of OH 119 ?m (e.g. NGC 1068; Spinoglio et al., 2005). Until we examine in detail all 5 ? 5 spaxels of the PACS data, the location of the OH emission in our objects will remain unclear. This type of analysis can be done in just a few select objects; this will be discussed in a future paper. Determining the origin of the silicate feature is also a challenge. The dust responsible for this feature may reside in the AGN torus, or on larger scale in the 79 300 200 100 0 BAT AGN Type 1 ? BAT AGN Type 2100 ULIRG Type 1 ULIRG Type 2 ?1 0 1 2 3 4 5 S9.7?m Figure 2.12 Total (absorption + emission) equivalent widths of OH 119 ?m as a function of the apparent strength of the 9.7 ?m silicate feature relative to the local mid-infrared continuum. Note that S9.7?m is a logarithmic quantity and can be interpreted as the apparent silicate optical depth. The strength of this absorption feature increases to the right. Also note that objects classified as LINERs have been excluded from this plot. Filled markers refer to Type 1 and open markers refer to Type 2. Blue squares and red circles represent BAT AGN and ULIRGs/PG QSOs, respectively. Vertical lines represent objects in which OH was not detected. Horizontal lines represent objects with a null S9.7?m detection. Dotted lines and dash-dotted lines refer to Type 1 and Type 2, respectively. Blue and red lines indicate BAT AGN and ULIRGs/PG QSOs, respectively. disk of the host galaxy, or some combination of these two. Figure 2.13 plots the ratio of the semi-minor and semi-major axes (a proxy for the inclination of the host galaxy disk) as a function of S9.7?m for BAT AGN. We have excluded the ULIRG/PG QSO sample from this particular analysis because these objects are undergoing or have undergone a major merger, and therefore, do not have a well-defined galactic plane or inclination. Visual inspection of Figure 2.13 suggests a weak trend between the inclination of the BAT AGN host galaxy and the depth of the 9.7 ?m silicate feature. First, we see that OH for nearly face-on 80 EW ?1OH [km s ] 1.0 Abs (BAT AGN) Type 1 Abs (BAT AGN) Type 2 Emi (BAT AGN) Type 1 Emi (BAT AGN) Type 2 Abs/Emi (BAT AGN) Type 1 Abs/Emi (BAT AGN) Type 2 Null (BAT AGN) Type 1 Null (BAT AGN) Type 2 0.5 NGC 4945 0.?0 1 0 1 2 3 4 5 6 S9.7?m Figure 2.13 Ratio of the semi-minor axis to the semi-major axis (a proxy for the inclination of the host galaxy disk) as a function of S9.7?m for the BAT AGN sam- ple. Squares, triangles, and circles represent BAT AGN in which OH is observed purely in absorption, purely in emission, composite absorption/emission, respec- tively. Diamonds represent objects in which OH was undetected. Filled points and dash-dotted lines indicate Type 1 while open points and dotted lines indicate Type 2 AGN. Horizontal lines represent objects with a null 9.7 ?m silicate feature detection. hosts (b/a & 0.8) is seen only in pure emission and the strength of S9.7?m is weak. Second, we find that nearly edge-on galaxies show the broadest range of silicate absorption strengths, including the most extreme case of NGC 4945. However, the lack of a strong trend between inclination and S9.7?m strength suggests that the dust responsible for the 9.7 ?m silicate feature is located not only in the plane of the host galaxy, but also in the nuclear torus. Our results are qualitatively consistent with Goulding et al. (2012), who invoke a clumpy torus paradigm of many individual optically thick clouds (Nenkova et al., 2002, 2008) and suggest that deeper silicate features (S9.7?m & 0.5) are due to dust distributed at radii much larger ( pc) than that predicted for a torus. 81 b/a 2.8 Conclusions We present the results of our analysis of Herschel/PACS spectroscopic obser- vations of 52 nearby (d < 50 Mpc) BAT AGN selected from the very hard X-ray (14-195 keV) Swift-BAT Survey of local AGN. We also include in our analysis the Herschel/PACS data on 38 ULIRGs and 5 PG QSOs from V13. The depth of the silicate feature at 9.7 ?m feature in all these objects is measured from archival Spitzer/IRS data. Our combined BAT AGN + ULIRG + QSO sample covers a range of AGN luminosity and SFR of & 3 dex and ? 3 dex, respectively. The main results from our analysis are: 1. The OH 119 ?m feature is detected in 42 of the 52 objects in our BAT AGN sample. Of these detections, OH 119 ?m is observed purely in emission for 25 targets, purely in absorption for 12 targets, and absorption+emission com- posite for 5 targets. 2. Evidence for molecular outflows (absorption profiles with median velocities more blueshifted than ?50 km s?1 and/or blueshifted wings with 84-percentile velocities less than ?300 km s?1) is seen in only four objects (NGC 7479, NGC 5506, NGC 7172, and IC 5063), corresponding to 24% of all targets where this search for outflows was possible. This outflow detection rate is significantly smaller than that found in ULIRGs by V13 (?70%). The best case for outflow is NGC 7479, an object with one of the highest infrared surface densities among our BAT AGN. 82 3. We find evidence for molecular inflows (absorption profiles with median veloc- ities more redshifted than 50 km s?1) in seven objects (Circinus, NGC 1125, NGC 3079, NGC 3281, and NGC 4945), corresponding to an inflow detection rate of ?40%, considerably higher than the rate measured among ULIRGs (?10%) but similar to the detection rate of neutral-gas (Na I D) inflows among IR-faint Seyfert galaxies. By far the best case for OH inflow among our BAT AGN is Circinus, where OH 119 ?m shows a distinct inverted P-Cygni profile. 4. The positive correlation between OH velocities and AGN luminosities reported in V13 is strengthened in the combined sample, but it is not seen in the BAT AGN sample. This suggests that luminous AGN play a dominant role in driving the fastest winds, but stellar processes will dominate when the AGN is weak or absent. 5. We confirm that the absorption strength of OH 119 ?m is a good proxy for dust optical depth in these systems. Our findings are consistent with most, but not all, of the OH 119 ?m emission originating near the AGN. Spatially resolved OH emission is seen in a few objects in our sample (e.g., NGC 1068) and may originate from a circumnuclear starburst where the number density and temperature are sufficiently high to collisionally excite the upper level of OH 119 ?m. 6. A comparison of the strength of the 9.7 ?m silicate feature with the inclination of the host galaxy disk and spectral type of the AGN confirm earlier results that the dust responsible for this feature is located not only in the nuclear 83 component, but also in the disk of the host galaxy. 84 Chapter 3: Constraints on the OH-to-H Abundance Ratio in Infrared- Bright Galaxies Derived from the Strength of the OH 35 ?m Absorption Feature 3.1 Introduction Galactic-scale outflows are driven by the collective effects of supernovae, winds from massive stars, and winds or jets from active galactic nuclei (AGN; e.g. Veilleux et al., 2005, and references therein). These outflows can disperse heavy elements, interstellar dust, and energy throughout a galaxy and into the galaxy halo; they may therefore play a significant (if not dominant) role in galaxy evolution (e.g. Benson et al., 2003; Bower et al., 2006; Croton et al., 2006; Fabian, 2012; Hopkins et al., 2008, 2006; King, 2010; Springel, 2005). They have been detected ubiquitously in both low- and high-redshift systems (e.g. Alexander et al., 2010; Banerji et al., 2011; Brusa et al., 2016; Cano-D??az et al., 2012; Carniani et al., 2015, 2016; Cicone et al., 2015; Cresci et al., 2015; Fo?rster Schreiber et al., 2014; Genzel et al., 2014; Greene et al., 2012; Harrison et al., 2014, 2012, 2016; Kornei et al., 2012; Maiolino et al., 2012; Martin et al., 2012; McElroy et al., 2015; Nesvadba et al., 2011; Tremonti et al., 2007; Veilleux et al., 2005; Weiner et al., 2009; Zakamska et al., 2016). They are 85 particularly significant in infrared luminous galaxy mergers where both starbursts and luminous AGN often co-exist together (e.g. Cicone et al., 2014; Gonza?lez-Alfonso et al., 2017; Janssen et al., 2016; Martin, 2005, 2006; Rupke et al., 2017, 2002, 2005a,b,c; Rupke & Veilleux, 2011; Spoon et al., 2013; Sturm et al., 2011; Veilleux et al., 2017, 2013). However, the central regions of these mergers are generally enshrouded in dust so infrared tracers are needed to peer through the dusty veil. Numerous studies have demonstrated that the infrared transitions of the hy- droxyl molecule, OH, are a powerful diagnostic tool with which to investigate the nature of cool molecular outflows. Fischer et al. (2010) first reported the remarkable detection of OH 79 and 119 ?m (hereafter, the OH ??doublet transitions will be referred for short only by their rounded wavelengths: OH 79 and OH 119) with P- Cygni profiles (incontrovertible evidence of the presence of an outflow) in the nearby QSO, Mrk 231. Results from Sturm et al. (2011) revealed the detection of massive molecular outflows traced by OH 79 for six galaxies and a tentative correlation be- tween OH outflow velocities and AGN luminosities. This correlation was confirmed by Veilleux et al. (2013) (and Spoon et al., 2013) who reported the detection of OH outflows in ?70% of 43 (24) ULIRGs + QSOs, and further strengthened when Stone et al. (2016) combined the 43 ULIRGs + QSOs of Veilleux et al. (2013) with OH 119 observations of 52 Swift/Burst Alert Telescope (BAT)-detected AGN. In order to evaluate the impact of these outflows on the evolution of their host galaxies, it is necessary to measure the sizes and energetics (e.g., mass-outflow rate, momentum flux, and mechanical power) of these outflows. Multi-transitional OH analyses have had considerable success constraining these quantities and the 86 dominant mechanisms driving these winds (Gonza?lez-Alfonso et al., 2014b, 2017). All of the dynamical quantities depend on the hydrogen column densities inferred from the OH features. These in turn depend on the OH?to?H abundance ratio, XOH, needed to infer the H column densities from the measured OH column den- sities. This abundance ratio is generally assumed to be ?2.5 ? 10?6, based on the value inferred from multitransition observations of OH in the Galactic Sgr B2, the Orion KL outflow, and in buried nuclei, as well as predictions of chemical models of photodissociation regions and of cosmic-ray- and X-ray-dominated regions (see Gonza?lez-Alfonso et al., 2017, and references therein). In this paper, we provide an independent constraint on XOH. We focus our at- tention on the OH 35 ?m (hereafter, OH 35) absorption feature covered by Spitzer- IRS. OH 35 is optically thinner than the other OH doublets in the far-infrared, and although it is more sensitive to extinction effects, OH 35 remains a reliable indicator of OH column densities. OH is assumed to be mixed with dust and given the high continuum optical depths (even in the far-IR; Gonza?lez-Alfonso et al., 2014b), the absorption is restricted to a ?skin? of matter. Our OH column densities therefore also provide a constraint on the abundance of OH relative to the dust, from which XOH (relative to H nuclei) is derived, assuming a gas?to?dust ratio. We are able to carry out this analysis on 15 local (z ? 0.06) (ultra-)luminous infrared galaxies (U/LIRGs), which have both Herschel-PACS observations of the OH 119 and/or 79 ?m transitions as well as Spitzer-IRS observations of the OH 35?m transition. Background information on the structure of the OH molecule is presented in Section 3.2. The galaxy sample, data reduction, and spectral analysis are described 87 Table 3.1. Spitzer -IRS Spectra: Observations Name PI AORKey(s) (1) (2) (3) Arp 220 Houck, James 16910080, 21247232 IRAS F04454-4838 Armus, Lee 20334080 IRAS F05189-2524 Calibration, IRS 04969216 IRAS F08572+3915 Calibration, IRS 20925696, 21457920, 24568064, 27381760, 28704256 IRAS F12224-0624 Bradford, Charles 11272960 IRAS 15250+3609 Houck, James 04983040 IRAS 17208-0014 Houck, James 04986624 IRAS F20551-4250 Houck, James 04990208 IRAS F23128-5919 Houck, James 04991744 Mrk 231 Calibration, IRS 16694016, 17102592, 17202688, 19318016, 21453568, 22157312, 24915456, 34294016 Mrk 273 Houck, James 04980224 NGC 2623 Houck, James 09072896 NGC 4418 Houck, James 04935168 UGC 5101 Houck, James 04973056 Zw453.062 Sturm, Eckhard 10512128 Note. ? Column 1: Galaxy name. Column 2: Principle Investi- gator. Column 3: AORKey. in Section 3.3. Section 3.4 presents the results of the OH 35 line profile fitting. Section 3.5 discusses the implications of our results. A summary of the results and conclusions is given in Section 4.10. 88 Table 3.2. Galaxy Properties Name Other Name za f15/f30 ?AGN logLbol logLSB logLAGN ?? MH Type [%] [L ] [L ] [L ] [km s ?1 ] mag (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) IRAS F04454-4838 0.0529 0.09b 42.0 11.84c 11.61 11.47 ? ? ? ? ? ? ? ? ? IRAS F05189-2524 0.0426 0.198 71.7 12.22 11.67 12.07 137 ?23.96 AGN 2 IRAS F08354+2555 NGC 2623 0.01846 0.07b 29.6 11.58c 11.05 11.43 ? ? ? ? ? ? L IRAS F08572+3915 0.0584 0.191 70.4 12.2 11.67 12.05 ? ? ? ?23.58 L IRAS F09320+6134 UGC 5101 0.0394 0.13 56.4 12.05 11.69 11.8 ? ? ? ? ? ? L IRAS F12224-0624 0.02636 0.04b 0 11.34c ? ? ? 11.34 ? ? ? ? ? ? L IRAS F12243-0036 NGC 4418 0.00727 0.128 55.7 11.06 10.71 10.81 ? ? ? ? ? ? AGN 2 IRAS F12540+5708 Mrk 231 0.0422 0.272 80.5 12.6 11.89 12.51 120 ?24.52 AGN 1 IRAS F13428+5608 Mrk 273 0.0378 0.081 34.2 12.21 12.03 11.74 285 ?24.32 AGN 2 IRAS 15250+3609 0.0554 0.095 42.3 12.1 11.86 11.73 150 ? ? ? L IRAS F15327+2340 Arp 220 0.0181 0.049 5.8 12.22 12.19 10.98 164 ? ? ? L IRAS 17208-0014 0.0428 0.038 < 5.0 12.45 12.43 < 11.15 229 ? ? ? L IRAS F20551-4250 0.043 0.132 56.9 12.11 11.74 11.87 140 ? ? ? L IRAS F23024+1916 Zw453.062 0.02510 0.07b 30.0 11.36c 10.84 11.21 ? ? ? ? ? ? L IRAS F23128-5919 ESO 148-IG 002 0.0446 0.154 63 12.09 11.66 11.89 ? ? ? ? ? ? H Note. ? Column 1: Galaxy name. Column 2: Another name. Column 3: Redshift. Column 4: f15/f30 values from Veilleux et al. (2009). Column 5: ?AGN, fractional contribution of the AGN to the bolometric luminosity based on the f15/f30 ratio. Column 6: Bolometric luminosity. Column 7: Starburst bolometric luminosity. Column 8: AGN luminosity. Column 9: ??, stellar velocity dispersion from Dasyra et al. (2006a,b, 2007). Column 10: MH , absolute H?band magnitudes from Veilleux (2006); Veilleux et al. (2009). Column 11: Optical spectral types, H for H II galaxies, L for LINER?like, AGN 2 and AGN 1 for type 2 AGN and type 1 AGN, respectively. For the spectral type, we adopted the classification from Veilleux et al. (1999, 2009). When not available, we used the values from NED/SIMBAD. aNASA/IPAC Extragalactic Database (NED); https://ned.ipac.caltech.edu/ bf15/f30 ratio is measured from low-resolution Spitzer -IRS spectra extracted from CASSIS. cValues for L(40 ? 500?m) and DL (the luminosity distance) used to calculate L(8 ? 1000?m) and Lbol (Sanders & Mirabel, 1996) are adopted from Armus et al. (2009) 89 3.2 Background: The OH Molecule The OH electronic ground state 2? is split due to spin-orbit coupling where spin angular momentum can be oriented either parallel or anti?parallel to the orbital angular momentum. Thus, the rotational states are split into two ladders: 2?3/2 and 2?1/2 (Storey et al., 1981), where the total angular momentum is given by J = N ? 1/2 (N is the rotational quantum number). Additionally, the orbital angular momentum can have two orientations with respect to the molecular axis, each having a slightly different energy. This ?-doubling further splits each J level into two levels of opposite parity. The cross-ladder 2? ? 23/2 ?1/2 J = 3/2? 5/2 ?-doublet transition at 34.60 and 34.63 ?m (see Fig.1 of Gonza?lez-Alfonso et al., 2014b) is an efficacious column density estimator. Within an optical depth region of ?35?m ? 0.5, we expect OH 35 to remain optically thin, enabling us to calculate the OH abundance for a given gas-to-dust ratio. Indeed, the spontaneous transition probabilities (described by the Einstein?A coefficient) of the intra-ladder transitions are much larger (? 0.1 s?1 for J = 3/2 to ? 10 s ?1 for J = 13/2) than the cross-ladder transitions (? 0.05? 10?4 s?1) from the same levels (Offer & van Dishoeck, 1992). Excitation of the cross- ladder transitions at 35 ?m decay through multiple branches (i.e. 35 ?m; 15-84-119 ?m; 48-119 ?m; 99-53 ?m; 99-163-79 ?m). Only 4% of all 35 ?m absorption events result in re-emission at 35 ?m (Bradford et al., 1999). Thus, OH 35 can provide a meaningful constraint on the true column density. Note that observations of OH 35 have also proved useful in studies of OH 90 megamasers (Darling, 2007; Darling & Giovanelli, 2006). Radiative pumping from the absorption of 35 ?m photons can invert the hyperfine population of the OH ground state. A 18 cm continuum source can then stimulate maser emission. Evi- dence in support of this radiative pump model was reported in Skinner et al. (1997) who observed OH 35 in Arp 220. They determined that the photon absorption rate in the 35 ?m transition could be efficient enough to power observed OH masers. 3.3 Sample Selection, Data Reduction, and Spectral Analysis 3.3.1 The Sample We have searched the Herschel Science Archive1 (HSA) and the Spitzer Her- itage Archive2 (SHA) for objects with (1) z . 0.06 (the OH 35 feature of more distant objects will be redshifted out of the Spitzer -IRS wavelength range), (2) PACS spectra covering the redshifted OH 79 and/or 119 ?m features, and (3) OH 35 ?m detection in the IRS long high (LH) spectra. There were 242 objects which met the first two criteria and only 15 of those objects satisfied all three criteria. Details of the Spitzer observations for our sample are listed in Table 3.1 and galaxy properties may be found in Table 3.2. All sources are morphologically classified as mergers or interacting systems (e.g. Veilleux et al., 2002). Our sample exhibits a range of interaction stages, from double nuclei systems with projected separation < 10 kpc (e.g. Mrk 273, IRAS F14348?1447, IRAS F20100?4156, IRAS F08572+3915), to mergers showing a sin- 1http://www.cosmos.esa.int/web/herschel/science-archive 2http://sha.ipac.caltech.edu/applications/Spitzer/SHA/ 91 gle nucleus with tidal tails (e.g. IRAS F05189?2524), to warm, quasar-dominated late stage mergers (e.g. Mrk 231). The bolometric luminosities of the galaxies are similar (log(Lbol/L ) ? 12.2), but the AGN contribution to the bolometric luminosity (i.e. the AGN fraction) essentially runs the full gamut from 0 to 1, resulting in AGN luminosities in the range of log(LAGN/L ) = 10.81? 12.51. The AGN fractions here are adopted from Veilleux et al. (2009), which are based on the 30-to-15 ?m continuum flux ratio, f30/f15. 3.3.2 Data Reduction All of the mid-infrared (5 ? 37 ?m), ?high resolution? (R ? 600) spectra used for this study were obtained with the Long High (LH) module of the Infrared Spec- trograph (IRS; Houck et al., 2004) on board the Spitzer Space Telescope (Werner et al., 2004). These spectra were acquired using a high-accuracy blue peak-up and observed in staring mode. The spectra were extracted using either the S17.2 or the S18.7 pipeline of the Spitzer Science Center (SSC). Data reduction follows the procedure outlined in Spoon et al. (2009) and in Spoon & Holt (2009), and starts at the DROOP level products created by the SSC pipeline. These products have been corrected for stray-light contributions, non-linear responsivity in the pixels, and drooping, but have not been flat fielded. For some observations, sky background images were then subtracted, and interpolation over bad pixels was performed us- 92 ing the IDL package IRSCLEAN1. The data analysis package SMART (Higdon et al., 2004) was then used to extract the one-dimensional target spectrum in ?full slit? (11.1?? ? 22.3??) mode. The same procedure was followed for a large number of observations of the calibration star ? Dra. The relative spectral response function (RSRF) was then created from the ratio of the observed spectrum of ? Dra to the stellar reference spectrum (Cohen et al., 2003, Sloan et al. 2005, private communication). The final spectrum was obtained by multiplying the object spectrum by the RSRF. For objects with more than one AOR Key, the final spectrum is the result of co-adding (no special weighting applied) the high-resolution spectra from each AOR key. For most of our objects, there is an artifact at? 34.8?35.0?m in the observer?s frame in the NOD 1 spectrum, but the artifact is not present in the NOD 2 spectrum. Inspection of the RSRF for NOD 1 shows that there is a 20% dip in the response at these wavelengths. It is possible that the dip in the NOD 1 spectrum therefore originates in the RSRF. The vertical light-green band in Figure 3.1 shows the region in velocity space where the dip occurs. Fortunately, its location does not affect the OH 35 and [Si II] spectral line structures. The wavelength calibration of our spectra is accurate to about 100 km s?1 (e.g. about 1/5 of an IRS resolution element of 500 km s?1). 1IRSCLEAN is available at http://irsa.ipac.caltech.edu/data/SPITZER/ docs/dataanalysistools/tools/irsclean/ 93 18 [S III] IRAS 17208-0014 16 k= 3 14 k= 2 12 k= 1 10 OH 35 [Si II] ?12000 ?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Figure 3.1 Example of twelve different continuum fits to the mid-infrared spectrum of IRAS F17208-0014. The black solid histograms are the data separated from each other by an arbitrary offset. Spline pivot points are anchored at wavelengths of 33.0 ?m, 33.8 ?m, 34.5 ?m, and 35.0 ?m. The blue, green, and red dots indicate the set of pivot points which are respectively the lower limits (v1), upper limits (v2), and best-fit-by-eye values (v3) used to fit a continuum spline. The top most set of fits is the result of using a third-order (k = 3) spline fit to each pivot point set. Below are the results for second-order (k = 2) and first-order (k = 1) spline fits. In each case, the blue, green, and red dotted lines show the results of the spline fits to the lower-limit, upper-limit, and ?best-fit-by-eye? continuum flux pivot points. The last spectrum at the bottom is fit with a polynomial. The magenta regions in this spectrum indicate the regions used to fit the first (solid blue line), second (solid green line), and third (solid red line) order polynomials to the continuum. The two dotted vertical lines in grey mark the rest wavelengths of the OH 35 doublet at systemic velocity. The dotted vertical lines in red mark the locations of the emission features [S III] 33.48 ?m and [Si II] 34.82 ?m. The vertical light-green band shows the region where a 20% dip in the detection response in the NOD 1 position occurs. 94 Flux Density [Jy] [S III] 3.4 [S III] 9.5 [S III] 46 Arp 220 IRAS F04454-4838 IRAS F05189-2524 3.2 44 9.0 3.0 42 8.5 40 2.8 8.0 38 2.6 36 2.4 7.5 OH 35 [Si II] OH 35 [Si II] OH 35 [Si II] 34 2.2 7.0 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 2.3 6.50 [S III] [S III] 5.50 [S III] IRAS F08572+3915 2.2 IRAS F12224-0624 IRAS 15250+3609 6.25 5.25 2.1 6.00 5.00 2.0 5.75 1.9 4.75 5.50 1.8 4.50 5.25 1.7 4.25 5.00 4.75 OH 35 [Si II] 1.6 OH 35 [Si II] 4.00 OH 35 [Si II] 1.5 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 8.0 [S III] [S III] [S III] 13 IRAS 17208-0014 IRAS F20551-4250 5.5 IRAS F23128-5919 7.5 12 5.0 7.0 11 4.5 6.5 10 4.0 6.0 9 OH 35 [Si II] OH 35 [Si II] 3.5 OH 35 [Si II] 5.5 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 10.0 12 24 [S III] [S III] [S III] Mrk 231 Mrk 273 9.5 NGC 2623 23 11 9.0 22 8.5 21 10 20 8.0 19 9 7.5 18 OH 35 [Si II] OH 35 [Si II] 7.0 OH 35 [Si II] 17 8 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 33 5.0 4.5[S III] [S III] [S III] NGC 4418 UGC 5101 Zw453.062 32 4.0 4.5 31 3.5 30 4.0 3.0 29 28 3.5 2.5 27 2.0 26 OH 35 [Si II] 3.0 OH 35 [Si II] OH 35 [Si II] 1.5 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 ?12000?10000 ?8000 ?6000 ?4000 ?2000 0 2000 4000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 3.2 Fits to the 33?35?m continua for all 15 objects in the sample. The grey shaded area shows the full range of the twelve different continuum fits described in Figure 3.1. Black dots are median values of the six fluxes at each pivot point, and the dotted line is the resultant third-order spline fit to those pivot points. The grey dotted vertical lines mark the rest wavelengths of the OH 35 doublet at systemic velocity. The red dotted vertical lines mark the locations of the emission features [S III] at 33.48 ?m and [Si II] at 34.815 ?m. The vertical light-green band shows the region where a 20% dip in the detection response in the NOD 1 position occurs. 95 Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] 3.3.3 Spectral Analysis To determine the extent to which continuum placement affects the measured properties (e.g. total integrated flux and equivalent width) of the OH 35 feature, we have fit twelve different baselines to the spectra. First-, second-, and third-order polynomials were fit via least squares minimization using wavelength regions which avoided spectral features (e.g. emission or absorption lines) and known artifacts. Splines were also fit to the continuum via four pivot points anchored at wavelengths of 33.0 ?m, 33.8 ?m, 34.5 ?m, and 35.0 ?m. Pivot point wavelengths were chosen to avoid known spectral features within this mid-infrared region. Three sets of continuum fluxes for the pivot points were chosen by eye. One set marks an upper limit on the continuum fluxes while another marks a lower limit on the continuum fluxes. The third set is a ?best fit? by visual inspection. To each set of pivot points, a first-, second-, and third-order spline was fit. Figure 3.1 shows an example of the continuum fits to IRAS F17208-0014. Figure 3.2 shows the range of continuum placements (grey shaded area) for each object. The black dots are median values of the six fluxes at each pivot point, and the dotted line is the resultant third-order spline fit to those pivot points. We choose this ?median? fit as the continuum for our sources and subtract it from each spectrum. The dotted vertical lines in red mark the locations of emission features [S III] at 33.48 ?m and [Si II] at 34.815 ?m. For both Figure 3.1 and Figure 3.2, the two dotted vertical lines in grey mark the rest wavelengths for both components of the OH 35 doublet at systemic velocity. These are separated by only 0.03 ?m or 96 ? 250 km s?1. Given the spectral resolution of the data (?500 km s?1), a single blended spectral feature is detected in our sources. Line profile fits of the OH 35 ??doublet were computed by using PySpecKit, a spectroscopic analysis and reduction toolkit for optical, infrared, and radio spectra (Ginsburg & Mirocha, 2011). The toolkit uses the Levenberg-Marquardt technique to solve the least-squares problem in order to find the best fit for the observa- tions. Profile fitting of the OH doublet followed a similar procedure as that outlined in Veilleux et al. (2013) and in Stone et al. (2016), in which the line profile was modeled using two Gaussian components, each characterized by their negative am- plitude, peak position, and standard deviation (or, equivalently FWHM). Gaussian parameters of the components were allowed to vary independently of each other. Note that these fits are only used to quantify the strength (i.e. total integrated flux and equivalent width) of the OH 35 absorption feature. The Gaussian components do not have a physical interpretation, nor do they account for each unresolved line in the doublet. 3.4 Results Figure 3.3 shows the fits to the continuum-subtracted OH 35 feature of each galaxy. The dashed blue lines indicate the Gaussian components and the magenta line is their sum. The total integrated fluxes and equivalent widths of the OH 35 feature measured from these spectral fits and their associated 1-? uncertainties based on the fits are listed in Table 3.3. 97 0.6 2 0.1Arp 220 IRAS F04454-4838 0.4 IRAS F05189-2524 IRAS F08572+3915 0.4 0.0 0.2 0 0.2 0.0 ?0.1 ?2 ?0.2 0.0 ?0.2 ?0.4 ? ?0.24 ?0.3 ?0.6 ?0.4 ?6 ?0.4 ?0.8 ?0.6 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 0.6 1.5 0.1 IRAS F12224-0624 0.4 IRAS 15250+3609 IRAS 17208-0014 IRAS F20551-42501.0 0.5 0.0 0.2 0.5 ? 0.0 0.0 0.1 0.0 ?0.2 ?0.5 ?0.2 ? ?1.0 ?0.50.4 ?0.3 ? ?1.50.6 ?0.4 ?1.0 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 1.0 1.5 1.5 1.5 IRAS F23128-5919 Mrk 231 Mrk 273 NGC 2623 0.5 1.0 1.0 1.0 0.0 0.5 0.5 0.5 ?0.5 0.0 0.0 0.0 ? ?0.51.0 ?0.5 ?1.0 ?0.5 ?1.5 ?1.5 ?1.0 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 6 1.5 2.5 NGC 4418 UGC 5101 2.0 Zw453.062 4 1.0 1.5 2 1.0 0.5 0 0.5 0.0 0.0?2 ?0.5 ?4 ?0.5 ?1.0 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 3.3 Two-Gaussian fits to the continuum-subtracted OH 35 line profiles of the 15 objects in our sample; see Section 3.3.3. In each figure, the solid black histogram is the data. Blue dashed lines indicate the two Gaussian components which best fit the line profile, and the magenta line is the sum of those two components. The grey dotted vertical lines mark the rest wavelengths of the OH 35 doublet at systemic velocity. The red dotted vertical line marks the location of the [Si II] emission line at 34.815 ?m. 98 Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] 8 8 8 8 IRAS 17208-0014 6 6 6 6 p3 v1k3 v2k3 v3k3 4 4 4 4 2 p2 2 v1k2 2 v2k2 2 v3k2 0 0 0 0 p1 v1k1 v2k1 v3k1 ?2 ?2 ?2 ?2 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] Figure 3.4 Examples of OH 35 line profile fits for each of the twelve different con- tinuum subtracted spectra in IRAS F17208-0014. Line colors and styles are the same as that for Figure 3.3. ?p1, p2, p3? indicate first-, second-, and third-order polynomial fits to the continuum, respectively. ?v1? refers to the ?lower-limit? pivot points in Figure 3.1. ?v2? refers to the ?upper-limit? pivot points, and ?v3? to the ?best-fit-by-eye? pivot points. ?k1, k2, k3? are the orders of the spline fitted to the continuum. For example, ?v2k3? is the third-order spline fit to the continuum using the ?upper-limit? pivot points. Our sources show the OH 35 line profile in absorption only. This is to be expected since essentially all OH molecules pumped to the upper 2?1/2, J = 5/2 level will relax by emitting a 99 ?m photon along the 2?1/2 ladder instead of emitting a 35 ?m photon. We note that this 99 ?m emission line is not observed in the PACS observations of our sample because it is redshifted into the 100 ?m gap of PACS. Figure 3.4 shows the twelve OH 35 line profile fits to each of the continuum- subtracted spectra of IRAS F17208-0014. ?p1, p2, p3? indicate first-, second-, and third-order polynomial fits to the continuum, respectively. ?v1? refers to the lower limit continuum flux pivot points in Figure 3.1. ?v2? refers to the upper limit continuum flux pivot points, and ?v3? refers to the ?best fit by eye? pivot points. ?k1, k2, k3? are the orders of the spline fitted to the continuum. For example, ?v2k3? is the third order spline fit to the continuum using the upper limit pivot points. The resultant fluxes and equivalent widths from all of the continuum fits 99 Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Flux Density [Jy] Integrated Flux (a) 8 6 4 2 0 0 500 1000 1500 2000 2500 3000 [Jy km s?1] Wvel (b) 6 4 2 0 0 25 50 75 100 125 [km s?1] Figure 3.5 Distributions of the measured OH 35 (a) integrated fluxes and (b) equiv- alent widths. are used to estimate the measurement errors. Section 3.4 shows the distributions of the measured OH 35 fluxes and equivalent widths in our sample. On average, we find that the uncertainties on the placement of the continuum translate into uncertainties of about ?15% on the final values of the equivalent width measurements listed in Table 3.3. Since the OH column density scales linearly with the OH 35 equivalent width, the typical uncertainties on the measurements of NOH and XOH are also ?15%. We note that the detection of OH 35 in UGC 05101 is marginal (with an uncertainty as high as ?64%), so it is excluded from these calculations and from the mean OH column density and mean OH?to?H abundance ratio in the following sections. 100 Table 3.3. OH 35 ?m Profile Properties Name Flux Wvel NOH XOH [Jy km s?1] [km s?1] [1017 cm?2] [10?6] (1) (2) (3) (4) (5) Arp 220 2977 ? 362 72 ? 9 1.69 ? 0.21 1.64 ? 0.20 IRAS F04454-4838 190 ? 15 64 ? 6 1.49 ? 0.15 1.44 ? 0.14 IRAS F05189-2524 524 ? 60 61 ? 7 1.42 ? 0.17 1.38 ? 0.16 IRAS F08572+3915 176 ? 22 30 ? 4 0.7 ? 0.09 0.68 ? 0.10 IRAS F12224-0624 109 ? 7 54 ? 4 1.26 ? 0.08 1.22 ? 0.08 IRAS 15250+3609 262 ? 23 52 ? 5 1.21 ? 0.12 1.17 ? 0.11 IRAS 17208-0014 1286 ? 292 108 ? 27 2.51 ? 0.63 2.44 ? 0.61 IRAS F20551-4250 633 ? 132 90 ? 20 2.09 ? 0.46 2.03 ? 0.45 IRAS F23128-5919 249 ? 27 60 ? 7 1.39 ? 0.16 1.35 ? 0.16 Mrk 231 1050 ? 186 51 ? 10 1.19 ? 0.22 1.15 ? 0.22 Mrk 273 531 ? 118 51 ? 12 1.19 ? 0.28 1.15 ? 0.28 NGC 2623 300 ? 9 36 ? 1 0.86 ? 0.02 0.83 ? 0.03 NGC 4418 1581 ? 403 53 ? 14 1.23 ? 0.32 1.20 ? 0.32 UGC 5101 67 ? 38 18 ? 11 0.42 ? 0.26 0.41 ? 0.26 Zw453.062 93 ? 8 42 ? 4 1 ? 0.08 0.95 ? 0.08 Note. ? Column 1: Galaxy name. Column 2: Total integrated flux. Col- umn 3: Equivalent width. Column 4: OH column density. Column 5: OH abundance relative to H. 101 log(N ?2OH/[cm ]) (a) 6 4 2 0 16.50 16.75 17.00 17.25 17.50 log(XOH) (b) 6 4 2 0 ?6.50 ?6.25 ?6.00 ?5.75 ?5.50 Figure 3.6 Distributions of the (a) OH column densities and (b) OH-to-H abundance ratios, XOH, inferred from the fits of the OH 35 feature. 3.4.1 OH Column Density The equivalent width of an absorption line is defined as ? Iv(c)? Iv Wv = dv, (3.1) Iv(c) where Iv(c) is the intensity of the continuum and Iv is the line intensity at frequency v. The optical depth is related to the line intensity via Iv = Iv(c) e ??v , (3.2) 102 ? where ? = n? ds (n is the volume density of absorbers and ? is the cross section for absorption). Therefore, we can relate the optical depth and the equivalent width, ? Wv = (1? e??v) dv. (3.3) For a foreground screen of homogeneous, isothermal gas illuminated by a back- ground source, in the optically thin limit (?  1), Equation 3.3 reduces to ? W? = NOH ?0 ?? d?, (3.4) where NOH is the ?ground state OH column density and ?? is the line profile function? defined such that ?? d? = 1 and ?0 holds the parameters of the transition. If the0 Einstein coefficients Au` for spontaneous emission and Bu` for stimulated emission are in units of [s?1] and [cm3 (J s2)?1], respectively (where u refers to the upper energy level and ` the lower energy level), we have the relations 8? h Au` = B3 u`, (3.5)? Bu` gu = B`u g`, (3.6) where g is the degeneracy of the level and h is the Planck constant. Thus, h A ?2u` gu ?0 = B`u = . (3.7) ? 8? g` Substituting ?0 in Equation 3.4 and noting that Wvel [cm s ?1] = W [s?1? ] ?? [cm], 103 we see that the number of molecules per unit area along the line of sight is Wvel8? g` NOH = . (3.8) A ?3u` gu Adopting Au` = 0.0174 s ?1 from Destombes et al. (1977) and Bradford et al. (1999), we calculate OH column densities using the empirically measured equivalent widths for our objects. Those column densities are shown in Figure 3.6 and listed in Ta- ble 3.3 with the associated 1-? standard deviation of the distribution. We find a mean of NOH = 1.31? 0.22? 1017 cm?2 for the objects in our sample. 3.4.2 XOH We can now calculate XOH, the OH abundance relative to H NOH XOH = . (3.9) NH First, we estimate the column density of H nuclei for a given ?35?m (the optical depth at 35 ?m due to dust) via (Mgas/Mdust) ?35?m NH = , (3.10) ?35?mmH where (Mgas/Mdust) is the gas-to-dust ratio by mass, ?35?m is the mass absorption coefficient of dust at 35 ?m, and mH is the H mass. From Figure 2 of Gonza?lez- Alfonso et al. (2014a) we adopt ? 2 ?135?m = 290 cm g , and we adopt a gas-to-dust 104 ratio of 100, appropriate for the nuclear regions of luminous and ultraluminous infrared galaxies (e.g. Wilson et al., 2008). We adopt an optical depth ?35?m = 0.5 because, in the inner regions, the OH excitation temperature will be in near equilibrium with the dust temperature. Moreover, for ?35?m > 0.5, OH absorption becomes harder to detect due to extinction at 35?m. Thus, we find NH ? 1? 1023 cm?2. Results for the OH column density from Equation 3.8 combined with the H column density from Equation 3.10 provide XOH. We calculate this abundance ratio for each of our sources (see Table 3.3) with the associated 1-? standard deviation of the distribution. We find a mean of XOH = 1.27 ? 0.21 ? 10?6 for our sample. Figure 3.6 shows the OH-to-H abundance ratios resulting from these calculations. We remark that our estimates for XOH are formally only lower limits because (1) some fraction of the OH molecules are in excited levels, so NOH is a lower limit to the column density of all OH states, (2) the covering fraction of the 35 ?m continuum source by this obscuring material may not be 100%, and (3) some of the spectra are not background-subtracted which reduces the equivalent width of the OH 35 feature1 1When comparing the optimally extracted spectra available from CASSIS with the spectra presented here, we find that the background subtraction does not have a noticeable effect on the equivalent widths. 105 3.5 Discussion 3.5.1 XOH: Comparison with the Literature Here we compare our empirically derived values of XOH with those found in the literature. Estimates of XOH are typically derived by fitting radiative transfer mod- els to multitransition observations of OH. For example, Goicoechea & Cernicharo (2002) modeled the first 20 rotational levels in OH and constrained the model with observations of Sgr B2. They report a range of X ?6OH = 2? 5? 10 (OH relative to H2). A similar study by Goicoechea et al. (2006) of the Orion KL outflow derived a lower value of XOH = 0.5 ? 1 ? 10?6. Following these two papers, the OH outflow studies of Sturm et al. (2011) and Gonza?lez-Alfonso et al. (2014b, 2017) adopted a value of XOH = 2.5 ? 10?6. Recent estimates of XOH in obscured nuclei (Falstad et al., 2015; Fischer et al., 2014; Gonza?lez-Alfonso et al., 2012) seem to favor the higher OH abundances. Our empirically derived value of XOH = 1.3 ? 0.4 ? 10?6 falls in the lower portion of this range of values. However, recall that our estimate is a lower limit, so it is formally consistent with all previously assumed and computed values of XOH in the literature. Our estimate is also consistent with Richings & Faucher-Giguere (2017)?s broad range of values for XOH (3.3?10?6 ? 2.2 ? 10?5), which is deduced from hydro- chemical simulations of AGN-driven winds. It should be noted, however, that the XOH values here and in Gonza?lez-Alfonso et al. (2014b, 2017) are relative to hydro- gen nuclei in both atomic and molecular forms, while Richings & Faucher-Giguere 106 (2017)?s abundance ratios are relative to molecular hydrogen, hence the larger val- ues. Finally, we note that our empirically determined values of XOH are consistent with the values of chemical models of photodissociation regions (the peak value in Sternberg & Dalgarno, 1995) and of cosmic-ray- and X-ray-dominated regions (Meijerink et al., 2011). 3.5.2 XOH: A Check on Radiative Transfer Models In recent years, radiative transfer models have been used to infer the main physical parameters of molecular outflows in several infrared-bright galaxies (e.g. Gonza?lez-Alfonso & Cernicharo, 1999; Gonza?lez-Alfonso et al., 2014b, 2017), and many of these properties depend upon the inferred H column density (e.g. outflowing mass, mass outflow rate, momentum flux, and energy flux). It is therefore important to verify that these models also predict realistic OH 35. Figure 3.7 shows the predicted OH 35 feature of the five ULIRGs (IRAS F05189-2524, IRAS F08572+3915, IRAS F20551-4250, Mrk 231, and Mrk 273) in common with the sample of Gonza?lez-Alfonso et al. (2017). In this paper, two ground-state lines (OH 79 and 119) and two excited lines (OH 65 and 84) were used to constrain the model. The green line is the OH 35 profile predicted by these models while the magenta line is the result from our fits to the observed feature (see Section 3.3.3). The agreement between the two is generally very good in terms of overall strength (equivalent width) of the feature. The predicted profiles are also largely consistent with the observed profiles, after taking into account the rather 107 1.15 1.15 1.15 IRAS F05189-2524 IRAS F08572+3915 IRAS F20551-4250 1.10 1.10 1.10 1.05 1.05 1.05 1.00 1.00 1.00 0.95 0.95 0.95 0.90 0.90 0.90 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Velocity [km s?1] 1.15 1.15 Mrk 231 Mrk 273 1.10 1.10 1.05 1.05 1.00 1.00 0.95 0.95 0.90 0.90 ?3000 ?2000 ?1000 0 1000 2000 3000 ?3000 ?2000 ?1000 0 1000 2000 3000 Velocity [km s?1] Velocity [km s?1] Figure 3.7 OH 35 spectra normalized to the continuum. Red and blue lines are the two Gaussian components of the fitted line and the magenta line is the sum of the two components. The green line is the profile predicted by the radiative transfer model described in Gonza?lez-Alfonso & Cernicharo (1999); Gonza?lez-Alfonso et al. (2017). The modeled line profiles are constrained by observations of four OH lines at 65, 79, 84, and 119 ?m. large uncertainties on the observed profiles associated with the continuum place- ment and the modest spectral resolution of the IRS data. The consistency between the model predictions and the observations adds some confidence in the energetics derived from these models. 108 3.6 Summary We have analyzed mid-infrared (5 ? 37 ?m) Spitzer?IRS spectra of 15 nearby (z . 0.06) ULIRGs focusing on the OH 35 ?m absorption feature. Several dif- ferent methods were used to fit these spectra and estimate the uncertainties on the measured line parameters. The measured equivalent widths of OH 35 im- plies an average OH column density and 1-? standard deviation to the mean of NOH = 1.31 ? 0.22 ? 1017 cm?2. We infer an OH-to-H abundance ratio of X = 1.27 ? 0.21 ? 10?6 when we assume a gas-to-dust ratio of 100, appropriate for the nuclear regions of luminous and ultraluminous infrared galaxies. This abundance ratio is formally a lower limit. It is consistent to within a factor of two with the value assumed in the radiative transfer models of OH outflows in these systems (XOH = 2.5 ? 10?6). We also find that the depths and profiles of OH 35 predicted by these models are largely consistent with the observed OH 35 spectra, providing support for the predicted energetics from these models. 109 Chapter 4: Far-Infrared Integral-Field Spectroscopy of Nearby Galactic Winds with Herschel-PACS 4.1 Introduction Large-scale galactic outflows driven by stellar processes and/or active galactic nuclei (AGNs) play a pivotal role in the life cycle of galaxies by heating the cold inter- stellar medium (ISM), suppressing or triggering star formation, fueling or quenching accretion onto the central black hole, and by influencing the chemical enrichment of galaxies (e.g. Fabian, 2012; Heckman & Thompson, 2017; Rupke, 2018; Veilleux et al., 2005, 2020). Although we now have a better understanding of how winds may significantly impact the formation and evolution of galaxies (e.g. we see this in semi- analytic, N-body, and hydrodynamical models of galaxy evolution where feedback is incorporated, (e.g. Naab & Ostriker, 2017; Somerville & Dave?, 2015; Zhang, 2018), there still exists a plethora of unanswered questions concerning the detailed physics and properties of galactic outflows. What is the primary mechanism of energy and momentum injection that powers an outflow? How can we constrain outflow prop- erties such as morphologies, lifetimes, spatial extent, radial-dependent velocities, mass outflow rates, and energetics? What is the efficiency of entraining material 110 from the ISM and ejecting it out of the disk? The answers to these questions can fundamentally measure the impact of outflows on galaxy evolution. However, it is challenging to constrain these properties because outflows involve multiple phases of the ISM (cold molecular, cool neutral atomic, warm and hot ionized; Heckman & Thompson, 2017) and hence, span a large range of temperatures (T ? 10 ? 107 K; Zhang, 2018), physical scales (a few parsecs to several kiloparsecs), and densities (n ? 10?2 ? 106 cm?3; Croxall et al., 2012; Draine, 2011). If we wish to gain a complete understanding of how winds contribute to the evolution of galaxies, it is crucial that we conduct detailed multi-phase ISM investi- gations of outflows and we must look to our local neighborhood where nearby galax- ies provide us the opportunity to study spatially resolved spectroscopic observations of outflows. Integral Field Spectroscopy (IFS) of the ionized (e.g. Harrison et al., 2014; Martin & Soto, 2016; Venturi et al., 2017), neutral (e.g. Rupke & Veilleux, 2011), and molecular (e.g. Pereira-Santaella et al., 2016; Rupke & Veilleux, 2013) ISM phases of outflows has made it possible to decompose the complex kinematics within galaxies. Not only can we utilize IFS to delineate disk rotation and outflow in a galaxy, but we can also gain insight into the spatial extent or morphological structure of the outflow. However, star formation regions and areas near the AGN in a galaxy are heav- ily enshrouded in dust which severely hampers UV and optical wavelength obser- vations. Fortunately, the far-infrared (FIR) provides us the opportunity to observe in these dusty regions without large extinction effects. We take advantage of the unprecedented combination of angular resolution, sensitivity, and spatial coverage of 111 the FIR spectroscopic integral field unit (IFU) of the Photodetector Array Camera and Spectrometer (Poglitsch et al., 2010, PACS;) on board the Herschel Space Ob- servatory (Pilbratt et al., 2010) to peer through the dusty veil obscuring the regions where stars form and black holes grow. In this paper, we present the main results of our analysis of the cool neu- tral atomic and the warm ionized ISM phases, as traced by FIR fine-structure line transitions in seven nearby (d < 16 Mpc) galaxies that are well known to harbor winds: Centaurus A (hereafter abbreviated as Cen A), Circinus, M 82, NGC 253, NGC 1068, NGC 3079, and NGC 4945. The proximity of these galaxies allows us to spatially resolve the morphologies and kinematics of the outflows (at the dis- tances for our sample, the physical sizes covered by a PACS 9.7??? 9.7?? spaxel are ? 0.2 ? 1 kpc). The OH 119 ?m doublet (hereafter, OH 119) has also been ob- served by PACS in this galaxy sample, which we present and compare with the observations of the ionized and neutral gas. The PACS data of the atomic gas for M 82 have been examined in detail in Contursi et al. (2013, hereafter C13). However, their observations cover a larger field of view (FOV), 2.5??2.5?, compared to our ?47??? 47?? FOV. Therefore, they were able to capture the outflow out to 1 kpc above and below the galaxy disk whereas we only capture the outflow out to a few hundred pc. Moreover, the majority of our FOV falls inside the starburst region of M 82. The observations from C13 were obtained in mapping mode instead of the pointed mode spectra presented here. Due to the spatial steps of the rasters used in their observations, the sky is Nyquist sampled and a spatial resolution of 6?? (? 130?270 pc, depending upon wavelength) 112 is achieved. In pointed mode, the beam is undersampled and has a native resolution of ? 10??. This is important to note because the sky will be undersampled which results in an underestimated source flux and the true morphology of an object will be degraded and difficult to reconstruct from a single pointed observation. Fortunately, we can increase confidence in the approximation in which the source morphology is reconstructed by comparing the interpolated observation with higher resolution data. Therefore, we re-examine PACS data of M 82 here for completeness and to validate our analysis on the other objects. This paper is organized as follows: The FIR atomic fine-structure lines are discussed in Section 4.2. The galaxies in our sample are described in Section 4.3. Descriptions of the observations, archival data, and data reductions are found in Section 4.4. We present the results of our analysis in Section 4.5 and Section 4.6. Section 4.7 and Section 4.8 outline the methods used to derive the gas outflow properties. The interpretation of the results and their implications are discussed in Section 4.9. Section 4.10 outlines our main conclusions. Appendix B contains the analysis of the molecular gas traced by OH 119. In this paper we adopt the standard convention that negative velocities with respect to systemic indicate approaching material. 4.2 Atomic Fine-Structure Emission Lines The gas phases in the PACS FIR range (55?210 ?m) cool through strong emis- sion lines arising from atomic fine-structure transitions and molecular transitions, 113 with the strongest atomic emission originating from carbon, nitrogen, and oxygen and their various ions. These atoms are collisionally excited and then de-excite through forbidden transitions, emitting photons and thus removing thermal energy from the gas. Indeed, the cooling radiation from these emission lines are so powerful that they provide almost all of the gas cooling in the cold neutral and warm ionized phases of the ISM, and hence the emission is easily detectable over large distances (for example, [O III] 88 has been observed out to z = 9.11; Hashimoto et al., 2018). The importance of these fine-structure transitions cannot be understated since they are the dominant mechanism for gas cooling and can therefore directly impact the star formation in a galaxy. Here we observe five atomic fine-structure line transi- tions within the PACS FIR waveband: [O I] 63 ?m, [O III] 88 ?m, [N II] 122 ?m, [O I] 145 ?m, and [C II] 158 ?m (hereafter, we drop the wavelength unit, ?m, for brevity). Because this suite of FIR lines span a wide range of ionization potentials (IP = 11.26 ? 35.12 eV), critical densities (ncrit ? 40 ? 105 cm?3), and excitation temperatures (Tex = 91? 326 K), the combination of these lines allows us to inves- tigate the physical conditions of the multiple phases of an outflow and probe the different regions of a galaxy from which these emission lines originate (see Table 4.1 for more details of the fine-structure lines). The low-excitation atomic lines ([O I] 63 , [O I] 145 , [C II] 158 ) sample pho- todissociation regions (PDRs) at the interfaces between molecular, atomic, and ionized gas phases in star-forming regions where intense far-ultra-violet (FUV) radiation photodissociates CO, resulting in bright emission of [O I] and [C II] 114 Table 4.1. Fine-structure Lines Line Transition ? ? IP Tex ncrit,H ncrit,e [?m] [GHz] [eV] [K] [cm?3] [cm?3] (1) (2) (3) (4) (5) (6) (7) (8) [O I] 63 3P ?31 P2 63.18 4744.77 0.0? 13.62 227 2? 105 ? ? ? [O III] 88 3P1?3P0 88.36 3393.01 35.12? 54.94 163 ? ? ? 510 [N II] 122 3P2?3P1 121.89 2459.38 14.53? 29.60 188 ? ? ? 375 [O I] 145 3P 30? P1 145.52 2060.07 0.0? 13.62 326 5? 104 ? ? ? [C II] 158 2P3/2?2P1/2 157.74 1900.54 11.26? 24.38 91 2? 103 47 Note. ? The information in Table 4.1 is based on Herrera-Camus et al. (2018a), Israel et al. (2017), Farrah et al. (2013) and the Leiden Atomic and Molecular Database (Scho?ier et al., 2005). Column 1: Atomic fine-structure line, Column 2: Electronic energy level transition, Column 3: Rest wavelength, Column 4: Rest frequency, Column 5: Ionization potential to create the species and the ionization potential to ionize the species, Column 6: Excitation temperature, Column 7: Critical density for collisions with H I and H2, Column 8: Critical density for collisions with electrons. (Hollenbach & Tielens, 1997; Sternberg & Dalgarno, 1995; Tielens & Hollenbach, 1985). [C II] 158 arises from both ionized and neutral gas due to its ionization potential of 11.2 eV. [C II] 158 is the dominant coolant in regions with densities nH? 10 ? 105 cm?3 and temperatures T ? 100 ? 300 K, and is the strongest emission line from cooler gas (T < 104 K) in galaxies (Carilli & Walter, 2013). In PDRs, the [O I] 63 and [O I] 145 line emission originates solely in the neutral regime. [O I] 145 is the faintest of the fine-structure lines, but its advantage is that it does not suffer from self-absorption like [O I] 63, which can be affected by self- absorption by relatively small amounts of foreground gas (N ? 2 ? 1020 cm?2H ; Liseau et al., 2006). [O I] 63 line emission can also originate in X-ray dominated regions (XDRs) near AGN (Dale et al., 2004). At higher densities (ne ? 105 cm?3) and higher temperatures (T > 200 K), [O I] 63 becomes the dominant cooling line 115 for neutral atomic gas instead of [C II] 158 (Meijerink et al., 2007). [O I] 63 line emission may also arise from shocks (Lutz et al., 2003). The high-excitation ionic emission line [O III] 88 traces moderate density (? 100 cm?3) clouds, H II regions of photoionization from young, massive stars (Spinoglio & Malkan, 1992; Voit, 1992), and the narrow line region (NLR) excited by AGN (D??az-Santos et al., 2017). The high abundance of oxygen makes [O III] 88 a pri- mary coolant of the warm ionized phase of the ISM, and its high ionization potential (35.12 eV) predominantly traces very energetic conditions near AGN (Ferkinhoff et al., 2010), hot OB stars, or shocks (Stasin?ska et al., 2015). [N II] 122 , a low- excitation line, also originates in H II regions and is solely excited in ionized gas due to its ionization potential of 14.53 eV, a value just above the ionization energy of hydrogen, 13.6 eV. Moreover, its low critical density (? 350 cm?3) means that it can easily be excited in the diffuse ionized phase of the ISM (Goldsmith et al., 2015; Herrera-Camus et al., 2016). 4.3 The Sample The seven galaxies in our sample (see Table 4.2 for more details about the galaxy sample) are well known to harbor winds and have been studied intensively across the entire accessible electromagnetic spectrum. The spatial scale of these studies span a range of a few pc to several hundred kpc. Two of our sample are classified as pure starbursts (?H II galaxies? in the table; M 82 and NGC 253) while the other five (Cen A, Circinus, NGC 1068, NGC 3079, and NGC 4945) harbor both 116 a narrow-line AGN and a starburst. M 82 is considered to be the archetypal starburst galaxy, hosting a substantial population of supernova remnants (Pedlar et al., 1999) and several star clusters (de Grijs, 2001). It is the energy from these populations? stellar processes that drives the well-known kpc-scale bipolar conically-shaped outflow along the galaxy?s minor axis. The inclination of M82 provides a direct sight line to the S part of the disk which makes the approaching S outflow cone clearer in visibility. The N outflow cone is receding from us and lies behind the galaxy disk where emission in the base of the outflow may be extinguished by the disk itself. Hard X-ray (E ? 6.7 keV) emission inside the central |r| < 200 pc, |z| < 100 pc of the galaxy arises from the wind fluid that drives the larger scale outflow (Griffiths et al., 2000). UV (e.g. Hoopes et al., 2005), mid-infrared (Engelbracht et al., 2006), and far-infrared/sub-mm (e.g. Alton et al., 1999; Roussel et al., 2010) images provide evidence of dust entrained in the hotter outflowing gas. Confined to the cone walls, warm ionized gas is seen as optical filamentary emission in H?, [N II], and [S II] (Shopbell & Bland-Hawthorn, 1998). Warm molecular hydrogen in the near-infrared roughly traces the H? and warm dust emission (Veilleux et al., 2009). Observations of CO show molecular line splitting along the minor axis providing further evidence of a conical outflow morphology (Leroy et al., 2015). 117 Table 4.2. Galaxy Properties Galaxy Name RA Dec Distance SFR log LFIR i PA Type Morph. [h m s] [? ? ??] [Mpc] [M ?1 ? ? yr ] [L ] [ ] [ ] (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) M 82 09 55 52.2 +69 40 46.60 3.6 10 10.67 81 65 SB I0 Cen A 13 25 27.62 ?43 01 08.81 3.8 2 9.91 75 278 AGN & SB S0 pec Circinus 14 13 09.95 ?65 20 11.87 4.2 4.7 9.92 65 216 AGN & SB SAb NGC 253 00 47 33.12 ?25 17 17.59 3.9 3 10.35 76 230 SB SABc NGC 1068 02 42 40.71 ?00 00 47.81 14.4 18 11.13 40 278 AGN & SB S0 pec NGC 3079 10 01 57.80 +55 40 47.24 16 10.64 84 166 AGN & SB SBc NGC 4945 13 05 27.48 ?49 28 05.57 3.8 0.4 10.28 75 45 AGN & SB SBcd Note. ? Column 1: Galaxy name, Column 2: Adopted RA of the galaxy center, Column 3: Adopted Dec of the galaxy center, Column 4: Luminosity distance, Column 5: Star formation rate, Column 6: FIR luminosity, Column 7: Galaxy inclination, Column 8: Position angle of the galaxy major axis, Column 9: Type, Column 10: Morphology classification. 118 Cen A is the prototypical Fanaroff-Riley I-type low-luminostiy galaxy, and at a distance of 3.8 Mpc (Harris et al., 2010), it is the nearest active radio galaxy. In the radio, Cen A hosts a collimated subparsec-scale jet and counterjet (Burns et al., 1983; Tingay et al., 1998), parsec-scale inner lobes (Clarke et al., 1992; Schreier et al., 1981), a kpc-scale middle lobe (Morganti et al., 1999), and giant outer radio lobes on scale of hundreds of kpc (McKinley et al., 2013). Optical images show evidence of a bipolar outflow on scales of ? 10 kpc (McKinley et al., 2018). H? (Blanco et al., 1975; McKinley et al., 2018) and [O III] ?5007 filaments are also seen on scales of the middle lobe. Circinus is the closest Seyfert 2 galaxy at 4.2 Mpc (Tully et al., 2009) and the second brightest AGN in the mid-infrared (after NGC 1068). The prominent ionization cone located NW of the galaxy nucleus contains extended optical filaments with outflowing velocities ? 100 ? 200 km s?1 out to ? 1 kpc (e.g. Marconi et al., 1994; Sharp & Bland-Hawthorn, 2010; Veilleux & Bland-Hawthorn, 1997; Veilleux et al., 2003). The ionization cone and outflow are traced on parsec scale in the NIR (Maiolino et al., 2000) and on kiloparsec scale in the X-Ray (Mingo et al., 2012). CO observations detect a ?NW cloud? spatially coincident with the ionized outflow cone with a blue-shifted velocity of ? 150 km s?1 and a ?far W cloud? moving along an H? filament (Zschaechner et al., 2016). Edge-brightened bipolar lobes are also seen in the radio which extend up to more than 2 kpc on either side of the galaxy plane (Elmouttie et al., 1998a). NGC 253 is a starburst-dominated galaxy at a distance of 3.9 Mpc (West- moquette et al., 2011), making it one of the two closest nuclear starburst galaxies 119 (M 82 being the other). Chandra observations (Strickland et al., 2002; Strickland & Stevens, 2000) show a limb-brightened conical outflow to the SE of the galaxy nucleus which is spatially aligned with the ?100?300 km s?1 outflow observed in H? (Westmoquette et al., 2011), and a northern X-ray outflow lobe that lies behind the disk of the galaxy. Outflows have been observed in Na I D (Heckman et al., 2000) and in OH (Sturm et al., 2011; Turner, 1985). A prominent molecular wind has been traced in CO with a mass outflow rate between ? 3 ? 9 M yr?1 (assuming optically thin CO emission; Bolatto et al., 2013) and ? 25? 50 M yr?1 (assuming optically thick CO emission; Krieger et al., 2019; Zschaechner et al., 2018). Even at the lower limit, these outflow rates (when compared to NGC 253?s star formation rate) are large enough to substantially affect the galaxy?s star formation activity. NGC 1068 is the archetypal Type 2 Seyfert galaxy and contains an AGN and a circumnuclear starburst. NGC 1068 also harbors a kiloparsec-scale radio jet (e.g. Wilson & Ulvestad, 1987). The NLR gas has been observed as an outflow with a biconical morphology (Cecil et al., 1990; Das et al., 2006). The northeast side of the cone is in front of the galactic disk (oriented towards us) while the Southwest side is behind (oriented away from us) and obscured by the galactic disk (Barbosa et al., 2014). Sub-mm interferometry of molecular lines in the circumnuclear disk strongly suggests the existence of a giant, AGN-driven outflow extending to ? 100 pc with a velocity of ? 100 ? 200 km s?1 (Garc??a-Burillo et al., 2014; Krips et al., 2011). 120 Table 4.3. PACS Fine-Structure Line Observations Object Name ObsId Line Spec. Res.a Ang. Res.b,c t dexp AOT Chop Throw Program ID (1) (2) (3) (4) (5) (6) (7) (8) (9) M 82 1342186799 [O I] 63 90 9.5 582 line large SDP esturm 3 M 82 1342186798 [O III] 88 125 9.5 986 line large SDP esturm 3 M 82 1342186798 [N II] 122 290 10.0 986 line large SDP esturm 3 M 82 1342186798 [O I] 145 250 11.0 986 line large SDP esturm 3 M 82 1342186798 [C II] 158 240 11.5 986 line large SDP esturm 3 Cen A 1342202590 [O I] 63 90 9.5 873 line large KPGT esturm 1 Cen A 1342202588 [O III] 88 125 9.5 11291 range med KPGT rguesten 1 Cen A 1342203444 [N II] 122 290 10.0 5663 range large KPGT rguesten 1 Cen A 1342202589 [O I] 145 250 11.0 2341 line large KPGT esturm 1 Cen A 1342202589 [C II] 158 240 11.5 2341 line large KPGT esturm 1 Circinus 1342191298 [O I] 63 90 9.5 1281 line large KPGT esturm 1 Circinus 1342191297 [O III] 88 125 9.5 1948 line large KPGT esturm 1 Circinus 1342191297 [N II] 122 290 10.0 1948 line large KPGT esturm 1 Circinus 1342191297 [O I] 145 250 11.0 1948 line large KPGT esturm 1 Circinus 1342191297 [C II] 158 240 11.5 1948 line large KPGT esturm 1 NGC 1068 1342191153 [O I] 63 90 9.5 3671 range large KPGT esturm 1 NGC 1068 1342203124 [O III] 88 125 9.5 3386 range large KPGT esturm 1 NGC 1068 1342191154 [N II] 122 290 10.0 3944 range large KPGT esturm 1 NGC 1068 1342191154 [O I] 145 250 11.0 3944 range large KPGT esturm 1 NGC 1068 1342203121 [C II] 158 240 11.5 3456 range large KPGT esturm 1 NGC 253 1342199414 [O I] 63 90 9.5 873 line large KPGT esturm 1 NGC 253 1342199415 [O III] 88 125 9.5 1975 line large KPGT esturm 1 NGC 253 1342199415 [N II] 122 290 10.0 1975 line large KPGT esturm 1 NGC 253 1342199415 [O I] 145 250 11.0 1975 line large KPGT esturm 1 NGC 253 1342199415 [C II] 158 240 11.5 1975 line large KPGT esturm 1 121 Table 4.3 (cont?d) Object Name ObsId Line Spec. Res.a Ang. Res.b,c t dexp AOT Chop Throw Program ID (1) (2) (3) (4) (5) (6) (7) (8) (9) NGC 3079 1342221391 [C II] 158 240 11.5 8045 range med DDT esturm 4 NGC 4945 1342212218 [O I] 63 90 9.5 2939 range large KPGT esturm 1 NGC 4945 1342212220 [O III] 88 125 9.5 1975 line large KPGT esturm 1 NGC 4945 1342212220 [N II] 122 290 10.0 1975 line large KPGT esturm 1 NGC 4945 1342212220 [O I] 145 250 11.0 1975 line large KPGT esturm 1 NGC 4945 1342212220 [C II] 158 240 11.5 1975 line large KPGT esturm 1 aSpectral resolutions are in units of km s?1 . bAngular resolutions are in units of arcsec. cEstimated from Fig.8 in the PACS Spectroscopy Performance and Calibration Guide Issue 3.0. dExposure times are in units of seconds. Note. ? Column 1: Galaxy name, Column 2: Observation ID, Column 3: Atomic fine-structure line, Column 4: Spectral Resolution of PACS at the emission line wavelength, Column 5: Angular resolution of PACS at the emission line wavelength, Column 6: Exposure time, Column 7: PACS Astronomical Observation Template, Column 8: Size of observation chopper throw, Column 9: Program ID. 122 NGC 3079 is well known for its spectacular super-bubbles. In the optical (e.g. Cecil et al., 2001; Veilleux et al., 1994), only the eastern bubble of NGC 3079?s double-lobed radio structure (Irwin et al., 2019) is visible. The galaxy disk is inclined such that the E side is nearest to us (Filippenko & Sargent, 1992; Yamauchi et al., 2004) and the counter bubble is located behind the disk where optical line emission is extinguished. The ionized outflow originating from the nucleus reaches velocities of up to ? 1500 km s?1 (Cecil et al., 2001). H? + [N II] emission is observed in alignment with the bridges and loops seen in radio (Duric et al., 1983; Filippenko & Sargent, 1992). There is a slow-moving (v ? 0.1c) parsec-scale radio jet (Middelberg et al., 2005) and copious amounts of hot X-ray gas that are aligned with the optical emission-line filaments (Cecil et al., 2002). NGC 4945 harbors a radio-quiet AGN and a nearly edge-on circumnuclear starburst disk with ? 83 pc radius (Marconi et al., 2000a). It is known for its high obscuration along the line of sight (NH ? 3.5?1024 cm?2; Puccetti et al., 2014) and is the brightest Seyfert 2 galaxy in the hard X-ray range (> 20 keV; Itoh et al., 2008), radiating at a variable rate of L/LEdd ? 0.1 (Madejski et al. 2000). H? emission tracing a central outflow cone was detected by Lehnert (1992); the cone has been imaged in both optical and near-infrared emission lines by Moorwood et al. (1996). In the soft X-ray (? 0.6 keV), the plume is interpreted as a mass-loaded wind launched from the nuclear starburst (Schurch et al., 2002). Heckman et al. (1990) detected line splitting in optical spectroscopic data and found the line emission to originate from the surface of an expanding conical structure. Line splitting starts around 70 pc and extends out to 700 pc. 123 [OI] 63 [OIII] 88 [NII] 122 [OI] 145 [CII] 158 45 69? 41? 00?? 37 28 45?? 20 69? 40? 30?? 12 4 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 34 ?43? 00? 45?? 27 ?43? 01? 00?? 20 ?15?? 14 7 ?30?? 0 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 41 ?65? 20? 00?? 33 ?15?? 25 16 ?30?? 8 ?45?? 0 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 37 ?25? 17? 00?? 29 ? 2215?? 15 ?30?? 7 0 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 32 ?0? 00? 30?? 25 19 ?45?? 13 ?0? 01? 00?? 6 0 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 62 49 37 25 13 0 30s 28s 13h 05m 26s 55 ?49? 27? 45?? 44 ?49? 28? 00?? 33 21 ?15?? 10 0 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.1 Signal-to-noise ratio maps. Contours are Contours are 0.1, 0.3, 0.5, 0.7, 0.8, and 0.9 of the peak value in each image. Black crosses mark the adopted galaxy center. 124 NGC 4945 NGC 3079 NGC 1068 NGC 253 Circinus Cen A M 82 [OI] 63 [OIII] 88 [NII] 122 [OI] 145 [CII] 158 26.84 15.60 3.48 2.06 16.69 69? 41? 00?? 21.79 12.65 2.83 1.67 13.62 16.73 9.69 2.19 1.27 10.55 45?? 11.68 6.74 1.55 0.88 7.48 69? 40? 30?? 6.62 3.78 0.90 0.48 4.42 175 pc 1.57 0.83 0.26 0.09 1.35 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 3.62 0.37 0.15 0.25 2.16 2.93 0.31 0.12 0.20 1.80 ?43? 01? 00?? 2.24 0.25 0.10 0.15 1.43 ?15?? 1.55 0.18 0.07 0.11 1.06 0.87 0.12 0.05 0.06 0.70 ?30?? 184 pc 0.18 0.06 0.03 0.01 0.33 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 6.90 1.70 0.60 0.53 4.30 ?65? 20? 00?? 5.57 1.38 0.49 0.43 3.48 ?15?? 4.24 1.06 0.37 0.32 2.66 ? ?? 2.90 0.74 0.26 0.22 1.8530 1.57 0.42 0.15 0.12 1.03 ?45?? 204 pc 0.24 0.10 0.03 0.01 0.21 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 11.58 2.70 2.77 1.45 8.97 ?25? 17? 00?? 9.31 2.17 2.23 1.16 7.26 ?15?? 7.04 1.64 1.70 0.88 5.56 4.77 1.11 1.16 0.59 3.85 ?30?? 2.50 0.59 0.62 0.31 2.14 189 pc 0.24 0.06 0.08 0.02 0.43 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 34s 33s 0h 47m 32s 3.30 2.08 0.53 0.20 1.77 ?0? 00? 30?? 2.69 1.69 0.44 0.16 1.49 ?45?? 2.08 1.29 0.35 0.12 1.21 1.47 0.90 0.25 0.08 0.93 ?0? 01? 00?? 0.85 0.51 0.16 0.05 0.65 698 pc 0.24 0.11 0.07 0.01 0.37 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 1.81 55? 41? 00?? 1.45 1.09 45?? 0.73 55? 40? 30?? 0.37 776 pc 0.01 10h 02m 00s 58s 10h 01m 56s ?49? 27? ?? 4.76 1.71 1.97 1.65 7.6845 3.85 1.38 1.59 1.32 6.20 ?49? 28? 00?? 2.93 1.04 1.20 0.99 4.72 2.02 0.71 0.81 0.67 3.25 ?15?? 1.10 0.38 0.43 0.34 1.77 184 pc 0.19 0.04 0.04 0.01 0.29 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.2 Total integrated emission line fluxes in 10?17 W m?2. Contours are 0.1, 0.3, 0.5, 0.7, and 0.9 of the peak flux in the image. Black crosses mark the adopted galaxy center. 125 NGC 4945 NGC 3079 NGC 1068 NGC 253 Circinus Cen A M 82 4.4 Observations, Archival Data, and Spectral Analysis The observations of the fine-structure lines, [O I] 63, [O III] 88, [N II] 122, [O I] 145, and [C II] 158 were performed with the Photoconductor Array Camera and Spectrometer (PACS; Poglitsch et al., 2010) on board the ESA Herschel Space Observatory (Pilbratt et al., 2010). These observations were done in pointed ob- serving mode with PACS line and range scan spectroscopy astronomical observ- ing templates (AOTs). A large chopper throw is used for a majority of the ob- servations (two medium throws were used for Cen A ObsID 134220258 and NGC 3079 ObsID 134221391). Observations of M 82 are from the Science Demonstration Phase SDP esturm 3 (PI: E. Sturm). Observations of Cen A, Circinus, NGC 253, and NGC 4945 are from Guaranteed Time Key Programs KPGT esturm 1 (PI: E. Sturm) and KPGT rguesten 1 (PI: R. Gu?sten). NGC 3079 is from the Director?s Discretionary Time DDT estrum 4 (PI: E. Sturm). More specific observation details are listed in Table 4.3. The FIR data were retrieved from the Herschel Science Archive (HSA)1 via the Herschel Interactive Processing Environment (HIPE v15.0.0; Ott, 2010). These data have been pipeline-processed at the Herschel Science Centre with the Standard Product Generation (SPG) software v14.2.0 up to the rebinnedCube task. The standard reduction steps include glitch masking, bad and noisy pixel masking, dark and background subtraction, spectral flatfielding, and flux calibration to Jy per spaxel. The uncertainty on the absolute flux calibration is of the order ? 6% and 1http://www.archives.esac.esa.int/hsa/whsa/ 126 the wavelength calibration uncertainties for the red and blue channel are ? 20 km s?1 and ? 40 km s?1, respectively (Poglitsch et al., 2010). For each spaxel in an observation, a second-order polynomial is fit to the continuum and subtracted from the spectrum. Two Gaussian components were then fit to the continuum-subtracted spectrum and line profile properties are estimated using the sum of the two components. It should be noted that NGC 1068 needed three Gaussian components to more accurately capture the broad wings exhibited in its emission lines, and in this case, the sum of the three Gaussian components is used to estimate line profile properties. Signal-to-noise ratio maps are presented in Table 4.3, where in general, the largest SNRs are observed in [C II] 158 and [O I] 63 . The velocity-integrated emission line flux and ratio maps are discussed in Section 4.5 and the kinematic analysis of these data is presented in Section 4.6. We note that the observed wavelength range of [N II] 122 contains a second pass ghost in most of the non-central PACS spectrometer spaxels. This ghost orig- inates from the [C II] 158 line and appears as a broad spectral feature shifted red- ward of the [N II] 122 line profile. This ghost lies close to the [N II] 122 line profile and leaves a small amount of observable continuum red-ward of [N II] 122 making continuum placement difficult. Therefore, we exclude the analysis of [N II] 122 in our discussion of the outflows in Section 4.9. 127 4.5 Velocity-Integrated Emission Line Flux and Ratio Maps Figure 4.2 and Figure 4.3 show the maps of the velocity-integrated line fluxes and line ratios, respectively. These emission line ratios are useful diagnostic tools that can probe the physical properties of the ISM. The spatial distribution of emis- sion ratios can reveal the excitation and ionization conditions in different regions of a galaxy. Moreover, it is possible to infer the ionization or excitation source of the gas. Line ratios constrain physical gas properties such as the strength of the surrounding radiation field, chemical abundances, local temperatures, and gas den- sities. To compute these line ratio maps, the total integrated flux maps were first smoothed to the same resolution and aligned to a common pixel grid. Here, we first briefly summarize the physical conditions traced by the following emission line ratios: [O I] 63/[C II] 158, [O III] 88/[O I] 63, [O III] 88/[N II] 122, [O III] 88/[C II] 158, [O I] 145/[O I] 63, and [C II] 158/[N II] 122, before discussing the results. It should be noted that [O I] 63 can be affected by self-absorption from cold foreground gas with column densities as low as NH ? 2? 1020 cm?2 (Liseau et al., 2006). Therefore, any conclusions drawn from ratio values dependent upon the [O I] 63 emission line flux should be made with care. In the ratios listed below, if self- absorption is present in the [O I] 63 spectrum, then the [O I] 63 / [C II] 158 values will be underestimated while the [O III] 88 / [O I] 63 and [O I] 145 / [O I] 63 values will be overestimated. In this galaxy sample, [O I] 63 self-absorption is only ob- vious in the spectra along the NE portion of the galaxy major axis of NGC 4945 (absorption is not obviously detected above and below the disk). 128 4.5.1 Constraints from the Emission Line Ratios [O I] 63 / [C II] 158 ? [C II] 158 and [O I] 63 are the dominant coolants in PDRs, so the relative strength of their emission is a proxy for the FUV heating intensity (a measure of how many FUV photons contribute to the gas heating) in these regions (Parkin et al., 2014). This ratio can also be used to discriminate between XDRs near AGN and classical PDRs. PDR emission is produced at the outer edges of molecular clouds, but at high densities (n ? 105 cm?3) in XDRs, the deeply penetrating X-ray photons can ionize [O I] in a larger volume of the molecular cloud, thereby enhancing the [O I] 63 emission (Meijerink et al., 2007). In XDRs, the [O I] 63 line emission can substantially exceed the emission of [C II] 158 and become the dominant cooling mechanism in the neutral gas (Herrera-Camus et al., 2018b). Thus, the ionization effects of the AGN can be traced by high ratios of [O I] 63 / [C II] 158 . [O III] 88 / [O I] 63 ? This ratio yields the ionized to neutral gas ratio in a region, but the ratio alone does not distinguish whether the source of ionization is the AGN or SB (Ferna?ndez-Ontiveros et al., 2016). Low ratios indicate the dominance of the [O I] 63 line emission from dense, neutral gas. [O III] 88 / [N II] 122 ? [O III] 88 and [N II] 122 line emission can arise in the NLR of AGN. Thus, their line ratio is a sensitive indicator of the strength of the ionization parameter, U , (the number of incident ionizing photons divided by the hydrogen gas density within the NLR of an AGN; Abel et al., 2009) of the NLR. In stellar H II regions, this ratio directly measures the effective temperature of the 129 stars responsible for ionization. The ratio reveals the stellar classification of the youngest stars in the region (Ferkinhoff et al., 2011). [O III] 88 / [C II] 158 ? This ratio is an excitation sensitive indicator. [O III] 88 line emission arises in the NLR of AGN or in H II regions where hot O and B stars produce photons energetic enough to ionize [O III] 88 . Therefore it traces regions of high radiation fields and gives a measure of the relative ionized to neutral gas abundances. [O I] 145 / [O I] 63 ? In the optically thin limit ([O I] 145 / [O I] 63 values . 0.1; Tielens & Hollenbach, 1985), and in the temperature range 100 ? 400 K and density range nH. 104 cm?3, the [O I] 145 / [O I] 63 line ratio is an effec- tive temperature tracer for the neutral gas in PDRs (Liseau et al., 2006; Tielens & Hollenbach, 1985). Higher values of this ratio indicates high gas temperatures (Ferna?ndez-Ontiveros et al., 2016; Kaufman et al., 1999). The gas temperature traced by this ratio is anti-correlated with the gas density traced by [S III] 33 ?m/ [S III] 18 ?m (Spinoglio et al., 2015). [C II] 158 / [N II] 122 ? [N II] 122 line emission arises solely from low- excitation photoionized gas, so the [C II] 158 / [N II] 122 ratio can quantify the fraction of [C II] 158 emission originating from the neutral gas in PDRs. 4.5.2 Results from the Emission Line Ratios M 82: Emission in all lines (see Figure 4.2) show elongation from E to W, however, [O I] 63 and [O III] 88 show a slightly more compact nuclear morphology. 130 [OI]63/[CII] [OIII]/[OI]63 [OIII]/[NII] [OIII]/[CII] [OI]145/[OI]63 [CII]/[NII] 2.03 0.71 4.91 1.23 0.13 8.31 69? 41? 00?? 1.70 0.61 4.37 1.02 0.11 7.02 45?? 1.37 0.50 3.84 0.81 0.08 5.73 ? ? ?? 1.04 0.40 3.31 0.60 0.06 4.4469 40 30 0.71 0.30 2.78 0.39 0.04 3.15 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 1.72 0.64 3.47 0.29 0.10 16.72 ?43? 01? 00?? 1.38 0.50 3.04 0.25 0.09 14.75 ?? 1.03 0.37 2.61 0.21 0.08 12.77?15 0.69 0.23 2.18 0.18 0.07 10.80 ?30?? 0.35 0.10 1.75 0.14 0.05 8.82 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 1.72 0.93 4.77 1.18 0.09 9.65 ?65? 20? 00?? 1.42 0.75 3.91 0.93 0.08 8.15 ?15?? 1.11 0.58 3.04 0.68 0.07 6.65 ?30?? 0.81 0.40 2.17 0.43 0.06 5.14 ?45?? 0.50 0.22 1.31 0.18 0.04 3.64 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s 12s 09s 14h 13m 06s ?25? 17? 00?? 1.41 0.30 1.53 0.34 0.17 7.53 1.18 0.25 1.29 0.28 0.14 6.30 ?15?? 0.94 0.21 1.04 0.22 0.11 5.07 ?30?? 0.70 0.16 0.80 0.16 0.08 3.84 0.46 0.12 0.56 0.10 0.06 2.62 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s ?0? 00? 30?? 1.96 0.95 4.22 1.21 0.08 7.01 ?40?? 1.56 0.78 3.37 0.94 0.07 6.02 ? ?? 1.15 0.62 2.53 0.68 0.06 5.0450 ? ? ? ?? 0.75 0.46 1.68 0.42 0.05 4.050 01 00 0.34 0.30 0.83 0.15 0.04 3.07 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s ?49? 27? 45?? 1.32 0.38 1.30 0.44 0.35 10.40 1.07 0.32 1.04 0.35 0.27 8.27 ?49? 28? 00?? 0.82 0.25 0.79 0.25 0.19 6.14 ?15?? 0.57 0.19 0.54 0.16 0.11 4.01 0.32 0.13 0.29 0.06 0.03 1.88 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.3 Maps of the emission line ratios. Contours are 0.1, 0.3, 0.5, 0.7, 0.8, and 0.9 of the peak value in each image. This morphology is emphasized in Figure 4.3 where high ratios of [O I] 63 /[C II] 158 , [O III] 88 /[N II] 122 , and [O III] 88 /[C II] 158 are in a compact region around the galaxy nucleus. The emission line ratios also indicate that the neutral gas traced by [O I] 63 and the ionized gas traced by [O III] 88 and [N II] 122 is typically stronger near the galaxy center, but further from the center, [O I] 145 and [C II] 158 emission begin to dominate. C13 shows that the bipolar outflow is best determined from the [O III] 88 /[O I] 63 ratio. Our smaller FOV makes a direct morphological compar- 131 NGC 4945 NGC 1068 NGC 253 Circinus Cen A M 82 [OI] 63 [OIII] 88 [NII] 122 [OI] 145 [CII] 158 ? ? ? ?? 13543 00 40 93 ?43? 01? 00?? 51 ? 920?? -32 ? ?? 184 pc40 -75 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 169 ?65? 20? 00?? 112 56 ?20?? ?40?? -57 204 pc -113 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s ?? 18115 131 69? 41? 00?? 81 45?? 31 30?? -18 175 pc 69? 40? 15?? -68 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 166 ?25? 17? 00?? 115 ?15?? 64 13 ?30?? -37 ? ?? 189 pc45 -88 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 201 ?0? 00? 30?? 137 ? 7345?? 10 ?0? 01? 00?? -53 698 pc ?15?? -116 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 168 15?? 109 55? 41? 00?? 51 45?? -7 30?? -65 776 pc 55? 40? 15?? -124 10h 02m 00s 10h 01m 57s 150 ?49? 27? 45?? 96 ?49? 28? 00?? 41 ? -1315?? -68 ?30?? 184 pc -123 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.4 Maps of the median velocities, v50 in units of km s ?1 . Contours are in eight equal steps between the minimum velocity and the maximum velocity in each image. 132 NGC 4945 NGC 3079 NGC 1068 NGC 253 M 82 Circinus Cen A [OI] 63 [OIII] 88 [NII] 122 [OI] 145 [CII] 158 ? 29843? 00? 40?? 238 ?43? 01? 00?? 178 ?20?? 119 59 ? ?? 184 pc40 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 289 ?65? 20? 00?? 231 173 ?20?? 115 ?40?? 57 204 pc 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s ?? 27315 219 69? 41? 00?? 164 45?? 109 30?? 54 175 pc 69? 40? 15?? 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 429 ?25? 17? 00?? 362 ?15?? 294 226 ?30?? 159 ? ?? 189 pc45 91 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 469 ?0? 00? 30?? 375 ?45?? 281 187 ?0? 01? 00?? 93 698 pc ?15?? 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 385 15?? 308 55? 41? 00?? 231 45?? 154 30?? 77 776 pc 55? 40? 15?? 10h 02m 00s 10h 01m 57s 520 ?49? 27? 45?? 430 ?49? 28? 00?? 341 ? ?? 25215 163 ?30?? 184 pc 73 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.5 Maps of the 1 ? ? line widths, W ?11? , in units of km s . Contours are 0.4, 0.5, 0.6, 0.7, 0.8, and 0.9 of the peak width in each image. 133 NGC 4945 NGC 3079 NGC 1068 NGC 253 M 82 Circinus Cen A ison difficult because our FOV encapsulates mostly the base of the outflow cones. C13 reports that the highest [O III] 88 / [O I] 63 ratios are highest in the N portion of the SB region, which is consistent with our findings. C13 also reports an asym- metry in the disk in the [O I] 145 / [O I] 63 line ratios. Our results are consistent with their findings with the higher ratios relegated to the E portion of the disk and the smaller ratios in the W portion of the disk. Cen A: In Figure 4.2 the adopted galaxy center of Cen A is spatially co- incident with the peak emission in all lines. Cen A exhibits a centrally condensed nuclear emission of neutral gas with a slightly shallower flux gradient towards the NE of the nucleus, as seen in the [O I] 63 and [O I] 145 maps. On the other hand, the ionized gas traced by [O III] 88 and [N II] 122 shows bright, elongated nuclear emission (oriented SE to NW) surrounded by diffuse emission that extends to the edges of the PACS FOV. Although [C II] 158 is seen as diffuse emission outside of the galaxy center similar to [O III] 88 and [N II] 122 , the brightest contours of [C II] 158 are not elongated SE to NW as seen in the ionized gas. Instead, the brightest contours of [C II] 158 emission are closer in morphology to the compact nuclei seen in the neutral gas. NE of the galaxy nucleus, in a location spatially coincident with ?Knot A?, there is a compact region of high [O III] 88 /[C II] 158 ratios. The highest [O I] 63 /[C II] 158 ratios in Figure 4.3 for Cen A are localized to the galaxy center, but there is some slight extension of higher ratios to the NE in the direction of the jet. There is also elongation of higher [C II] 158 /[N II] 122 ratios to the NE in the direc- tion of the jet. In general, the neutral gas traced by [O I] 63 dominates the central re- 134 gions (as seen in the low [O III] 88 /[O I] 63 ratios and the high [O I] 63 /[C II] 158 ratios). Circinus: The nuclear neutral gas emission seen in [O I] 63 and [O I] 145 shows a compact spatial distribution with a shallower flux gradient extending to the SW of the galaxy center. The ionized gas traced by [O III] 88 in contrast shows elon- gated emission with a shallower flux gradient extending to the NW. We see that the emission in [N II] 122 and [C II] 158 are the most spatially extended of all the lines and show a preferential elongation along the north ? south axis. The well-known ionization cone in Circinus is most clearly identified in the [O III] 88 /[N II] 122 , [O III] 88 /[C II] 158 , and [C II] 158 /[N II] 122 emission line ratio maps shown in Figure 4.3. The [O III] 88 /[N II] 122 ratios are, on average, a factor of 2 greater inside the cone than those in the disk. The high [O I] 63 /[C II] 158 ratios localized to the central region of Circinus reveals the effects of the AGN on the nearby gas ([O I] 63 line emission can be particularly strong in XDRs or NLRs near AGN). NGC 253: The emissions for all of the fine-structure lines in NGC 253 lie in a elongated morphology along the NE ? SW axis, coincident with the 2.2 ?mm stellar bar detected by Scoville et al. (1985). There is a steep flux gradient for all lines towards the SW. Most of the [O I] 63, [O III] 88, and [O I] 145, emis- sion is concentrated within the central 10???10??. In contrast, the [C II] 158 and [N II] 122 emission distributions extend further above and below the galaxy major axis. Again we see the effects of the AGN radiation on the gas in the elevated values of the line ratios of [O I] 63 /[C II] 158 and [O III] 88 /[C II] 158 (although 135 these ratios also show some elongation towards the E along the galaxy disk) and the low ratios of [C II] 158 /[N II] 122 found in compact nuclear regions in NGC 253. The [O I] 63 /[C II] 158 and [O III] 88 /[C II] 158 ratios also indicate the presence of neutral gas (as traced by [C II] 158 ) in the N outflow cone of NGC 253. NGC 1068: All line emission in NGC 1068 is elongated along the NE ? SW axis and aligned with the galaxy?s 2.5 kpc near-infrared nuclear bar (Schinnerer et al., 2000). The peak line emission for [O I] 63, [O III] 88, and [O I] 145 are local- ized to a compact region near the galaxy nucleus while [C II] 158 and [N II] 122 emission is much more diffuse. The outflow in NGC 1068 is most obvious in the [O III] 88 /[O I] 63 ratio (see Figure 4.3) where the largest ratios are in a compact region NE of the galaxy nu- cleus in the direction of the known ionization cone and outflow. The effects of the AGN radiation are seen in the nuclear compact morphologies of high ratios in [O I] 63 /[C II] 158 , [O III] 88 /[N II] 122 , and [O III] 88 /[C II] 158 . Also clear from those ratios is that the neutral gas traced by [C II] 158 increases in emission further away from the galaxy center. The higher ratios in [O I] 145 /[O I] 63 N of the nucleus and the higher ratios of [C II] 158 /[N II] 122 S of the nucleus appear to trace molecular arms in NGC 1068. NGC 3079: [C II] 158 line emission in NGC 3079 is diffuse and elongated in the direction of the galaxy major axis. East of the nucleus, a steep flux gradient is observed along the galaxy minor axis in the direction of the optical bubble (Cecil et al., 2001; Veilleux et al., 1994). NGC 4945: For all lines in NGC 4945, peak line emission is in a compact 136 region near the nucleus. The [O III] 88, [N II] 122, and [C II] 158 line emission is elongated N to S where [N II] 122 and [C II] 158 are more diffuse than [O III] 88 . The neutral gas traced by [O I] 63 is diffuse but the flux decreases more uniformly with increasing distance from the nucleus, whereas the steeper flux gradients of [O III] 88, [N II] 122, and [C II] 158 lie in the direction of the galaxy minor axis. The [O I] 145 emission is elongated to the S and the SE of the nucleus. It should be noted that there is significant self-absorption in [O I] 63 for NGC 4945. We do not correct for this absorption, so the [O I] 63 fluxes reported here should be considered lower limits. The [O III] 88 /[N II] 122 ratios in NGC 4945 (see Figure 4.3) show high val- ues in a compact region inside the NW outflow cone and low ratios in the SE cone, indicating a higher ionization in the NW cone. The [O III] 88 /[O I] 63 ratios are highest in the disk along the NE direction. As mentioned earlier in this section, the [O I] 63 spectrum of NGC 4945 suffers from self-absorption along the NE portion of the disk. So the higher [O III] 88 /[O I] 63 values are most likely due to the self- absorption as opposed to higher ionization. The [O I] 145 /[O I] 63 ratio shows high values in a compact region around the nucleus which also extends SE in the direction of the outflow. Lower [O I] 145 /[O I] 63 ratios in the NW cone suggest a lower tem- perature for the neutral gas compared to the SE cone. The [O I] 63 self-absorption is not obviously observed in regions above and below the galaxy disk. Therefore, even though the [O I] 145 /[O I] 63 values near the nucleus are most likely overesti- mates, there is greater confidence in the accuracy of the [O I] 145 /[O I] 63 values in the NW and SE outflow cones of NGC 4945. The [O III] 88 /[C II] 158 ratios show 137 low values in a compact region around the galaxy center indicating higher ioniza- tion near the AGN with [C II] 158 line emission increasing with distance from the nucleus. The [C II] 158 /[N II] 122 ratios are lower in a compact region around the nucleus compared to a compact region in the NW outflow cone with higher values, indicating a higher [C II] 158 emission in the NW outflow. 4.6 Gas Kinematics The widths and velocities of the emission line profiles in each spaxel are de- scribed by the non-parametric characteristics v50 and W1? . v50 is the median velocity of the fitted emission line profile, i.e. 50% of the emission is produced at velocities below v50 . Zero velocity corresponds to the rest wavelength at the galaxy?s systemic velocity. W1? is the width of the line profile within 1-? standard deviation of v50 . Figure 4.4 and Figure 4.5 show the v50 and W1? maps of the sample galaxies, respectively. These maps are analyzed in Section 4.8 and the results of this analysis are discussed in Section 4.9 where they are compared to the multi-phase outflows known to exist in these galaxies. 4.7 PDR Modeling It is possible that the emission of the neutral ISM component in the outflows of our galaxies originate from photodissociation regions (PDRs). PDRs are the warm, dense surfaces of molecular clouds that are exposed to the far-ultraviolet (FUV) photons (6 eV < h? < 13.6, Hollenbach & Tielens, 1997, 1999) that escape from H II 138 Solution A Solution B Solution C Solution D ?43? 00? 45?? 1780 1780 56200 56200 1426 1426 48505 48505 ?43? 01? 00?? 1072 1072 40811 40811 ?15?? 718 718 33116 33116 ?30?? 364 364 25422 25422 med: 17 mean: 188 med: 562 mean: 709 med: 17800 mean: 20636 med: 31600 mean: 31497 10 10 17727 17727 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 1225 1225 100000 100000 ?65? 20? 00?? 982 982 86213 86213 ?15?? 739 739 72427 72427 ? ?? 496 496 58641 5864130 253 253 44855 44855 ?45?? med: 178 mean: 179 med: 562 mean: 630 med: 31600 mean: 36449 med: 31600 mean: 49763 10 10 31069 31069 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 15?? 3160 3160 100000 100000 2530 2530 86200 86200 69? 41? 00?? 1900 1900 72400 72400 45?? 1270 1270 58600 58600 69? 40? 30?? 640 640 44801 44801 med: 562 mean: 581 med: 1780 mean: 1476 med: 31600 mean: 39309 med: 56200 mean: 66155 10 10 31001 31001 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 1065 1065 178000 178000 ?25? 17? 00?? 854 854 145960 145960 ?? 643 643 113920 113920?15 432 432 81880 81880 ?30?? 221 221 49840 49840 ?45?? med: 100 mean: 151 med: 562 mean: 663 med: 56200 mean: 53769 med: 56200 mean: 66020 10 10 17800 17800 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 3160 3160 178000 178000 ?0? 00? 30?? 2530 2530 145960 145960 1900 1900 113920 113920 ?45?? 1270 1270 81880 81880 ?0? 01? 00?? 640 640 49840 49840 ?15?? med: 56 mean: 117 med: 316 mean: 831 med: 31600 mean: 35731 med: 31600 mean: 45638 10 10 17800 17800 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 3160 562000 ?49? 27? 45?? 2530 451600 ?49? 28? 00?? 1900 341200 ?15?? 1270 230800 640 120400 ?30?? med: 562 mean: 774 med: 178000 mean: 227626 10 10000 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.6 Maps of the hydrogen density, nH in each object. Units are in cm ?3. The ?low? density range is defined as 0 < n < 104 cm?3H and the ?high? density range is defined as 104 < nH< 10 7 cm?3. See Section 4.7.4 for details about the determination of these density ranges and definitions of Solutions A?D. Solution A is the low density solution with only the [C II] 158 flux correction (see Section 4.7.1 and Section 4.7.4), Solution B is the low density solution with both the fluxes of [C II] 158 and [O I] 63 corrected (see Section 4.7.1, Section 4.7.2 and Section 4.7.4), Solution C is the high density solution with only the [C II] 158 flux correction, and Solution D is the high density solution with both the fluxes of [C II] 158 and [O I] 63 corrected. As discussed in Section 4.7.4, the high density PDR solutions ?C? and ?D? lead to unphysical ISRFs. Therefore, we have excluded them from the analysis of the outflow. 139 NNGGCC 44994455 NNNNGGGGCCCC 1111000066668888 NNNNGGGGCCCC 222255553333 MMMM 88882222 CCCCiiiirrrrcccciiiinnnnuuuussss CCCCeeeennnn AAAA Solution A Solution B Solution C Solution D ?43? 00? 45?? 1780 1780 1.79 1.79 1426 1426 1.63 1.63 ?43? 01? 00?? 1072 1072 1.47 1.47 ?15?? 718 718 1.32 1.32 ?30?? 364 364 1.16 1.16 med: 56 mean: 301 med: 316 mean: 390 med: 1.00 mean: 1.00 med: 1.78 mean: 1.76 10 10 1.00 1.00 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 5620 5620 1.83 1.83 ?65? 20? 00?? 4496 4496 1.67 1.67 ?15?? 3372 3372 1.50 1.50 ?30?? 2248 2248 1.33 1.33 1124 1124 1.17 1.17 ?45?? med: 562 mean: 1058 med: 1000 mean: 1165 med: 1.00 mean: 1.00 med: 1.78 mean: 1.39 0 0 1.00 1.00 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 15?? 5620 5620 1.83 1.83 4496 4496 1.66 1.66 69? 41? 00?? 3372 3372 1.50 1.50 45?? 2248 2248 1.33 1.33 69? 40? 30?? 1124 1124 1.17 1.17 med: 1780 mean: 2186 med: 1780 mean: 1902 med: 1.00 mean: 1.00 med: 1.78 mean: 1.72 0 0 1.00 1.00 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 31600 31600 1.78 1.78 ?25? 17? 00?? 25286 25286 1.54 1.54 18972 18972 1.29 1.29 ?15?? 12658 12658 1.05 1.05 ?30?? 6345 6345 0.81 0.81 ?45?? med: 1000 mean: 3389 med: 3160 mean: 4572 med: 1.00 mean: 0.96 med: 1.00 mean: 1.18 31 31 0.56 0.56 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 10000 10000 1.78 1.78 ?0? 00? 30?? 8000 8000 1.62 1.62 6000 6000 1.46 1.46 ?45?? 4000 4000 1.31 1.31 ?0? 01? 00?? 2000 2000 1.15 1.15 ?15?? med: 316 mean: 891 med: 316 mean: 593 med: 1.00 mean: 1.00 med: 1.00 mean: 1.23 0 0 0.99 0.99 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 42s 41s 40s 2h 42m 39s 17800 17 ?49? 27? 45?? 14243 14 ?49? 28? 00?? 10687 11 ?? 7130 8?15 3574 5 ?30?? med: 3160 mean: 4722 med: 5 mean: 7 17 1 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.7 Maps of the strength of the UV radiation field, G0, for the low density and high density limits PDR solutions. Definitions of the Solutions are the same as those in Figure 4.6. 140 NNGGCC 44994455 NNNNGGGGCCCC 1111000066668888 NNNNGGGGCCCC 222255553333 MMMM 88882222 CCCCiiiirrrrcccciiiinnnnuuuussss CCCCeeeennnn AAAA regions in star-forming regions or the accretion disks in AGN. The absorption of FUV photons by gas and dust is converted to FIR radiation including intense emission of [C II] 158 and [O I] 63 transitions, mainly excited by collisions with H2 molecules and H I (nH & 103 ? 104 cm?3;Liseau et al., 2006), whose line emissions dominate the gas cooling in PDRs (Tielens & Hollenbach, 1985). These two transitions are the brightest PDR emission lines in the Herschel -PACS spectral range and thus, are sensitive probes to the physical conditions of the gas in a PDR. In this section, we use the fluxes of [O I] 63, [O I] 145, and [C II] 158 and the total infrared continuum emission (TIR ? 30 ? 1000 ?m) to derive simple PDR models and infer the spatially-averaged properties of the gas by comparing observed emission line ratios to theoretically predicted line ratios. The models we employ here were first presented by Hollenbach et al. (1991); Tielens & Hollenbach (1985); Wolfire et al. (1990). Kaufman et al. (2006, 1999) has since updated these models to include more recent values of atomic and molecular data, current chemical rate coefficients, and grain photolelectric heating rates. These models assume that the light-emitting material is a plane-parallel, semi-infinite slab illuminated from one side. We use the PDR Toolbox1 (Pound & Wolfire, 2008), a publicly available implementation of the model diagnostics to constrain the intensity of the ambient UV radiation field (G0 or ISRF; conventionally expressed in units of the local Galactic FUV flux = 1.6?10?3 erg cm?2 s?1 , a.k.a Habing flux, Habing, 1968) and the hydrogen nucleus density (nH). Before we can estimate the PDR parameters of our sample, we must first consider a flux correction for [C II] 158 and [O I] 63 . 1http://dustem.astro.umd.edu/pdrt/ 141 4.7.1 [C II] 158 Emission The [C II] 158 line emission originates from both the ionized and neutral ISM phases due to the transition?s low ionization potential (11.26 eV), low transition temperature (91 K), and low critical density for collisions (2 ? 103 cm?3). The PDR model, however, only considers the [C II] 158 flux arising from the neutral gas. Therefore, we must correct for the fraction of [C II] 158 flux that is emitted from the ionized gas. We can derive this fraction using the [N II] 122 flux whose higher ionization potential (14.53 eV) implies that the nitrogen emission originates solely from the ionized gas. We assume that the [C II] 158 flux from the ionized gas scales directly with that of the [N II] 122 line using the relation valid for dense HII regions and for a Milky Way C/N abundance ratio equal to 3.8 (Rubin et al., 1993, 1988): [CII]ionized = 1.1? [NII]. (4.1) 4.7.2 [O I] 63 Emission The PDR models employed here assume that the line emission originating from the irradiated side of a cloud is optically thin. However, when a spectral line is optically thick, as is the case with [O I] 63 , and the cloud is lit from behind, we will not observe emission of that line. This means we only see the [O I] 63 flux if the irradiated side of a cloud is facing us. Kaufman et al. (1999) states that we will observe half of the [O I] 63 emission produced with the other half radiating away from our line of sight. We therefore correct for this geometrical effect by multiplying 142 the observed [O I] 63 flux by a factor of two. Most likely, the true [O I] 63 flux correction factor lies somewhere between one and two. Therefore, we estimate the PDR solutions with and without the [O I] 63 flux correction. The final PDR results represent the range of values that the observed gas may have. As mentioned in Section 4.5, [O I] 63 can suffer from self-absorption which would also affect the amount of the emission detected. There is heavy absorption in [O I] 63 in NGC 4945 (see Section 4.5.2), therefore we do not compute a PDR solution for NGC 4945 using the [O I] 63 flux. The other objects do not show obvious absorption in [O I] 63 , therefore we apply only a geometrical correction to the [O I] 63 emission. 4.7.3 Total Infrared Emission We estimate the TIR emission (8? 1000?m) using the specific flux values at [O I] 63 , [O III] 88 , [O I] 145 , and [C II] 158 from the continua fitted in Section 4.4. Assuming optically thin emission and a fixed grain emissivity (? = 1), we fit the continuum fluxes with a modified blackbody and extract the calculated flux density at 60?m and 100?m. We then compute the FIR emission using the equation from Helou et al. (1988) : FIR [Wm?2] =1.2(6? 10 ?14 ) (4.2) ? 2.58? I60?m[Jy] + I100?m[Jy] . Using the IR bolometric correction from Dale et al. (2001) we approximate 143 the TIR emission via: ( ) TIR log = a0 + a1x+ a 2 2x + a3x 3 + a x44 , (4.3) FIR where x = log[I60?m/I100?m] and a ? [0.2738, ?0.0282, 0.7281, 0.6208, 0.9118] assuming a redshift z = 0. 4.7.4 Hydrogen Density and UV Radiation Field The solutions from the PDR toolbox are degenerate. Therefore, a predicted emission line ratio may be produced by either a low density environment with a strong radiation field or a high density environment with a weak radiation field. For our sample of galaxies, inspection of the ?2 map in the nH?G0 parameter space shows a delineation between the degenerate solutions of the PDR modeling. Therefore, we also modeled the PDR regions within a ?low? density range, 0 < n < 104 cm?3H and a ?high? density range, 104 < nH< 107 cm?3 Here we present the hydrogen density (nH; see Figure 4.6) and far-ultraviolent interstellar radiation field (ISRF; G0; see Figure 4.5.2) solutions from the PDR modeling. Hereafter, for brevity and clarity, the four PDR solutions will be re- ferred to as Solutions A, B, C, and D. Solution A is the low density solution with only the [C II] 158 flux correction. The [C II] 158 flux correction attempts to remove the portion of [C II] 158 emission that originates from ionized gas by us- ing Equation 4.1. Solution B is the low density solution with both the fluxes of [C II] 158 and [O I] 63 corrected. The [O I] 63 flux is corrected using a geometrical 144 factor of two as described in Section 4.7.2. Solution C is the high density solution with only the [C II] 158 flux correction. Solution D is the high density solution with both the fluxes of [C II] 158 and [O I] 63 corrected. Setting a high density limit (n > 104H cm ?3) results in a ISRF ? 1G0 across the entire FOV in the galaxy sample (NGC 4945 is the exception with a peak ISRF of ? 17G0 around the galaxy center). Radiation fields this weak and this extended in the central regions of these galaxies are not physical since much higher radiation fields are expected near an AGN or a circumnuclear starburst. Although NGC 4945 does show a higher ISRF (. 17G0) around the nucleus compared to the other galaxies, it is still too small to be realistic. Because the radiation fields modeled at a high density limit produce unrealistic ISRFs, we will only discuss the low-density solutions A and B in the following analysis. 4.7.5 Caveats We have derived the spatially-averaged neutral gas properties of our galaxy sample via a simple and uniform treatment of the integrated line emission fluxes. However, the physical parameters inferred from any PDR model should not be taken too literally, since they are dependent upon the assumptions adopted and the microphysics implemented in a model that oversimplifies a much more complex set of physical conditions. Moreover, due to the discretization of the parameter space in the PDR models, we are wary of the pixel-to-pixel variations in the final (nH, G0) solutions. In particular, if the input flux uncertainties had been lower, it is possible 145 that the final (nH, G0) results would have been different than the results presented here. This is seen in the final PDR solution maps as a hot/cold pixel in the middle of an otherwise well defined morphological structure. This is corrected by interpolating these pixels with surrounding neighbors. 4.8 Methods to Derive the Outflow Properties In this section the methods for estimating the physical and kinematic outflow properties in the galaxy sample are outlined. The spectral observation cubes are first used to model the disk rotation in each galaxy. The resulting velocity field is then subtracted from the observed emission line profile properties and the residuals are used to define the spatial location of the outflow. The emission line fluxes within the wind region are used to estimate line luminosities. The line luminosities in turn, are used to compute the masses and kinetic energies in the neutral and ionized gas in the outflow. 4.8.1 Beam Smearing One of the greatest difficulties in deriving galaxy kinematics from velocity fields is overcoming the effect of beam smearing (Bosma, 1978). The radial velocities in a galaxy vary on spatial scales smaller than the beam size of the observing instrument. Effectively, this means that the velocities at different radii will be blended, thus flattening the observed velocity field gradients, decreasing the slope of the derived rotation curve at the galaxy center, and broadening the width of the observed line 146 profiles. This is a significant problem since the broadening effect can be incorrectly attributed to the gas velocity dispersion. Fortunately, we can employ methods that utilize the full 3D data cube (two spatial dimensions and one spectral dimension) of a galaxy and build a model that directly incorporates the instrument?s 2D point spread function and spectral resolu- tion. To do this we use the software package 3DBarolo1 (Di Teodoro & Fraternali, 2015) which simulates the observational data in a cube by building a ?tilted-ring? model that best fits the data. Accounting for the instrumental contribution to the observed data will result in a disk model which more accurately captures the true kinematics of the gas. The details of the ?tilted-ring? model and the employment of 3DBarolo are discussed in the following section. 4.8.2 Tilted Ring Model The velocity field of a disk galaxy can be described by a ?tilted-ring? model (Rogstad et al., 1974), where the disk is built from a series of concentric annuli that increase in radius with the distance from the galaxy center. This model assumes that the line emitting material is confined to a thin disk and that the kinematics are dominated by rotational motion. Emission of the gas in each ring is described by geometric parameters (centroid, radius, width, scale height, inclination angle along the line of sight of the observer, and position angle of the major axis) and kinematic parameters (systemic velocity, rotational velocity, and velocity dispersion). For 3DBarolo, the instrumental spectral and spatial resolutions are also inputs so that 1http://editeodoro.github.io/Bbarolo/, version 1.4 147 the final model accounts for both instrumental effects. In reproducing the observed data cube, 3DBarolo assumes that all of the velocity dispersion is due to rotation and turbulence. Outflows are known to exist along the minor axes of the sample galaxies, so this assumption means that the velocity dispersion of the disk model will be overestimated, with the largest errors at the center of the galaxy where line broadening suffers the most from beam smearing. We attempt to mitigate this effect by assuming a constant velocity dispersion across the disk and estimate the value of this velocity dispersion using data outside the galaxy center where the beam smearing is the least severe. The data are fit in two stages. In the first stage, the gas velocity dispersion and the rotational velocity are left as free parameters. The remaining parameters are input as ?correct? values and held constant. Afterwards, the mean of the fitted gas velocity dispersions of the three outmost rings is then computed. For the second stage, this mean dispersion is input as a fixed parameter and only the rotational velocity is left free. The resultant disk model is the assumed galaxy velocity field. 4.8.3 Defining the Spatial Location of the Outflow The kinematics of the outflow can be delineated from those of the disk by examining the residuals in v50 and in W1? between the observed data and the mod- eled disk (?v50 and ?W1?, respectively). Recall from Section 4.8.2, the modeled galaxy velocity field accounts for the instrumental spatial and spectral resolutions. Therefore, in the residuals, the instrumental effects have been removed. In this 148 galaxy sample, excess line widths are a more reliable indicator of outflow than the v50 residuals. For example, the outflow of M 82 is known to lie along the galaxy minor axis. Inspection of the ?W1? maps (see Column 3 in Figure 4.8b) shows an obvious SE to NW trend consistent with the direction of the outflow, but no such trend is observed in the ?v50 maps (see Column 3 in Figure 4.8a). For this reason ?W1? is used to define the spatial location of the wind. The excess line widths are assumed to trace non-circular motions in the gas. These motions can be due to the presence of morphological features in a galaxy, such as spiral arms, a nuclear bar, a warped disk, and/or an outflow. Ideally, in the absence of such features, save the outflow, the spatial location of the wind would be defined as those regions where ?W ?11? > 0 km s . However, these features do exist in our galaxy sample and for some, the projected morphologies of those features overlap with each other, making decomposition of the wind from the features difficult. Therefore, when defining the spatial location of the wind, it is important to carefully choose a minimum threshold in ?W1? which would account for gas motions not related to the outflow. For various thresholds, features in the ?W1? maps were compared to the physical features of the galaxies observed at other wavelengths. For example, for [C II] 158 in NGC 1068, a region of excess line width SW of the nucleus more likely traces a molecular spiral arm because the region is elongated N to S along the arm (see Column 3 of Figure 4.19b) and because the S outflow cone lies behind the galaxy disk where even in the FIR the S cone may be extinguished. A thresh- old of ?W1? > 25 km s ?1 was chosen to minimize the inclusion of regions where 149 excess line widths were spatially coincident with features unrelated to the known outflow. Ultimately, this threshold does mitigate the detection of gas motions un- related to the wind, but it does not remove those motions completely. Although a higher threshold would more effectively exclude gas motions unrelated to the out- flow, such a threshold might also exclude the outflow all together. For example, the detection of a [C II] 158 outflow in M 82 disappears completely with a threshold of ?W1? > 50 km s ?1 (see Column 5 of Figure 4.8b), however, [C II] 158 is known to exist in the outflow of M 82 (C13). This is a probable indicator that a thresh- old of ?W ?11? > 50 km s is too harsh a limit. We include W1? residual maps with ?W > 50 km s?11? as reference for a lower limit to the spatial extent of the wind in the sample galaxies (see Column 5 of subfigure (b) in Figure 4.8, Figure 4.11, Figure 4.14, Figure 4.16, Figure 4.19, Figure 4.22, Figure 4.24), but the following analysis defines the spatial location of the wind as those regions where ?W1? > 25 km s?1. 4.8.4 Line Luminosities The relation from Solomon et al. (1992) is used to estimate the emission line luminosities from the observed line fluxes via Lline = 1.04? 10?3 Sline ?v (1 + z)?1 ?restD2L, (4.4)L where Sline ?v is the velocity integrated flux in Jy km s ?1, ?rest = ?obs(1 + z) is the rest frequency in GHz, and DL is the luminosity distance in Mpc (see Table 4.2 for 150 galaxy details). 4.8.5 Mass of the Neutral Atomic Wind We follow Hailey-Dunsheath et al. (2010) to estimate the neutral atomic gas mass in the outflow by using the inferred [C II] 158 luminosity and the PDR solu- tions (see Section 4.7). Assuming the line emission is optically thin, the minimum atomic gas mass can be estimated by ( )( M L ? ) 4 atomic [C II] 158 1.4? 10 = 0.7 M (7 L XC)+ (4.5) 1 + 2e?91K/T? + ncrit/n , 2e?91K/T where L[C II] 158 is the [C II] 158 line luminosity in solar units, XC+ is the carbon abundance per hydrogen atom, n is the gas density, and ncrit is the critical density. We assume that the PDR surface temperature, TS, derived in Section 4.7 is repre- sentative of the region?s gas temperature and we adopt XC+ = 1.6 ? 10?4 (Sofia et al., 2004) which assumes a C+ abundance typical of PDRs in the Milky Way. The atomic outflow mass estimated here (see Table 4.4) is derived only from the gas associated with [C II] 158 emission and is not an estimate of the total neutral atomic gas in the outflow. Also, a lower XC+ value will result in a more massive wind. Therefore, we emphasize that these mass estimates should be considered lower limits. 151 4.8.6 Mass of the Ionized Wind Following Ferkinhoff et al. (2010), the minimum ionized gas mass required to produce the observed [O III] 88 line emission can be estimated in the high-density (ncrit/n 1), high-temperature (Teff > 40, 000 K) limit via M 2ionized gt 4? ?DL mH 1= Fline ? , (4.6) M gu A h? Xelement Xion where Fline is the emission line flux, gt = ?giexp (??Ei/kT ) is the partition func- tion, gu is the statistical weight of the upper transition level, DL is the source luminosity distance, A is the Einstein A coefficient, ? is the rest frequency of the emission line, mH is the mass of hydrogen, Xelement is the relative abundance of the emission line species to hydrogen (e.g. O++/H+, which, for a minimum mass is the total abundance ratio O/H), and Xion is the fractional abundance of a given species in a particular ionization state (e.g. O++/O) . We adopt the gas-phase abundance ratio XO = 5.9? 10?4 from Savage & Sembach (1996). It is important to note that in our gas mass calculations we are assuming that all of the oxygen is in the form of O++. The ionization potential of O++ (in this case, 35.12 eV for [O III] 88) is much higher than that of O+ so it is unlikely that all of the oxygen will be doubly ionized. Therefore, we include an ionization correction term, Xion, to estimate a more realistic gas-phase abundance ratio, XO++ . For example, if the stellar T ++eff ? 36, 000 K, the fraction of O to the total oxygen abundance will only be ? 15%, and the ionized wind mass will be about 7 times larger. The 152 computed masses of the ionized gas in the wind are listed in Table 4.4. 4.8.7 Kinetic Energy of the Wind The kinetic energy in the wind may be described as the sum of the ?bulk? kinetic energy and the ?turbulent? kinetic energy of the outflowing gas (Rupke et al., 2005a; Veilleux et al., 2020) via: [ ( ) ]2 1 2 W1?,windKEwind = Mwind v50,wind + 3 . (4.7)2 2 The estimates of the kinetic energies in the wind are listed in Table 4.4. 4.9 Properties of the Outflows and Multi-Phase Comparisons In this section we discuss the results from Section 4.8 and directly compare the physical properties of the neutral and ionized outflows within each sample galaxy. For each object, the first figure shows the observed velocity field and 1?? line width (v50 and W1? , respectively) measured from the line profile fitting outlined in Section 4.4 and Section 4.6. This figure also shows the modeled v50 and W1? results from Section 4.8.2, followed by the residuals between the data and the model. Lastly, the figure shows the regions where the 1-? line width residuals (?W1?) exceed 25 and 50 km s?1 to delineate where the outflow is believed to be significant. Hereafter, the residual velocities and line widths inside the regions where ?W > 25 km s?11? are defined as ?v50 wind and ?W1?wind, respectively. The second figure for each object 153 Table 4.4. Wind Properties Solution A Solution B [O III] 88 Galaxy Name M KE M KE M KE (1) (2) (3) (4) (5) (6) (7) Outflow Definition: ?W ?11? > 25 km s M 82 86.1 34.7 17.4 6.13 3.56 1.62 Cen A 18.2 4.35 1.57 0.27 0.05 0.02 Circinus ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 253 106 35.9 10.9 3.57 0.40 0.32 NGC 1068 ? ? ? ? ? ? ? ? ? ? ? ? 8.95 20.8 NGC 4945 ? ? ? ? ? ? ? ? ? ? ? ? 0.36 0.22 Outflow Definition: ?W1? > 50 km s ?1 M 82 ? ? ? ? ? ? ? ? ? ? ? ? 1.22 0.70 Cen A ? ? ? ? ? ? ? ? ? ? ? ? 0.01 0.01 Circinus ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? NGC 253 23.6 12.9 1.62 0.93 0.26 0.28 NGC 1068 ? ? ? ? ? ? ? ? ? ? ? ? 5.49 19.0 NGC 4945 ? ? ? ? ? ? ? ? ? ? ? ? 0.07 0.13 Note. ? Masses of the wind are in units of 106 M . Kinetic energies of the wind are in units of 1053 erg. ?Solution A? and ?Solution B? refer to the low density limit (nH < 10 4 cm?3) PDR results from Section 4.7.4 used to compute the neutral outflow mass (see Section 4.8.5) and kinetic energy (see Section 4.8.7). Outflow masses and kinetic energies listed in the top table are estimated using fluxes, radial velocities, and 1?? line widths in the spatial regions where the 1?? line width residuals between the observed and modeled velocity field (?W1? ) is larger than 25 km s?1 (see Section 4.8.3 for further details). The bottom table is the same as that above except the 1?? line width residual threshold is 50 km s?1. This higher threshold illustrates the lower limit of the outflow masses and kinetic energies. Column 1: Galaxy Name. Column 2: Neutral gas mass in the outflow derived in the low density limit with only the [C II] 158 flux corrected, Column 3: Kinetic energy of the neutral gas in the outflow derived in the low density limit with only the [C II] 158 flux corrected, Column 4: Neutral gas mass in outflow derived in the low density limit with the [C II] 158 and [O I] 63 fluxes corrected, Column 5: Kinetic energy of the neutral gas in the outflow derived in the low density limit with the [C II] 158 and [O I] 63 fluxes corrected, Column 6: Ionized gas mass in the outflow estimated from the [O III] 88 flux (see Section 4.8.6). Column 7: Kinetic energy of the ionized gas in the outflow estimated from the [O III] 88 flux (see Section 4.8.7). 154 compares these outflow maps with existing photometric and kinematic data at other wavelengths. The final figure for each object presents maps of the neutral and ionized gas masses and the corresponding kinetic energies associated with the outflows. As mentioned in Section 4.4, the continuum placement of [N II] 122 is difficult due to the proximity of the second pass ghost to the [N II] 122 line profile. Therefore, [N II] 122 is excluded from the following analysis because of the high uncertainty in its line profile measurements. Also, [O I] 63 is not used to model the NGC 4945 galaxy disk due to the presence of strong self-absorption in the line profiles which is not seen in the other galaxies. Additionally, the wind kinematics for NGC 3079 only include the gas traced by [C II] 158 since this is the only fine structure line observed in NGC 3079. 4.9.1 M 82 The M 82 v50 maps in all lines (see Figure 4.8a, Column: 1) show that the E side of the disk is receding from us (largest velocity at ? 145 km s?1 ) and the W side of the disk is approaching us (largest velocity at ? ?55 km s?1 ). Although the direction of the disk rotation and the lower approaching radial velocity in this PACS data is consistent with the analysis in C13, the maximum receding velocity reported here is ? 45 km s?1 greater than that reported in C13. The M 82 ?v50 wind maps (see Column 4 of Figure 4.8a) show the NW outflow in all lines and indicate that the cone is receding from us with ?v50 wind up to ? 65 km s?1 (the largest recession velocity is in [O I] 145 ). The S outflow is approaching us, but is seen only in the ionized 155 M 82 (a) ?v50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) 145 145 84 84 84 69? 41? 00?? 82 82 46 46 46 45?? 19 19 7 7 7 ? ? ?? -43 -43 -30 -30 -30 69 40 30 -107 -107 -68 -68 -68 145 145 84 84 84 69? 41? 00?? 82 82 46 46 46 45?? 19 19 7 7 7 -43 -43 -30 -30 -30 69? 40? 30?? -107 -107 -68 -68 -68 145 145 84 84 84 69? 41? 00?? 82 82 46 46 46 45?? 19 19 7 7 7 -43 -43 -30 -30 -30 69? 40? 30?? -107 -107 -68 -68 -68 145 145 84 84 84 69? 41? 00?? 82 82 46 46 46 45?? 19 19 7 7 7 -43 -43 -30 -30 -30 69? 40? 30?? 175 pc -107 -107 -68 -68 -68 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 223 223 84 84 84 69? 41? 00?? 171 171 35 35 35 45?? 119 119 -13 -13 -13 ? ? ?? 67 67 -63 -63 -6369 40 30 15 15 -112 -112 -112 223 223 84 84 84 69? 41? 00?? 171 171 35 35 35 45?? 119 119 -13 -13 -13 ? ? ?? 67 67 -63 -63 -6369 40 30 15 15 -112 -112 -112 223 223 84 84 84 69? 41? 00?? 171 171 35 35 35 45?? 119 119 -13 -13 -13 ? ? 67 67 -63 -63 -6369 40 30?? 15 15 -112 -112 -112 223 223 84 84 84 69? 41? 00?? 171 171 35 35 35 45?? 119 119 -13 -13 -13 ? ? ?? 67 67 -63 -63 -6369 40 30 175 pc 15 15 -112 -112 -112 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s 56s 52s 9h 55m 48s Figure 4.8 M 82: Results from modeling the disk velocity field with 3DBarolo (which accounts for both the instrumental spectral and spatial resolu- tions) and the location of the outflow based on excess line broadening. Color bar values are in units of km s?1 . For each line, left to right: observed data, model re- sult, data - model residual, spatial location of the wind in regions where ?W1? > 25 km s?1 , and spatial location of the wind in regions where ?W1? > 50 km s?1 . Con- tours in (a) are in five equal steps between the minimum and maximum velocities in each image. Contours in (b) are 0.3, 0.5, 0.7, and 0.9 of the peak value in each image. The solid blue line marks the galaxy major axis. The magenta cross marks the adopted galaxy center. 156 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 Figure 4.9 M 82: PACS contours (magenta) overlaid on the total integrated intensity contours of SiO(2-1) (black, Garc??a-Burillo et al., 2001) and the radio continuum at 4.8 GHz (grey scale, Wills et al., 1999). gas traced by [O III] 88 with ?v50 wind velocities ranging from ? ?30 km s?1 to ? ?10 km s?1 . In general, the ?v50 wind velocities increase with increasing (projected) distance from the galaxy nucleus in [O III] 88 , [O I] 145 , and [C II] 158 . This is not the case with [O I] 63 , however, where the ?v50 wind velocities are greatest near the nucleus and decrease further up above the disk. The largest residual velocity dispersions in the wind region (Column 4 of Fig- ure 4.8b) are in [O I] 145 with ?W1?wind ranging from ? 25? 85 km s?1 , where the velocity dispersions increase as the (projected) distance from the nucleus increases. The range of ?W1?wind velocities is similar between [O I] 63 and [O III] 88 (?W1?wind? 30 ? 65 km s?1 ). Unlike [O I] 145 , the largest residual velocity dispersions in [O I] 63 are localized in a compact area near the galaxy center and decrease with increasing distance from the center. [O III] 88 contains a similar compact region near 157 M 82 Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 318.19 160.10 254.76 128.12 69? 41? 00?? 191.34 96.14 40?? 127.92 64.16 64.49 32.18 69? 40? 20?? Mtot: 8606.21 M KEtot: 3466.50 ergs Mtot: ? KEtot: ? 1.07 0.20 318.19 160.10 254.76 128.12 69? 41? 00?? 191.34 96.14 40?? 127.92 64.16 64.49 32.18 69? 40? 20?? Mtot: 1735.93 M KEtot: 612.59 ergs Mtot: ? KEtot: ? 1.07 0.20 4.99 1.83 4.08 1.48 69? 41? 00?? 3.18 1.13 40?? 2.27 0.78 1.36 0.43 69? 40? 20?? Mtot: 930.73 M KEtot: 326.43 ergs Mtot: ? KEtot: ? 0.45 0.08 4.99 1.83 4.08 1.48 69? 41? 00?? 3.18 1.13 40?? 2.27 0.78 1.36 0.43 69? 40? 20?? Mtot: 957.11 M KEtot: 336.14 ergs Mtot: ? KEtot: ? 0.45 0.08 1.53 1.53 1.44 1.44 1.24 1.24 1.17 1.17 69? 41? 00?? 0.95 0.95 0.90 0.90 40?? 0.66 0.66 0.63 0.63 0.37 0.37 0.36 0.36 69? 40? 20?? Mtot: 356.38 M KEtot: 162.25 ergs Mtot: 122.23 M KEtot: 70.32 ergs 0.08 0.08 0.08 0.08 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s 55s 9h 55m 50s Figure 4.10 Mass and KE in the wind derived from the PDR solutions (see Sec- tion 4.7.4 for definitions of Solutions A, B, C, and D) and the [O III] 88 flux. Units are in 104M and 1051 erg for the masses and KEs, respectively. 158 [[OOIIIIII]] 8888 SSoolluuttiioonn DD SSoolluuttiioonn CC SSoolluuttiioonn BB SSoolluuttiioonn AA the nucleus as seen in [O I] 63 , however, although the dispersions in [O III] 88 initially decrease with increasing distance from the nucleus, at about r ? 10 ??, ?W1?wind increases at larger radii. [C II] 158 appears to have a near constant ?W1?wind? 40 km s?1 along the entire length of the collimated structure. C13 estimate the velocity dispersions are ? 50? 100 km s?1 larger inside the outflow cones compared to the disk. Our observed data (even in its smaller FOV) are consistent with this assessment. The sputtering of dust grains by shocks will release gas-phase silicon. The silicon is rapidly oxidized and forms SiO (Flower et al., 1996). Therefore, SiO is a good tracer of shocks, since it is thought to be formed in shocks. Figure 4.9 shows PACS contours (magenta) overlaid on the SiO (black and white contours) observed by Garc??a-Burillo et al. (2001) and the 4.8 GHz continuum (greyscale) observed by Wills et al. (1999). The SiO chimney appears to originate in the galaxy disk and extends to the NE beyond the PACS FOV. All lines trace this chimney out to ? 20?? (? 350 pc) above the disk and consistently indicate it is receding from us (see Figure 4.9a - Figure 4.9d). Residuals in [O III] 88 are the most tightly spatially correlated with the chimney while in [O I] 145 the residuals are the most extended W and E of the chimney. The residuals in the wind of M 82 are consistent with the idea that at the base of the outflow, gas flows out of the galaxy disk in filamentary channels instead of being launched uniformly across the starburst region. The largest masses and KEs in the neutral outflow lie on the N edge of the outflow for Solution A, while the largest masses and KEs from Solution B lie inside the galaxy disk (see Figure 4.10). The morphology of the ionized mass distribution 159 follows that of the [O III] 88 flux distribution (i.e. the overall morphology of the ionized gas mass in the outflow is ellipsoidal and lies spatially coincident with the galaxy disk). The highest KEs in the ionized outflow coincide with the portion of the SiO chimney that intersects the galaxy plane. C13 estimate the mass of the neutral gas in the outflow to be 2? 8? 107 M ; a value ? 4 ? 15 times lower than the cold molecular gas reported in Walter et al. (2002, 3.3 ? 108 M ) and 3 orders of magnitude higher than the warm molecular gas reported in Veilleux et al. (2009). C13 estimated a KE of ? 1 ? 5 ? 1054 erg in the neutral outflow; an order of magnitude lower than that of H? (Shopbell & Bland-Hawthorn, 1998, 3? 1055 erg) and 2 orders of magnitude larger than the KE measured in H2. 4.9.2 Cen A The velocity field shown in the first column of Figure 4.11a indicates that the E side of the gaseous disk at the center of Cen A is approaching us, while the W side of the disk is receding from us, consistent with the velocity field observed in the optical ionized (Bland et al., 1987) and molecular gas (Espada et al., 2009). The velocities range from v50? ?60 to +105 km s?1 for all lines. The wind region in this object, defined by the excess line broadening or residual velocity dispersions, lies along the galaxy minor axis in the direction of the well-known radio jet. Bulk velocities of the gas near the galaxy nucleus are mostly systemic with ?v50 wind? ?10 km s?1 for all lines (see Figure 4.11a.1 - 160 Cen A (a) ?v50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) 100 100 49 49 49 ?43? 01? 00?? 60 60 27 27 27 ?? 20 20 5 5 5?15 -19 -19 -15 -15 -15 ?30?? -37 -37 -37 -60 -60 100 100 49 49 49 ?43? 01? 00?? 60 60 27 27 27 ?? 20 20 5 5 5?15 -19 -19 -15 -15 -15 ?30?? -37 -37 -37 -60 -60 100 100 49 49 49 ?43? 01? 00?? 60 60 27 27 27 ?? 20 20 5 5 5?15 -19 -19 -15 -15 -15 ?30?? -37 -37 -37 -60 -60 100 100 49 49 49 ?43? 01? 00?? 60 60 27 27 27 20 20 5 5 5 ?15?? -19 -19 -15 -15 -15 ?30?? -37 -37 184 pc -37 -60 -60 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 283 283 100 100 100 ?43? 01? 00?? 216 216 38 38 38 ?? 150 150 -22 -22 -22?15 84 84 -84 -84 -84 ?30?? 17 17 -145 -145 -145 283 283 100 100 100 ?43? 01? 00?? 216 216 38 38 38 150 150 -22 -22 -22 ?15?? 84 84 -84 -84 -84 ?30?? 17 17 -145 -145 -145 283 283 100 100 100 ?43? 01? 00?? 216 216 38 38 38 ?? 150 150 -22 -22 -22?15 84 84 -84 -84 -84 ?30?? 17 17 -145 -145 -145 283 283 100 100 100 ?43? 01? 00?? 216 216 38 38 38 ?? 150 150 -22 -22 -22?15 84 84 -84 -84 -84 ?30?? 184 pc 17 17 -145 -145 -145 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s Figure 4.11 Cen A: Results from modeling the disk velocity field with 3DBarolo. Symbols and plots are the same as those in Figure 4.8. 161 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [OI] 63 [OIII] 88 [OI] 145 [CII] 158 Figure 4.12 Cen A: PACS contours overlaid on 8.4 GHz image from Hardcastle et al. (2003). Cen A Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 39.73 13.90 ?43? 00? 40?? 31.88 11.14 ?43? 01? 00?? 24.03 8.37 ?? 16.17 5.60?20 8.32 2.83 ?? Mtot: 1814.68 M KEtot: 434.73 ergs Mtot: ? KE? tot : ? 40 0.46 0.07 39.73 13.90 ?43? 00? 40?? 31.88 11.14 ?43? 01? 00?? 24.03 8.37 16.17 5.60 ?20?? 8.32 2.83 ?? Mtot: 157.17 M KEtot: 27.17 ergs Mtot: ? KEtot: ??40 0.46 0.07 0.78 0.15 ?43? 00? 40?? 0.65 0.12 ?43? 01? 00?? 0.52 0.10 ?? 0.39 0.07?20 0.26 0.05 ?? Mtot: 74.73 M KEtot: 11.80 ergs Mtot: ? KEtot: ??40 0.13 0.02 0.78 0.15 ?43? 00? 40?? 0.65 0.12 ?43? 01? 00?? 0.52 0.10 ?? 0.39 0.07?20 0.26 0.05 ? ?? Mtot: 76.97 M KEtot: 12.28 ergs Mtot: ? KEtot: ? 40 0.13 0.02 0.04 0.04 0.03 0.03 ?43? 00? 40?? 0.03 0.03 0.02 0.02 ?43? 01? 00?? 0.03 0.03 0.02 0.02 ?20?? 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.01 ?? Mtot: 5.22 M KEtot: 2.17 ergs Mtot: 1.16 M KE : 1.09 ergs? tot40 0.01 0.01 0.01 0.01 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s 30s 28s 13h 25m 26s Figure 4.13 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. 162 ?W1? ?v50 [[OOIIIIII]] 8888 SSoolluuttiioonn DD SSoolluuttiioonn CC SSoolluuttiioonn BB SSoolluuttiioonn AA Figure 4.11a.4). In [O III] 88 , there is a compact feature ? 15??NE of the nu- cleus with v50 residuals ? ?30 km s?1 in the E and ? +45 km s?1 in the W. This feature is not seen in [C II] 158 and shows little bulk motion in either [O I] 63 or [O I] 145 (?v ?150 wind? +10 km s ). The residual velocity dispersions in the wind region (Figure 4.11b.1 - Fig- ure 4.11b.4) near the galaxy nucleus for [O III] 88 and [C II] 158 are spatially com- pact with ?W1?wind . 30 km s?1 . Also near the nucleus, the residual velocity dispersions in [O I] 63 and [O I] 145 are elongated towards the NW in the direction of the jet with ?W ?11?wind? 30?50 km s . The wind region observed in [O I] 145 is the most extended along the major axis compared to the wind region mapped by the other three emission lines. The largest residual velocity dispersions in the compact feature NE of the nucleus is ?W1?wind? 90 km s?1 in [O III] 88 , while moderate velocity dispersions of ? 30 km s?1 are seen in [O I] 63 and [O I] 145 . In [O I] 145 an elongated feature is seen S of the galaxy nucleus with ?v50 wind? 25 km s?1 and ?W1?wind? 95 km s?1 . It is unlikely that this feature is a component of the outflowing gas as it spatially coincides with the edge of the putative warped disk seen in CO, Pa-?, and the MIR (Espada et al., 2009; Marconi et al., 2000b; Quillen et al., 2006; Radomski et al., 2008). Figure 4.12 shows PACS contours of ?v50 wind (top row) and ?W1?wind (bottom row) overlaid on the 8.4 GHz images of Cen A from Hardcastle et al. (2003). NE of the nucleus, within r . 1 kpc, a compact feature is seen in [O I] 63 , [O III] 88 , and [O I] 145 . This feature is spatially coincident with a complex of bright emission knots seen in the radio and X-ray and which are known to be associated with the 163 outflow (e.g. Hardcastle et al., 2003; Kraft et al., 2002). For PDR Solution A (see Figure 4.13), the largest neutral outflow masses (? 4? 105M ) are located to the SW of the nucleus in a thin structure elongated from NW to SE. This is in striking contrast to PDR Solution B where the largest neutral masses (? 4?103M ) are compact and localized around the galaxy nucleus. Figure 4.13 also shows a compact region of ionized gas around the nucleus that contains ?50 M and a structure to the NE with mass ?30 M . The largest KE of 9? 1048 erg lies in a concentrated region to the NE with more moderate energies of ? 2 ? 1048 centered close to the nucleus. This region of large energy spatially corresponds to knot ?A? found in Kraft et al. (2002). 4.9.3 Circinus The v50 maps of Circinus (see Column 1 of Figure 4.14a) range in velocity between ?100 km s?1 and confirm data at other wavelengths that the N side of the disk is approaching us and the S side of the disk is receding from us consistent with observations in H? (Elmouttie et al., 1998b) and in CO (Zschaechner et al., 2016). Circinus lacks a significant detection of an outflow in any of the fine-structure lines. Although the [O III] 88 residuals NW of the nucleus lie within the ioniza- tion cone, the residuals also overlap with a molecular spiral arm seen in CO (Izumi et al., 2018). Lying on the PACS FOV and spatially coincident with the molec- ular spiral arm, it is difficult to attribute this residual line width to the outflow. [O I] 145 residuals in the wind lie SE of the nucleus and are more likely associated 164 Circinus ?v (a) 50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) ? ? ? ?? 128 12865 20 00 76 76 76 70 70 ?15?? 41 41 41 12 12 5 5 5 ?30?? -45 -45 -29 -29 -29 ?45?? -103 -103 -64 -64 -64 ? ? ? ?? 128 12865 20 00 76 76 76 70 70 ?? 41 41 41?15 12 12 5 5 5 ?30?? -45 -45 -29 -29 -29 ?45?? -103 -103 -64 -64 -64 ? 128 12865? 20? 00?? 76 76 76 70 70 ?? 41 41 41?15 12 12 5 5 5 ?30?? -45 -45 -29 -29 -29 ?45?? -103 -103 -64 -64 -64 ?65? 20? 00?? 128 128 76 76 76 70 70 ?? 41 41 41?15 12 12 5 5 5 ?30?? -45 -45 -29 -29 -29 ?45?? 204 pc -103 -103 -64 -64 -64 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 ? ? ? 255 255 81 81 8165 20 00?? 195 195 23 23 23 ?15?? 135 135 -34 -34 -34 ?30?? 75 75 -92 -92 -92 ?45?? 15 15 -150 -150 -150 ? 255 255 81 81 8165? 20? 00?? 195 195 23 23 23 ?15?? 135 135 -34 -34 -34 ?30?? 75 75 -92 -92 -92 ?45?? 15 15 -150 -150 -150 ? ? ? ?? 255 255 81 81 8165 20 00 195 195 23 23 23 ?15?? 135 135 -34 -34 -34 ?30?? 75 75 -92 -92 -92 ?45?? 15 15 -150 -150 -150 ? ? 255 255 81 81 8165 20? 00?? 195 195 23 23 23 ?15?? 135 135 -34 -34 -34 ?30?? 75 75 -92 -92 -92 204 pc ?45?? 15 15 -150 -150 -150 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s Figure 4.14 Circinus: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. 165 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 Circinus Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 16.87 4.76 ?65? 20? 00?? 13.55 3.82 10.23 2.88 ?20?? 6.90 1.93 ?40?? 3.58 0.99 Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.26 0.05 16.87 4.76 ?65? 20? 00?? 13.55 3.82 10.23 2.88 ?20?? 6.90 1.93 ?40?? 3.58 0.99 Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.26 0.05 1.78 1.39 ?65? 20? 00?? 1.44 1.11 1.11 0.84 ?20?? 0.77 0.57 ?40?? 0.43 0.29 Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.09 0.02 1.78 1.39 ?65? 20? 00?? 1.44 1.11 1.11 0.84 ?20?? 0.77 0.57 ?40?? 0.43 0.29 Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.09 0.02 0.18 0.18 0.08 0.08 ?65? 20? 00?? 0.15 0.15 0.06 0.06 0.12 0.12 0.05 0.05 ?20?? 0.08 0.08 0.04 0.04 ?40?? 0.05 0.05 0.03 0.03 Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.01 0.01 0.02 0.02 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s 12s 14h 13m 08s Figure 4.15 Circinus does not have a discernible outflow in either the neutral or ionized gas. 166 [[OOIIIIII]] 8888 SSoolluuttiioonn DD SSoolluuttiioonn CC SSoolluuttiioonn BB SSoolluuttiioonn AA with the circumnuclear starburst ring (see Figure 5 of Mingo et al., 2012). 4.9.4 NGC 253 For NGC 253, the v50 maps (see Column 1 of Figure 4.16a) range in velocity from ? ?80 km s?1 to ? 130 km s?1 for all lines. The E side of the disk is approach- ing us while the W side of the disk is receding from us, consistent with observations in H? (Hlavacek-Larrondo et al., 2011). In Figure 4.16(a, b), the wind traced by [O I] 63 is a column oriented along the galaxy minor axis. The column is approximately ? 5??wide to the NE of the nucleus. That width increases as distance from the nucleus increases, with maximum widths of ? 15?? and ? 30?? for the regions N and S of the galaxy center, respectively. The largest v (? 40 km s?150 ) are located in the S portion of the column while the v50 in the N portion of the column are nominal (? 0? 5 km s?1 ). This suggests that the S cone traced by [O I] 63 is receding but there is little (if any) bulk motion of the gas in the N cone along the line of sight. The velocity dispersion across the column is nearly uniform with W1?? 20 km s?1 . The wind in [O III] 88 lies in two distinct regions to the NW and SE of the nucleus. The NW feature with diameter ? 15?? is located ? 10?? from the galaxy center while the W edge of the ? 20?? ? 15?? SE feature is adjacent to the nucleus. Figure 4.16a shows the largest approaching velocities (? ?40 km s?1 ) are con- centrated in the NW feature. The largest velocity dispersions (? 200 km s?1 ) are prominent in the NW feature while more moderate velocity dispersions (. 90 167 NGC 253 (a) ?v50 ?v50v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) ?25? 17? 00?? 130 130 69 69 69 79 79 40 40 40 ?15?? 28 28 11 11 11 ?30?? -22 -22 -17 -17 -17 -73 -73 -46 -46 -46 ?25? 17? 00?? 130 130 69 69 69 79 79 40 40 40 ?15?? 28 28 11 11 11 ?30?? -22 -22 -17 -17 -17 -73 -73 -46 -46 -46 ?25? 17? 00?? 130 130 69 69 69 79 79 40 40 40 ?15?? 28 28 11 11 11 ?30?? -22 -22 -17 -17 -17 -73 -73 -46 -46 -46 ?25? 17? 00?? 130 130 69 69 69 79 79 40 40 40 ?15?? 28 28 11 11 11 ?30?? -22 -22 -17 -17 -17 189 pc -73 -73 -46 -46 -46 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 ?25? 17? 00?? 414 414 211 211 211 337 337 147 147 147 ?15?? 260 260 83 83 83 ?30?? 183 183 19 19 19 106 106 -43 -43 -43 ?25? 17? 00?? 414 414 211 211 211 337 337 147 147 147 ?15?? 260 260 83 83 83 ?30?? 183 183 19 19 19 106 106 -43 -43 -43 ?25? 17? 00?? 414 414 211 211 211 337 337 147 147 147 ?15?? 260 260 83 83 83 ?30?? 183 183 19 19 19 106 106 -43 -43 -43 ?25? 17? 00?? 414 414 211 211 211 337 337 147 147 147 ?15?? 260 260 83 83 83 ?30?? 183 183 19 19 19 189 pc 106 106 -43 -43 -43 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s Figure 4.16 NGC 253: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. 168 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [OI] 63 [OIII] 88 [OI] 145 [CII] 158 (a) (b) (c) (d) (e) (f) (g) (h) Figure 4.17 NGC 253: PACS contours (yellow), contours of receding and approach- ing CO outflow (magenta and blue, respectively) from Bolatto et al. (2013). Com- posite image from Heesen et al. (2011) which shows H? in red from Westmoquette et al. (2011), ?20 cm continuum (green), and soft X-ray in blue from Hardcastle et al. (2010). km s?1 ) are found in the SE feature. The wind traced by [O I] 145 is concentrated to the SE of the nucleus with receding velocities 50 & v50 & 0 km s?1 (see Figure 4.16a) and velocity dispersions W1?. 115 km s?1 (see Figure 4.16b). The feature is ? 20?? in diameter on the E side while the W side tapers to the SW for ? 8??. As seen in [O III] 88, the W edge of the [O I] 145 SW feature lies adjacent to the nucleus. Interestingly, not only does this SW feature contain similar velocity dispersions (? 115 km s?1 ) in [O III] 88 and in [O I] 145 , but the extent and location of those velocity dispersions are spatially coincident in the emission lines. No feature is seen in the NW in [O I] 145 . Figure 4.16a shows that the majority of [C II] 158 is approaching us with the largest velocities (v50? ?45) located NW of the nucleus. The velocity dispersions of [C II] 158 are moderate (? 40? 80 km s?1 ; see Figure 4.16b) in the NW and SE 169 ?W1? ?v50 NGC 253 Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 125.25 40.83 55.21 29.25 ?25? 17? 00?? 100.30 32.68 44.28 23.45 ?? 75.35 24.54 33.34 17.65?15 50.40 16.39 22.41 11.86 ?30?? 25.45 8.24 11.48 6.06 ? ?? Mtot: 10611.05 M KEtot: 3586.39 ergs M45 tot: 2362.56 M KEtot: 1286.40 ergs 0.50 0.10 0.55 0.26 125.25 40.83 55.21 29.25 ?25? 17? 00?? 100.30 32.68 44.28 23.45 ?? 75.35 24.54 33.34 17.65?15 50.40 16.39 22.41 11.86 ?30?? 25.45 8.24 11.48 6.06 ? ?? Mtot: 1092.11 M45 KEtot: 357.49 ergs Mtot: 161.71 M KEtot: 93.31 ergs 0.50 0.10 0.55 0.26 2.01 0.49 0.75 0.33 ?25? 17? 00?? 1.64 0.40 0.63 0.28 ?? 1.27 0.31 0.52 0.23?15 0.90 0.22 0.40 0.18 ?30?? 0.53 0.13 0.28 0.13 ? ?? Mtot: 387.03 M KEtot: 121.75 ergs M45 tot: 57.22 M KEtot: 32.38 ergs 0.16 0.04 0.16 0.09 2.01 0.49 0.75 0.33 ?25? 17? 00?? 1.64 0.40 0.63 0.28 ?? 1.27 0.31 0.52 0.23?15 0.90 0.22 0.40 0.18 ?30?? 0.53 0.13 0.28 0.13 ? ?? Mtot: 400.28 M KEtot: 125.78 ergs Mtot: 59.39 M KEtot: 33.55 ergs45 0.16 0.04 0.16 0.09 0.27 0.27 0.25 0.25 ?25? 17? 00?? 0.22 0.22 0.20 0.20 ?? 0.17 0.17 0.15 0.15?15 0.11 0.11 0.10 0.10 ?30?? 0.06 0.06 0.05 0.05 ? ?? Mtot: 40.03 M KEtot: 32.31 ergs Mtot: 26.18 M KEtot: 28.31 ergs45 0.01 0.01 0.01 0.01 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s 34s 0h 47m 32s Figure 4.18 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. 170 [[OOIIIIII]] 8888 SSoolluuttiioonn DD SSoolluuttiioonn CC SSoolluuttiioonn BB SSoolluuttiioonn AA features. The length of the SE feature is detected along the entire S edge of the FOV while the width of the feature tapers from the largest width of ? 10?? and decreases towards the SE. The NW feature is roughly 15??? 10?? and lies along the edge of the N FOV. Unlike the [O I] 63 and [O I] 145 the SE feature traced by [C II] 158 does not lie adjacent to the galaxy center. The greatest amount of ionized gas is found in the SE lobe with masses at ? 350 M near the galaxy center and then decreasing in the SE direction with increasing radius from the nucleus. The ionized gas masses in the NW lobe are considerably smaller than in the southern lobe, with masses ranging from ? 10 to ? 80 M . We see that the largest energies (? 2 ? 1049 ergs) in the ionized wind are located in long, thin structure in the SE lobe. We see that the moderate KE are more extended in the SE lobe. 4.9.5 NGC 1068 The v50 maps of NGC 1068 (see Column 1 in Figure 4.19a) indicate that the galaxy disk is receding in the W and approaching in the E. However, the bulk mo- tions of the wind, as determined by the v50 residuals (Column 4 in Figure 4.19a), are oriented in the opposite direction of the disk rotation with approaching ve- locities in the W and receding velocities in the E. In [O I] 63 , the largest ap- proaching ?v50 wind velocities (? ?85 km s?1 ) are in a compact feature just W of the nucleus while moderate receding velocities are in the E (? 35 km s?1 ). [O III] 88 ?v50 wind residuals show moderate approaching values in the W (? ?20 171 NGC 1068 (a) ?v50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) ?0? 00? 30?? 157 157 93 93 93 ?40?? 89 89 51 51 51 ? ?? 21 21 8 8 850 -46 -46 -33 -33 -33 ?0? 01? 00?? -114 -114 -76 -76 -76 ?0? 00? 30?? 157 157 93 93 93 ?40?? 89 89 51 51 51 ? ?? 21 21 8 8 850 -46 -46 -33 -33 -33 ?0? 01? 00?? -114 -114 -76 -76 -76 ?0? 00? 30?? 157 157 93 93 93 ?40?? 89 89 51 51 51 ? ?? 21 21 8 8 850 -46 -46 -33 -33 -33 ?0? 01? 00?? -114 -114 -76 -76 -76 ?0? 00? 30?? 157 157 93 93 93 ?40?? 89 89 51 51 51 ? 21 21 8 8 850?? -46 -46 -33 -33 -33 ?0? 01? 00?? 698 pc -114 -114 -76 -76 -76 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 ?0? 00? 30?? 454 454 205 205 205 ?40?? 364 364 133 133 133 ? 274 274 60 60 6050?? 184 184 -12 -12 -12 ?0? 01? 00?? 93 93 -85 -85 -85 ?0? 00? 30?? 454 454 205 205 205 ?40?? 364 364 133 133 133 ? ?? 274 274 60 60 6050 ? ? ?? 184 184 -12 -12 -12 ?0 01 00 93 93 -85 -85 -85 ?0? 00? 30?? 454 454 205 205 205 ?40?? 364 364 133 133 133 ?50?? 274 274 60 60 60 184 184 -12 -12 -12 ?0? 01? 00?? 93 93 -85 -85 -85 ?0? 00? 30?? 454 454 205 205 205 ?40?? 364 364 133 133 133 ? ?? 274 274 60 60 6050 184 184 -12 -12 -12 ?0? 01? 00?? 698 pc 93 93 -85 -85 -85 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s 42s 41s 2h 42m 40s Figure 4.19 NGC 1068: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. 172 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[OOII]] 6633 Figure 4.20 NGC 1068: PACS contours overlaid on 349 GHz continuum and CO(3-2) residual mean-velocity field from Garc??a-Burillo et al. (2014). 173 NGC 1068 Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 858.12 513.98 ?0? 00? 30?? 689.84 411.53 ?? 521.57 309.09?45 353.30 206.64 ?0? 01? 00?? 185.03 104.20 Mtot: ? KEtot: ? Mtot: ? KE?? tot : ? ?15 16.76 1.76 858.12 513.98 ?0? 00? 30?? 689.84 411.53 521.57 309.09 ?45?? 353.30 206.64 ?0? 01? 00?? 185.03 104.20 Mtot: ? KEtot: ? Mtot: ? KE?? tot : ? ?15 16.76 1.76 82.20 50.15 ?0? 00? 30?? 66.45 40.20 50.69 30.26 ?45?? 34.93 20.32 ?0? 01? 00?? 19.17 10.38 M ?? tot : ? KEtot: ? Mtot: ? KEtot: ? ?15 3.42 0.44 82.20 50.15 ?0? 00? 30?? 66.45 40.20 ?45?? 50.69 30.26 34.93 20.32 ?0? 01? 00?? 19.17 10.38 Mtot: ? KEtot: ? Mtot: ? KE?? tot : ? ?15 3.42 0.44 3.27 3.27 3.27 3.27 ?0? 00? 30?? 2.66 2.66 2.67 2.67 ?? 2.05 2.05 2.06 2.06?45 1.44 1.44 1.45 1.45 ?0? 01? 00?? 0.83 0.83 0.84 0.84 Mtot: 895.16 M KEtot: 2075.06 ergs Mtot: 548.77 M KEtot: 1900.36 ergs ?15?? 0.22 0.22 0.24 0.24 42s 2h 42m 40s 42s 2h 42m 40s 42s 2h 42m 40s 42s 2h 42m 40s Figure 4.21 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. 174 [[OOIIIIII]] 8888 SSoolluuttiioonn DD SSoolluuttiioonn CC SSoolluuttiioonn BB SSoolluuttiioonn AA km s?1 ) and fast receding velocities in the E (? 65 km s?1 ). [O I] 145 and [C II] 158 show mostly systemic velocities in the W (? ?5 km s?1 ). However, [O I] 145 shows fast receding velocities in the E (? 90 km s?1 ) while [C II] 158 shows moderate receding velocities of ? 30 km s?1 . The largest residual velocity dispersions in the wind (?W1?wind? 200 km s?1 ; see Figure 4.19b) are detected in [O III] 88 and are in a central compact region that is elongated towards the NE along the direction of the nuclear bar. This region is spatially coincident with the molecular outflow seen in CO(3-2) (Garc??a-Burillo et al., 2014). It is also spatially coincident with H? and [O III] 5007 line emission which extends NE of the nucleus (Cecil et al., 1990; Veilleux et al., 2003). In the NLR, the deprojected intrinsic ionized gas velocities associated with optical [NII] emission are ? 1500 km s?1 with respect to systemic (Cecil et al., 1990). In [O I] 63, [O I] 145, and [C II] 158, there are two knots which flank the galaxy nucleus. The W knot contains the larger ?W1?wind residuals with ?W1?wind 170 km s?1 in [O I] 63 and ?W ?11?wind? 100 km s in [O I] 145. The E knot contains ?W1?wind residuals of ? 120 km s?1 and ? 70 km s?1 for [O I] 63 and [O I] 145, re- spectively. In [C II] 158, the knots have similar dispersion residuals with ?W1?wind? 70 km s?1 . The ?W1?wind knots seen in [O I] 63, [O I] 145, and [C II] 158 are most likely non-circular motions due to the presence of NGC 1068?s nuclear bar and spiral arms (features which are not accounted for in our simple disk velocity field model), as opposed to an outflow. The ionized gas mass follows the morphological structure of the [O III] 88 flux (i.e. an ellipse spanning from the NE to the SW) with the greatest masses (? 4000 175 M ) concentrated around the galaxy center. The largest energies are concentrated around the nucleus, while there may be some structure following the spiral arms. 4.9.6 NGC 3079 Column 1 of Figure 4.22a shows the [C II] 158 v50 map of NGC 3079. The range of velocities are between ? ?120 and ? +160 km s?1 and indicate that the gas on the N side is approaching us while the gas on the S side is receding from us, consistent with observations of H? (Veilleux et al., 1999) and H2O maser spots in the disk (Yamauchi et al., 2004). The [C II] 158 v50 residuals in the wind region east of the nucleus (Figure 4.22a, Column 4) show redshifted velocities in the N and blueshifted velocities in the S. W of the nucleus, most of the redshifted velocities lie close to the galaxy plane while blueshifted velocities begin to appear further below the plane at ? 10??. The [C II] 158 ?W1?wind map (Figure 4.22b, Column 4) shows an elongated fea- ture ? 10?? in length just S of the nucleus with ?W1?wind? 100 km s?1 . A more extended structure lies E to W around the elongated feature and has ?W1?wind? 75 km s?1 . The contours of these [C II] 158 residuals are overlaid on the 1.4 GHz image from Sebastian et al. (2019) (see the top row of Figure 4.23), where the double-lobed radio morphology of the wind is outlined by bright filaments. At the base of the W lobe, where the filaments are brightest, ?v50 wind appears to trace the lobe?s edges. The velocities along the southern edge are receding with ?v ?150 wind? 35 km s and 176 NGC 3079 (a) ?v50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) 153 153 74 74 74 55? 41? 00?? 80 80 776 pc36 36 36 45?? 8 8 -1 -1 -1 -63 -63 -39 -39 -39 55? 40? 30?? -136 -136 -77 -77 -77 10h 02m 00s 10h 02m 00s 10h 02m 00s 10h 02m 00s 10h 02m 00s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 370 370 107 107 107 55? 41? 00?? 309 309 55 55 776 pc 55 45?? 247 247 3 3 3 186 186 -47 -47 -47 55? 40? 30?? 124 124 -99 -99 -99 10h 02m 00s10h 01m 57s 10h 02m 00s10h 01m 57s 10h 02m 00s10h 01m 57s 10h 02m 00s10h 01m 57s 10h 02m 00s10h 01m 57s Figure 4.22 NGC 3079: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. the velocities along the northern edge are approaching with ?v50 wind? ?15 km s?1 . The middle row of Figure 4.23 shows the same contours as above, but overlaid on the continuum subtracted H?+ [N II] image from Cecil et al. (2001) where only the eastern bubble is visible (the western lobe lies behind the disk of galaxy and is therefore extinguished). In the E lobe, there is a lack of radio emission along the southern edge of the bubble, extending out to ? 1 kpc above the disk. However, there is optical emission along this edge where approaching velocities in [C II] 158 are ?v50 wind? ?35 km s?1 and where the largest velocities of ?W1?wind (? 100 km s?1 ). Cecil et al. (2001) has shown that this filament aligns with the axis of the jet ob- served at 8 GHz (Trotter et al., 1998). The larger dispersion residuals observed here are most likely due to the presence of this jet interacting with the ISM of the galaxy disk. At ? 1 kpc, the optical filament appears to disperse at the location where the 177 [[CCIIII]] 115588 [[CCIIII]] 115588 Figure 4.23 Top row: PACS contours of the [C II] 158 v50 and W1? residuals in the wind in NGC 3079 overlaid on 1.4GHz observations from Sebastian et al. (2019). Middle row: PACS contours of the [C II] 158 v50 and W1? residuals in the wind in NGC 3079 overlaid on H?+ [N II] image from Cecil et al. (2001). Bottom row: Chandra image (blue) + HST (red and green). 178 radio emission begins and extends ? 1 kpc to the top of the bubble. The ?v50 wind along this filament is ? ?20 km s?1 . The base of the northern edge is seen in both optical and radio where receding velocities are ?v ?150 wind? 45 km s . The bottom row of Figure 4.23 shows the [C II] 158 ?W1?wind contours over- laid on the HST image from above (red and green) and the Chandra data fromCecil et al. (2001). Notice that in the E, the (soft) X-ray filaments spatially correlate with the optical filaments, while in the W, the (soft) X-ray emission fills the second lobe, but lacks any of the filamentous structures seen at 1.4 GHz. 4.9.7 NGC 4945 Column 1 of Figure 4.24a shows the v50 maps of NGC 4945 which range in velocity from ? ?90 km s?1 to ? +100 km s?1 for [O III] 88 , [O I] 145 , and [C II] 158 . As discussed in Section 4.8.3, the [O I] 63 emission in NGC 4945 shows significant self-absorption along the galaxy major axis and is excluded from this analysis. The v50 maps indicate that the NE part of the disk is receding from us and the SW part of the disk is approaching us, consistent with previous observations (Venturi et al., 2017). The maps of the residuals in the wind in Column 4 of Figure 4.24 show in [O III] 88 a SE to NW collimated structure aligned with NGC 4945?s biconical out- flow. The presence of the SE lobe is clearly seen in the [O I] 145 wind residuals, with perhaps some detection of the inner NW lobe. No outflow is detected in [C II] 158 . In Figure 4.25, the contours of the wind residuals (magenta lines) are overlaid 179 NGC 4945 ?v (a) 50 ?v50 v50 Data v50 Model ?v50 (?W1? > 25) (?W1? > 50) 119 119 86 86 86 ?49? 27? 45?? 68 68 43 43 43 ?49? 28? 00?? 17 17 0 0 0 ?15?? -33 -33 -42 -42 -42 -84 -84 -85 -85 -85 119 119 86 86 86 ?49? 27? 45?? 68 68 43 43 43 ?49? 28? 00?? 17 17 0 0 0 ?15?? -33 -33 -42 -42 -42 -84 -84 -85 -85 -85 119 119 86 86 86 ?49? 27? 45?? 68 68 43 43 43 ?49? 28? 00?? 17 17 0 0 0 ?15?? -33 -33 -42 -42 -42 184 pc -84 -84 -85 -85 -85 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s (b) W1? Data W1? Model ?W1? ?W1? > 25 ?W1? > 50 ?49? 27? 45?? 505 505 278 278 278 ? ? ?? 404 404 178 178 178?49 28 00 303 303 78 78 78 ?15?? 202 202 -21 -21 -21 101 101 -120 -120 -120 ?49? 27? 45?? 505 505 278 278 278 ? ? ?? 404 404 178 178 178?49 28 00 303 303 78 78 78 ?15?? 202 202 -21 -21 -21 101 101 -120 -120 -120 ?49? 27? 45?? 505 505 278 278 278 ? ? ?? 404 404 178 178 178?49 28 00 303 303 78 78 78 ?15?? 202 202 -21 -21 -21 184 pc 101 101 -120 -120 -120 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.24 NGC 4945: Results from modeling the disk velocity field with 3DBarolo. Symbols, units, and plots are the same as those in Figure 4.8. 180 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 [[CCIIII]] 115588 [[OOII]] 114455 [[OOIIIIII]] 8888 Figure 4.25 Top row: PACS contours overlaid on the [NII] residual velocity field of NGC 4945 from Venturi et al. (2017). Bottom row: PACS contours overlaid on the [NII] W70 of NGC 4945 from Venturi et al. (2017). NGC 4945 Mwind KEwind Mwind KEwind (?W1? > 25) (?W1? > 25) (?W1? > 50) (?W1? > 50) 56.09 7.71 ?49? 27? 45?? 44.93 6.18 ?49? 28? 00?? 33.77 4.65 ? 22.62 3.1215?? 11.46 1.58 ?30?? Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.30 0.05 0.66 0.11 ?49? 27? 45?? 0.56 0.09 ?49? 28? 00?? 0.47 0.07 ? ?? 0.38 0.0615 0.29 0.04 ?30?? Mtot: ? KEtot: ? Mtot: ? KEtot: ? 0.20 0.03 0.19 0.17 0.09 0.17 ?49? 27? 45?? 0.15 0.14 0.07 0.14 ?49? 28? 00?? 0.11 0.10 0.06 0.10 ? ?? 0.08 0.07 0.04 0.0715 0.04 0.04 0.02 0.04 ?30?? Mtot: 36.37 M KEtot: 22.16 ergs Mtot: 7.29 M KEtot: 13.00 ergs 0.01 0.00 0.01 0.00 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s 30s 28s 13h 05m 26s Figure 4.26 Mass and KE in the wind derived from the PDR solutions and the [O III] 88 flux. Symbols, units, and plots are the same as those in Figure 4.10. 181 [[OOIIIIII]] 8888 SSoolluuttiioonn CC SSoolluuttiioonn AA on the optical [N II] 6583 A? images from Venturi et al. (2017) (top row) and the X-ray images from Marinucci et al. (2012) (bottom row). Residual velocities and dispersions in the ionized gas as traced by [O III] 88 are largest along the edges of the outflow lobes, this is most evident in Figure 4.25(a,c), with receding velocities up to ?v ? +95 km s?150 wind and up to ?v50 wind? +50 km s?1 for the SE and NW lobes, respectively. [O III] 88 residual dispersions at the base of the SE lobe are in the range ?W1?wind? 100 km s?1 to ? 160 km s?1 , while ?W1?wind ranges from ? 90? 100 km s?1 on the edges of the NW lobe. The [O I] 145 v50 wind residuals (see Column 1 of Figure 4.24a) are located mostly below the plane of the galaxy. In the optical, the SE lobe only appears at ? 15?? from the nucleus, but in [O I] 145 , the strongest feature of the SE lobe begins to emerge at ? 10??, demonstrating the advantage of the FIR to penetrate further through the galaxy disk. The [O I] 145 feature in the SE lobe is approaching with ?v50 wind? ?90 km s?1 and contains large dispersions in the range of ?W1?wind? 250? 280 km s?1 . The NW outflow cone in NGC 4945 opens towards us and lies in front of the galactic disk, while the SE cone is oriented in the opposite geometry (Venturi et al., 2017) and is partially obstructed by foreground dust located in the galactic plane (Heckman et al., 1990; Koornneef, 1993). The kinematics of the SE lobe as seen in [O III] 88 and [O I] 145 are consistent with the MUSE [N II] observations from Venturi et al. (2017). The center of the SE lobe is approaching us (seen in [O I] 145 ) and the edges are receding (seen in [O III] 88 ). The ?v50 wind of the ionized gas in the NW lobe however, only indicates that the outflow is receding from us. Venturi et al. 182 (2017) suggests that the edges of the NW lobe are approaching and the inner region of the NW lobe is receding, however we do not detect any approaching velocities in the NW lobe. There is no significant detection of the neutral gas in the outflow of NGC 4945 (see Figure 4.26), but there is a prominent detection of the ionized gas. The largest ionized masses are concentrated around the galaxy nucleus while the largest KE in the outflow lie S of the nucleus in a ? 20?? long structure elongated E to W. The most eastern part of this structure spatially coincides with the SE outflow cone. Some western portions of this structure may also be attributed to the outflow, but the most W edge of the structure may be contamination from disk emission. 4.10 Summary We have analyzed archival Herschel -PACS data of five fine structure lines ([O III] 88 , [O I] 63 , [N II] 122 , [O I] 145 , and [C II] 158 ) in seven nearby galaxies (Cen A, Circinus, M 82, NGC 253, NGC 1068, NGC 3079, and NGC 4945). While the galactic-scale outflows in these objects have been studied extensively at wave- lengths spanning the entire electromagnetic spectrum, they are relatively unexplored in the FIR wavelength regime, 55 ? 210 ?m , studied here. Thanks to the unique combination of angular resolution, sensitivity, and two-dimensional coverage in the FIR of Herschel PACS, we have been able to spatially and kinematically resolve the neutral atomic and ionized gas in the outflow of each galaxy in our sample. The main results are summarized as follows: 183 1. We have derived velocity-integrated fluxes, radial velocities, and 1?? line width maps of all five atomic fine-structure lines. Radial velocity maps indicate that the kinematics within the galaxy disks are dominated by rotation. For some objects (e.g. M 82 and Cen A), the observed line dispersion maps alone can delineate the collimated outflow from the galaxy disk. 2. Analysis of the emission line ratio maps confirm earlier results in the lit- erature (e.g. the prominent ionization cone in Circinus, the jet in Cen A, and the ionized outflow in NGC 1068). These ratios are insightful tools to diagnose the physical conditions of the ionized and neutral gas within the outflows and near the AGN or SB regions. Compact shapes of the line ratios near the galaxy center reveal effects of ionizing radiation (e.g. [O I] 63 / [C II] 158 ). 3. Modeling of the PDRs near the AGN and/or SB and within the outflow has constrained the physical parameters of the neutral atomic gas, namely the UV interstellar radiation field, G0, and the hydrogen gas density, nH. In general, the highest densities are compact structures that surround the galaxy center. Exceptions are seen in M 82 and NGC 253, where the structures are more extended and irregular in shape. For the most part, the largest ISRFs are also compact structures centered on the galaxy nucleus; exceptions are M 82 where the shape of the highest radiation is extended and elongated in the direction of the galaxy disk, and in NGC 4945 where the extended ISRF is elongated to the E. Additionally, although the shape of G0 in NGC 253 is compact and spherical, it is off center to the W. Two objects show obvious structures in G0 and/or nHwhich are spatially coincident with the known outflow. Higher densities and radiation are seen elongated NE in the direction of 184 the jet in Cen A. For NGC 4945, the highest densities are found not around the galaxy center, but are located SE of the nucleus in the S outflow. 4. We have modeled the gas velocity field of the galaxy disks via a ?tilted-ring? model. Subsequently, the excess residual between the observed and modeled 1-? line widths was used to define the spatial location of the wind (e.g. regions where the 1-? line width residuals were > 25 km s?1 ). 5. We have derived properties of the outflow such as bulk motions, masses, and kinetic energies based on our definition of a wind. We find that excess line widths are a better indicator of outflow compared to excess radial velocities. 6. We have compared our results with other wavelength data and find that our definition of a wind results in features (morphological and kinematic) consistent with the literature. For example, bulk motions in the wind region of M 82 indicate that the N cone is receding from us and the S cone is approaching us, in agreement with the literature. Inspection of the excess line widths inside the wind region show that larger line width residuals coincide with known features in an outflow. For example, the largest excess line widths in NGC 3079 spatially coincide with the known radio jet and an optical filament E of the nucleus. Morphology of the wind is also consistent with the literature. M 82 shows line width residuals extending SE to NW in a collimated feature which spatially coincides with a known SiO chimney. 7. We have demonstrated the advantage of the FIR wavelength range over that of the optical and UV when exploring SB or AGN regions. This is most obvious in NGC 3079 where the outflow W of the nucleus lies behind the disk and is extinguished in the optical, but the superbubble is detected in [C II] 158. 185 8. For completeness, we also present the results of the analysis of the molecular gas traced by OH 119 in Appendix B. We have pushed the boundaries of the capabilities of Herschel -PACS in this work, but higher resolution observations of AGN and SB regions are necessary to clearly disentangle the outflow from the galaxy disk. JWST will be able to observe out to ?28 ?m but with better than ?1?? angular resolution, and therefore out to distances an order of magnitude larger (?100 Mpc; z ? 0.03). There are many ionized, neutral, and molecular gas diagnostics in the MIR and this will allow studies of the multi-phase winds in exquisite details. Moreover, line ratios in the MIR, such as [O IV] 25.6 / [Ne II] 12.8 and [Ne V] 14.3 / [Ne II] 12.8 will nicely complement the FIR, providing powerful diagnostics to discriminate AGNs from star formation activity (e.g. Veilleux et al., 2009). 186 Chapter 5: Summary and Future Work The goal of this thesis was to study and explore the cool component of galactic- scale winds in nearby SBs and AGNs. This research is based on archival FIR and MIR spectroscopic data obtained with Herschel?PACS and Spitzer -IRS, respec- tively. In this chapter, we briefly summarize the main results of the work discussed in this thesis. 5.1 Main Results In Chapter 2 we presented the results of our analysis of spectroscopic data of OH at 119 ?m obtained with Herschel?PACS of 52 local (d < 50 Mpc) BAT AGN selected from the very hard X-ray (14 ? 195 keV) 58 month Swift?BAT Survey. These data were combined with Herschel?PACS observations of 43 nearby (z < 0.3) ULIRG/QSOs from (Veilleux et al., 2013). The inclusion of the BAT AGN sample with the ULIRG/QSO sample extended the range of AGN properties (luminosities, star formation rates, and stellar masses) to lower values by 1 ? 2 orders of magnitude compared to the ULIRG/QSO sample alone. We measure the frequency of occurrence of the wind in the BAT AGN. We also look for correlations between the properties of the wind and that of the host galaxy. 187 Evidence for molecular outflows as traced by OH at 119 ?m, was seen in only four of the 52 BAT AGN. This corresponds to a 24% outflow detection rate in the BAT AGN where the search for outflows was possible (i.e. 12 objects were seen in pure absorption and 5 were seen with absorption+emission composite profiles). Evidence for molecular inflows is seen in seven objects, corresponding to a ? 40% inflow detection rate in the BAT AGN. When the BAT AGN are combined with the ULIRG/QSOs, the total sample covers a range of ? 3 dex in SFR and AGN luminosity and ? 2 dex in stellar mass. The combined sample shows positive trends between outflow velocities and the host galaxy properties: stellar mass, SFR, and AGN luminosity. The correlation between wind velocity and AGN luminosity is the strongest of these three. Our results strongly suggest that at higher AGN luminosities (log(LAGN/L & 11.5) the AGN dominates over star formation in driving the outflow. At lower luminosities the AGN may not have the energetics required to drive fast molecular outflows, but stellar processes could dominate if the AGN were weak or absent. It is clear that SF plays a crucial role in driving winds, but the presence of an AGN is needed to drive the fastest winds in our sample. In Chapter 3 we presented the analysis of Spitzer?IRS observations of OH at 35 ?m in 15 nearby (z . 0.06) U/LIRGs. The estimation of molecular out- flow properties, such as mass, from radiative?transfer models, (such as those in Gonza?lez-Alfonso et al., 2017), are dependent on the estimation of the OH?to?H abundance ratio, XOH. Thus, constraining the value of this parameter will result in more accurate predictions when employed. Therefore, we exploited the fact that the 188 ground?state OH absorption feature at 35 ?m is optically thinner than the other OH doublets in the FIR to provide an independent constraint on the estimation of XOH. We computed the OH column densities in our sample and compared them to the hydrogen column density for a typical optical depth at 35 ?m of ? 0.5. Assum- ing a gas?to?dust ratio of ? 125, we find a mean of XOH = 1.01? 0.15? 10?6 for our sample. We attempt to verify that the radiative transfer models from Gonza?lez- Alfonso et al. (2017) predict a realistic OH 35 line profile by comparing the shape of the predicted line profile with the shape of the line profile fitted to observed data. The agreement between the two is generally very good in terms of overall strength (equivalent width) of the feature, and this adds some confidence in the energetics derived from these models. In Chapter 4 we presented the analysis of archival Herschel -PACS data of five atomic fine structure lines ([O I] 63, [O III] 88, [N II] 122, [O I] 145, and [C II] 158) in seven nearby galaxies (M 82, Cen A, Circinus, NGC 253, NGC 1068, NGC 3079, and NGC 4945). Included is also the preliminary analysis of OH 119 in these objects. We derived velocity?integrated fluxes, radial velocities, and 1-? line width maps of the fine structure lines in each object. For some objects (e.g. M 82) the observed line dispersion maps alone can delineate the collimated outflow from the galaxy disk. We computed emission line ratios and detect well known features in some of the objects (e.g. the ionization cone in Circinus). We employ PDR models to estimate the strength of the radiation field and the hydrogen density in our objects. For some objects higher densities and radiation field strengths are spatially coincident with the known outflow (e.g. the jet in Cen A). We model the gas 189 velocity fields in our sample with 3DBarolo and use the excess residual between the observed and modeled line widths to define the spatial location of the wind. Once the wind is defined, residual radial velocities can trace bulk motions of the outflow. For example, in M 82, for all atomic lines, the N cone is receding from us and the S cone is approaching us. The significant advantage of the FIR over optical or UV wavelengths was underscored while defining the wind location. In particular for NGC 3079, the FIR captures the W superbubble which lies behind the disk and is extinguished in the optical. We also estimate masses and kinetic energies for the neutral and ionized gas in the wind for objects with a wind detection. We also compare the results of our analysis with other wavelength data and find remarkable consistency in the comparisons. For example, the largest residual line widths in NGC 3079 spatially coincide with the known radio jet indicating that the larger residuals are due to the presence of the jet interacting with the ISM of the galaxy disk. The preliminary analysis of OH 119 in our sample shows unambiguous extended outflows in NGC 253 in the form of P?Cygni profiles and unambiguous extended inflows in Circinus and NGC 4945 in the form of inverted P?Cygni profiles. 5.2 Future Work We have pushed the boundaries of the capabilities of Herschel -PACS in Chap- ter 4, but higher resolution observations of AGN and SB regions are necessary to clearly disentangle the outflow from the galaxy disk. The high resolution (? 1??) and sensitivity of the James Webb Space Telescope (JWST) will be able to resolve the 190 SB and/or AGN activity inside the central spaxel of our PACS observations, thus providing the opportunity to study the energy mechanisms of winds in great detail. In the MIR alone, there are many ionized, neutral, and molecular gas diagnostics that will allow for studies of multiphase winds and the MIR range of JWST will also be a nice compliment to the FIR since the combination of these wavelength ranges will expand the possible line ratios such as [O IV] 25.6 / [O III] 88 which has been proposed as a powerful diagnostic to discriminate AGNs from SF activity. Also of particular interest will be SPace Infrared telescope for Cosmology and Astrophysics (SPICA; Swinyard et al., 2009) whose unprecedented spectroscopic sensitivity in the FIR will be able shed light on the role of molecular outflows at high z (currently, only a few detections have been reported). At higher redshifts, the star formation rate density and AGN activity increase which points to an increasing importance of feedback processes earlier in the Universe. SPICA will be able to measure outflow properties of distant AGN with unprecedented accuracy and depth, and provide us with a greater understanding of galaxy formation and evolution. 191 Appendix A: Spitzer MIR SPECTRA 10.0 CenA Circinus 10.00 ESO 005-G004 100.0 1.00 1.0 10.0 0.10 0.1 1.0 0.01 10.0 10.0 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.0 IC 5063 MCG-05-23-016 1.0 MCG-06-30-015 1.0 1.0 0.1 0.1 0.1 10.0 10.0 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] 1.00 Mrk 18 1.00 NGC 1052 NGC 1365 10.0 0.10 0.10 1.0 0.01 0.01 0.1 10.00 10.00 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] NGC 2110 1.00 NGC 2655 10.00 NGC 2992 1.00 1.0 0.10 0.10 0.1 0.01 0.01 10.0 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.1 Mid-infrared (5-37?m) spectra used to measure S9.7?m . The dashed line is the continuum calculated from the cubic spline interpolation fitted to the pivot points shown as black dots. Red dots show fcont(9.7?m) (located on dashed continuum line) and fobs(9.7?m) (located on the solid black line or the observed flux density). The blue line shows the integration range used to calculate the flux and total equivalent width of the 9.7 ?m silicate feature (see Table 2.4 and Table 2.5). 192 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy 10.00 NGC 3079 NGC 3081 NGC 3227 1.00 1.00 1.0 0.10 0.10 0.01 0.01 0.1 10.00 10.00 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.0 NGC 3281 10.0 NGC 3783 NGC 4051 1.0 1.0 1.0 0.1 0.1 0.1 10.0 10.0 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] NGC 4102 10.0 NGC 4151 1.0 NGC 4258 10.0 1.0 1.0 0.1 0.1 0.1 10.0 10.0 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.0 NGC 4388 NGC 4395 1.0 NGC 4593 0.100 1.0 0.010 0.1 0.001 0.1 10.0 10.000 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.2 (Continued) 193 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy 1.000 NGC 4939 1.00 NGC 4941 100.00 NGC 4945 10.00 0.100 0.10 1.00 0.010 0.10 0.001 0.01 0.01 10.000 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] NGC 5273 10.0 NGC 5506 NGC 5728 1.00 0.10 1.0 0.10 0.01 10.00 10.0 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] 1.00 NGC 5899 NGC 6221 10.0 NGC 6300 10.00 1.00 0.10 1.0 0.10 0.01 0.1 10.00 10.00 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] 1.00 NGC 6814 10.00 NGC 7172 1.00 NGC 7213 1.00 0.10 0.10 0.10 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.3 (Continued) 194 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy NGC 7314 10.00 NGC 747910.00 NGC 7582 10.0 1.00 1.00 1.0 0.10 0.10 0.01 0.01 0.1 10.00 10.00 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] F00509+1225 F01572+0009 1.000 F05024-1941 1.00 1.0 0.100 0.10 0.010 0.001 0.1 0.01 10.0 10.00 10.000 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.0 F05189-2524 10.000 07251-0248 1.0 F07598+6508 1.000 1.0 0.100 0.010 0.1 0.001 0.1 10.0 10.000 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.00 F08572+3915 10.00 09022-3615 10.00 F09320+6134 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.4 (Continued) 195 Jy Jy Jy Jy Jy JyJy Jy Jy Jy Jy Jy 10.00 F10565+2448 1.0 F11119+3257 F12072-0444 1.00 1.00 0.10 0.10 0.1 0.01 0.01 10.00 10.0 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.000 F12112+0305 100.00 F12243-0036 1.0 F12265+0219 1.000 10.00 0.100 1.00 0.010 0.10 0.001 0.01 0.1 10.000 10.00 10.0 Wavelength [?m] Wavelength [?m] Wavelength [?m] F12540+5708 13120-5453 F13428+5608 10.00 10.00 10.00 1.00 1.00 0.10 0.10 1.00 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] F13451+1232 10.00 F14348-1447 10.00 F14378-3651 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.5 (Continued) 196 Jy Jy Jy Jy Jy Jy Jy Jy Jy JyJy Jy F14394+5332 F15206+3342 10.00 F15250+3609 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] F15327+2340 F15462-0450 F16504+0228 10.00 10.00 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] F17208-0014 F19297-0406 19542+1110 10.00 10.00 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] 10.00 F20551-4250 F22491-1808 F23128-5919 1.00 1.00 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] Figure A.6 (Continued) 197 Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy Jy F23365+3604 1.00 PG 1351+640 1.00 PG 1440+356 1.00 0.10 0.10 0.10 0.01 0.01 0.01 10.00 10.00 10.00 Wavelength [?m] Wavelength [?m] Wavelength [?m] 1.00 PG 1613+658 1.00 PG 2130+099 0.10 0.10 0.01 10.00 10.00 Wavelength [?m] Wavelength [?m] Figure A.7 (Continued) 198 Jy Jy Jy Jy Jy Appendix B: Results on OH 119 ?m PACS observations of the OH 119.233, 119.441 ?m doublet were retrieved and reduced in the same manner as the atomic FIR data (see Section 4.4). The PACS resolution at 119 ?m is ? 270 km s?1 . Details of the OH observations are in Table B.1. To reduce noise, the OH data have been smoothed with a Gaussian kernel of width 0.05 ?m (this step was excluded for the spectra in NGC 253). For each spaxel, a first-order spline was fit to the continuum and subtracted from the spectra. Profile fitting of the OH doublet followed a similar procedure from Veilleux et al. (2013) and Stone et al. (2016). The OH profile is modeled using four Gaussian components (two for each line of the doublet). The separation between the two lines of the doublet was set to 0.208 ?m in the rest frame (? 520 km s?1 ) and the amplitude and standard deviation were fixed to be the same for each component in the doublet. We characterize the OH line profile properties in the same manner as the atomic lines (e.g. v50 and W1? ) and present here our preliminary results of the spectral analysis of OH 119 for each object. Four figures are presented for each object (see Figure B.1). The top left shows the PACS IFU footprint on the sky (black lines and squares) and the outline of the ionized wind traced by [O III] 88 (white line). 199 Table B.1. PACS OH 119 ?m Observations Object Name ObsId texp (s) Proposal ID M 82 1342232257 958 OT1 shaileyd 1 Cen A 1342225989 976 OT1 shaileyd 1 Circinus 1342225147 958 OT1 shaileyd 1 NGC 253 1342237604 976 OT1 shaileyd 1 NGC 1068 1342203128 3330 KPGT esturm 1 NGC 3079 1342221391 8045 DDT esturm 4 NGC 4945 1342247792 958 OT1 shaileyd 1 Note. ? Column 1: Galaxy Name. Column 2: Obser- vation Id. Column 3: Exposure time. Column 4: Proposal ID. Top right shows the spline fits to the continuum while bottom left shows the line profile fits to the continuum subtracted spectrum. Finally, bottom right shows maps of the line profile properties for the absorption and/or emission components fitted to the spectra. The 1-? line width maps have not been corrected for instrumental broadening. When pure emission and/or pure absorption (e.g. M 82 and NGC 1068) is observed in the OH 119 line profile, the rotation of the galaxy disk is discernible in the v50 maps (see Figure B.1 and Figure B.5). The rotation is consistent with the results from the analysis of the ionized and neutral gas traced by the fine structure lines. However, when there is a combination of absorption and emission components in an OH line profile, the physical interpretation of v50 is not so clear. For all of objects, there is no detection of an outflow in the central spaxel. Only Circinus (see Figure B.3) shows a wind in the central spaxel in the form of an unambiguous inflow (e.g. an inverted P-Cygni profile is observed). Cen A shows no 200 discernible detection of a wind in OH 119. NGC 1068 is seen in pure emission with no detection of an outflow. The largest line widths (? 500 km s?1 ) in M 82 (see Figure B.1) appear in pure absorption in spaxels (3,0) and (2,3). The absorption lines have blue-shifted wings and are spatially coincident with the ionized outflow detected in [O III] 88, suggesting there is also outflow in the molecular gas. NGC 253 shows detections of unambiguous outflow outside of the central spaxel (see Figure B.4) which extend N, S, and NW of the nucleus. The P-Cygni profiles lie in good spatial agreement with the ionized outflow traced by [O III] 88 . NGC 4945 (see Figure B.7) shows pure absorption above and below the disk at the edges of the PACS FOV. Unambiguous inflow is seen mostly S of the nucleus in spaxels (2,1), (4,2), and (4,3). These spaxels are spatially coincident with the edges of the ionized outflow traced by [O III] 88 . There is also unambiguous inflow W of the nucleus in spaxel (3,3), but the spaxel lies outside of the ionized outflow. There is also a possible detection of a blue absorption wing E of the nucleus in spaxel (1,1). 201 M82 1342232257 OH 119 ?m 37 48 17 20 4x4 11536 4x3 4x2 4x1 4x0 46 69?41'15" 1635 19110 4x1 4x0 4434 15 18 105 42 33 17 100 40 14 32 4x2 3x1 3x0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 33 3x4 160 3x3 3x2 26 3x1 3x018 125 00" 4x3 32 25 150 17 3x2 2x1 31 1204x4 2x0 24 140 16 30 115 23 3x3 29 130 15 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 2x2 2003x4 1x1 50 60 1x0 95.023 2x4 2x3 195 2x2 2x1 2x0 58 92.5 48 190 40'45" 2x3 22 90.056 185 46 87.5 2x4 1x2 21 1800x1 540x0 4485.0175 52 82.5 20 42 1x3 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 300076 240 1x4 0x2 14 1x4 30 1x3 1x2 1x1 115 1x0 74 230 13 29 110 30" 0x3 72 22028 12 105 70 0x4 27 210 10011 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 11.0 31 22 61 35 10.5 0x4 0x3 0x2 0x1 0x0 30 60 34 10.0 21 9.5 59 33 15" 20 29 9.0 58 32 8.5 19 28 57 31 8.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 9h55m57s 54s 51s 48s Velocity [km s?1] M82 1342232257 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] M82 1342232257 OH 119 ?m 3.51 -1.87 3.51 4 4 4 0.0 0 0 4x4 4x3 4x2 4x1 4x0 -16.23 -20.54 3.170.0 0.0 ?0.5 ?1 3 3 3 ?5 ? ?0.2? 21.0 ?0.2 -35.97 -39.20 2.83 ?3 ?0.4 ? 2 2 21.5 ?10 ?4 ?0.6 ?0.4 -55.71 -57.86 2.48 ?2.0 ? ?5 1 1 115 ? ?0.82.5 ?0.6 -75.45 -76.52 2.14 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0 0 0 0 00.0 0.0 0.03x4 3x3 ?2 3x2 3x1 3x0 -95.19 -95.19 1.80 ?5 ?0.5 ? ?0.20.2 ?4 4 3 2 1 0 4 3 2 1 0 4 3 2 1 0 ? ?10 ?1.0 ?0.4 0.4 ?6 ? ?8?0.6 15 ? ?0.6 1.5 v50,abs [km/s] v50,emi [km/s] ?10 ?0.8 ?0.8 ?20 ?2.0 119 61 ?12 ?1.0 4 4 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.4 85 150 0 0.0 0 2x4 2x3 2x2 2x1 2x0 3 3 0.3 ? ?20.5 ?5 ?1 52 -31 0.2 ?1.0 ?4 2 2 0.1 ?10 ?2 ? 19 -77?1.5 6 1 1 0.0 ? ?15 ? ?32.0 8 -13 -123 ?0.1 ? 0 02000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.0 0.0 0 0.0 -47 -169 0.8 0.8 1x4 1x3 1x2 1x1 1x0?0.5 ?5 ?2.5 4 3 2 1 0 4 3 2 1 0 0.6 0.6 ?1.0 ?10 ?5.0 0.4 0.4 ?1.5 ?7.5 W1?,abs [km/s] W1?,emi [km/s]?15 0.2 0.2 ?2.0 ? 515 846? 10.020 0.0 4 4 0.0 ?2.5 ?12.5 ?25 470 744 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.8 3 3 0.8 0.0 0.0 0x4 0.8 0x3 0x2 0x1 0x0 426 642 0.6 0.6 ?0.5 ?0.5 0.6 2 2 0.4 0.4 ?1.0 382 541 0.4 ?1.0 0.2 ?1.5 1 1 0.2 0.2 ?1.5 338 439 0.0 ?2.0 0.0 0.0 0 0 ?2.0 ?2.5 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 294 338 Velocity [km s?1] 4 3 2 1 0 4 3 2 1 0 Figure B.1 Top left: PACS IFU footprint of OH 119 (black squares) overlaid on the 22 ?m WISE image of M 82. The white contour marks the spatial ex- tent of the [O III] 88 wind. The black contour outlines the PACS footprint of the [O III] 88 observation. The black cross marks the adopted galaxy center and the blue line marks the galaxy major axis. Top right: Spline fits to the OH 119 contin- uum (blue dashed lines). Black lines are the observed data. Magenta areas indicate the regions used to fit the continuum. Blue dots mark the pivot points used to fit the spline. Bottom left: Line profile fitting results of the continuum-subtracted spectra. Solid blue lines indicate gaussian absorption components. Solid red lines indicate gaussian emission components. Vertical dashed blue (red) lines mark the v16 , v50 , and v84 velocities in absorption (emission). Bottom right: Top row, from left to right shows the total velocity-integrated flux of the fitted OH line profiles, the total flux in the absorption components only, and the total flux in the emission components only. Middle row shows v50 for the absorption (left) and emission com- ponents (right). Bottom row shows the 1?? line widths of the absorption (left) and emission (right) components. 202 Flux [Jy] Flux [Jy] CENA 1342225989 OH 119 ?m 5.5 5.5 4.5 5.0 0x0 0x1 5.5 0x2 5.0 0x3 0x4 5.0 4.0 4.5 0x4 5.0 4.5 3.5 4.5 4.0 4.0 4.5 3.0 -43?00'45" 4.03.50x3 3.54.0 2.5 3.0 3.5 1x4 3.03.5 2.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 8.0 10.0 12.5 0x2 8.51x3 1x0 9.5 1x1 6.512.0 1x2 1x3 1x47.5 8.0 2x4 9.0 11.5 6.07.0 7.5 8.5 11.0 5.5 6.5 10.5 7.0 8.0 01'00" 0x1 1x2 2x3 5.06.0 10.0 6.57.5 0x0 3x4 4.59.5 6.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 28 15.0 7.5 9.0 11.0 2x2 3x3 2x0 14.5 2x1 2x2 2x3 2x427 7.01x1 1x0 4x4 8.5 14.0 10.5 6.5 8.0 13.5 26 10.0 13.0 6.0 7.5 25 9.5 12.5 15" 3x2 4x3 5.57.02x1 9.012.0 24 2x0 5.0?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 6.5 8.0 13.5 9.5 8.0 4x2 3x0 3x1 3x2 3x3 3x46.0 13.0 7.5 7.5 9.0 3x1 12.5 5.5 7.0 3x0 7.0 8.512.0 6.5 5.0 6.5 11.5 8.0 11.0 6.0 6.0 4.5 7.5 30" 4x1 10.5 5.54x0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 4.0 4.0 6.5 4x0 4x1 5.0 4x2 6.5 4x3 4x4 3.5 3.5 6.0 4.5 6.0 3.0 3.0 5.5 4.0 5.5 2.5 5.0 2.5 3.5 5.0 2.0 4.5 2.0 3.0 4.5 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 13h25m30s 28s 26s Velocity [km s?1] CENA 1342225989 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] CENA 1342225989 OH 119 ?m 0.13 -0.14 0.13 0.15 0 0 0 0.075 0.05 0.10 0x0 0x1 0.05 -0.77 -0.99 0.130.050 0.10 0x2 0x3 0x4 0.05 0.00 1 1 1 0.025 0.05 0.00 -1.68 -1.84 0.12 0.00 0.000 0.00 ?0.05 2 2 2 ?0.05 ?0.025 ?0.05?0.05 -2.59 -2.70 0.12 ?0.10 ? ?0.100.050 ? 3 3 30.10 ? ?0.100.15 ?0.075 -3.49 -3.55 0.11 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.10 0.10 0.05 4 4 4 0.0 1x0 1x1 0.05 1x2 1x3 0.1 1x4 0.00 -4.40 -4.40 0.110.05 ? ?0.10.05 0.00 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0.00 0.0 ?0.10 ?0.05 ?0.2 ?0.05 ?0.15 ?0.10 v50,abs [km/s] v50,emi [km/s] ? ?0.10.3 ?0.20 ?0.10 ? 284 -870.15 ?0.25 ?0.4 ?0.2 0 0 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 208 -126 0.00 0.05 0.0 0.0 2x0 2x1 2x2 2x3 2x4 1 10.0 ?0.25 0.00 132 -166 ?0.1 ?0.2 ?0.50 ?0.1 ?0.05 2 2 ? ?0.75 ?0.100.2 56 -206 ?0.4 ?0.2 ?1.00 ?0.15 3 3 ?0.3 ?1.25 ?0.3 ?0.20 ? -19 -2460.6 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ? 4 41000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.1 0.1 0.10 -95 -286 3x0 3x1 0.0 3x2 0.0 3x3 3x4 0.0 0 1 2 3 4 0 1 2 3 4 0.05 ?0.2 0.0 ?0.1 0.00 ?0.1 ? W [km/s] W [km/s]0.05 ?0.4 1?,abs 1?,emi ? ?0.10.2 ?0.2 ? 873 2500.10 ?0.6 ?0.3 0 0?0.15 ?0.3 ?0.2 758 239 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.05 1 1 0.075 0.04 0.10 0.10 4x0 4x1 4x2 4x3 4x4 643 229 0.050 0.02 0.000.05 0.05 2 2 0.025 0.00 0.00 ?0.05 529 218 0.00 0.000 ?0.02 ?0.05 ?0.10 3 3 ?0.025 ? ?0.05 ?0.100.04 414 208 ?0.15 ?0.050 ?0.15 4 4 ?0.06 ?0.10 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 300 198 Velocity [km s?1] 0 1 2 3 4 0 1 2 3 4 Figure B.2 Cen A. Lines and symbols are the same as those in Figure B.1. 203 Flux [Jy] Flux [Jy] CIRCINUS 1342225147 OH 119 ?m 17.5 9.0 4.0 4.0 15.0 0x0 0x1 17.0 0x2 8.5 0x3 0x43.5 3.5 14.5 16.5 8.0 3.0 3.0 14.0 16.0 7.5 15.5 2.5 13.5 0x4 2.5 7.015.0 2.0 13.0 2.0 14.5 6.5 -65?20'00" 0x3 12.5 1.5?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 38 4.5 5.5 0x2 1x0 1x1 1x2 1x3 1x4 4.0 1x4 5.0 33 37 13 1x3 3.54.5 36 0x0 32 120x1 3.04.0 35 1x2 31 2.5112x4 3.5 34 2.0 15" 2x3 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 1x0 967.5 4.5 1x1 2x0 2x1 2x2 17.5 2x3 2x4 2x2 37 94 17.0 4.03x4 7.0 16.5 3.5 6.5 92 3x3 36 16.0 3.0 2x0 6.0 90 15.5 2x1 35 2.53x2 88 15.04x4 5.5 2.0?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 30" 5.04x3 50.0 16.516.56.5 3x0 3x1 3x2 3x3 4.5 3x4 49.5 16.0 3x0 16.0 4.03x1 6.04x2 49.0 15.515.5 3.5 48.5 5.5 15.0 3.0 15.0 48.0 5.0 14.5 2.5 4x0 47.514.54x1 14.0 2.04.5 47.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 6.0 22.5 45" 12.5 4.022.0 5.5 4x0 4x1 4x2 10.0 4x3 4x4 12.0 3.5 21.5 5.0 9.511.5 3.0 21.0 11.0 4.5 9.0 2.5 20.5 10.5 4.0 8.520.0 2.0 10.0 3.5 19.5 8.0 1.5 9.5 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 14h13m15s 12s 09s 06s Velocity [km s?1] CIRCINUS 1342225147 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] CIRCINUS 1342225147 OH 119 ?m 2.77 -0.08 2.77 0 0 0 0.3 0.05 0.0750.25 0x0 0x1 0x2 0.4 0x3 0x4 1.88 -0.75 2.24 0.20 0.00 0.050 1 1 1 0.3 0.2 0.15 ?0.05 0.025 0.99 -1.42 1.71 0.2 0.10 ?0.10 0.000 2 2 2 0.1 0.1 0.05 ?0.15 ?0.025 0.10 -2.09 1.18 0.00 ?0.20 ?0.050 0.0 0.0 3 3 3 ?0.05 ?0.075 ?0.1 -0.80 -2.76 0.64 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.0 0.4 4 4 4 0.3 0.6 1x0 1x1 1x2 1x3 0.40.8 1x4 0.3 -1.69 -3.43 0.11 0.2 0.6 0.3 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0.2 0.4 0.4 0.2 0.1 0.1 0.2 0.2 0.1 v50,abs [km/s] v50,emi [km/s] 0.0 0.0 0.0 892 3770.0 0.0 ?0.1 0 0 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.0 638 2470.5 2x0 0.6 1 10.15 0.4 2x1 2x2 2x3 2x4 0.5 0.4 384 117 0.4 0.10 0.2 0.30.0 2 2 0.2 0.2 0.05 0.0 ? 130 -110.5 0.1 3 3 0.0 0.00 ?0.2 ?1.0 0.0 -123 -141 ?0.2 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ? 4 42000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.2 0.6 -377 -271 0.20 0.10 3x0 3x1 0.1 3x2 3x3 3x4 0.2 0 1 2 3 4 0 1 2 3 4 0.15 0.0 0.4 0.05 0.10 ?0.1 0.1 W [km/s] W [km/s] 0.05 ?0.2 0.2 1?,abs 1?,emi 0.00 ?0.3 0.0 402 574 0.00 ?0.4 0.0 0 0 ?0.05 ?0.05 ?0.1 361 492 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1 1 0.4 0.5 4x0 4x1 0.3 0.3 4x2 4x3 4x4 321 411 0.2 0.4 0.3 2 2 0.2 0.2 0.3 0.2 0.1 281 330 0.2 0.1 0.1 0.1 3 3 0.1 0.0 241 249 0.0 0.0 0.0 0.0 4 4 ? ?0.10.1 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 201 168 Velocity [km s?1] 0 1 2 3 4 0 1 2 3 4 Figure B.3 Circinus. Lines and symbols are the same as those in Figure B.1. 204 Flux [Jy] Flux [Jy] NGC253 1342237604 OH 119 ?m 10.0 20.0 17.5 -25?16'45" 9.50x0 19.5 0x1 0x2 17.09.5 0x3 0x4 21 16.5 9.019.0 9.0 18.5 16.0 8.5 20 8.5 18.0 15.5 8.0 8.0 17.5 19 15.0 7.5 0x0 17.0 14.5 7.5 7.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 29 40 130 13 17'00" 0x2 0x3 1x0 1x1 75 1x2 1x3 1x428 380x1 1250x4 127027 36120 1x0 26 115 65 34 11 1x2 1x3 25 110 32 60 10 105 1x1 1x4 24 30?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 550 90 19 2x0 32.0 922x0 2x1 2x2 2x3 18 2x415" 31.5 90 855002x2 2x3 31.0 88 1780 2x1 30.52x4 86 450 16 30.0 84 75 3x0 1529.5 82 400 14 29.0 70 3x2 803x3 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 170 3x1 3x4 10 16 95 4x0 3x0 3x1 3x2 3x3 24 3x4 30" 16015 909 22 4x2 14 150 854x3 84x1 204x4 13 140 80 7 12 75 18 130 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 11 27 28 16 12 4x0 4x1 26 4x2 4x3 4x4 10 15 45" 25 2611 9 1424 24 10 23 13 8 22 22 12 9 7 21 11 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 0h47m35s 34s 33s 32s 31s Velocity [km s?1] NGC253 1342237604 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] NGC253 1342237604 OH 119 ?m 7.53 -0.35 12.38 0.2 0 0 0 0.2 0.2 0.6 0.4 -86.24 -92.54 10.04 0x0 0x1 0x2 0x3 0.1 0x4 0.1 1 1 1 0.0 0.1 0.4 0.0 0.2 -180.01 -184.74 7.69 ?0.1 ?0.1 2 2 2 0.0 0.2 0.0 ?0.2 ?0.2 -273.78 -276.93 5.35 ?0.3 ? ?0.2 3 3 3? 0.3 0.0 0.1 ?0.4 -367.54 -369.12 3.01 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.25 0 4 4 40.0 1 0 1x0 1x1 1x2 1x3 1.00 1x4 -461.31 -461.31 0.66 ? ? 0 2 0.5 ?5 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 ? 0.75? 14 ?1.0 ?10 ? ?2 0.506 v50,abs [km/s] v50,emi [km/s]?8 ?3 0.25 ?1.5 ?15 9 321 ?10 ?4 0.00 0 0 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 -28 201 0.4 00 2.5 1.5 2x0 2x1 2x2 2x3 2x4 1 1 ?25 0.0 1.0 ?2 -66 81 0.2 ?50 ?2.5 0.5 2 2 ?4 ?75 ?5.0 0.0 -103 -390.0 ?6 ?100 ?7.5 ? 3 30.5 ?0.2 -141 -159?125 ?10.0 ?1.0 ?8 ?2000 ? 4 41000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 5 1 -179 -279 0.0 1.5 1.5 3x0 3x1 3x2 3x3 3x40 ? 0 1 2 3 4 0 1 2 3 42.5 0 1.0 ?5 ?5.0 1.0 ?1 ?10 ?7.5 W1?,abs [km/s] W1?,emi [km/s] 0.5 0.5 ? ?10.015 ?2 371 605 ?12.5 0 0 0.0 0.0 ?20 ?3 ?15.0 328 523 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.5 1 1 0.3 0.0 4x0 0.6 4x1 4x2 4x3 4x4 285 442 1.0 0 0.2 ?0.5 2 2 0.4 0.5 ? 243 3601 0.1 0.2 0.0 ?1.0 3 3 ?0.5 ?2 200 279 0.0 0.0 ?1.5 ?1.0 4 4 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 158 198 Velocity [km s?1] 0 1 2 3 4 0 1 2 3 4 Figure B.4 NGC 253. Lines and symbols are the same as those in Figure B.1. 205 Flux [Jy] Flux [Jy] NGC1068 1342203128 OH 119 ?m -0?00'15" 7.0 6.56.0 1611.0 4x4 4x3 4x2 6.5 4x1 4x0 5.5 6.0 10.5 15 5.0 6.0 10.0 5.5 4.5 5.5 9.5 5.0 14 4.0 9.0 5.0 4.5 4x4 3.5 8.5 4.513 4.0 4x3 4x1 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30004x2 11.0 20.5 12.030" 10.5 3x4 3x3 24 3x2 17.5 3x1 3x020.0 11.5 17.0 4x0 10.0 2319.5 11.03x4 3x1 16.59.5 19.0 22 10.53x3 16.03x2 9.0 18.5 10.021 15.5 8.5 18.0 9.5 15.0 3x0 8.0 20 17.5 9.0 2x4 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30002x1 9.0 14.5 54 20.5 45" 2x3 14.02x2 8.5 2x4 2x3 2x2 21 2x1 20.0 2x052 13.5 8.0 20 19.5 13.0 50 2x0 7.5 19.01x4 1912.5 481x1 7.0 18.512.0 18 1x3 1x2 6.5 46 18.0 11.5 6.0 44 17 17.5 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 11.0 31 19 0x4 1x0 5.0 1x4 10.5 1x3 1x2 1x1 16.5 1x001'00" 300x3 0x1 4.5 10.0 18 16.00x2 29 15.5 4.0 9.5 17 28 9.0 15.0 3.5 0x0 8.5 27 14.5 3.0 16 8.0 14.026 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 4.0 21.5 0x4 5.5 0x3 9.5 0x2 0x1 6.0 0x0 21.0 3.5 15" 5.0 9.0 5.520.5 3.0 8.5 4.5 20.0 5.0 2.5 8.0 19.5 4.0 4.5 2.0 7.5 19.0 3.5 4.0 18.5 7.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 2h42m43s 42s 41s 40s 39s Velocity [km s?1] NGC1068 1342203128 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] NGC1068 1342203128 OH 119 ?m 23.92 23.92 0.6 0.8 4 4 4 0.4 0.4 4x4 0.6 4x3 4x2 4x1 4x0 19.18 19.18 0.6 0.3 0.3 0.4 3 3 3 0.4 0.4 0.2 0.2 14.44 14.44 0.2 0.2 0.1 2 2 2 0.2 0.1 0.0 9.71 9.71 0.0 0.0 0.0 0.0? 1 1 10.1 ?0.2 4.97 4.97 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.0 0 0 0 0.8 0.8 3x4 3x3 2.00.8 3x2 3x1 0.6 3x0 0.23 0.23 0.6 0.6 0.6 1.5 4 3 2 1 0 4 3 2 1 0 4 3 2 1 0 0.4 0.4 0.4 0.4 1.0 0.2 0.2 0.2 0.2 v50,abs [km/s] v50,emi [km/s]0.5 199 0.0 0.0 0.00.0 0.0 4 4 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.0 0.8 122 0.6 2.06 2x4 0.8 2x3 2x2 2x1 2x0 3 3 0.6 1.5 45 0.4 0.6 4 0.4 2 2 1.0 0.4 0.2 -30 2 0.2 0.2 0.5 1 1 0.0 0.00.0 -107 0 0.0 ? ? 0 02000 1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 1.25 -184 0.3 0.6 0.6 1x4 1x3 1.00 1x2 2.0 1x1 1x0 4 3 2 1 0 4 3 2 1 0 0.2 0.4 0.75 1.5 0.4 0.50 1.0 W1?,abs [km/s] W1?,emi [km/s]0.2 0.1 0.2 0.25 0.5 631 0.0 0.0 0.00 0.0 4 4 0.0 568 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.3 0.5 3 3 0.8 0.15 0x4 0x3 0.4 0x2 0x1 0x0 505 0.05 0.2 0.6 0.10 0.3 2 2 0.1 0.4 0.05 4420.00 0.2 1 1 0.00 0.1 0.2 ? 0.00.05 379 0.0 ?0.05 0.0 0 0 ?0.1 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 316 Velocity [km s?1] 4 3 2 1 0 4 3 2 1 0 Figure B.5 NGC 1068. Lines and symbols are the same as those in Figure B.1. 206 Flux [Jy] Flux [Jy] NGC3079 1342221391 OH 119 ?m 3.0 4.0 1.5 55?41'15" 1.0 0x0 0x1 4.52.5 0x2 0x3 0x43.5 1.0 4.0 0.5 2.0 3.0 0x4 0.5 3.50.0 1.5 2.5 0x3 0.0?0.5 3.01.0 2.0 ?0.5 ?1.0 0x2 0.5 2.5 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 1x4 2.53.0 6.5 8.5 1.0 1x0 1x1 1x2 1x3 1x4 8.0 2.0 00" 0x0 1x3 2.5 6.00x1 0.5 7.52.0 5.5 1.5 7.0 1x2 0.0 1.5 5.0 1.0 6.5 2x4 ?0.5 1.0 4.5 6.0 0.5 4.0 1x0 2x3 0.5 ? 5.51.0 0.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 1x1 4.0 2.01.5 30 10 2x2 2x0 3.5 2x1 2x2 2x3 1.5 2x41.0 28 3x4 0.5 3.0 9 1.0 40'45" 26 2x0 3x3 0.0 2.5 0.524 8 2x1 ?0.5 2.0 0.022 3x2 7?1.0 1.5 20 ?0.5 4x4 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30004.51.5 1.5 4.0 17 4x3 3x0 3x1 3x2 4.0 3x3 3x43x0 1.0 1.0163.5 3x1 3.50.5 4x2 15 0.53.0 3.0 30" 0.0 14 0.0 2.5 2.5 ?0.5 13 ?0.5 4x0 2.0 2.04x1 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30002.0 4.5 2.5 1.5 4.5 4x0 4x1 4x2 4x3 4x4 1.5 2.04.0 1.0 4.0 1.5 1.0 3.5 0.5 3.5 1.0 0.5 3.0 0.0 3.0 15" 0.52.50.0 ?0.52.5 0.0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 10h02m00s 01m58s 56s 54s Velocity [km s?1] NGC3079 1342221391 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] NGC3079 1342221391 OH 119 ?m 0.23 -0.12 0.23 0.10 0.025 0 0 00.10 0.02 0x0 0x1 0x2 0.000 0x3 0x4 -5.28 -5.55 0.19 0.05 0.05 0.05 0.00? 1 1 10.025 ?0.02 -10.78 -10.99 0.160.00 0.00 ?0.050 0.00 2 2 2 ?0.075 ?0.04 ?0.05 -16.29 -16.43 0.13 ?0.05 ?0.05 ?0.100 ?0.06 3 3 3 ?0.10 ?0.125 ?0.08 -21.80 -21.87 0.09 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.050 4 4 40.075 0.0 1x0 0.0 1x1 1x2 0.0 1x3 1x4 0.025 -27.30 -27.30 0.06 0.050 ?0.1 ?0.2 ?0.2 0.000 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0.025 ?0.025 ?0.2 0.000 ?0.4 ?0.4 ?0.050 v50,abs [km/s] v50,emi [km/s] ?0.025 ?0.3 ?0.6 ?0.6 ?0.075 340 362 ?0.050 ?0.4 ?0.100 ?0.8 0 0 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.2 241 246 0.15 0 2x0 2x1 2x2 0.0 2x3 0.05 2x4 1 1 0.10 0.0 ?2 142 131 ?0.2 0.00 0.05 2 2 ?0.1 ? ?0.44 43 15 0.00 ?0.05 ?0.6 ? 3 3? 0.20.05 ?6 ?0.10 ?0.8 -56 -99 ?0.10 ?2000 ? 4 41000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.100 0.0 0.10 -155 -215 3x0 3x1 3x2 0.0 3x3 3x4 0.05 0.075 0 1 2 3 4 0 1 2 3 4 ?0.5 0.05 0.050 ?0.1 0.00 ?1.0 0.025 W [km/s] W [km/s]0.00 1?,abs 1?,emi ?0.2 0.000 756 745?0.05 ?1.5 ?0.05 ? 0 00.025 ?0.3 ?2.0 642 629 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 0.10 1 1 0.050 0.05 4x0 0.05 4x1 0.05 4x2 4x3 4x4 528 513 0.025 0.05 0.00 2 2 0.00 0.00 0.000 414 397 0.00 ?0.05 ?0.025 ?0.05 ?0.05 3 3 ?0.050 ?0.05 ?0.10 300 281 ?0.10 ?0.10 ?0.075 4 4 ?0.15 ?0.10 ?0.100 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 186 166 Velocity [km s?1] 0 1 2 3 4 0 1 2 3 4 Figure B.6 NGC 3079. Top left: The white contour marks the spatial extent of the [C II] 158 wind. Lines and symbols are the same as those in Figure B.1. 207 Flux [Jy] Flux [Jy] NGC4945 1342247792 OH 119 ?m 26 21 40 0x0 0x1 0x2 0x3 8 0x4 6 25 20 39 19 5 24 7 0x4 23 38 18 4 6 0x3 22 17-49?27'45" 373 21 16 5 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 0x2 421x4 12 101x0 1x1 160 1x2 41 1x3 1x476 11 1x3 9 10 150 40 0x0 74 80x1 9 39140 1x2 7722x4 8 38130 6 7 70 37 28'00" 2x3 120 5 1x0 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 300052 13 1x1 12 52 2x2 2x0 2x1 550 2x2 2x3 12 2x4 3x4 50 11 50 500 10 48 3x3 10450 48 46 9 2x0 8 4002x1 83x2 46 444x4 350 7 6 15" ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30004x3 16 17 180 7 3x0 3x0 3x1 3x2 52 3x3 3x43x1 16 144x2 6 160 5015 5 12 14 48 140 4 13 46 4x0 104x1 3 12 120 44 30" ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 30009 19 30 12 4x0 4x1 4x2 22 4x3 4x4 18 11 28 8 17 21 10 26 16 20 7 9 15 24 19 8 14 22 6 13 18 7 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 ?3000?2000?1000 0 1000 2000 3000 13h05m30s 28s 26s 24s Velocity [km s?1] NGC4945 1342247792 OH 119 ?m Total Flux [10?17 Wm?2] Abs Total Flux [10?17 Wm?2] Emi Total Flux [10?17 Wm?2] NGC4945 1342247792 OH 119 ?m 17.90 -4.13 17.90 3.0 0 0 0 3.0 1.25 0.5 1.0 0x0 0x1 0x2 2.52.5 0x3 0x4 -129.14 -146.76 14.44 1.00 0.8 0.0 1 1 12.0 2.0 0.75 0.6 -276.18 -289.40 10.98 1.5 ? 1.50.5 0.50 0.4 2 2 2 1.0 1.0 ?1.0 0.2 -423.23 -432.04 7.520.25 0.5 0.5 ?1.5 3 3 3 0.00 0.0 0.0 0.0 -570.27 -574.67 4.05 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 3.0 0 2.5 4 4 40 2.5 1x0 1x1 1x2 0.0 1x3 1x4 -717.31 -717.31 0.59 2.0 ?1 2.0 ?10 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 ?0.5 1.5 1.5 ?2 1.0 1.0 ?20 ? ?1.0 v3 50,abs [km/s] v50,emi [km/s] 0.5 0.5 224 495 0.0 ?4 ?30 ?1.5 0.0 0 0 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 192 347 0 5 1 2x0 2x1 2x2 2x3 2x4 1 1 3 34 ?50 161 199 0 3 2 2 2 2 ?100 2 130 51 ?1 1 1 ? 1150 3 3 ?2 0 99 -96 0 ? 0200 ? ? 4 42000 1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 5 2.5 0 68 -244 2.5 1 3x0 3x1 3x2 3x3 4 3x4 ?10 0 0 1 2 3 4 0 1 2 3 42.0 2.0 1.5 1.5 ?20 ?1 3 ?30 ?2 2 W1?,abs [km/s] W1?,emi [km/s]1.0 1.0 0.5 0.5 ? ?3 40 346 4611 ? ?4 0 00.0 0.0 50 0 ?5 322 410 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 2.0 1.0 1 1 1.0 2.5 4x0 4x1 1 4x2 4x3 4x4 298 359 0.8 1.5 0.5 2.0 0 2 2 0.6 1.0 ?1 0.0 1.5 274 308 0.4 ?2 1.0 0.5 ? 3 30.5 0.2 ?3 0.5 250 257 0.0 ?1.0 0.0 ?4 0.0 4 4 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 ?2000 ?1000 0 1000 2000 227 206 Velocity [km s?1] 0 1 2 3 4 0 1 2 3 4 Figure B.7 NGC 4945. Lines and symbols are the same as those in Figure B.1. 208 Flux [Jy] Flux [Jy] Bibliography Abel, N. P., Dudley, C., Fischer, J., Satyapal, S., & van Hoof, P. A. M. 2009, ApJ, 701, 1147 Aguirre, A., Hernquist, L., Schaye, J., et al. 2001, ApJ, 561, 521 Alatalo, K., Blitz, L., Young, L. M., et al. 2011, ApJ, 735, 88 Alexander, D. M., Swinbank, A. M., Smail, I., McDermid, R., & Nesvadba, N. P. H. 2010, in Astronomical Society of the Pacific Conference Series, Vol. 427, Accretion and Ejection in AGN: a Global View, ed. L. Maraschi, G. Ghisellini, R. Della Ceca, & F. Tavecchio, 74 Alton, P. B., Davies, J. I., & Bianchi, S. 1999, A&A, 343, 51 Armus, L., Mazzarella, J. M., Evans, A. S., et al. 2009, PASP, 121, 559 Baldry, I. K., Glazebrook, K., Brinkmann, J., et al. 2004, ApJ, 600, 681 Banerji, M., Chapman, S. C., Smail, I., et al. 2011, MNRAS, 418, 1071 Barbosa, F. K. B., Storchi-Bergmann, T., McGregor, P., Vale, T. B., & Rogemar Riffel, A. 2014, MNRAS, 445, 2353 Barnes, J. E., & Hernquist, L. 1996, ApJ, 471, 115 Bassani, L., Dadina, M., Maiolino, R., et al. 1999, ApJS, 121, 473 Baumgartner, W. H., Mushotzky, R., & Tueller, J. 2011, in Bulletin of the American Astronomical Society, Vol. 43, American Astronomical Society Meeting Abstracts #217, #125.01 Bell, E. F., & de Jong, R. S. 2001, ApJ, 550, 212 Benson, A. J., Bower, R. G., Frenk, C. S., et al. 2003, ApJ, 599, 38 Bertone, S., De Lucia, G., & Thomas, P. A. 2007, MNRAS, 379, 1143 Blanco, V. M., Graham, J. A., Lasker, B. M., & Osmer, P. S. 1975, ApJ, 198, L63 209 Bland, J., Taylor, K., & Atherton, P. D. 1987, MNRAS, 228, 595 Blanton, M. R., & Moustakas, J. 2009, ARA&A, 47, 159 Bolatto, A. D., Warren, S. R., Leroy, A. K., et al. 2013, Nature, 499, 450 Bosma, A. 1978, in IAU Symposium, Vol. 77, Structure and Properties of Nearby Galaxies, ed. E. M. Berkhuijsen & R. Wielebinski, 28 Bower, R. G., Benson, A. J., Malbon, R., et al. 2006, MNRAS, 370, 645 Bradford, C. M., Stacey, G. J., Fischer, J., et al. 1999, in ESA Special Publication, Vol. 427, The Universe as Seen by ISO, ed. P. Cox & M. Kessler, 861 Bravo-Guerrero, J., & Stevens, I. R. 2017, MNRAS, 467, 3788 Brusa, M., Perna, M., Cresci, G., et al. 2016, A&A, 588, A58 Burns, J. O., Feigelson, E. D., & Schreier, E. J. 1983, ApJ, 273, 128 Calzetti, D., Wu, S.-Y., Hong, S., et al. 2010, ApJ, 714, 1256 Cano-D??az, M., Maiolino, R., Marconi, A., et al. 2012, A&A, 537, L8 Carilli, C. L., & Walter, F. 2013, ARA&A, 51, 105 Carniani, S., Marconi, A., Maiolino, R., et al. 2015, A&A, 580, A102 ?. 2016, A&A, 591, A28 Cecil, G., Bland, J., & Tully, R. B. 1990, ApJ, 355, 70 Cecil, G., Bland-Hawthorn, J., & Veilleux, S. 2002, ApJ, 576, 745 Cecil, G., Bland-Hawthorn, J., Veilleux, S., & Filippenko, A. V. 2001, ApJ, 555, 338 Chevalier, R. A., & Clegg, A. W. 1985, Nature, 317, 44 Cicone, C., Brusa, M., Ramos Almeida, C., et al. 2018, Nature Astronomy, 2, 176 Cicone, C., Feruglio, C., Maiolino, R., et al. 2012, A&A, 543, A99 Cicone, C., Maiolino, R., Sturm, E., et al. 2014, A&A, 562, A21 Cicone, C., Maiolino, R., Gallerani, S., et al. 2015, A&A, 574, A14 Cielo, S., Bieri, R., Volonteri, M., Wagner, A. Y., & Dubois, Y. 2018, MNRAS, 477, 1336 Clarke, D. A., Burns, J. O., & Norman, M. L. 1992, ApJ, 395, 444 210 Cohen, M., Megeath, S. T., Hammersley, P. L., Mart??n-Luis, F., & Stauffer, J. 2003, AJ, 125, 2645 Cole, S., Norberg, P., Baugh, C. M., et al. 2001, MNRAS, 326, 255 Combes, F. 2014, Proceedings of the International Astronomical Union, 10, 182189 Combes, F., & Charmandaris, V. 2000, A&A, 357, 75 Contursi, A., Poglitsch, A., Gra?cia Carpio, J., et al. 2013, A&A, 549, A118 Cooper, J. L., Bicknell, G. V., Sutherland, R. S., & Bland-Hawthorn, J. 2008, ApJ, 674, 157 ?. 2009, ApJ, 703, 330 Costa, T., Sijacki, D., & Haehnelt, M. G. 2014, MNRAS, 444, 2355 Crenshaw, D. M., Kraemer, S. B., Boggess, A., et al. 1999, ApJ, 516, 750 Cresci, G., Mainieri, V., Brusa, M., et al. 2015, ApJ, 799, 82 Croton, D. J., Springel, V., White, S. D. M., et al. 2006, MNRAS, 365, 11 Croxall, K. V., Smith, J. D., Wolfire, M. G., et al. 2012, ApJ, 747, 81 Dale, D. A., Helou, G., Brauher, J. R., et al. 2004, ApJ, 604, 565 Dale, D. A., Helou, G., Contursi, A., Silbermann, N. A., & Kolhatkar, S. 2001, ApJ, 549, 215 Darling, J. 2007, in IAU Symposium, Vol. 242, Astrophysical Masers and their Environments, ed. J. M. Chapman & W. A. Baan, 417?426 Darling, J., & Giovanelli, R. 2006, AJ, 132, 2596 Das, V., Crenshaw, D. M., Kraemer, S. B., & Deo, R. P. 2006, AJ, 132, 620 Dasyra, K. M., Tacconi, L. J., Davies, R. I., et al. 2006a, ApJ, 638, 745 ?. 2006b, ApJ, 651, 835 ?. 2007, ApJ, 657, 102 de Grijs, R. 2001, Astronomy and Geophysics, 42, 4.12 Destombes, J. L., Marliere, C., Baudry, A., & Brillet, J. 1977, A&A, 60, 55 Di Matteo, T., Springel, V., & Hernquist, L. 2005, Nature, 433, 604 Di Teodoro, E. M., & Fraternali, F. 2015, MNRAS, 451, 3021 Diamond-Stanic, A. M., Moustakas, J., Tremonti, C. A., et al. 2012, ApJ, 755, L26 211 D??az-Santos, T., Armus, L., Charmandaris, V., et al. 2017, ApJ, 846, 32 Dinerstein, H. L., Lester, D. F., & Werner, M. W. 1985, ApJ, 291, 561 Draine, B. T. 2011, Physics of the Interstellar and Intergalactic Medium Dunn, J. P., Crenshaw, D. M., Kraemer, S. B., & Trippe, M. L. 2008, AJ, 136, 1201 Duric, N., Seaquist, E. R., Crane, P. C., Bignell, R. C., & Davis, L. E. 1983, ApJ, 273, L11 Elbaz, D., Daddi, E., Le Borgne, D., et al. 2007, A&A, 468, 33 Elmouttie, M., Haynes, R. F., Jones, K. L., Sadler, E. M., & Ehle, M. 1998a, MNRAS, 297, 1202 Elmouttie, M., Koribalski, B., Gordon, S., et al. 1998b, MNRAS, 297, 49 Engelbracht, C. W., Kundurthy, P., Gordon, K. D., et al. 2006, ApJ, 642, L127 Espada, D., Matsushita, S., Peck, A., et al. 2009, ApJ, 695, 116 Fabian, A. C. 2012, ARA&A, 50, 455 Fabian, A. C., Sanders, J. S., Haehnelt, M., Rees, M. J., & Miller, J. M. 2013, MNRAS, 431, L38 Falstad, N., Gonza?lez-Alfonso, E., Aalto, S., et al. 2015, A&A, 580, A52 Farrah, D., Lebouteiller, V., Spoon, H. W. W., et al. 2013, ApJ, 776, 38 Ferkinhoff, C., Hailey-Dunsheath, S., Nikola, T., et al. 2010, ApJ, 714, L147 Ferkinhoff, C., Brisbin, D., Nikola, T., et al. 2011, ApJ, 740, L29 Ferna?ndez-Ontiveros, J. A., Spinoglio, L., Pereira-Santaella, M., et al. 2016, ApJS, 226, 19 Feruglio, C., Maiolino, R., Piconcelli, E., et al. 2010, A&A, 518, L155 Filippenko, A. V., & Sargent, W. L. W. 1992, AJ, 103, 28 Fischer, J., Abel, N. P., Gonza?lez-Alfonso, E., et al. 2014, ApJ, 795, 117 Fischer, J., Sturm, E., Gonza?lez-Alfonso, E., et al. 2010, A&A, 518, L41 Flower, D. R., Pineau des Forets, G., Field, D., & May, P. W. 1996, MNRAS, 280, 447 Fo?rster Schreiber, N. M., Genzel, R., Newman, S. F., et al. 2014, ApJ, 787, 38 Garc??a-Burillo, S., Mart??n-Pintado, J., Fuente, A., & Neri, R. 2001, ApJ, 563, L27 212 Garc??a-Burillo, S., Combes, F., Usero, A., et al. 2014, A&A, 567, A125 Genzel, R., Fo?rster Schreiber, N. M., Rosario, D., et al. 2014, ApJ, 796, 7 Ginsburg, A., & Mirocha, J. 2011, PySpecKit: Python Spectroscopic Toolkit, As- trophysics Source Code Library, , , ascl:1109.001 Goicoechea, J. R., & Cernicharo, J. 2002, ApJ, 576, L77 Goicoechea, J. R., Cernicharo, J., Lerate, M. R., et al. 2006, ApJ, 641, L49 Goldsmith, P. F., Y?ld?z, U. A., Langer, W. D., & Pineda, J. L. 2015, ApJ, 814, 133 Gonza?lez-Alfonso, E., & Cernicharo, J. 1999, in ESA Special Publication, Vol. 427, The Universe as Seen by ISO, ed. P. Cox & M. Kessler, 325 Gonza?lez-Alfonso, E., Fischer, J., Aalto, S., & Falstad, N. 2014a, A&A, 567, A91 Gonza?lez-Alfonso, E., Fischer, J., Gracia?-Carpio, J., et al. 2012, A&A, 541, A4 ?. 2014b, A&A, 561, A27 Gonza?lez-Alfonso, E., Fischer, J., Sturm, E., et al. 2015, ApJ, 800, 69 Gonza?lez-Alfonso, E., Fischer, J., Spoon, H. W. W., et al. 2017, ApJ, 836, 11 Goulding, A. D., Alexander, D. M., Bauer, F. E., et al. 2012, ApJ, 755, 5 Greene, J. E., Zakamska, N. L., & Smith, P. S. 2012, ApJ, 746, 86 Griffiths, R. E., Ptak, A., Feigelson, E. D., et al. 2000, Science, 290, 1325 Gu?ltekin, K., Richstone, D. O., Gebhardt, K., et al. 2009, ApJ, 698, 198 Habing, H. J. 1968, Bull. Astron. Inst. Netherlands, 19, 421 Hailey-Dunsheath, S., Nikola, T., Stacey, G. J., et al. 2010, ApJ, 714, L162 Hardcastle, M. J., Worrall, D. M., Kraft, R. P., et al. 2003, ApJ, 593, 169 Harris, G. L. H., Rejkuba, M., & Harris, W. E. 2010, PASA, 27, 457 Harrison, C. M., Alexander, D. M., Mullaney, J. R., & Swinbank, A. M. 2014, MNRAS, 441, 3306 Harrison, C. M., Alexander, D. M., Swinbank, A. M., et al. 2012, MNRAS, 426, 1073 Harrison, C. M., Alexander, D. M., Mullaney, J. R., et al. 2016, MNRAS, 456, 1195 Hashimoto, T., Laporte, N., Mawatari, K., et al. 2018, Nature, 557, 392 213 Heckman, T. M., Alexandroff, R. M., Borthakur, S., Overzier, R., & Leitherer, C. 2015, ApJ, 809, 147 Heckman, T. M., Armus, L., & Miley, G. K. 1990, ApJS, 74, 833 Heckman, T. M., Lehnert, M. D., Strickland, D. K., & Armus, L. 2000, ApJS, 129, 493 Heckman, T. M., & Thompson, T. A. 2017, arXiv e-prints, arXiv:1701.09062 Heesen, V., Beck, R., Krause, M., & Dettmar, R. J. 2011, A&A, 535, A79 Helou, G., Khan, I. R., Malek, L., & Boehmer, L. 1988, ApJS, 68, 151 Herrera-Camus, R., Bolatto, A., Smith, J. D., et al. 2016, ApJ, 826, 175 Herrera-Camus, R., Sturm, E., Gracia?-Carpio, J., et al. 2018a, ApJ, 861, 94 ?. 2018b, ApJ, 861, 95 Higdon, S. J. U., Devost, D., Higdon, J. L., et al. 2004, PASP, 116, 975 Hlavacek-Larrondo, J., Carignan, C., Daigle, O., et al. 2011, MNRAS, 411, 71 Hollenbach, D. J., Takahashi, T., & Tielens, A. G. G. M. 1991, ApJ, 377, 192 Hollenbach, D. J., & Tielens, A. G. G. M. 1997, ARA&A, 35, 179 ?. 1999, Reviews of Modern Physics, 71, 173 Hoopes, C. G., Heckman, T. M., Strickland, D. K., et al. 2005, ApJ, 619, L99 Hopkins, P. F., Cox, T. J., Keres?, D., & Hernquist, L. 2008, ApJS, 175, 390 Hopkins, P. F., Somerville, R. S., Hernquist, L., et al. 2006, ApJ, 652, 864 Houck, J. R., Roellig, T. L., van Cleve, J., et al. 2004, ApJS, 154, 18 Hutton, S., Ferreras, I., Wu, K., et al. 2014, MNRAS, 440, 150 Irwin, J., Wiegert, T., Merritt, A., et al. 2019, AJ, 158, 21 Israel, F. P., Gu?sten, R., Meijerink, R., Requena-Torres, M. A., & Stutzki, J. 2017, A&A, 599, A53 Itoh, T., Done, C., Makishima, K., et al. 2008, PASJ, 60, S251 Izumi, T., Wada, K., Fukushige, R., Hamamura, S., & Kohno, K. 2018, ApJ, 867, 48 Janssen, A. W., Christopher, N., Sturm, E., et al. 2016, ApJ, 822, 43 214 Jarrett, T. H., Chester, T., Cutri, R., Schneider, S. E., & Huchra, J. P. 2003, AJ, 125, 525 Kaufman, M. J., Wolfire, M. G., & Hollenbach, D. J. 2006, ApJ, 644, 283 Kaufman, M. J., Wolfire, M. G., Hollenbach, D. J., & Luhman, M. L. 1999, ApJ, 527, 795 Kennicutt, Jr., R. C. 1998, ApJ, 498, 541 Kimm, T., Somerville, R. S., Yi, S. K., et al. 2009, MNRAS, 394, 1131 King, A. R. 2010, in Astronomical Society of the Pacific Conference Series, Vol. 427, Accretion and Ejection in AGN: a Global View, ed. L. Maraschi, G. Ghisellini, R. Della Ceca, & F. Tavecchio, 315 King, A. R., Zubovas, K., & Power, C. 2011, MNRAS, 415, L6 Koornneef, J. 1993, ApJ, 403, 581 Kormendy, J., & Ho, L. C. 2013, ARA&A, 51, 511 Kormendy, J., & Richstone, D. 1995, ARA&A, 33, 581 Kornei, K. A., Shapley, A. E., Martin, C. L., et al. 2012, ApJ, 758, 135 ?. 2013, ApJ, 774, 50 Koss, M., Mushotzky, R., Baumgartner, W., et al. 2013, ApJ, 765, L26 Koss, M., Mushotzky, R., Veilleux, S., et al. 2011, ApJ, 739, 57 Kraft, R. P., Forman, W. R., Jones, C., et al. 2002, ApJ, 569, 54 Krieger, N., Bolatto, A. D., Walter, F., et al. 2019, ApJ, 881, 43 Krips, M., Mart??n, S., Eckart, A., et al. 2011, ApJ, 736, 37 Krug, H. B., Rupke, D. S. N., & Veilleux, S. 2010, ApJ, 708, 1145 Laine, S., & Beck, R. 2008, ApJ, 673, 128 Lehnert, M. D., & Heckman, T. M. 1996, ApJ, 462, 651 Leroy, A. K., Walter, F., Martini, P., et al. 2015, ApJ, 814, 83 Liseau, R., Justtanont, K., & Tielens, A. G. G. M. 2006, A&A, 446, 561 Lutz, D., Sturm, E., Genzel, R., et al. 2003, A&A, 409, 867 Lutz, D., Berta, S., Contursi, A., et al. 2015, ArXiv e-prints, arXiv:1511.02075 Madau, P., Ferrara, A., & Rees, M. J. 2001, ApJ, 555, 92 215 Maiolino, R., Alonso-Herrero, A., Anders, S., et al. 2000, ApJ, 531, 219 Maiolino, R., Gallerani, S., Neri, R., et al. 2012, MNRAS, 425, L66 Malkan, M. A., Gorjian, V., & Tam, R. 1998, ApJS, 117, 25 Marconi, A., & Hunt, L. K. 2003, ApJ, 589, L21 Marconi, A., Moorwood, A. F. M., Origlia, L., & Oliva, E. 1994, The Messenger, 78, 20 Marconi, A., Oliva, E., van der Werf, P. P., et al. 2000a, A&A, 357, 24 Marconi, A., Schreier, E. J., Koekemoer, A., et al. 2000b, ApJ, 528, 276 Marinucci, A., Risaliti, G., Wang, J., et al. 2012, MNRAS, 423, L6 Martin, C. L. 2005, ApJ, 621, 227 ?. 2006, ApJ, 647, 222 Martin, C. L., Shapley, A. E., Coil, A. L., et al. 2012, ApJ, 760, 127 ?. 2013, ApJ, 770, 41 Martin, C. L., & Soto, K. T. 2016, ApJ, 819, 49 Matsuta, K., Gandhi, P., Dotani, T., et al. 2012, ApJ, 753, 104 McCormick, A., Veilleux, S., Mele?ndez, M., et al. 2018, MNRAS, 477, 699 McElroy, R., Croom, S. M., Pracy, M., et al. 2015, MNRAS, 446, 2186 McKinley, B., Briggs, F., Gaensler, B. M., et al. 2013, MNRAS, 436, 1286 McKinley, B., Tingay, S. J., Carretti, E., et al. 2018, MNRAS, 474, 4056 Meijerink, R., Spaans, M., & Israel, F. P. 2007, A&A, 461, 793 Meijerink, R., Spaans, M., Loenen, A. F., & van der Werf, P. P. 2011, A&A, 525, A119 Mele?ndez, M., Mushotzky, R. F., Shimizu, T. T., Barger, A. J., & Cowie, L. L. 2014, ApJ, 794, 152 Mele?ndez, M., Veilleux, S., Martin, C., et al. 2015, ApJ, 804, 46 Middelberg, E., Krichbaum, T. P., Roy, A. L., Witzel, A., & Zensus, J. A. 2005, Astronomical Society of the Pacific Conference Series, Vol. 340, Approaching NGC 3079 with VLBI, ed. J. Romney & M. Reid, 140 Middelberg, E., Roy, A. L., Nagar, N. M., et al. 2004, A&A, 417, 925 216 Mingo, B., Hardcastle, M. J., Croston, J. H., et al. 2012, ApJ, 758, 95 Moorwood, A. F. M., van der Werf, P. P., Kotilainen, J. K., Marconi, A., & Oliva, E. 1996, A&A, 308, L1 Morganti, R., Frieswijk, W., Oonk, R. J. B., Oosterloo, T., & Tadhunter, C. 2013, A&A, 552, L4 Morganti, R., Killeen, N. E. B., Ekers, R. D., & Oosterloo, T. A. 1999, MNRAS, 307, 750 Morganti, R., Veilleux, S., Oosterloo, T., Teng, S. H., & Rupke, D. 2016, A&A, 593, A30 Mushotzky, R. F., Shimizu, T. T., Mele?ndez, M., & Koss, M. 2014, ApJ, 781, L34 Naab, T., & Ostriker, J. P. 2017, ARA&A, 55, 59 Nardini, E., Reeves, J. N., Gofford, J., et al. 2015, Science, 347, 860 Nenkova, M., Ivezic?, Z?., & Elitzur, M. 2002, ApJ, 570, L9 Nenkova, M., Sirocky, M. M., Nikutta, R., Ivezic?, Z?., & Elitzur, M. 2008, ApJ, 685, 160 Nesvadba, N. P. H., Polletta, M., Lehnert, M. D., et al. 2011, MNRAS, 415, 2359 Netzer, H., Lutz, D., Schweitzer, M., et al. 2007, ApJ, 666, 806 Offer, A. R., & van Dishoeck, E. F. 1992, MNRAS, 257, 377 Ott, S. 2010, in Astronomical Society of the Pacific Conference Series, Vol. 434, Astronomical Data Analysis Software and Systems XIX, ed. Y. Mizumoto, K.-I. Morita, & M. Ohishi, 139 Parkin, T. J., Wilson, C. D., Schirm, M. R. P., et al. 2014, ApJ, 787, 16 Pedlar, A., Muxlow, T. W. B., Garrett, M. A., et al. 1999, MNRAS, 307, 761 Peng, Y.-j., Lilly, S. J., Kovac?, K., et al. 2010, ApJ, 721, 193 Pereira-Santaella, M., Colina, L., Garc??a-Burillo, S., et al. 2016, A&A, 594, A81 Pilbratt, G. L., Riedinger, J. R., Passvogel, T., et al. 2010, A&A, 518, 7. http: //arxiv.org/abs/1005.5331 Poglitsch, A., Waelkens, C., Geis, N., et al. 2010, A&A, 518, L2. http://arxiv. org/abs/1005.1487 Pound, M. W., & Wolfire, M. G. 2008, in APSC, Vol. 394, Astronomical Data Analysis Software and Systems XVII, ed. R. W. Argyle, P. S. Bunclark, & J. R. Lewis, 654 217 Puccetti, S., Comastri, A., Fiore, F., et al. 2014, ApJ, 793, 26 Quillen, A. C., Brookes, M. H., Keene, J., et al. 2006, ApJ, 645, 1092 Radomski, J. T., Packham, C., Levenson, N. A., et al. 2008, ApJ, 681, 141 Renzini, A., & Peng, Y.-j. 2015, ApJ, 801, L29 Richings, A. J., & Faucher-Giguere, C.-A. 2017, ArXiv e-prints, arXiv:1706.03784 Roche, P. F., Whitmore, B., Aitken, D. K., & Phillips, M. M. 1984, MNRAS, 207, 35 Rogstad, D. H., Lockhart, I. A., & Wright, M. C. H. 1974, ApJ, 193, 309 Roussel, H., Wilson, C. D., Vigroux, L., et al. 2010, A&A, 518, L66 Rubin, K. H. R., Prochaska, J. X., Koo, D. C., et al. 2014, ApJ, 794, 156 Rubin, K. H. R., Weiner, B. J., Koo, D. C., et al. 2010, ApJ, 719, 1503 Rubin, R. H., Dufour, R. J., & Walter, D. K. 1993, ApJ, 413, 242 Rubin, R. H., Simpson, J. P., Erickson, E. F., & Haas, M. R. 1988, ApJ, 327, 377 Rupke, D. 2018, Galaxies, 6, 138 Rupke, D., Gu?ltekin, K., & Veilleux, S. 2017, ArXiv e-prints, arXiv:1708.05139 Rupke, D. S., Veilleux, S., & Sanders, D. B. 2002, ApJ, 570, 588 ?. 2005a, ApJS, 160, 115 ?. 2005b, ApJS, 160, 87 ?. 2005c, ApJ, 632, 751 Rupke, D. S. N., & Veilleux, S. 2011, ApJ, 729, L27 ?. 2013, ApJ, 768, 75 Rupke, D. S. N., & Veilleux, S. 2015, ApJ, 801, 126 Sanders, D. B., & Mirabel, I. F. 1996, ARA&A, 34, 749 Savage, B. D., & Sembach, K. R. 1996, ARA&A, 34, 279 Scheuer, P. A. G. 1974, MNRAS, 166, 513 Schiano, A. V. R. 1985, ApJ, 299, 24 Schinnerer, E., Eckart, A., Tacconi, L. J., Genzel, R., & Downes, D. 2000, ApJ, 533, 850 218 Scho?ier, F. L., van der Tak, F. F. S., van Dishoeck, E. F., & Black, J. H. 2005, A&A, 432, 369 Schreier, E. J., Burns, J. O., & Feigelson, E. D. 1981, ApJ, 251, 523 Schurch, N. J., Roberts, T. P., & Warwick, R. S. 2002, MNRAS, 335, 241 Schwartz, C. M., & Martin, C. L. 2004, ApJ, 610, 201 Scoville, N. Z., Soifer, B. T., Neugebauer, G., et al. 1985, ApJ, 289, 129 Sebastian, B., Kharb, P., O?Dea, C. P., Colbert, E. J. M., & Baum, S. A. 2019, ApJ, 883, 189 Sharp, R. G., & Bland-Hawthorn, J. 2010, ApJ, 711, 818 Shimizu, T. T., Mele?ndez, M., Mushotzky, R. F., et al. 2016, MNRAS, 456, 3335 Shimizu, T. T., Mushotzky, R. F., Mele?ndez, M., Koss, M., & Rosario, D. J. 2015, MNRAS, 452, 1841 Shopbell, P. L., & Bland-Hawthorn, J. 1998, ApJ, 493, 129 Skinner, C. J., Smith, H. A., Sturm, E., et al. 1997, Nature, 386, 472 Smajic?, S., Fischer, S., Zuther, J., & Eckart, A. 2012, A&A, 544, A105 Sofia, U. J., Lauroesch, J. T., Meyer, D. M., & Cartledge, S. I. B. 2004, ApJ, 605, 272 Solomon, P. M., Downes, D., & Radford, S. J. E. 1992, ApJ, 398, L29 Somerville, R. S., & Dave?, R. 2015, ARA&A, 53, 51 Somerville, R. S., Hopkins, P. F., Cox, T. J., Robertson, B. E., & Hernquist, L. 2008, MNRAS, 391, 481 Spinoglio, L., & Malkan, M. A. 1992, ApJ, 399, 504 Spinoglio, L., Malkan, M. A., Smith, H. A., Gonza?lez-Alfonso, E., & Fischer, J. 2005, ApJ, 623, 123 Spinoglio, L., Pereira-Santaella, M., Dasyra, K. M., et al. 2015, ApJ, 799, 21 Spoon, H. W. W., Armus, L., Marshall, J. A., et al. 2009, ApJ, 693, 1223 Spoon, H. W. W., & Holt, J. 2009, ApJ, 702, L42 Spoon, H. W. W., Marshall, J. A., Houck, J. R., et al. 2007, ApJ, 654, L49 Spoon, H. W. W., Farrah, D., Lebouteiller, V., et al. 2013, ApJ, 775, 127 219 Springel, V. 2005, MNRAS, 364, 1105 Stasin?ska, G., Izotov, Y., Morisset, C., & Guseva, N. 2015, A&A, 576, A83 Sternberg, A., & Dalgarno, A. 1995, ApJS, 99, 565 Stone, M., Veilleux, S., Mele?ndez, M., et al. 2016, ApJ, 826, 111 Storey, J. W. V., Watson, D. M., & Townes, C. H. 1981, ApJ, 244, L27 Strateva, I., Ivezic?, Z?., Knapp, G. R., et al. 2001, AJ, 122, 1861 Strickland, D. K., & Heckman, T. M. 2007, ApJ, 658, 258 Strickland, D. K., Heckman, T. M., Weaver, K. A., Hoopes, C. G., & Dahlem, M. 2002, ApJ, 568, 689 Strickland, D. K., & Stevens, I. R. 2000, MNRAS, 314, 511 Sturm, E., Rupke, D., Contursi, A., et al. 2006, ApJ, 653, L13 Sturm, E., Gonza?lez-Alfonso, E., Veilleux, S., et al. 2011, ApJ, 733, L16 Swinyard, B., Nakagawa, T., Merken, P., et al. 2009, Experimental Astronomy, 23, 193 Tadhunter, C., Morganti, R., Rose, M., Oonk, J. B. R., & Oosterloo, T. 2014, Nature, 511, 440 Tielens, A. G. G. M., & Hollenbach, D. 1985, ApJ, 291, 722 Tingay, S. J., Jauncey, D. L., Reynolds, J. E., et al. 1998, AJ, 115, 960 Tommasin, S., Spinoglio, L., Malkan, M. A., & Fazio, G. 2010, ApJ, 709, 1257 Tremonti, C. A., Moustakas, J., & Diamond-Stanic, A. M. 2007, ApJ, 663, L77 Trotter, A. S., Greenhill, L. J., Moran, J. M., et al. 1998, ApJ, 495, 740 Tully, R. B., Rizzi, L., Shaya, E. J., et al. 2009, AJ, 138, 323 Turner, B. E. 1985, ApJ, 299, 312 Turner, T. J., & Pounds, K. A. 1989, MNRAS, 240, 833 Vasudevan, R. V., & Fabian, A. C. 2007, MNRAS, 381, 1235 ?. 2009, MNRAS, 392, 1124 Vasudevan, R. V., Fabian, A. C., Gandhi, P., Winter, L. M., & Mushotzky, R. F. 2010, MNRAS, 402, 1081 Veilleux, S. 1991a, ApJS, 75, 357 220 ?. 1991b, ApJS, 75, 383 ?. 1991c, ApJ, 369, 331 Veilleux, S. 2006, in Astronomical Society of the Pacific Conference Series, Vol. 357, Astronomical Society of the Pacific Conference Series, ed. L. Armus & W. T. Reach, 231 Veilleux, S., & Bland-Hawthorn, J. 1997, ApJ, 479, L105 Veilleux, S., Bolatto, A., Tombesi, F., et al. 2017, ApJ, 843, 18 Veilleux, S., Cecil, G., & Bland-Hawthorn, J. 2005, ARA&A, 43, 769 Veilleux, S., Cecil, G., Bland-Hawthorn, J., et al. 1994, ApJ, 433, 48 Veilleux, S., Kim, D.-C., & Sanders, D. B. 1999, ApJ, 522, 113 Veilleux, S., Kim, D. C., & Sanders, D. B. 2002, ApJS, 143, 42. http://arxiv. org/abs/astro-ph/0207401 Veilleux, S., Maiolino, R., Bolatto, A. D., & Aalto, S. 2020, arXiv e-prints, arXiv:2002.07765 Veilleux, S., Shopbell, P. L., Rupke, D. S., Bland -Hawthorn, J., & Cecil, G. 2003, AJ, 126, 2185 Veilleux, S., Rupke, D. S. N., Kim, D.-C., et al. 2009, ApJS, 182, 628 Veilleux, S., Mele?ndez, M., Sturm, E., et al. 2013, ApJ, 776, 27 Venturi, G., Marconi, A., Mingozzi, M., et al. 2017, Frontiers in Astronomy and Space Sciences, 4, 46 Ve?ron-Cetty, M.-P., & Ve?ron, P. 2006, A&A, 455, 773 Voit, G. M. 1992, ApJ, 399, 495 Walter, F., Weiss, A., & Scoville, N. 2002, ApJ, 580, L21 Walter, F., Bolatto, A. D., Leroy, A. K., et al. 2017, ApJ, 835, 265 Weiner, B. J., Coil, A. L., Prochaska, J. X., et al. 2009, ApJ, 692, 187 Werner, M. W., Roellig, T. L., Low, F. J., et al. 2004, ApJS, 154, 1 Westmoquette, M. S., Smith, L. J., Gallagher, J. S., & Walter, F. 2013, MNRAS, 428, 1743 Westmoquette, M. S., Smith, L. J., & Gallagher, III, J. S. 2011, MNRAS, 414, 3719 Whitaker, K. E., Franx, M., Leja, J., et al. 2014, ApJ, 795, 104 221 Wills, K. A., Redman, M. P., Muxlow, T. W. B., & Pedlar, A. 1999, MNRAS, 309, 395 Wilson, A. S., & Ulvestad, J. S. 1987, ApJ, 319, 105 Wilson, C. D., Petitpas, G. R., Iono, D., et al. 2008, ApJS, 178, 189 Winter, L. M., Veilleux, S., McKernan, B., & Kallman, T. R. 2012, ApJ, 745, 107 Wolfire, M. G., Tielens, A. G. G. M., & Hollenbach, D. 1990, ApJ, 358, 116 Wright, E. L. 2006, PASP, 118, 1711 Yamauchi, A., Nakai, N., Sato, N., & Diamond, P. 2004, PASJ, 56, 605 Zakamska, N. L., Hamann, F., Pa?ris, I., et al. 2016, MNRAS, 459, 3144 Zhang, D. 2018, Galaxies, 6, 114 Zschaechner, L. K., Walter, F., Bolatto, A., et al. 2016, ApJ, 832, 142 Zschaechner, L. K., Bolatto, A. D., Walter, F., et al. 2018, ApJ, 867, 111 Zubovas, K., & Nayakshin, S. 2014, MNRAS, 440, 2625 222