Photoionization calculations of the radiation force due to spectral lines in AGNs. Submitted to ApJ

Similar documents
Astrophysical Quantities

ACTIVE GALACTIC NUCLEI: FROM THE CENTRAL BLACK HOLE TO THE GALACTIC ENVIRONMENT

Simulations of Winds. Daniel Proga University of Nevada, Las Vegas Princeton University

Astrophysics of feedback in local AGN and starbursts

Accretion Disks. 1. Accretion Efficiency. 2. Eddington Luminosity. 3. Bondi-Hoyle Accretion. 4. Temperature profile and spectrum of accretion disk

AGN Physics of the Ionized Gas Physical conditions in the NLR Physical conditions in the BLR LINERs Emission-Line Diagnostics High-Energy Effects

Opacity and Optical Depth

arxiv:astro-ph/ v1 4 Jan 2007

Warm absorbers from torus evaporative flows(??) Tim Kallman (NASA/GSFC) + A. Dorodnitsyn (GSFC) + D. Proga (UNLV)

VII. Hydrodynamic theory of stellar winds

Astronomy. Astrophysics. A non-hydrodynamical model for acceleration of line-driven winds in active galactic nuclei. G. Risaliti 1,2 and M.

Francesco Tombesi. University of Rome, Tor Vergata NASA - Goddard Space Flight Center University of Maryland, College Park

AGN feedback from accretion disk winds

The Stellar Opacity. F ν = D U = 1 3 vl n = 1 3. and that, when integrated over all energies,

RELATIVISTIC SPECTROSCOPY OF BLACK HOLES

Monte Carlo Simulator to Study High Mass X-ray Binary System

Ultra-fast disk wind from a high accretion rate black hole 1H

X-raying the interstellar medium at high latitudes

How do disks transfer angular momentum to deliver gas onto compact objects? How do accretion disks launch winds and jets?

arxiv:astro-ph/ v1 25 Sep 2006

Radiative Transfer and Stellar Atmospheres

AG Draconis. A high density plasma laboratory. Dr Peter Young Collaborators A.K. Dupree S.J. Kenyon B. Espey T.B.

Diffuse Interstellar Medium

Quasar Winds: the 4 th element

Stars AS4023: Stellar Atmospheres (13) Stellar Structure & Interiors (11)

A study of accretion disk wind emission

Broadband X-ray emission from radio-quiet Active Galactic Nuclei

Studying AGN winds and feedback with Athena. Massimo Cappi. INAF/IASF-Bologna. 4. Future

The Powerful Black Hole Wind in the Luminous Quasar PDS 456

Some HI is in reasonably well defined clouds. Motions inside the cloud, and motion of the cloud will broaden and shift the observed lines!

Electromagnetic Spectra. AST443, Lecture 13 Stanimir Metchev

Variable X-ray absorption in the mini-bal QSO PG

Astronomy 421. Lecture 14: Stellar Atmospheres III

Redshifted Broad Absorption Troughs in Quasars

PHOTOIONIZATION X-RAY BINARIES INSTABILITY OF WINDS IN STEFANO BIANCHI. June 8 th 2017 The X-ray Universe 2017 Rome, Italy

X-ray Radiation, Absorption, and Scattering

(c) Sketch the ratio of electron to gas pressure for main sequence stars versus effective temperature. [1.5]

Discovery of Associated Absorption Lines in an X-Ray Warm Absorber:[ITAL]Hubble Space Telescope[/ITAL] Observations of PG

Theory of optically thin emission line spectroscopy

Advanced Stellar Astrophysics

WINDS OF HOT MASSIVE STARS III Lecture: Quantitative spectroscopy of winds of hot massive stars

Physics and Chemistry of the Interstellar Medium

AC Fabian, M. Cappi, J Sanders. Cosmic Feedback from AGN

arxiv:astro-ph/ v1 17 Dec 2001

arxiv:astro-ph/ v3 27 May 2005

Astrophysics of Gaseous Nebulae and Active Galactic Nuclei

Accretion disks. AGN-7:HR-2007 p. 1. AGN-7:HR-2007 p. 2

Thermal and Ionization Aspects of Flows from Hot Stars: Observations and Theory

X-ray signatures of AGN outflows: multi-dimensional radiative transfer simulations

Three Dimensional Radiative Transfer in Winds of Massive Stars: Wind3D

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

THE OBSERVATION AND ANALYSIS OF STELLAR PHOTOSPHERES

X-ray Emission from O Stars

Overview spherical accretion

Broad X-ray Absorption lines from the Wind in PDS 456

X-ray Emission from O Stars

X-ray Emission from O Stars. David Cohen Swarthmore College

arxiv:astro-ph/ v1 23 Dec 2005

THE PHYSICS AND EVOLUTION OF ACTIVE GALACTIC NUCLEI

Low-Ionization BAL QSOs in Ultraluminous Infrared Systems

arxiv:astro-ph/ v1 4 Oct 2004

Black Hole Formation in the Early Universe

The Physical Basis of the L x L bol Empirical Law for O-star X-rays

Supernovae. Supernova basics Supernova types Light Curves SN Spectra after explosion Supernova Remnants (SNRs) Collisional Ionization

Nuclear X-ray Emission and Mass Outflow From Broad Lined Radio Galaxies (Lorentz Center, Leiden 2009)

Stellar atmospheres: an overview

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

Energy transport: convection

Observational Appearance of Black Hole Wind Effect of Electron Scattering

AGN Central Engines. Supermassive Black Holes (SMBHs) Masses and Accretion Rates SMBH Mass Determinations Accretion Disks

A Global View: AGN Ionized Winds

arxiv: v1 [astro-ph.co] 16 Dec 2011

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

SISD Training Lectures in Spectroscopy

Active Galactic Nuclei

Lecture 6: Continuum Opacity and Stellar Atmospheres

ASTR-1010: Astronomy I Course Notes Section IV

ASTRONOMY QUALIFYING EXAM August Possibly Useful Quantities

Simulations of Mass Outflows from Accretion Powered Sources

Outflowing disk formation in B[e] supergiants due to rotation and bi-stability in radiation driven winds

Protostars 1. Early growth and collapse. First core and main accretion phase

Simultaneous X-ray and Radio Observations of Seyferts, and Disk-Jet Connections

Stellar Winds: Mechanisms and Dynamics

The Chandra Survey of Outflows in AGN with Resolved Spectroscopy (SOARS)

Supernovae. Supernova basics Supernova types Light Curves SN Spectra after explosion Supernova Remnants (SNRs) Collisional Ionization

Molecular hydrogen in the chromosphere IRIS observations and a simple model

The optical Fe II emission lines in Active Galactic Nuclei

X-ray Radiation, Absorption, and Scattering

Section 11.5 and Problem Radiative Transfer. from. Astronomy Methods A Physical Approach to Astronomical Observations Pages , 377

Calculating Radiative Recombination Continuum From a Hot Plasma

THE INVERSE-COMPTON X-RAY SIGNATURE OF AGN FEEDBACK

Notes on Photoionized Regions Wednesday, January 12, 2011

Narrow (UV) Absorption Line Outflows from Quasars

Ay Fall 2004 Lecture 6 (given by Tony Travouillon)

Hubble Space Telescope ultraviolet spectroscopy of blazars: emission lines properties and black hole masses. E. Pian, R. Falomo, A.

A Unified Model for AGN. Ryan Yamada Astro 671 March 27, 2006

Model Atmospheres. Model Atmosphere Assumptions

The Broad Line Region - An Innermost Torus

Atomic Structure & Radiative Transitions

Substellar Atmospheres II. Dust, Clouds, Meteorology. PHY 688, Lecture 19 Mar 11, 2009

Transcription:

Draft version December 6, 2018 Typeset using LATEX twocolumn style in AASTeX62 Photoionization calculations of the radiation force due to spectral lines in AGNs Randall C. Dannen, 1 Daniel Proga, 1 Timothy R. Kallman, 2 and Tim Waters 3 can determine the ionization and temperature. Such is thought to be the case for the broad and narrow line regions (BLRs and NLRs) of AGNs, the source of prominent spectral features formed through such processes as line emission, absorption, and scattering. From the thermodynamics point of view, the radiation is responsible for heat supply to these regions. Detailed photoionization, radiative transfer, and energy balance calculations are required to create models that can account for the presence of ions responsible for producing the observed spectral features of AGNs. To explain the observed spectra, such calculations rearxiv:1812.01773v1 [astro-ph.ga] 5 Dec 2018 1 Department of Physics & Astronomy University of Nevada, Las Vegas 505 S. Maryland Pkwy Las Vegas, NV, 8915-002, USA 2 NASA Goddard Space Flight Center Greenbelt, MD 20771, USA 3 Theoretical Division, Los Alamos National Laboratory (Received December 6, 2018; Revised December 6, 2018; Accepted 2019) Submitted to ApJ ABSTRACT One of the main physical mechanisms that could drive mass outflows in AGNs is radiation pressure on spectral lines. Although this mechanism is conceptually straightforward to understand, the actual magnitude of the radiation force is challenging to compute because the force depends on the physical conditions in the gas, as well as the strength, spectral energy distribution (SED), and geometry of the radiation field. We present results from our photoionization and radiation transfer calculations of the so-called force multiplier, M, using the same radiation field to compute the gas photoionization and thermal balance. Our method is general and can be used for an arbitrary SED. Here we focus on describing results for two SEDs that correspond to a Type 1 and Type 2 AGN. We use the photoionization code XSTAR and take into account the most up-to-date and complete atomic data and line list. Our main results are the following: 1) for a fixed value of the optical depth parameter, M is not a monotonic function of the photoionization parameter ξ. Although M starts to decrease with ξ for ξ > 1 as shown by others, this decrease in our calculations is relatively gradual and in fact M can increase by a factor of few at ξ 00 (details depend on the assumed SED). The main dynamically relevant effect of this behavior is that the multiplier can stay larger than 1 for ξ as high as 00; 2) at these same ξ for which the multiplier is higher than in previous calculations, the gas is thermally unstable to isobaric perturbations. We discuss implications of our results in the context of AGN winds that are observed in the UV and X-ray bands, and we confirm previous findings that the force multiplier depends on very many lines, not a few very optically thick lines. Keywords: galaxies: active - methods: numerical - hydrodynamics - radiation: dynamics 1. INTRODUCTION Active galactic nuclei (AGNs) are examples of astrophysical objects that emit photons at very high rates and over a very broad range of energies. The radiation energy density can be significant compared to the density of any other form of energy in the regions where the AGN radiation can penetrate. If the gas in these regions has a significant opacity, the radiative processes Corresponding author: Randall Dannen randall.dannen@unlv.edu

2 Dannen et al. quire certain assumptions or results from other types of models. This additional level of complexity comes from the fact that observed lines are relatively broad and can have profiles of rather complex shapes (e.g., P-Cygni like line profiles) which indicate non-thermal line broadening most likely due to the Doppler effect associated with supersonic motion (e.g., Krolik 1999). Supersonic motion could be a result of thermal driving, that is, the work done by the gas against the surroundings and the gravity of the supermassive black hole as a consequence of the first law of thermodynamics (e.g., see Lamers & Cassinelli 1999, for a review of the basic physics of winds). As mentioned above, the AGN radiation acts as a strong heat source but may also supply momentum. The work done by the radiation force f rad has been shown to be the main mechanism for driving supersonic gas flows in objects such as OB stars (e.g., Castor et al. 1975, hereafter CAK) and cataclysmic variables (e.g, Pereyra et al. 1997; Proga et al. 1998), and it could also be important in AGNs (Arav & Li 199; Murray et al. 1995; Proga et al. 2000; Proga 2007). A self-consistent and complete treatment of the most important effects of radiation on gas dynamics requires time consuming calculations. Therefore, most theoretical studies of gas flows that are driven by thermal pressure neglect or simplify the treatment of the radiation force, whereas studies of radiation pressure driven flows neglect or simplify the radiation heating and cooling rates, Γ and Λ, respectively. In some applications, one could justify neglecting effects of the radiation force by referring to the rule of thumb that radiation can heat (cool), but frequently finds it difficult to push (Shu 1992). However, under some circumstances, radiation can effectively push gas, for example, when the total opacity of the gas, κ tot, is dominated by the contribution from photon scattering (no energy transfer) rather than from photon absorption (heating). OB stars are examples of objects where the total opacity in their upper atmospheres and winds is dominated by contributions from spectral line transitions, which mostly scatter photons, hence their winds are driven by the so-called line force, f rad,l [e.g., CAK and see Chapter 8 in Lamers & Cassinelli (1999) for an overview]. Another example includes situations where the gas temperature does not change much downstream (i.e., gas is nearly isothermal). In these cases, the rate of energy change of the gas due to the radiation force (i.e., the rate of work done by radiation force, vf rad, where v is the flow velocity) can be higher than the rate of heat deposition, Q L = Γ Λ. 1.1. Motivation This paper is part of a series in which we have been studying AGN gas flows where both radiation source terms are included in a progressively more self-consistent manner (e.g., Proga et al. 2000; Proga 2007; Waters & Proga 2016). To properly include radiation heating and cooling and pressure, one needs to compute selfconsistently the coupling between radiation and matter, (i.e., the gas opacity and emissivity) from the underlying spectral energy distribution (SED) of the electromagnetic radiation. The top panel of Fig. 1 shows examples of Type 1 and Type 2 AGN SEDs adopted from Mehdipour et al. (2015). In the bottom panels of the figure, we mark the fraction of the total energy contained in various portions of the electromagnetic spectrum. We note that the comparison of ratios between the hard and soft UV and X-ray fractional energy is likely an important factor. That is, the quantity of the radiation is clearly important, but the shape of the radiation field is also important. For example, modeling of AGN observations requires spectra that are of a certain hardness (e.g., Mehdipour et al. 2015, and references therein). However, only a fraction of the physically relevant SED energy range can be observed. The behavior in the unobservable EUV region, for example, must be assumed. In addition, the SED also affects the gas thermal stability (e.g., Kallman & McCray 1982; Krolik 1999; Mehdipour et al. 2015; Dyda et al. 2017). In particular, and AGN2 both have regions of isobaric thermal instability, (we return to this point in section 3). In addition, the deficit of soft photons in AGN2 allows for isochoric instability, which leads to the formation of non-isobaric clouds (Waters & Proga 2018). Dyda et al. (2017, D17 hereafter) described a general method for the self-consistent modeling of the outflow hydrodynamics that result from irradiation of an astrophysical object by a radiation field with an arbitrary strength and SED. D17 used the photoionization code XSTAR (Kallman & Bautista 2001) to calculate Γ and Λ as a function of gas photoionization parameter ξ = πf X n, (1) and gas temperature, T, where F X is the integrated flux from 0.1 Ry-00 Ry and n hydrogen nucleon number density. D17 explored several SEDs: those of Type 1 and Type 2 AGN from Fig. 1, as well as SEDs for hard and soft state X-ray binaries, bremsstrahlung and blackbody. This general method was applied to study the hydrodynamics of 1-D spherical winds heated by a uniform radiation field using the magnetohydrodynamic (MHD) code Athena++ (Gardiner & Stone 2005, 2008).

Photoionization calculations of the radiation force due to spectral lines in AGNs 3 0 log[ J ] Fraction of Total Energy 2 6 1 0 1 2 3 5 6 log[energy] [ev] 0. 0. 0.3 0.2 0.1 0.0 IR Optical UV X-ray 0.3 0.2 0.1 0.0 AGN2 IR Optical UV X-ray AGN2 Figure 1. Top panel: SEDs of two AGN types used in our calculations: represents an SED of a Type I AGN (solid line) while AGN2 represents an SED of a Type II AGN (dashed line). We normalized the SEDs using the maximum emission as a unit. We use different color shading to show the IR, optical, UV, and X-ray energy bands. The two vertical yellow dashed lines mark the energy interval used to calculate ξ. Bottom panels: Fraction of the total energy of the SED in each energy component band for and AGN2, left and right panel, respectively. We also show the fraction of the total energy used in computing ξ (i.e., the fraction of the total energy between the vertical yellow dashed lines in the top panel). D17 found that in all the stated cases a wind settles into a transonic, steady state. The wind is at near radiative heating equilibrium (i.e., L 0) until adiabatic cooling becomes important and the flow temperature can be significantly smaller than the one corresponding to the radiative equilibrium value for a given ξ. D17 s results also show how the efficiency with which the radiation field transfers energy to the wind is dependent on the SED of the external source, particularly the relative flux of soft X-rays. This soft X-ray dependence is related to the thermal stability of the gas (Field 1965), namely a relative deficit of the soft X-rays leads to more unstable gas (see also Kallman & McCray 1982) which in turn increases heat deposition and the flow velocity. Overall, D17 s results demonstrate how detailed photoionization calculations are essential to properly capture the flow dynamics. In this paper, we make the next step in our development of a self-consistent comprehensive model of astrophysical winds and use the same photoionization code XSTAR as in D17, to compute not only the heat deposition but also the line force as a function of ξ and T. Here we limit our presentation of the results for the two AGN SEDs shown in fig. 1. The data tables for and AGN2 from D17 (heating and cooling) and presented in this work (force multipliers), along with sample code for utilizing them can be found on our project webpage 1. Our ultimate goal is not only to compute the line force but also to self-consistently combine the radiation driving and thermal driving due to the same radiation field. Moreover, we are now equipped to explore the parameter space and identify regions where the radiation force dominates over thermal driving in determining the mass loss rate or terminal velocity (or both). In Section 2, we summarize the key elements of our calculations. We present the results from our calculations in Section 3, and a discussion of our results and our conclusions are presented in Section. 2. NUMERICAL METHODS In this section, we describe our process for calculating the force multiplier. Our method of computing the line force is based on the classic method due to CAK. More specifically, we follow (Stevens & Kallman 1990, SK90 heareafter) who used XSTAR version 1 to calculate the state of the gas. Here we use XSTAR version 2.37 along with the most up-to-date lists of spectral lines, using the AGN SEDs from Fig. 1 as inputs to the code. In addition, we use the same radiation field to compute both f rad,l and L. SK90 considered an X-ray binary system where black body radiation from a star drives a 1 http://www.physics.unlv.edu/astro/xstartables.html

Dannen et al. UV 6 6 6 5 5 5 Number of Lines 0 3 Energy [ev] 0 20 5 0 5 0 5 gf [Oscillator Strength] 0 0 5 5 20 25 Ionization Degree Figure 2. Histograms illustrating properties of the lines used in our calculations. Left panel: Solid lines indicate the number of lines as function of energy in units of ev for the atomic database used in this work (black line) and used in SK90 (green line). For emphasis, we shade in blue the energy range for UV photons and in red for the X-rays. Middle panel: Number of lines as a function of oscillator strength gf for the entire line list (black line), including only the UV lines (blue line), and only X-ray lines (red line). Right panel: Number of lines as a function of ionization degree for the atomic database used in this work (black line), UV lines only (blue line), X-ray lines only (red line), and the atomic database used in SK90 (green line). stellar wind which was irradiated by X-rays emitted by a companion. Unlike SK90, we assume the radiation field for both the ionizing flux and line driving is the same, as this is more appropriate for modeling gas dynamics in AGNs. Other authors have also explored the line force due to AGNs (e.g. Arav & Li 199; Chelouche & Netzer 2003; Everett 2005; Chartas et al. 2009; Saez & Chartas 2011) but they used a different photoionization code (e.g. CLOUDY) and different atomic data sets and line lists. The line list that we use is a combination of the XSTAR atomic data set and the atomic data curated by Robert L. Kurucz 2. We take special care when merging these atomic data sets to not double count any lines. If a line was found in both data sets, we prioritize the XSTAR atomic data for X-ray lines and high energy UV lines and Kurucz s data set otherwise. Information about the distribution of lines as a function of energy, oscillator strength, and ionization degree is shown in Fig. 2. This figure also includes information about the atomic data set used by SK90. Our current atomic data set contains over two million lines, covering a wider range of energies and ionization degrees and allowing for a more complete calculation of the force multiplier compared to previous 2 http://kurucz.harvard.edu studies, especially due to X-rays lines and lines from highly ionized plasma. We adopt the elemental abundances as Mehdipour et al. (2016) in both models. The ionization balance is determined by the external radiation field rather than by the LTE assumption (i.e. Saha ionization balance), using the photoionization code XSTAR to determine the ion abundances as a function of ξ. The gas temperature is also function of ξ, which entails an implicit assumption that the gas is optically thin. In the two top panels of Fig. 3, we show the thermal equilibrium temperature as a function of ξ (black solid lines) and we shaded regions where the condition for isobaric thermal instability is satisfied (Field 1965). The force per unit mass due to an individual line can be computed as f L = F ν ν D c κ L τ L ( 1 e τ L ), (2) where κ L is the line s opacity, τ L the line s optical depth, ν D = ν 0 v th /c is the line s thermal Doppler width, with v th being defined as the thermal speed of the ion that a given atomic line belongs to, and F ν is the specific flux (Castor 197). The optical depth for specific line in a static atmosphere is τ L = r ρκ L dr, (3)

Photoionization calculations of the radiation force due to spectral lines in AGNs 5 M(t = 6, ) max 0 0 5 9 7 5 2 0 5 8 7 6 5 AGN2 0 Total UV X-ray 0 5 9 UV X-ray 7 5 2 0 5 8 7 T [K] 6 5 Figure 3. Summary of results from our photoionization and line force calculations. Top panels: The force multiplier as a function of ξ for an optically thin gas (i.e., t = 6 ) for and AGN2, left and right panel, respectively. This low t = 6 is our proxy for M max. The solid blue, dashed red, and dash-dotted green line correspond to the force multiplier due to all lines, only UV lines, and only X-ray lines, respectively (the ordinate on the left side). The solid black line represents the thermal equilibrium temperature determined by XSTAR corresponding to L = 0 (the ordinate on the right side of each of the top two panels). The shaded regions indicate the parameter space where the gas is thermally unstable to isobaric linear perturbations [i.e., [ log T/ log ξ] L 1; (e.g., Barai et al. 2012)]. Bottom panel: The opacity of the single strongest UV and X-ray line as a function of ξ (dashed red and dash dotted green lines, respectively). The line opacity is in units of electron scattering opacity (see eq. 8). while in an expanding atmosphere is τ L = ρκ L l Sob. () Here, l Sob v th / dv/dl is the so-called Sobolev length. CAK defined a local optical depth parameter t = σ e ρ l Sob. (5) We take v th to be the proton thermal speed at 50,000 K. For a given line, we can define the opacity accounting for stimulated emission, κ L = πe2 m e c gf N L/g L N U /g U ρ ν D, (6) where κ L is the opacity in units cm 2 g 1 (where all other symbols have their conventional meaning). We assume a Boltzmann distribution when determining the level populations. The exicted levels are overpopulated using this distribution when compared to results from XSTAR where the level populations are computed by solving more general statistical equilibrium equations. Our simplified calculations of the level populations overestimate the force multiplier. It is conventional to rewrite the optical depth as where η is a rescaling of line opacity τ L = ηt, (7) η = κ L /σ e, (8) We can now write an expression for the total acceleration due to lines as a L = σ ef M(t, ξ), (9) c where M(t, ξ) is the total force multiplier given by M(t, ξ) = lines ν D F ν F 1 ( 1 e ηt ). () t

6 Dannen et al. 3 2 1 0 0 0 3 2 1 0 0 0 3 2 1 0 0 M(t, ) log( )=-2 log( )=5 8 6 2 log( )=-2 M(t, ) 3 2 1 0 0 8 6 log( )=5 t 2 log(t)=-8 log(t)=1 2 1 0 1 2 3 5 AGN2 log(t)=-8 log(t)=1 2 1 0 1 2 3 5 Figure. The main results from our photoionization and line force calculations. Top panels: These correspond to. The left panel shows our results as a function of t for a fixed ξ. The solid lines show results for a fixed ξ [from top to bottom, ξ increases from 2 (violet region) to 5 (red region)], and the ξ levels are separated by log(ξ) = 1. The right panel shows results of our force multiplier calculations as function of ξ for a fixed t. The solid lines show results for a fixed t [from top to bottom, t increases from 8 (violet region) to (red region)], and the t levels are separated by log(t) = 1. Bottom panels: Same as the top panels for AGN2. Although, t depends on the temperature through vth, the temperature value is arbitrary and has no direct role in M ; our calculations of M just require one be specified (see Gayley 1995, for a detailed discussion of this point). CAK found that M increases with decreasing t and it saturates as t approaches zero (i.e., gas becomes optically thin even for the most opaque lines). We will refer to the saturated value of M as Mmax. CAK also showed that for OB stars, Mmax can be as high as 2000 (see also Gayley 1995). This means that the gravity can be overcome by the radiation force even if the total luminosity, L, is much smaller than the Eddington luminosity, LEdd = πcgm/σe. In other words, the radiation force can drive a wind when L0 Mmax > 1, where L0 L/LEdd is the so-called Eddington factor. The key result of SK90 was to show how M changes not only as a function of t but also as a function of ξ. In particular, they found that Mmax increases gradually from 2000 to 5000 as ξ increases from 1 to 3 and then drops to 0.1 at ξ 00. The line force becomes negligible for ξ > 0 because then Mmax. 1. The force multiplier depends also on other gas and radiation properties, for example gas metalicity (e.g., CAK) and column density, NH (e.g., Stevens 1991). However, here we concentrate only on the effects due to t and ξ for a given SED and chemical abundance (the same as the ones in Mehdipour et al. 2016). 3. RESULTS In Fig., we show how the force multiplier changes as a function of ξ and t. The left hand side panels show that M is a mostly monotonic function of t that saturates at very small t for all ξ, while at large t, M is a power law of t as first found by CAK. The right hand side panels show that M decreases slightly with increasing ξ for small ξ and a fixed t. However, for ξ & 1, M decreases significantly with increasing ξ (the details depend on t). The decrease is not monotonic; there is a resurgence in the line force for ξ 0 00 for the case. This bump is also a feature found in SK90 (see Fig. 2 there). No such feature was shown in the work of Everett (2005), Chartas et al. (2009), and Saez & Chartas (2011); their models indicate there should be a dramatic

Photoionization calculations of the radiation force due to spectral lines in AGNs 7 M(t 6, ) M(t 6, ) 0 H He N C O Ne Fe Ni 0 5 0 HeII CIV OVIII FeII FeXXIII FeXXIV FeXXVI 2 0 5 AGN2 0 0 5 0 2 0 5 Figure 5. Main contributors to the total force multipliers. The left panels correspond to and the right panels correspond to AGN2. The solid black line in each panel represents the total force multiplier for t 6 (our proxy for M max). The top panels shows the contribution to M max for various elements of interest (see the legend in the bottom left corner of the top left panel). Similarly, the bottom panels show the contribution of various ions (see the legend in the bottom right corner of the bottom left panel). decrease of M in that region. Our calculations are nevertheless in agreement with the finding made by these recent studies showing that M(t, ξ) can be larger than 00 for low values of ξ. The force multiplier is a sum of opacities due to various lines. To better understand its properties, in Fig. 5 we show the contributions to the force multiplier from various ions as function of ξ for t = 6 [M(t = 6, ξ) is our proxy for M max so this figure illustrates how M max changes with ξ]. Similar to SK90, we find that Fe and He are the dominant contributors to the force multiplier. Table 1 lists the top five contributors to the force multiplier for four values of ξ. The top contributors for low values of ξ contribute < 5% to the total force multiplier, agreeing with previous findings (e.g. Abbott 1982, Gayley 1995, & Puls et al. 2000). Comparing to Fig. 5, we notice that although He contributes a substantial amount to M at ξ 0, it is absent from our table. Thus, this table illustrates an important result: the force multiplier is due to the contributions of many weak lines rather than a few very strong lines. Returning to Fig. 3, we demonstrate how M depends on the energy of the photons. We present the UV and X-ray contributions to the force multiplier (top panels). The UV band contributes the majority of the line force with the X-ray energy band only being important for a narrow window in (i.e., for ξ between 200 and 00). We find a similar trend for the maximum opacity of a single line (bottom panels). We note that even though opaque IR lines are present in our line list, we omit them from these two panels, because there are comparatively few IR photons (see Fig. 1) compared to UV and X-ray. The line opacity is weighted by the continuum flux, hence IR lines contribute very little to the total force multiplier. We finish this section with a couple of observations based on a comparison of the results from the force multiplier calculations and the heating and cooling calculations that we carried out using the same code and input parameters. For example, the significant increases of the equilibrium temperature with ξ (the solid black lines in the top panels of Fig. 3) occurs for the photoionization parameter range where M max (solid blue lines) strongly

8 Dannen et al. AGN2 log(ξ) 0 M total (t, ξ) = 522 log(ξ) 0 M total (t, ξ) = 1635 Ion Wavelength (Å) M L(t, ξ) Ion Wavelength (Å) M L(t, ξ) H I 1216 5.051 H I 1216 5.992 H I 26.768 H I 26 5.769 C IV 1550 1.550 C IV 1550 1.80 N V 1239 1.355 C III 977 1.562 O VI 37 1.22 N V 1239 1.520 log(ξ) 1 M total (t, ξ) = 123 log(ξ) 1 M total (t, ξ) = 117 H I 1216.771 H I 1216 7.522 Ne VI 01 1.02 H I 26.023 O VI 150 0.95 C IV 158 2.227 O VII 21 0.72 N V 1239 1.907 C V 0 0.6812 O VI 38 1.89 log(ξ) 2 M total (t, ξ) = 2 log(ξ) 2 M total (t, ξ) = 9.2 O VIII 19 0.60 H I 6563 0.09 Si X 303 0.399 O VI 32 0.171 Fe XI 180 0.30 Fe XI 180 0.096 Fe XII 202 0.330 Fe X 17 0.089 Fe XIV 28 0.290 Fe XII 187 0.089 log(ξ) 3 M total (t, ξ) = 20 log(ξ) 2.6 M total (t, ξ) = 22 Fe XXIII 777 7.328 Fe XXIII 1157 6.200 Fe XXII 1157 3.56 Fe XXIII 135.725 Fe XXII 13 2.60 Fe XXIII 3739 3.909 Fe XXII 3738 2.172 Fe XXIII 9587 2.193 Fe XXII 9586 1.561 O VIII 1303 0.16 Table 1. Table showing the top five contributing lines to the force multiplier for (left column) and AGN2 (right column) for fixed t = 6 and selected values of ξ. Note that the value of ξ in the bottom section of the table differs in the each column; it is intended to correspond to the bump in the M max distribution at large ξ (see the black lines in fig. 5). decreases with increasing ξ. This is expected because spectral lines are major contributors to the gas cooling at small and intermediate ξ. Therefore, the decrease in the overall number of line transitions with increasing ξ manifests itself in a decreases in M as well as the increase in the gas equilibrium temperature (i.e., the increase in the slope of the log T vs log ξ relation). In these two panels, we also shaded regions to mark the ξ values for which the gas is thermally unstable (where the slope is greater than one). These regions closely coincide with those corresponding to the resurgence in the line force. As we will discuss in the next section, this overlap may strongly impact the significance of the line driving for high photoionizaton parameters where M is nominally still larger than one.. CONCLUSIONS & DISCUSSION In the present paper, we have investigated the line force that results from exposing gas to the AGN radiation field. In our calculations, we have used the most complete and up-to-date line list. We confirm the wellknown result that the line force is a strong function of the photoniozation parameter for high values of the parameter and explore how this sensitivity depends on the AGN SED. We computed the force for several SEDs and here presented results for the SEDs that are representative of Type I and Type II AGNs. Our most significant results are the following: For a fixed value of the optical depth parameter, t, the force multiplier is not a monotonic function of the photoionization parameter. While M first decreases with ξ for ξ > 1 as shown by others (e.g., SK90), this decrease is not as strong for AGN Type I/II SEDs, and in fact we found that M can increase again at ξ 30 and ξ 900 for and ξ 50 and ξ 300 for AGN2. The main consequence of this behavior is that the multiplier can stay larger than 1 for ξ approaching. At the very same range of ξ for which the multiplier is higher than in previous calculations, i.e.

Photoionization calculations of the radiation force due to spectral lines in AGNs 9 ξ, gas is thermally unstable to constant pressure perturbations. The basic requirement for line driving to win over gravity is M max L > 1. The first result above implies that line-driving can be dynamically important over a relatively wide range of ξ, wider than explored in previous line-driven disk wind simulations performed, for example, by Proga et al. (2000), Proga & Kallman (200) and Proga (2007). This might indicate that the previous simulations underestimated the strength of the line force. We say this with caution, since the above requirement is just a necessary and not a sufficient condition for producing an appreciable line-driven wind. The second result above implies that this potential increase of the line force might not be physically realized because it occurs for ξ where the gas is locally thermally unstable, hence only a relatively small amount of gas (e.g., as measured by the column density) may be a subject to this strong force. However, the effects of thermal instability in dynamical flows (e.g., Balbus & Soker 1989) may alter this conclusion. The bump at ξ 0 00 may have interesting implications for the dynamics of a radiatively driven wind. That is, we expect different phases of acceleration with distinct lines to be tracked for each stage with Eddington fraction as low as 0.01. Note, however, that if the overall decrease in the number of lines, including coolants, leads to significant heating (i.e., runaway heating and thermal instability), the flow might be thermally driven rather than line-driven. For these reasons it is premature to state if our new results for the line force have significant consequences in explaining UV or X-ray absorbing outflows such as BALs, warm absorbers or ultra-fast outflows (e.g., Weymann et al. 1991; Reynolds & Fabian 1995; Chartas et al. 2002; Crenshaw et al. 2003; Tombesi et al. 20; Giustini et al. 2011; Hamann et al. 2013; Kaastra et al. 201; Nardini et al. 2015; McGraw et al. 2017; Arav et al. 2018). Therefore, in our follow-up paper, we will present results from our radiation-hydrodynamical s imulations of outflows experiencing radiative heating and line-force from the same radiation flux. We thank Sergei Dyda for useful discussions. Support for Program number HST-AR-1579.001-A was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. This work also was supported by NASA under ATP grant NNX1AK. REFERENCES Abbott, D. C. 1982, ApJ, 259, 282 Arav, N., & Li, Z.-Y. 199, ApJ, 27, 700 Arav, N., Korista, K. T., & de Kool, M. 2002, ApJ, 566, 699 Arav, N., Liu, G., Xu, X., et al. 2018, ApJ, 857, 60 Balbus, S. A., & Soker, N. 1989, ApJ, 31, 611 Barai, P., Proga, D., & Nagamine, K. 2012, MNRAS, 2, 728 Bautista, M. A., & Kallman, T. R. 2001, ApJS, 13, 139 Castor, J. L. 197, MNRAS, 169, 279 Castor, J. I., Abbott, D. C., & Klein, R. I. 1975, ApJ, 195, 157 (CAK) Chartas, G., Brandt, W. N., Gallagher, S. C., & Garmire, G. P. 2002, ApJ, 579, 169 Chartas, G., Saez, C., Brandt, W. N., Giustini, M., & Garmire, G. P. 2009, ApJ, 706, 6 Chelouche, D., & Netzer, H. 2003, MNRAS, 3, 233 Crenshaw, D. M., Kraemer, S. B., & George, I. M. 2003, ARA&A, 1, 117 Dyda, S., Dannen, R., Waters, T., & Proga, D. 2017, MNRAS, 67, 161 (D17) Gayley, K. G. 1995, ApJ, 5, Everett, J. E. 2005, ApJ, 631, 689 Field, G. B. 1965, ApJ, 12, 531 Gardiner, T. A., & Stone, J. M. 2005, Journal of Computational Physics, 205, 509 Gardiner, T. A., & Stone, J. M. 2008, Journal of Computational Physics, 227, 123 Giustini, M., Cappi, M., Chartas, G., et al. 2011, A&A, 536, A9 Hamann, F., Chartas, G., McGraw, S., et al. 2013, MNRAS, 35, 133 Kallman, T. R., & McCray, R. 1982, ApJS, 50, 263 Kallman, T., & Bautista, M. 2001, ApJS, 133, 221 Kaastra, J. S., Kriss, G. A., Cappi, M., et al. 201, Science, 35, 6 Krolik, J. H. 1999, Active galactic nuclei : from the central black hole to the galactic environment /Julian H. Krolik. Princeton, N. J. : Princeton University Press, c1999.,

Dannen et al. Lamers, H. J. G. L. M., & Cassinelli, J. P. 1999, Introduction to Stellar Winds, by Henny J. G. L. M. Lamers and Joseph P. Cassinelli, pp. 52. ISBN 0521593980. Cambridge, UK: Cambridge University Press, June 1999., 52 McGraw, S. M., Brandt, W. N., Grier, C. J., et al. 2017, MNRAS, 69, 3163 Mehdipour, M., Kaastra, J. S., Kriss, G. A., et al. 2015, A&A, 575, A22 Mehdipour, M., Kaastra, J. S., & Kallman, T. 2016, A&A, 596, A65 Murray, N., Chiang, J., Grossman, S. A., & Voit, G. M. 1995, ApJ, 51, 98 Nardini, E., Reeves, J. N., Gofford, J., et al. 2015, Science, 37, 860 Noebauer, U. M., & Sim, S. A. 2015, MNRAS, 53, 3120 Pereyra, N. A., Kallman, T. R., & Blondin, J. M. 1997, ApJ, 77, 368 Proga, D. 2007, ApJ, 661, 693 Proga, D., Ostriker, J. P., & Kurosawa, R. 2008, ApJ, 676, 1-112 Proga, D., Stone, J. M., & Drew, J. E. 1998, MNRAS, 295, 595 Proga, D., Jiang, Y.-F., Davis, S. W., Stone, J. M., & Smith, D. 201, ApJ, 780, 51 Proga, D., & Kallman, T. R. 200, ApJ, 616, 688 Proga, D., Stone, J. M., & Kallman, T. R. 2000, ApJ, 53, 686 Proga, D., & Waters, T. 2015, ApJ, 80, 137 Puls, J., Springmann, U., & Lennon, M. 2000, A&AS, 11, 23 Reynolds, C. S., & Fabian, A. C. 1995, MNRAS, 273, 1167 Saez, C., & Chartas, G. 2011, ApJ, 737, 91 Shu, F. H. 1992, Physics of Astrophysics, Vol. II, by Frank H. Shu. Published by University Science Books, ISBN 0-935702-65-2, 76pp, 1992., II, Smith, R. K., Ackermann, M., Allured, R., et al. 2016, Proc. SPIE, to appear Stevens, I. R. 1991, ApJ, 379, 3 Stevens, I. R., & Kallman, T. R. 1990, ApJ, 365, 321 (SK90) Tombesi, F., Cappi, M., Reeves, J. N., et al. 20, A&A, 521, A57 Waters, T., & Proga, D. 2016, MNRAS, 60, L79 Waters, T., & Proga, D. 2018, arxiv:18.11500 Weymann, R. J., Morris, S. L., Foltz, C. B., & Hewett, P. C. 1991, ApJ, 373, 23