Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Draft version April 26, 2024

Typeset using LATEX twocolumn style in AASTeX63

Position dependent radiation fields near accretion disks


Kara Smith ,1, 2 Daniel Proga ,1, 2 Randall Dannen ,1, 2 Sergei Dyda ,3 and Tim Waters 4

1 Department of Physics & Astronomy


University of Nevada, Las Vegas
4505 S. Maryland Pkwy
arXiv:2404.16175v1 [astro-ph.GA] 24 Apr 2024

Las Vegas, NV, 89154-4002, USA


2 Nevada Center for Astrophysics

University of Nevada, Las Vegas


4505 S. Maryland Pkwy.
Las Vegas, NV 89154, USA
3 Department of Astronomy

University of Virginia
530 McCormick Rd
Charlottesville, VA 22904, USA
4 Theoretical Division, Los Alamos National Laboratory,

NM 87545, USA

Submitted to ApJ

ABSTRACT
In disk wind models for active galactic nuclei (AGN) outflows, high-energy radiation poses a signif-
icant problem wherein the gas can become overionized, effectively disabling what is often inferred to
be the largest force acting on the gas: the radiation force due to spectral line opacity. Calculations of
this radiation force depend on the magnitude of ionizing radiation, which can strongly depend on the
position above a disk where the radiation is anisotropic. As our first step to quantify the position and
direction dependence of the radiation field, we assumed free streaming of photons and computed energy
distributions of the mean intensity and components of flux as well as energy-integrated quantities such
as mean photon energy. We find a significant dependence of radiation field properties on position, but
this dependence is not necessarily the same for different field quantities. A key example is that the
mean intensity is much softer than the radial flux at many points near the disk. Because the mean
intensity largely controls ionization, this softening decreases the severity of the overionization problem.
The position dependence of mean intensity implies the position dependence of gas opacity, which we
illustrate by computing the radiation force a fluid element feels in an accelerating wind. We find that
in a vertical accelerating flow, the force due to radiation is not parallel to the radiation flux. This
misalignment is due to the force’s geometric weighting by both the velocity field’s directionality and
the position dependence of the mean intensity.

Keywords: galaxies: active - methods: numerical - hydrodynamics - radiation: dynamics

1. INTRODUCTION tures can be well-explained as consequences of disk ac-


The key properties of active galactic nuclei (AGN), cretion onto a supermassive black hole (SMBH). Spec-
such as compactness, high luminosity, time variability, tral features such as broad emission lines (BELs) have
spectral energy distribution (SED), and spectral fea- for decades been used as a basis for classifying AGN,
while broad absorption lines (BALs) in the ultraviolet
(UV) are a spectacular manifestation for gas not only ac-
Corresponding author: Kara Smith creting onto an SMBH but also flowing outward. Even
smith99@unlv.nevada.edu though it has been shown the gas responsible for line
2 Smith, Proga, Dannen, Dyda, & Waters

emission/absorption is irradiated by the SMBH accre- in terms of computational resources. In this paper, we
tion disk (e.g., see Krolik 1999, for a textbook review), use our code to explore variables that contribute to line
the structure and dynamics of the region where these driving.
broad lines are formed (the broad line region, BLR) re- The paper is organized as follows. In §2, we give a
mains elusive. basic overview of the theory behind radiation fields and
Spectral features indicate that these accretion disks line-driven winds from accretion disks. In §3, we discuss
commonly produce outflows of matter in the form of the parametrization of the disk and the computation
jets and winds. The latter is one of the fundamental of radiation fields. In §4, we discuss the results of the
mechanisms by which the central SMBH interacts with study; including comparisons between SEDs at different
its host galaxy (e.g., Giustini & Proga 2019; Laha et al. points relative to the disk (§4.1), the sensitivity of mean
2020, and references therein). Therefore, the gas out- photon energy, ⟨hν⟩ and mean ionizing intensity, Jion ,
flows are important not only for the inner workings of to changes in the coronal radiation (§4.3), and finally,
AGN but also for AGN feedback (Fabian 2012). The the differences in the direction between the radiation
mechanisms launching and driving these accretion disk flux and the radiation force for a vertical velocity field
winds include thermal driving, magnetic driving, and (§4.2). The paper closes with a summary and discussion
radiation pressure due to photon scattering by spectral of the results in §5.
lines (‘line driving’; e.g., see for a review Proga 2007,
and references therein). While each of these three mech- 2. PRELIMINARIES
anisms requires its own methods of study and computa- 2.1. From Stars to Disks: Necessary Changes in
tion, a commonality between them is that each depends Perspective.
on the radiation fields of the accretion disk as the source
Models of outflows in AGN are informed by the rel-
of ionizing radiation.
atively well-understood winds from O stars (for a text-
One major problem for any AGN wind model is ove-
book review see Lamers & Cassinelli 1999, and refer-
rionization of the gas, which can both suppress or in-
ences therein). However, one must remember that stars
hibit outflows and fail to reproduce observed line pro-
are simpler than disks when discussing the radiation
files, namely lines produced from C iv, Si v, and N v.
field. For example, stellar photospheres are relatively
Even a clumpy magnetically driven outflow not shielded
uniform in temperature; therefore, under free-streaming
from the powerful AGN radiation will be overionized
conditions, the specific intensity, Iν , above the star is di-
and not reproduce observed BALs (de Kool & Begel-
rection independent inside the solid angle subtended by
man 1995). In a line driving scenario, the problem is
the star, Ω∗ . Further, when free-streaming is applica-
much more severe; an overionized gas cannot be accel-
ble, Iν equals the intensity of the stellar photosphere,
erated by line driving at all (e.g., Stevens & Kallman
Iν,0 . Therefore, calculations that require integration
1990; Proga et al. 2000; Dannen et al. 2019).
over Ω∗ can be simplified, e.g., when calculating the
Estimates of the ionizing radiation based on con-
mean intensity, Jν = (4π)−1 Ω∗ Iν dΩ, one can remove
H
straints of the SED and luminosity obtained by a distant
the term for specific intensity from the geometric inte-
observer show that for the so-called unobscured AGN,
gral, yielding the well-known relation Jν = Iν,0 W(r),
without shielding, the wind would be overionized (Mur- p
where W(r) = 21 (1 − 1 − (r∗ /r)2 ) is the geometric di-
ray et al. 1995). However, a fluid element in the wind
lution factor and r∗ is the stellar radius. For r ≈ r∗ ,
is not a distant observer; it sees the radiation source as
W(r) ≈ 1/2 and for large distances, W(r) ≈ (r∗ /2r)2
a nearby observer. Regardless of which driving mecha-
(see Lamers & Cassinelli 1999).
nism is favored or dominant, understanding the radia-
Disks are flattened structures that are not uniform
tion fields seen by the fluid element in the wind is fun-
in temperature. Not only the distance, r but also the
damental to determining if the wind is as overionized as
viewing angle, θ, measured from the normal of the disk
the distant observer implies.
are necessary variables to consider in describing any ob-
To understand the radiation fields as seen by the out-
served properties of the disk. A nearby observer is taken
flow, we compute mean intensity, flux, mean photon en-
to be a fluid element with a radius less than the disk’s
ergy, and the line force as a function of position. We find
outer radius, rout . For fluid elements very near to the
that a simple multi-temperature blackbody treatment,
disk, the mean intensity does approach Jν → 21 Iν,0 , as it
which a distant observer sees, is woefully inadequate for
does for stars. However, the specific intensity above the
describing the radiation very near to the disk. We have
disk is not direction-independent within the solid angle
developed a code capable of accurately computing radi-
subtended by the disk, ΩD . Instead, the disk photo-
ation fields near the disk and can do so inexpensively
spheric temperature and specific intensity are typically
Disk Radiation Fields 3

a decreasing function of the radius along the disk, rD .


In a standard accretion disk theory (Shakura & Sunyaev σe
Frrad = M (t) Fr , (2)
1973), the disk photospheric temperature, T (rD ), fol- c
−3/4
lows a rD scaling for large rD (a similar scaling can where M (t) is the force multiplier, which is a function of
also be derived for an irradiated disk) Assuming that optical depth parameter t, σe is the electron scattering
the disk radiates as a blackbody at every point, Iν,0 (rD ) coefficient, and c is the speed of light. CAK defined t
−3
follows the famous rD scaling for large rD , so near the using the Sobolev approximation (Rybicki & Hummer
disk, Jν scales as 1/r3 rather than 1/r2 . 1978), t ≡ σe vth ρ|dvr /dr|−1 , where vth is the thermal
If the disk were of constant specific intensity, then the velocity of the gas, and ρ is the density. Notice that
radial radiation flux for a large r and given θ would this definition of t assumes a radial flow; additionally,
be Fν,r = Iν,0 AD cos θ/r2 , where AD is the area of the CAK’s theory uses radial force and flux as analogs for
disk. The factor of cos θ accounts for the geometrical their total values because the theory is developed for
foreshortening effect. Because disks have non-uniform optically thin gas around a spherically symmetric star.
specific intensity profiles, we must integrate them over CAK’s theory has been successfully extended to a disk
the entire disk surface. A simple way to do this is to (Proga et al. 1998, 1999, hereafter PSD98 and PSD99).
assume that the angular distance between points on the However, the gas velocity is no longer radial, so instead.
disk is negligible and only to integrate over radii (see,
e.g., Frank et al. 2002). This is only a good approxima- −1
dvℓ
tion for the radial flux as seen by a very distant observer, t ≡ ρσe vth , (3)
dℓ
though, so we denote it Fν,dist :
where ℓ is the distance from the emitting point on the
2π cos θ
Z rout disk along a given line of sight defined by the unit vector
Fν,dist = Iν,0 (rD ) · rD drD , (1) n̂ and vℓ is the velocity along the line of sight: vℓ =
r2 rin
v · n̂. To compute the radiation force, we must integrate
where rin is the radius of the disk’s inner edge. This the intensity over all such lines of sight, and because
estimate of flux is not adequate for computing proper- dvℓ /dℓ is direction dependent, M(t) cannot be removed
ties from the point of view of a nearby fluid element, as from this integral. Thus, we have an expression for the
non-radial flux components comprise a significant por- radiation force,
tion of the total flux near the disk. In fact, as we ap- I
proach arbitrarily close to the disk, the radiation flux is rad σe I
F = M(t) · n̂dΩ, (4)
dominated by local radiation. Each point on the disk ΩD c
radiates a single-temperature blackbody, not the multi-
where I is the frequency integrated specific intensity of
temperature blackbody produced by Equation (1). Even
the disk. In a sense then, Equation (2) is the “distant
if a fluid element receives substantial radiation from the
observer” analog of Equation (4), where I is non-zero
entire disk, assuming negligible angular differences be-
for only one direction: purely radial. CAK developed a
tween points on the disk can lead to inaccurate esti-
correction to account for the changes in velocity along
mates. Finally, the scaling in Equation (1) is 1/r2 , but
a non-radial line of sight. This correction, elaborated
the flux should be proportional to 1/r3 near the disk,
upon by Friend & Abbott (1986) and Pauldrach et al.
as noted above. Therefore, the distant observer approx-
(1986), uses the symmetry of stars to integrate over all
imation (Equation 1) cannot be applied to points close
lines of sight analytically. The line force remains purely
to the disk. However, it is still useful for benchmark-
radial, though, also due to symmetry. For a disk, we
ing numerical calculations, as demonstrated by previous
cannot use such symmetry; instead, we must use fast
studies (e.g., see (Proga et al. 1998)).
numerical methods to account for sufficiently many lines
of sight to the disk to sample ΩD . One might assume
2.2. Line Driven Disk Winds that, generally, the line force has the same direction as
One of the primary quantities derived from the ra- the radiation flux – as is done to derive Equation (2) –
diation field is the radiation force due to lines, F rad . even when that direction is not purely radial, and that
Winds can be launched when continuum photons inter- the biggest term of dvℓ /dℓ suffices to compute the under-
act with gas through scattering in spectral lines (Castor lying velocity shear tensor (Proga et al. 1998; Nomura
et al. 1975, hereafter CAK). CAK expresses the radial et al. 2020). In section §4.2, however, we show that the
component of the radiation force, Frrad , in terms of the force and flux are not aligned near a disk of non-uniform
radial radiation flux, Fr , as brightness.
4 Smith, Proga, Dannen, Dyda, & Waters

The CAK line-driven wind theory informed earlier et al. 2017; Proga et al. 2022). We find that, with-
studies of line-driven winds from disks around cata- out attenuation of the central source, when the central
clysmic variables (Proga et al. 1998, 1999, CVs;) and source has just 5% the luminosity of the disk, all po-
in AGN (Murray et al. 1995; Proga & Kallman 2004). sitions above the disk will suffer from runaway heating
In CVs, launching and driving winds are robust, as the effects, and, even at only 1%, a significant portion of
winds are optically thin to continuum photons and the points will overheat as well.
radiation is dominated by UV, similar to the O stars
modeled by CAK. In AGN, however, the acceleration of 3. CALCULATING RADIATION AND RADIATION
a wind is less robust. The wind can be accelerated to FORCE
high velocities only if both the disk continuum radia- To evaluate the properties of the radiation field, it is
tion is strong in the UV and the wind can be shielded helpful to start with the fundamental equation governing
from the ionizing radiation from the inner disk and the coupling between radiation and matter as a function
the X-ray corona (Dyda et al. 2023). Because AGN of position along a line of sight, ℓ, and time, t, which is
winds are much denser than those of O stars and CVs, the radiation transfer (RT) equation:
gas launched from large radii can be accelerated when
shielded by a dense flow launched from smaller radii that 1 ∂Iν
+ n̂ · ∇Iν = jν − χν Iν , (5)
fail to escape the disk due to overionization (Proga et al. c ∂t
2000, hereafter PSK).
where jν is the emissivity and χν the total opacity. Here,
In PSK, the radiation transfer of the ionizing radia-
we consider a wind with conditions similar to those
tion was treated in a simplified way, namely by account-
studied by CAK. Specifically, we focus on the time-
ing only for the attenuation of direct radiation. Such a
independent solutions and assume that the wind emis-
treatment results in a lower limit
R ∞ for the mean intensity sion is negligible compared to the photospheric emis-
of ionizing radiation Jion ≡ νH Jν dν, where νH cor-
sion so that we may set jν to zero. In addition, we
responds to 13.6 eV and Jν is the mean intensity as a
consider that the opacity due to scattering, χscν , and ab-
function of frequency. It does not account for scattered −1
sorption, χabs sc abs
ν , (i.e., χν = χν +χν with units cm ) for
radiation found by Sim et al. (2010) and Higginbottom
continuum processes. This corresponds to the situation
et al. (2014). The latter studies used Monte Carlo pho-
where continuum photons emitted at the photosphere
toionization calculations and showed that radiation scat-
can stream freely through the wind. Again, following
tering can allow ionizing photons to propagate ‘around’
CAK, we will account for RT in spectral lines by using
the shielding inner flow and thereby increase the ioniza-
the Sobolev approximation. This approximation takes
tion of the outer outflow. The techniques used by Sim
advantage of the fact that when a photon is emitted
et al. (2010) and Higginbottom et al. (2014) are compu-
at a certain frequency, ν, in the atmosphere, it inter-
tationally expensive, and as such, it is not yet feasible
acts with the outflowing gas through a specific line only
to use them in the context of radiation-hydrodynamical
when the frequency is very close to the line center fre-
simulations, wherein we would need to compute line dis-
quency, ν0 . This resonance point, which is determined
tributions for each grid point at each time step. It is
by the Doppler shift, occurs when the photon frequency
necessary, then, to find computationally inexpensive and
matches the line center frequency. Under these assump-
sufficiently accurate estimators of the ionization state of
tions, Iν can be treated as constant along any line of
the gas.
sight, and we need only calculate the intensity of the
The main three quantities that can be used to es-
photosphere. So, for both the corona and the disk, we
timate ionization are the above mentioned Jion , gas
let Iν = Iν,0 for continuum photons that do not interact
number density n, and mean photon energy, ⟨hν⟩ ≡
R∞ via lines, and Iν = Iν,0 exp(−τν ) for the photons that
0
hνJ ν /J dν, where h is Planck’s constant. In opti-
do interact, where τν is the optical depth due to the
cally thin plasmas, the ionization state is determined by
absorption line in the wind.
the ratio of the first two quantities, as shown by Tarter
Because Iν = Iν,0 for continuum photons, calculations
et al. (1969), referred to as the ionization parameter,
of moments of intensity can be simplified and, there-
ξ ≡ (4π)2 Jion /n. The mean photon energy is vital for
fore, sped up. To do this, we use the methods described
high values of ξ, where the main heating is Compton
in PSD99 and adopted by Dyda et al. (2017). We use
heating, which heats the gas to the Compton temper-
a spherical-polar coordinate system (r1 , θ1 , ϕ1 ) centered
ature, TC ≡ ⟨hν⟩ /(4kB ). If TC > 106 K, the gas will
on a point in the wind, W Let the point W ′ be the point
undergo runaway heating and could even become ther-
on the disk vertically below W . The coordinate system’s
mally unstable (e.g., Kallman & McCray 1982; Dyda
zenith is directed along W C – where C is the center of
Disk Radiation Fields 5

the disk – and ϕ1 = 0 the half-plane opposite W C from Table 1. Values of the accretion disk and corona parameters
W ′ . We also introduce a shorthand µ = cos θ1 . We can used in this paper.
write the expressions for n̂ and Ω in (r1 , θ1 , ϕ1 )-space as Parameter Value(s)
M 108 M⊙
Γ 0.7
p p
n̂ = (µ, − 1 − µ2 cos ϕ1 , 1 − µ2 sin ϕ1 ) (6)
rout 100 rin
and x 0.001-0.25
dΩ = sinθ1 dϕ1 dθ1 = − dϕ1 dµ. (7) β -1.1
Equations for the zeroth moment, the mean intensity,
Jν , and the first moment, or the flux, Fν , are quite sim-
ple in this coordinate system: We obtain an explicit expression for Iν,0 following
I PSK’s model of radiation from the disk and corona. We
1 separate the intensity into that from each surface: the
Jν (r, θ) = Iν,0 (r, θ, µ, ϕ1 ) dΩ (8)
4π Ω(r,θ) intensity of the disk, Iν,D , and that of the corona, Iν,∗ .
and For the disk, we assume a geometrically thin, optically
thick disk with a temperature profile given by Shakura
I & Sunyaev (1973). The disk extends from some inner
Fν (r, θ) = Iν,0 (r, θ, µ, ϕ1 )n̂(µ, ϕ1 )dΩ. (9) radius, rin , to some outer radius, rout . We take rin to
Ω(r,θ) be rISCO = 6GM/c2 , where M is the mass of the cen-
In Equations (8) and (9), we explicitly show Jν , Fν , Ω, tral black hole. Because the exact shape of the central
Iν,0 , and n̂ as functions of r, θ, µ and ϕ1 . For the sake of source is not well known, we assume a spherical extended
brevity, we omit these explicit dependencies in the rest corona with a radius also equal to rISCO . The total lu-
of this paper. minosity of the disk, LD , is given by
To compute integrals over dΩ numerically, we use the
GM Ṁ
Gauss-Legendre quadrature, in which we divide the an- LD = , (12)
gular dµdϕ1 -space into a discrete ∆µ∆ϕ1 -sum: 2rin

1 XX where Ṁ is the mass accretion rate. We parametrize the


Jν = Iν,0 (µi , ϕ1 j )ωi ωj (10) accretion rate in terms of the Eddington fraction, Γ ≡
4π i j
Ṁ /ṀEdd , where ṀEdd = 8πcrin /σe is the Eddington
and accretion rate.
We also assume the corona radiates with a power-law
XX spectrum – i.e. the luminosity Lν,∗ ∝ ν β for some con-
Fν = Iν,0 (µi , ϕ1 j )n̂(µi , ϕ1 j )ωi ωj , (11)
stant β – when 0.1 keV < hν < 511 keV (Sim et al.
i j
2010). We normalize the corona luminosity by setting,
where µi and ϕ1 j are the roots of Legendre polynomials as an input parameter, the ratio x ≡ L∗ /LD . We include
while ωi and ωj are the associated weights (for a review further details about the computation of the intensity of
of Gauss-Legendre quadrature as well as other methods the corona in §3.1. The accretion disk system is param-
of numerical integration see Press et al. 2007). In this eterized by M , Γ, rout , x, and β. We list the values used
paper, we used a 15x15 grid of quadrature points, which in our calculations in Table 1.
is a relatively low number of grid points, especially com-
pared to older methods such as the one used in PSD98. 3.1. Evaluation of Intensity
With such a grid, on a single 12th Gen Intel i7-12700H To develop an expression for the intensity of the disk,
Core (i.e., a modern laptop), computing SEDs for Jν we begin with the surface brightness of the disk, Q. The
and Fν sampled at 300 frequencies takes 0.1 seconds total surface brightness is the sum of two effects:
per location, (r, θ). Using Gauss-Laguerre Quadrature
(again, see Press et al. 2007) with 50 frequency quadra- Q = Qint + Qirr , (13)
ture points, we can accurately compute J and F in about
0.02 seconds per location. This computational speed where Qint is the intrinsic brightness of the disk due to
makes feasible the goal of incorporating these methods internal processes and Qirr is the brightness due to ir-
in a time-dependent radiation-hydrodynamical simula- radiation by an outside source (in our case, the black
tion. hole corona). The intrinsic brightness is given by the
6 Smith, Proga, Dannen, Dyda, & Waters

Shakura-Sunyaev model and the brightness due to irra- 3.2. Radiation Force due to Lines
diation is given by PSD98: Computing the radiation force due to a line requires
3
 r  the local line absorption coefficient, κL – which is related
rin rin
Qint = Q0 · 3 1 − (14) to the opacity in Equation 5 by χl,ν = ρκL ϕ(ν), where
rD rD ρ is the density of the gas and ϕ(ν) is the assumed line

  s  2
 absorption profile – and the local intensity, Iν .
xQ0  rin rin rin 
Qirr = arcsin − 1− ,
3π rD rD rD
∆νD κL
(15) dFLrad = Iν n̂dΩ, (21)
c
3
where Q0 = 3GM Ṁ /8πrin . Assuming that the disk ra-
diates as a blackbody at every point, we can use the
where ∆νD ≡ ν0 vth /c is the Doppler width of the line.
Stefan-Boltzmann law to write temperature as a func-
To integrate Equation (21) over dΩ, we cannot simply
tion of brightness and, thereby, as a function of radius
use Iν = Iν,0 , as this is not a continuum process; we
along the disk.
must include an escape probability to account for the
 1/4 interaction of photons with spectral lines along a line
Qint (rD ) + Qirr (rD )
T (rD ) = , (16) of sight (Owocki & Puls 1996). For a derivation of this
σ escape probability from ϕ(ν), see Lamers & Cassinelli
where σ is the Stefan-Boltzmann constant. (1999). With the escape probability included, we have
As per our assumption that the disk radiates as a an integral for the radiation force due to lines in terms
blackbody, we take the intensity from a surface element of the intensity of the disk photosphere.
at radius rD to be the blackbody intensity of that sur-
face element.
σe ∆νD 1 − e−τL
I
FLrad = Iν,0 n̂dΩ, (22)
Iν,D (rD ) ≡ Bν (T (rD )), (17) Ω c τL

where rD is given by
where τν is the local optical depth, which we use the
Sobolev approximation in three dimensions (see Rybicki
2
rD = r12 + r2 − 2rr1 µ (18) & Hummer 1978) to compute.
r cos θ
r1 = p ,
µ cos θ − 1 − µ2 cos ϕ1 sin θ Z ∞ −1
Sob dvℓ
where r1 is the distance from W to the disk along the τL = ρκL ϕ(ν) dr = ρκL vth , (23)
r dℓ
line of sight. Connecting back to Equations (8) and (9),
Iν,D is a function of r, θ, µ and ϕ1 through the fact that
rD is a function of those four variables. where ϕ(ν) is the aforementioned assumed line absorp-
We then model the corona’s specific luminosity ac- tion profile.
cording to a power-law spectrum with cutoffs at hνX = To compute the total radiation force due to lines, we
0.1 keV and hνγ = 511 keV . follow CAK and define a variable η ≡ κL /σe so that
τL = ηt. Substituting in t and η into Equation (21), we

xL ν β get an equation of radiation force due to a single line:
0 νX < ν < νγ
Lν,∗ = (19)
0 otherwise
σe ∆νD 1 − e−ηt
I
where L0 is a normalization constant such that x = FLrad = · · Iν,0 n̂dΩ. (24)
L∗ /LD . The intensity of the corona is then given in Ω c t
terms of the luminosity,

Lν,∗ Then, by summing FLrad over all lines, we obtain the full
Iν,∗ = , (20) radiation force due to all lines:
4π 2 r∗2
where r∗ is the radius of the corona. For our purposes X I σe ∆νD 1 − e−ηt
rad
r∗ = rin . F = · Iν,0 n̂dΩ. (25)
Ω c t
lines
Disk Radiation Fields 7

3.3. The Distribution of Lines


We can use photoionization codes such as XSTAR
(Kallman & Bautista 2001) to compute the abundances
of ions and opacities of lines for any given SED, which
allows us to compute the sum in Equation (25). But, as
mentioned before, doing so in the context of radiation-
hydrodynamical simulations is computationally expen-
sive. So, instead, we again follow CAK and describe
a statistical distribution of lines, where the number of
lines of a given opacity and frequency is a function of
that opacity and frequency. To recover the power law
given in CAK, M (t) = kt−α for some constant k, we
have
1
dN = N0 η α−2 dηdν, (26)
ν
where α is the CAK parameter. Owocki et al. (1988,
hereafter OCR), add an exponential drop-off to limit
the effect of very strong driving lines. With this modi-
fication, we have
1
dN = N0 η α−2 e−η/ηmax dηdν, (27)
ν
where ηmax is the maximum opacity for which lines are
physically relevant.
We are interested in finding the distribution of lines
relative to energy. CAK introduced a frequency distri-
bution 1/ν, which allows for the frequency dependency
of the force multiplier to cancel out entirely. However,
based on photoionization calculations using the methods
outlined by Dannen et al. (2019), we see that the line
distribution is not a simple power-law. For example, in
Figure 1, we show that the distribution of lines actually
traces the energy distribution of the mean radiation in-
tensity. This informs our second modification to the line
distribution, which is to use Jν /J to weight our lines by
frequency rather than the classic 1/ν. Combining the
modifications of OCR with our own, we arrive at


dN = N0 η α−2 e−η/ηmax dηdν, (28)
J
We can replace the sum in Equation (25) with an in-
tegral over dN and analytically integrate over dη to find
Figure 1. Comparisons of the energy distribution of black- our, now ν dependent, force multiplier.
body radiation (black-dashed curves) and single line force
(τmax + 1)1−α − 1
 
multipliers (colored vertical bars) for four blackbody temper- νJν −α
atures: Tr = 105 K, 2 × 105 K, 4 × 105 K, 6 × 105 K. The ver- M(t, ν) = · kt · 1−α , (29)
J τmax
tical bars represent the 1,000 strongest individual line force
multipliers – normalized to the maximum force multiplier, where τmax = ηmax t. Accounting for the SED depen-
Ml,max – for fixed values of log ξ = 2 and t = 10−10 (Eq. 3) dence of the line distribution, not only must the force
as a function of energy. These 1,000 strongest lines have multiplier be integrated over the solid angle of the disk
been binned into quartiles, red being the upper quartile, yel- but also over all frequencies. We therefore have
low the second, green the third, and blue the weakest set of Z ∞I
lines. The black dashed line in each panel represents the as- rad σe Iν,0
F = M(t, ν) · n̂dΩdν (30)
sumed blackbody SED (temperature indicated in each panel) 0 Ω c
that we used for both determining the ionization balance and
computing M . The minimum force multiplier, Ml,min is also
listed in each panel.
8 Smith, Proga, Dannen, Dyda, & Waters

as the full expression for the radiation force due to lines.

4. RESULTS
4.1. Local vs Distant Radiation Fields
We compute energy distributions Jν (r, θ), Fν,r (r, θ),
and Fν,θ (r, θ) for r ranging from 2rin to 100rin and θ
ranging from 0◦ to 90◦ . In Figure 2 we show example
SEDs for three locations, (r, θ), above the disk: (100,
40), (40, 80), (12, 85). We selected these three loca-
tions as illustrative representations of the three distinct
classes of spectra we find above the disk.
Specifically, the top panel of Figure 2 shows the
spectra at large radius and intermediate theta, (r =
100 rin , θ = 40◦ ). In this regime, both the spectra of
4πJν (blue solid line) and Fν,r (orange solid line) overlap
with each other. Moreover, they agree well with the ap-
proximation for the flux seen by a distant observer given
in Equation (1), Fν,dist (r, θ) (red dashed line). Agree-
ment between these three SEDs is consistent with the
expectations discussed in §2. First, the θ component of
the flux (green solid line) will disagree with 4πJν , Fν,r ,
and Fν,dist because it is weighted by nθ . And further,
Fν,θ will be much weaker than Fν,r , which we also see,
Fν,θ peaks an order of magnitude lower than Fν,r . For
reference, we plot a curve for a blackbody spectrum with
the temperature of the point projected down onto the
disk. We see that Fν,θ peaks at similar energies to the
peak of the blackbody spectrum, but the spectrum is
broader, again due to the entire disk contributing to the
flux.
In the case that we are very near the disk, shown in the
middle panel of Figure 2, (r = 12 rin , θ = 85◦ ), we see
that Fν,θ not only peaks at the same energy as the black-
body spectrum, but is very similar to it at all energies,
save for a small, high-energy tail. The Jν spectrum is
the most similar, in fact, aside from the high energy tail,
4πJν = 2Fν,θ . It is to be expected that contribution to
Jν is dominated by Fν,θ as radiation from any location
not immediately below the point (r, θ) will be diminished
by geometric foreshortening. The other two spectra are
quite different from the Fν,θ , 4πJν , and local blackbody
spectra. While Fν,r spectrum is relatively narrow, it
peaks at higher energies than Jν and Fν,θ because Fν,r Figure 2. Comparison of the energy distributions of the disk
is weighted by nr , and so is dominated by the high en- radiation field properties at three locations: (100 rin , 40◦ ),
ergy radiation from the inner disk. On the other hand, (12 rin , 85◦ ), (40 rin , 80◦ ), top to bottom. Each panel shows
the Fν,dist spectrum is still a broad multi-temperature 4πJν (blue), Fν,r (orange), Fν,θ (green), Fν,dist (red dashed),
and the blackbody spectrum of the point projected verti-
blackbody, as it’s intensity varies uniformly, indepen-
cally onto the disk (black dashed). The area below Fν,dist is
dent of frequency. shaded to make it more apparent at what energies a spec-
From these two cases, it may seem reasonable to as- tral property is greater or less than that seen by a distant
sume that as one moves further from the disk, Jν broad- observer.
ens, Fν,r broadens and increases in power, and Fν,θ
broadens slightly but decreases in power. However, the
Disk Radiation Fields 9

Figure 3. Illustrating the misalignment of the line force and the radiation flux. In a vertical velocity field above a disk, the
line force (black) is not parallel to the radiation flux (red), as it would be in a spherically symmetric wind case.

bottom panel of Figure 2, (40 rin , θ = 80◦ ), shows k, α, vth , and ρ, we are then able to compute the radia-
that a simple interpolating function between the nar- tion force using Equation (30). The radiation force is a
row spectum near the disk and broad distant observer function of three quantities: the flux from the disk, the
spectrum does not reproduce the spectra at intermediate velocity gradient of the accelerating gas, and the lines
positions. At points with intermediate radii and fairly available for line driving. The weighting effects of the
high viewing angles, Fν,r experiences contributions from velocity field cause the force to be biased toward the
a larger portion of the inner disk, broadening its spec- direction of greatest acceleration, as it has the greatest
trum. The part of the disk closer to the point in the value of kt−α . The weighting effects of line availabil-
wind contributes more to the spectrum and is at lower ity cause the force to be more biased by local radiation
energies. This leads to a low energy bump in the Fν,r than the flux alone would imply. For example, when the
spectrum. The Jν spectrum has a more prominent bump point in the wind, W , is near enough to the disk, local
that covers both the Fν,r and Fν,θ peaks. radiation is the dominant contributor to Jν (see Fig. 2).
The position dependences of Jν are important for our Because the available lines are determined by Jν , these
discussion of the radiation force due to lines, as they lines will be primarily in frequencies produced by the
determine what lines are available for line driving, as disk directly below W , which creates a vertical bias in
discussed in §3.3. This means that in some sense F rad the force. The biases due to both the velocity field and
is doubly position dependent. It depends on the position the local SED cause a misalignment between the flux
like SEDs do directly through its integration over Ω. But and the force based on the strength of the directional
also, it is position dependent on what lines are available force multiplier.
through Jν . In the bottom panel of Figure 2, we see Figure 3 shows an example of this misalignment. The
that the Jν spectrum peaks at a very different position red arrows show the radiation flux purely from the disk
than the Fν,r spectrum, but relatively close to the peak (x = 0). The black arrows show the radiation force
of the Fν,θ spectrum. This biases the force toward the θ due to this flux in a vertical velocity field, which is the
direction, as there will be more lines that the radiation velocity field we expect to see very near the disk. This
comprising Fν,θ can interact with. Such a bias creates a velocity field is defined by a power-law of the vertical
misalignment between the radiation force and the flux. distance. In cylindrical coordinates (rD , z, ϕD )
The velocity of the gas further biases which component  δ
of the flux predominantly contributes to the line force. z
vz = v0 , (31)
In the following subsection, we discuss these two effects rD
and determine their relative contribution to causing the p
misalignment. where v0 = GM/rD , and in spherical coordinates
r
GM 1
4.2. The Misalignment of F rad and F vz = r−1/2 . (32)
sin θ tanδ θ
For a given velocity field, we can compute dvℓ /dℓ for
each line of sight to the disk, and with gas parameters In a Keplerian disk, vϕ = v0 giving us an expression for
10 Smith, Proga, Dannen, Dyda, & Waters

Figure 4. Position dependence of log10 (Jion ) and log10 (⟨hν⟩) for different corona intensities. The top two rows of panels show
contours of Jion in units of erg s−1 cm−2 , both on its own (top) and scaled by W(r) (middle). The bottom panel shows ⟨hν⟩
in units of eV. Each column is computed for a coronal luminosity fraction x. The colors are for visual contrast only. We
demonstrate the position dependence of these quantities as well as the change in that position dependence due to the luminosity
of the corona.
v in spherical coordinates: causing this misalignment; but the effect of accounting
for the frequency-weighting of lines is not insignificant.
v = vz cos θ, − sin θ, tanδ θ ,

(33)

which we can use to compute M(ν, t).


Figure 3 shows the misalignment for δ = 1/2. As one 4.3. Position Dependencies of Frequency Integrated
moves vertically upwards from the disk surface, the ra- Radiation Fields
diation flux is significantly influenced by its radial com- Along with the radiation force due to lines, other
ponent. But, due to the vertical velocity field, the radia- frequency-integrated properties help determine the state
tion force remains close to vertical for much longer than of the gas. The properties of the disk that we focus on
the flux. This allows the gas to potentially be launched in this study are ⟨hν⟩ and Jion . These two quantities
higher than if the force were parallel to the flux. largely control the heating and ionization of the gas sur-
As mentioned, this misalignment comes from a mix- rounding the disk, as discussed in §2. As spectra change
ture of two effects: the velocity field and the position de- with a position near the disk, so do these frequency-
pendence of the line distribution. To determine which integrated quantities. In Figure 4 we show log10 (Jion )
is the dominant effect, we computed the directions of (top), log10 (Jion /W(r)) (middle), and log10 (⟨hν⟩) (bot-
the radiation force with the classical 1/ν weighting of tom).
frequencies (for which all frequency dependence of the In the case where there is no corona (x = 0), we see
force multiplier cancels). We found the difference be- a simple trend: as we move radially outward, both Jion
tween the misalignment obtained using Equation (29) and ⟨hν⟩ decrease. For large enough r and θ, Jion /W(r)
and that obtained using the 1/ν weighting. The points is radius independent, consistent with a W(r) scaling.
with the largest difference between the two methods dif- Moving toward the disk along constant r < 20 rin , Jion
fered by about 10%. We conclude that the directionality and ⟨hν⟩ both decrease as θ increases. At radii larger
caused by the velocity gradient is the dominant effect than 20 rin , both increase slightly before decreasing.
Disk Radiation Fields 11

Figure 5. Relations between the energy integrated properties of the disk radiation field with and without coronal radiation.
The top panels show ⟨hν⟩ vs J and the bottom show ⟨hν⟩ vs Jion . The left column shows each point colored according to r. The
right shows each point colored according to θ. The vertical lines are an analytic estimate for ⟨hν⟩ based on the assumption that
J∗ /JD ≈ x; dashed is x = 0.01, dotted is x = 0.05, and dash-dotted is x = 0.25. We do not show these vertical lines for x = 0
and x = 0.001 because the approximation only holds when x ≫ ⟨hν⟩D / ⟨hν⟩∗ ∼ 10−3 for our disk parameters. The maximum
value of ⟨hν⟩ in each panel is ⟨hν⟩∗ ≈ 42 keV.

When a corona is introduced, the shapes of the con- spectively. We rewrite mean photon energy as
tours of constant Jion are mainly unaffected. The values 1 xJ
of Jion are increased due to a term proportional to W(r) ⟨hν⟩ = ⟨hν⟩D + ⟨hν⟩∗ , (34)
xJ + 1 xJ + 1
added by the corona. This term never varies by more
where ⟨hν⟩D and ⟨hν⟩∗ are the mean photon energies
than a factor of 21 W(r) across all θ, contributing a rela-
of the disk and corona, respectively. This form of the
tively constant factor at any given radius.
equation is rather useful because ⟨hν⟩∗ is a constant; it
The effect of a corona on ⟨hν⟩ is much more com-
is approximately 42 keV for our disk parameters. As r
plicated than on Jion . Here we introduce a variable
increases, xJ approaches x and ⟨hν⟩D approaches a con-
xJ ≡ J∗ /JD , where J∗ and JD are the frequency-
stant ⟨hν⟩dist – the mean photon energy of the disk as
integrated mean intensities of the corona and disk, re-
seen by a distant observer – which for our disk parame-
ters is ∼25 eV.
12 Smith, Proga, Dannen, Dyda, & Waters

For constant θ ≲ 40◦ , as r increases, first ⟨hν⟩ de- ergy distributions of mean intensity and flux, the line
creases. This is the regime wherein the radiation from force, the mean ionizing intensity, and the mean photon
the outer disk becomes more relevant, softening the energy.
mean intensity spectrum. Then, at large r, ⟨hν⟩ be- Typically, the energy distributions of mean inten-
gins increasing with r. As r increases, xJ approaches x, sity and flux components differ, and their position de-
and so ⟨hν⟩ approaches it’s limit: x/(x + 1) · 4.2 × 104 + pendence is non-trivial. However, certain distributions
1/(x + 1) · 25 eV. For our four values of x: 0.001, 0.01, are relatively straightforward to estimate in two limit-
0.05, and 0.25, these limits are 67 eV, 440 eV, 2 keV, ing cases. Firstly, at the locations near the disk, the
and 8.4 keV, respectively. θ component of the flux is characterized by a single-
The behavior when θ ≳ 40◦ is the opposite of its be- temperature blackbody distribution. This flux origi-
havior in the x = 0 case. As one moves closer to the disk nates from a small, almost constant temperature disk
while maintaining constant r, ⟨hν⟩ increases. The cause area directly below. Secondly, at locations far from the
is that as one approaches the disk, the mean intensity disk, the mean intensity and radial flux distributions
spectrum experienced loses its high energy tail due to resemble multi-temperature blackbody spectra originat-
the foreshortening effect discussed in §4.1, so the total ing from the entire disk. Using our methods, we can
mean intensity of the disk decreases. The corona does recover the dependencies in these two regions very well
not suffer any foreshortening and, therefore, becomes (e.g., compare the solid green and dashed black curves
the dominant contributor to ⟨hν⟩. in the middle panel of Fig. 2 and the solid blue and or-
In Figure 5 we show the relation between J, Jion , and ange curves to the dashed red in the top panel of that
⟨hν⟩ more directly. We can see that the change in con- figure). In addition, we find that the geometric dilution
tour shape present in Figure 4 translates to a reversal factor well captures the radius dependence of the mean
in the correlation of J and ⟨hν⟩. Namely, when there intensity for r ≳ 15 rin and θ ≳ 40◦ .
is no corona, when one fixes a value of either r or θ We also illustrated how the direction of the radiation
and allows the other to vary, we find a positive corre- force can be affected by the position-dependent mean in-
lation between J and ⟨hν⟩. However, when a corona is tensity spectrum. In accordance with detailed photoion-
added, most fixed r and θ yield negative correlations. ization calculations (see Fig. 1), we made the assump-
Further, from Equation (34), if xJ ≫ ⟨hν⟩∗ / ⟨hν⟩D ∼ tion that the line opacity’s energy distribution scales like
10−3 we can ignore the disk contribution to ⟨hν⟩ and Jν and showed that the line force can be more perpen-
use ⟨hν⟩ = xJ /(xJ + 1) ⟨hν⟩∗ . In Figure 5, we show dicular to the disk than the radiation flux vector (see
x/(x + 1) ⟨hν⟩∗ , which, if one assumes xJ ≈ x as it Fig. 3).
does at large distances, should be a good estimate of The wind’s thermodynamics and observational prop-
⟨hν⟩. This benchmark is represented by three vertical erties, regardless of its launching mechanism, are de-
lines at x/(x + 1) ⟨hν⟩∗ for x = 0.01, 0.05, and 0.25 termined by its ionization and radiative heating and
(dashed, dotted, and dash-dotted lines, respectively). cooling rates. The temperature and thermal stability
In all three cases, this benchmark overestimates ⟨hν⟩ of highly ionized gas depend on the mean photon en-
for most points near the disk, indicating xJ ̸≈ x; rather, ergy, ⟨hν⟩. Therefore, we calculated ⟨hν⟩ for different
xJ < x, indicating that near the disk, the coronal con- cases and found that the mean intensity is much softer
tribution to ⟨hν⟩ is less than one would assume from the than the radial flux at many points near the disk. This
disk and coronal luminosities. softening reduces the severity of the overionization prob-
lem as we discussed in the introduction. We also found
5. CONCLUDING REMARKS that near the disk, solely geometrical factors can cause
the value of ⟨hν⟩ to vary by up to a factor of 3 when
We have investigated how the properties of the radia-
the central source’s X-ray luminosity is zero. However,
tion field near an accretion disk change with a position
even a small increase in the X-ray luminosity can in-
due to purely geometric effects in the case of a standard,
crease ⟨hν⟩ by several orders of magnitude. We note
flat Shakura-Sunyaev disk surrounded by optically thin
that this dependence on the central luminosity of the X-
gas. As explained in §2, quantifying this position depen-
rays can be reduced by attenuation caused by a bulging
dence is a necessary prerequisite for accurately modeling
disk or an inner disk wind. We will asses the impact of
important processes in radiation-hydrodynamical simu-
such intervening gas and the misalignment of the line
lations, such as radiative heating and cooling, and the
force and flux vectors. by carrying out time-dependent
radiation force due to lines. We described our numerical
radiation-hydrodynamical simulations of disk winds. To
methods for computing, in a position-dependent man-
do so, we will incorporate our radiation field methods in
ner, various radiation field characteristics, including en-
Disk Radiation Fields 13

a radiation-MHD code like Athena++ . Our treatment 0.8%, 1%. These errors become progressively worse as
in this work assumes a geometric optics formalism where one approaches risco , particularly near θ = 90◦ , where
radiation propagates along linear rays. We neglect spe- they can grow as large as 50%. However at such posi-
cial and general relativistic effects. For example, we do tions that these differences are large the assumption of
not account for Doppler boosting (both Doppler beam- linear radiation propagation is violated and so a wholly
ing and the Doppler shift) due to emission from a rela- more detailed relativistic treatment is necessary to ac-
tivistic disk or compute gravitational deflection of pho- curately describe radiation fields near the black hole.
tons. Using the treatment presented in Equation (4)
of Jiang (2022) we estimated the percent difference in Acknowledgments
J and Jν between including and not including Doppler
boosting for the three spectra shown in Figure 2. For Support for this work was provided by the National
the points (12r∗ , 85◦ ), (40r∗ , 80◦ ), and (100r∗ , 40◦ ), we Aeronautics and Space Administration under TCAN
find the errors in J to be, respectively, 3%, 0.7%, and grant 80NSSC21K0496. D.P. and T.W. acknowledge
0.6%. For Jν , we found the maximum errors to be 3%, support from NASA grant HST-GO- 16196.

REFERENCES
Castor, J. I., Abbott, D. C., & Klein, R. I. 1975, ApJ, 195, Murray, N., Chiang, J., Grossman, S. A., & Voit, G. M.
157, doi: 10.1086/153315 1995, ApJ, 451, 498, doi: 10.1086/176238
Dannen, R. C., Proga, D., Kallman, T. R., & Waters, T. Nomura, M., Ohsuga, K., & Done, C. 2020, MNRAS, 494,
2019, ApJ, 882, 99, doi: 10.3847/1538-4357/ab340b 3616, doi: 10.1093/mnras/staa948
de Kool, M., & Begelman, M. C. 1995, ApJ, 455, 448, Owocki, S. P., Castor, J. I., & Rybicki, G. B. 1988, ApJ,
doi: 10.1086/176594 335, 914, doi: 10.1086/166977
Dyda, S., Dannen, R., Waters, T., & Proga, D. 2017, Owocki, S. P., & Puls, J. 1996, ApJ, 462, 894,
MNRAS, 467, 4161, doi: 10.1093/mnras/stx406 doi: 10.1086/177203
Dyda, S., Davis, S. W., & Proga, D. 2023, arXiv e-prints, Pauldrach, A., Puls, J., & Kudritzki, R. P. 1986, A&A, 164,
arXiv:2310.18557, doi: 10.48550/arXiv.2310.18557 86
Press, W. H., Teukolsky, S. A., Vetterling, W. T., &
Fabian, A. C. 2012, ARA&A, 50, 455,
Flannery, B. P. 2007, Numerical Recipes 3rd Edition:
doi: 10.1146/annurev-astro-081811-125521
The Art of Scientific Computing, 3rd edn. (USA:
Frank, J., King, A., & Raine, D. J. 2002, Accretion Power
Cambridge University Press)
in Astrophysics: Third Edition
Proga, D. 2007, in Astronomical Society of the Pacific
Friend, D. B., & Abbott, D. C. 1986, ApJ, 311, 701,
Conference Series, Vol. 373, The Central Engine of
doi: 10.1086/164809
Active Galactic Nuclei, ed. L. C. Ho & J. W. Wang, 267
Giustini, M., & Proga, D. 2019, A&A, 630, A94,
Proga, D., & Kallman, T. R. 2004, ApJ, 616, 688,
doi: 10.1051/0004-6361/201833810
doi: 10.1086/425117
Higginbottom, N., Proga, D., Knigge, C., et al. 2014, ApJ, Proga, D., Stone, J. M., & Drew, J. E. 1998, MNRAS, 295,
789, 19, doi: 10.1088/0004-637X/789/1/19 595, doi: 10.1046/j.1365-8711.1998.01337.x
Jiang, Y.-F. 2022, ApJS, 263, 4, —. 1999, MNRAS, 310, 476,
doi: 10.3847/1538-4365/ac9231 doi: 10.1046/j.1365-8711.1999.02935.x
Kallman, T., & Bautista, M. 2001, ApJS, 133, 221, Proga, D., Stone, J. M., & Kallman, T. R. 2000, ApJ, 543,
doi: 10.1086/319184 686, doi: 10.1086/317154
Kallman, T. R., & McCray, R. 1982, ApJS, 50, 263, Proga, D., Waters, T., Dyda, S., & Zhu, Z. 2022, ApJL,
doi: 10.1086/190828 935, L37, doi: 10.3847/2041-8213/ac87b0
Krolik, J. H. 1999, Active galactic nuclei : from the central Rybicki, G. B., & Hummer, D. G. 1978, ApJ, 219, 654,
black hole to the galactic environment doi: 10.1086/155826
Laha, S., Reynolds, C. S., Reeves, J., et al. 2020, Nature Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 500, 33
Astronomy, doi: 10.1038/s41550-020-01255-2 Sim, S. A., Proga, D., Miller, L., Long, K. S., & Turner,
Lamers, H. J. G. L. M., & Cassinelli, J. P. 1999, T. J. 2010, MNRAS, 408, 1396,
Introduction to Stellar Winds doi: 10.1111/j.1365-2966.2010.17215.x
14 Smith, Proga, Dannen, Dyda, & Waters

Stevens, I. R., & Kallman, T. R. 1990, ApJ, 365, 321, Tarter, C. B., Tucker, W. H., & Salpeter, E. E. 1969, ApJ,
doi: 10.1086/169486 156, 943, doi: 10.1086/150026

You might also like