Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Adaptive Analytical Ray Tracing of Black Hole Photon Rings

Alejandro Cárdenas-Avendaño ,1, 2 Alexandru Lupsasca ,1, 3 and Hengrui Zhu 1

1
Princeton Gravity Initiative, Princeton University, Princeton, New Jersey 08544, USA
2
Programa de Matemática, Fundación Universitaria Konrad Lorenz, 110231 Bogotá, Colombia
3
Department of Physics & Astronomy, Vanderbilt University, Nashville, Tennessee 37212, USA
Recent interferometric observations by the Event Horizon Telescope have resolved the horizon-scale emission
from sources in the vicinity of nearby supermassive black holes. Future space-based interferometers promise
to measure the “photon ring”—a narrow, ring-shaped, lensed feature predicted by general relativity, but not yet
observed—and thereby open a new window into strong gravity. Here we present AART: an Adaptive Analytical
Ray-Tracing code that exploits the integrability of light propagation in the Kerr spacetime to rapidly compute
high-resolution simulated black hole images, together with the corresponding radio visibility accessible on very
long space-ground baselines. The code samples images on a nonuniform adaptive grid that is specially tailored to
the lensing behavior of the Kerr geometry and is therefore particularly well-suited to studying photon rings. This
numerical approach guarantees that interferometric signatures are correctly computed on long baselines, and the
modularity of the code allows for detailed studies of equatorial sources with complex emission profiles and time
arXiv:2211.07469v3 [gr-qc] 22 Feb 2023

variability. To demonstrate its capabilities, we use AART to simulate a black hole movie of a stochastic, non-
stationary, non-axisymmetric equatorial source; by time-averaging the visibility amplitude of each snapshot, we
are able to extract the projected diameter of the photon ring and recover the shape predicted by general relativity.

I. INTRODUCTION However, these observations have thus far not provided any
evidence for the presence of a lensed photon ring [23]. They
According to general relativity, black holes display unique
are instead dominated by n = 0 photons [6, 7], which form the
lensing behavior: for instance, any two spatial points outside
image layer that is more sensitive to the astrophysics of the
the event horizon are connected by infinitely many light rays,
flow than to purely relativistic effects [24–26]. As a result, the
each of which executes a different number of orbits around
bounds placed on possible deviations from general relativity
the black hole under its extreme gravitational pull [1–5]. As a
are on the order of several percent [27, 28] and comparable to
result, images of a black hole surrounded by a non-spherical,
the constraints derived from gravitational-wave observations
optically thin emission region decompose into a sequence of
of stellar-mass, binary black holes [29], or x-ray spectroscopy
superimposed layers indexed by photon half-orbit number n,
measurements of low-mass binaries and active galactic nuclei
with each layer consisting of a full lensed image of the main
[30]. By contrast, future detections of orbiting (n ≥ 1) photons
emission [5–11]. The direct (n = 0) layer typically displays a
will open a new window into strong gravity and enable higher-
central dark area—the black hole—encircled by the weakly
precision probes of the Kerr geometry, since it is this orbiting
lensed, primary image of the accretion flow onto the hole,
light which forms the part of the image—the photon ring—
whose details depend sensitively on astrophysical conditions.
that belongs to the black hole itself, rather than to its plasma.
On the other hand, the higher-n layers arise from photons
on highly bent trajectories that are strongly lensed to form There are three major obstacles to measuring a photon ring.
a series of narrow “photon rings.”1 These rings are usually First, since the photon rings are exponentially narrow (in n)
stacked on top of the broader n = 0 emission and their shape features, resolving them requires interferometric observations
rapidly converges (exponentially fast in n) to that of the “Kerr on exponentially long baselines [7]. At the current observing
critical curve”2 [15]: a theoretical curve in the observer sky frequency of 230 GHz, even Earth-spanning baselines are too
corresponding to the apparent image of asymptotically bound short to resolve the first ring, but it should become accessible
photon orbits. In contrast to the astrophysics-dependent n = 0 to next-generation space-based interferometers. In particular,
image, this “n → ∞ photon ring” is completely determined SALTUS (the Single Aperture Large Telescope for Universe
by the Kerr black hole—depending only on its mass, spin, and Studies) is a bold proposal—currently a contender for NASA’s
inclination—and delineates its cross-sectional area in the sky. upcoming Probe mission—to launch within the next decade a
Recently, 1.3 mm interferometric observations by the Event spacecraft far enough to access the first two rings of M87*.
Horizon Telescope have resolved the horizon-scale emission Optical depth poses a second hurdle: even though photons
from sources in the immediate vicinity of two nearby super- could in principle circumnavigate the black hole indefinitely
massive black holes: M87* [16], the central compact object (albeit unstably), in practice, those that traverse its emission
at the core of our neighboring galaxy Messier 87, and Sgr A* region multiple times are eventually reabsorbed by the matter
[17], our own black hole at the center of the Milky Way. Their they intersect—an effect that cuts off image layers past some
reconstructed images display a central brightness depression n > 0. Nevertheless, since absorptivity decreases with photon
within a thick ring consistent with theoretical expectations for energy, the first few rings still ought to be present in images
the direct image of the surrounding accretion flow [18–22]. taken at sufficiently high frequencies. Simple models suggest
that at 230 GHz, the n = 1 ring is always visible while the
n = 2 ring may be only marginally observable, whereas both
1 For an animation of this effect, see https://youtu.be/4-DvyMPs-gA. rings should be clearly visible at 345 GHz [11]. State-of-the-
2 Spherical emission creates a “shadow” inside of this curve [11–13], which art simulations of general-relativistic magnetohydrodynamic
is often called the “shadow edge” even when its interior is not dark [14]. (GRMHD) flows [31] also confirm these expectations [32].
2

For this reason, SALTUS is slated to simultaneously observe In other words, observations of the n = 2 photon ring can
at both frequencies. M87* makes for a particularly exciting in principle disentangle gravitational and astrophysical effects
prospective target because a measurement of its n = 2 ring that are otherwise commingled in the direct image. Moreover,
diameter could deliver a stringent test of the Kerr hypothesis, GLM simulated interferometric data of the kind that could be
which predicts a definite shape for its higher-n rings: photons collected by a mission like SALTUS, and were able to extract
orbiting just outside the horizon of a black hole can probe its this ring shape from the visibility amplitude on space-ground
extreme gravity and carry away information about its space- baselines. Their experimental forecast achieved a sub-percent
time geometry, encoded in the observable shape of the rings level of precision for the resulting test of the Kerr hypothesis,
that these photons produce in their observer’s sky [33–35]. suggesting that M87* holds exceptional promise as a target for
Time variability introduces a third significant complication. such a test in practice. This analysis was recently reviewed in
While time-averaged GRMHD-simulated movies have shown depth and extended to an even larger selection of equatorial
that the photon rings are persistent, sharp features that come to disk models, supporting these conclusions [10].
dominate observations with very-long-baseline interferometry While these early results are encouraging, demonstrating
(VLBI) after averaging over sufficiently long timescales [7], it the feasibility of a ring shape measurement requires further
remains to be understood how clearly visible the rings will be theoretical work. Crucially, even though the GLM analysis
to a realistic, near-future, space-VLBI mission like SALTUS, did include realistic instrument noise, it only considered time-
which will be limited in the number of snapshots it can collect. averaged images of equatorial disks. The latter limitation was
In each snapshot, such an interferometer—with a single space recently tackled with a study of geometric thick-disk models
leg—can only sample the radio visibility on one space-ground [11], but to date a detailed investigation of source fluctuations
baseline, thereby only measuring the projected diameter dϕ of and time variability has yet to be carried out.
the photon ring at one angle ϕ in the image.3 To compensate The present work is the first attempt to remedy this lacuna.
for its sparse baseline coverage, the instrument can observe at The main obstruction is technical: as high-order photon rings
regular intervals along its orbit around the Earth, eventually are exponentially narrow compared to the overall structure of
filling in every angle ϕ in the Fourier domain, with each dϕ a black hole image, resolving them in the image plane incurs
thus sampled twice per orbit. The baseline lengths over which a large computational cost. More precisely, their presence
the ring signature dominates the signal fix the orbital radius in the image introduces a large separation of scales between
(about lunar distance for the n = 2 ring of M87*) and hence the pixel grid size (which must be large enough to capture
the orbital period (∼ 1 month), which in turn sets the cadence the entire field of view) and the pixel spacing (the grid must
of these snapshots: ∼ 40MM87∗ , or roughly every two weeks. achieve a sufficiently fine resolution to see the narrow rings).
As it is evidently impractical to maintain coherence over such While a brute-force approach—pumping millions of pixels in
timescales, the snapshots must be incoherently time-averaged; the grid—can overcome this scale separation for a handful of
moreover, only ∼ 24 snapshots of dϕ can be sampled per year. images, it becomes intractable when dealing with a black hole
In a single snapshot, the “clean” interferometric signature movie consisting of several hundred snapshots, in which case
of the ring—a periodic ringing in the visibility amplitude— adaptive ray tracing is necessary [36, 37] (and less wasteful).
is typically “polluted” by noise from both the instrument and Here, we present a numerical tool4 designed to efficiently
from astrophysical fluctuations (plasma flares, emission ropes, compute high-resolution “slow-light” movies of generic (non-
or other ring mimickers), which can obscure the signal. The stationary and non-axisymmetric) equatorial sources around a
key question is then: Can the interferometric signature of a Kerr black hole, together with their associated radio visibility
photon ring—and hence its projected diameter—be recovered on very long baselines; we present example outputs in Fig. 1.
from an incoherent time-average over N ≈ 20 snapshots of its The code was developed with the intent to maximize speed
visibility amplitude on very long space-ground baselines? while still guaranteeing the accuracy of its output, particularly
An affirmative answer to this question would open the door the radio visibility on very long baselines, which encodes the
to a consistency test of the Kerr hypothesis—a cornerstone of high-frequency components of a snapshot’s Fourier transform
general relativity (GR) in the strong-field regime—via space- and is therefore extremely sensitive to its most minute image
VLBI measurements of the photon ring shape. The paper [35] features. The code’s structure is highly modular and most of
(henceforth: GLM) took the first steps toward establishing its individual components reproduce pre-existing capabilities;
the viability of such a test for M87*. GLM examined a its main novelty is arguably to combine all these routines into
range of models of stationary, axisymmetric, equatorial disks one single and convenient-to-use (we hope!) package.
that reproduce the time-averaged observational appearance One new and important technique from which AART derives
of GRMHD-simulated flows, and found that the observable much of its power deserves special mention: it turns the very
shape of their n = 2 ring always follows a specific functional feature of photon rings that makes them so difficult to fully
form, independent of the source model. They concluded that resolve—namely, their thinness—to its advantage. It does so
this ring shape is a robust prediction of strong-field GR. by decomposing the full image into multiple layers labeled by
half-orbit number n, and ray tracing in each layer the image of
the nth photon ring with exponentially high (in n) resolution.
3 The angle in the image corresponds to the space element’s baseline angle in
the Fourier plane [33–35], while the baseline length determines the index n
of the subring whose interferometric signature dominates the signal [7, 10]. 4 The code is publicly available at https://github.com/iAART/aart.
3

α (M )
−10 −5 0 5 10 −10 −5 0 5 10 −10 −5 0 5 10 −10 −5 0 5 10

10

5
β (M )

−5

−10
Visibility Amplitude (Jy)

100
V (0) = 0.53 Jy V (0) = 0.66 Jy V (0) = 0.83 Jy V (0) = 0.91 Jy - ϕ = 0◦
- ϕ = 90◦
10−1

10−2

10−3

10−4

10−5
0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500

Baseline Length u (Gλ)

FIG. 1. Top: Snapshots of a non-stationary, non-axisymmetric, equatorial source around a Kerr black hole ray-traced with AART. Most image
features in these snapshots arise from fluctuations in the source, with the exception of the strikingly bright and narrow photon ring: a persistent,
sharp feature that comes to dominate the image formed by time-averaging many of these snapshots (Fig. 2). These images were produced using
the parameters listed in Table II at regular time intervals of 250M . Bottom: The corresponding visibility amplitudes for spin-aligned (blue)
and spin-perpendicular (red) cuts across each of the above images. The black hole spin is a/M = 94% and the observer inclination is θo = 17◦ .

More precisely, the lensing behavior of a black hole forces By contrast, the adaptiveness provided by the lensing bands
the nth photon ring to lie within an exponentially small (in n) is determined by the Kerr geometry alone and results in the
region of the image plane, dubbed the nth lensing band, with same, uniform grid for every image of a given black hole spin
each band completely fixed by the Kerr geometry [10, 35]. A and inclination. These grids therefore need be computed only
key innovation of AART is to first compute (once and for all) once, and provided that their resolution increases at the proper
the lensing bands associated with a given black hole spin and rate set by the demagnification factor, they cannot miss any
inclination, and then to ray trace, for the nth layer, only pixels feature that is already resolved in the direct n = 0 image.
lying in the nth lensing band. Since each lensing band contains Finally, high-n layers comprise photons that execute many
an exponentially demagnified image of the main emission, by orbits in the Kerr photon shell [39] where their radial potential
also increasing the resolution in each band exponentially,5 the almost develops a double root and geodesic integrals diverge
code is guaranteed to resolve the source with roughly the same logarithmically, leading to a growing risk of numerical error.
effective resolution in every layer. In other words, AART adapts To minimize error, AART performs analytical ray tracing using
its ray-tracing resolution (grid spacing) to each layer, but also an exact solution of the Kerr null geodesic equation recently
adjusts the ray-tracing region (grid size), so as to resolve the given in terms of Legendre elliptic integrals and Jacobi elliptic
increasingly fine features present in higher layers using only a functions [40], similar in spirit to previous implementations
fixed number of pixels per layer. based on Carlson symmetric forms [41, 42].
Other ray tracers use adaptive mesh refinement to increase To summarize: AART uses lensing bands to prevent a large
pixel density in regions where they detect long geodesic path separation of scales (between the grid size and its spacing)
lengths [36] or large gradients [37, 38]. This results in a new, and thereby maintain its speed as it increases the pixel density
non-uniform grid for every new snapshot, which is recursively in higher-n photon rings, which is necessary to ray trace the
refined until a desired criterion is met, or else the number of fine image features that (due to the nonlocal character of the
recursions exceeds a hard-set limit. Therefore, small image Fourier transform) influence the visibility on long baselines;
features can sometimes be missed when this limit is hit. moreover, the rings are ray traced analytically to avoid errors.
These features of AART are specially tailored to the photon
ring and its interferometric signature; with this tool in hand, it
5 Successive subrings are demagnified by an analytically known, angle-and- becomes feasible to investigate the effects of time variability
spin-dependent factor e−γ(ϕ) , where γ(ϕ(r̃)) is the Lyapunov exponent that on measurements of the n = 2 ring shape, and we can begin to
governs the instability of nearly bound photons at orbital radius r̃ [5, 7]. answer the key experimental question posed above.
4

10 100
- ϕ = 0◦ - ϕ = 90◦
0.30

0.25

10 −1 0.20

(mJy)
5 0.15

Visibility Amplitude (Jy)


0.10

0.05
10−2 425 430 435 440 445
β (M )

10−3

-5
10−4

-10 10−5
-10 -5 0 5 10 0 100 200 300 400 500
α (M ) Baseline Length u (Gλ)

FIG. 2. Left: Time average of 100 ray-traced snapshots of inoisy equatorial profiles uniformly sampled over a time interval of 1000M. The
inoisy parameters take the “best-guess” values for M87* presented in Table II. Hence, this image is directly comparable to the time-averaged
image of M87* presented in Fig. 1 of Ref. [7], for which the underlying GRMHD simulation’s parameters were chosen to be consistent with the
2017 Event Horizon Telescope observations of M87*. Right: Visibility amplitudes of all 100 snapshots along cuts parallel (ϕ = 0◦ , blue) and
perpendicular (ϕ = 90◦ , red) to the black hole spin axis. The solid lines correspond to the incoherently time-averaged amplitudes. Embedded is
a panel zooming into the average visibility amplitude in the baseline range u ∈ [425, 445] Gλ, with the best-fit ring signature overlaid (black).

To study the effects of source fluctuations on the observed Correspondingly, the incoherently time-averaged visibility
visibility amplitude, we call upon both AART and inoisy [43]: amplitude is dominated on long baselines by the perfectly
a code that can rapidly generate realizations of a 2D Gaussian clean interferometric signature of the n = 2 ring. In particular,
random field with Matérn covariance. Such a field can provide we find that the signal in the range u ∈ [425, 445] Gλ—which
a simple statistical model for a generic (non-stationary and a satellite at lunar distance from the Earth could access with
non-axisymmetric) stochastic source in the equatorial plane 345 GHz observations—exactly follows the periodic ringing
of a Kerr black hole, with a prescribed two-point function. pattern predicted for a thin ring (black overlay in panel inset).
Since a “realistic” choice of autocorrelation structure is not yet As we show in Fig. 3, the projected diameter dϕ of the n = 2
known (and will require additional research into the physics of ring can then be extracted from the periodicity of this ringing
the plasma), our goal will be to vary the statistics of the model in the visibility amplitude |V(u, ϕ)|. The top panels display the
within a wide range of “reasonable” possibilities (informed by ring diameter dϕ inferred from an average over N snapshots,
GRMHD simulations) so as to parameterize our uncertainty in with N = 5, 10, 20, and finally 100, by which time the smooth
the expected variability of signals from sources such as M87*. shape of the ring has emerged. The bottom panels display the
Completing such a parameter survey is a large undertaking relative deviation from this fiducial shape that is induced by
beyond the scope of this first paper. Here, we will be content astrophysical fluctations, whose noise clearly averages out of
with a proof-of-concept demonstration that AART is up to the the image; it is encouraging to see this noise is also beat down
task for such a study. To showcase its capabilities, we ray in measurements of the ring shape using only a few snapshots.
trace 100 snapshots of an inoisy source with statistics set Of course, whether these conclusions are likely to hold for
by the “best-guess” parameters for M87* (listed in Table II). M87* has yet to be established, and a systematic investigation
Sample snapshots and their associated visibility amplitude are of astrophysical fluctuations remains to be done. In a soon-to-
shown in Fig. 1, while Fig. 2 includes all of the snapshots, be-released paper, we will initiate such a study by repeating
with the left panel displaying their time-averaged image and this analysis for multiple models of M87* in which we vary
the right panel all the individual visibility amplitudes together the parameters both of the black hole (its spin and inclination)
with their incoherent time average (solid lines). As expected, and of the source (the inoisy model). We also hope to report
we find that the astrophysical fluctuations wash out from the on the n = 1 ring’s signature and how it may encode the spin.
time average, leaving an image that is visibly dominated by a To guide the reader, we now give a summary of the rest of
prominent photon ring with clear n = 1 and n = 2 subrings. the paper, which—like AART—is written in a modular way.
5

38.4
N =5 N = 10 N = 20 N = 100
38.3
dϕ (µas)

38.2
38.1
38.0
37.9
0.08
∆(%)

0
-0.08
0◦ 45◦ 90◦ 135◦ 180◦ 0◦ 45◦ 90◦ 135◦ 180◦ 0◦ 45◦ 90◦ 135◦ 180◦ 0◦ 45◦ 90◦ 135◦ 180◦
Baseline Angle ϕ

FIG. 3. Top: The projected diameter dϕ of the n = 2 photon ring inferred from an incoherent time average over N snapshots of the source. The
green curves are best fits of the GR-predicted shape [Eq. (127)] to the synthetic data. To show how the fits improve with increasing N, each
panel displays the data of its predecessor in the background. Bottom: The difference between the best-fitting curve at a given N and the “true”
shape of the ring (solid gold line in last panel) to which the signal converges as N → ∞. The parameters of the model are listed in Table II.

Summary We model these fluctuations using Gaussian random fields and


describe their statistical properties in great detail. In Sec. IV,
we use inoisy to simulate a variable source. With AART, we
In Sec. II, we review the problem of light propagation in the
ray trace its instantaneous snapshots (Fig. 12) and compute its
Kerr spacetime. We write down the null geodesic equation and
light curve (Fig. 13). Great care must be taken in the choice
its exact analytical solution as it is implemented in AART, and
of resolution and field of view used in each layer, and Sec. V
describe the key concept of lensing bands (Fig. 4). We then
discusses these issues in depth (Figs. 15, 16, and 14).
illustrate the lensing behavior of the black hole by plotting its
“transfer functions”: mappings of directions in the observer We can then produce movies of a stochastic, non-stationary,
sky to the spacetime points where the corresponding light rays non-axisymmetric source. In Sec. VI, we compute movies of
intersect the equatorial plane. The transfer functions for polar the associated visibility amplitude and use this synthetic data
coordinates (rs , φs ) in the plane are shown in Figs. 5, 6, and 9 to reconstruct the projected diameter dϕ of the n = 2 ring,
for the n = 0, n = 1, and n = 2 images, respectively. Likewise, before concluding with a brief discussion of future prospects
the time lapse ∆t between source and observer is also shown in in Sec.VII. We relegate some details to Apps. A, B, and C.
Figs. 7 and 8 for the n = 0 and n = 1 images, respectively. The Throughout the paper, we work in (−, +, +, +) metric signature
nth image always fills out the nth lensing band, as expected. with geometric units in which GN = c = 1. Our conventions
for Legendre elliptic integrals are listed in App. A of Ref. [46].
We also give an approximate formula for light bending that
was derived by Beloborodov [44] for nonrotating black holes
in the weak-deflection regime. We express his result in a very
II. THEORETICAL FRAMEWORK
simple form [Eq. (43)] that proves to be remarkably accurate
for the computation of n = 0 images of axisymmetric sources,
for most black hole spins and inclinations. This observation, The special integrability properties of the Kerr spacetime
which is illustrated in Fig. 10, highlights the fact that n = 0 reduce its geodesic equation to a problem of quadratures [47],
photons barely carry an imprint of the black hole spin, as they resulting in elliptic integrals that are expressible in Legendre
do not spend enough time near it to be strongly affected by its normal form [40, 48–50]. Modern computers can evaluate the
gravitational field. Beloborodov’s approximation can thus be Legendre elliptic integrals very fast and to arbitrary precision,
regarded as a generalization to all inclinations of the “just add reducing the computational cost of ray tracing in Kerr. In this
one” prescription for the transfer function of a spin-aligned section, we review the exact solution of the Kerr null geodesic
observer, for whom the impact parameter ρ is simply related equation in Legendre form [5, 40] as well as the approximate
to emission radius rs by adding one: ρ ≈ rs + M [5, 45]. solution derived in Schwarzschild by Beloborodov [44], and
AART can also parallel transport linear polarization. We check we use them to plot the transfer functions mapping Bardeen’s
that Beloborodov’s approximation is adequate for ray tracing coordinates in the observer sky [15] to the equatorial plane.
n = 0 polarimetric images (Fig. 11), which can be done using
simple algebraic equations that we write down explicitly.
In Sec. III, we describe our model of equatorial emission. A. Null geodesics of the Kerr exterior
First, we review stationary and axisymmetric equatorial disk
profiles that can reproduce the time-averaged observational Astrophysical, rotating black holes of mass M and angular
appearance of M87* in GRMHD simulations [9, 35]. Then, momentum J = aM are subject to the Kerr bound |a| ≤ M and
we add in astrophysical fluctuations with prescribed statistics. are described by the Kerr geometry.
6

The Kerr metric is written in Boyer-Lindquist coordinates Both potentials have exactly four (not always real) roots
(t, r, θ, φ) in terms of functions Σ(r, θ) = r2 + a2 cos2 θ and
∆(r) = r2 − 2Mr + a2 . The roots of ∆(r) define the radii of the
r
A B √ 
outer and inner event horizons: r1 = −z − − − z2 + , θ1 = arccos u+ , (11)
2 4z
√ r
r± = M ± M 2 − a2 . (1) A B √ 
r2 = −z + − − z2 + , θ2 = arccos u− , (12)
2 4z
With respect to Mino time τ, a photon with four-momentum r
pµ follows a null geodesic xµ (τ) obtained by solving A
r3 = z − − − z2 − ,
B  √ 
θ3 = arccos − u− , (13)
2 4z
dxµ Σ
= − pµ .
r
(2) A B  √ 
dτ pt r4 = z + − − z2 − , θ4 = arccos − u+ , (14)
2 4z
The resulting trajectory is independent of the photon energy
−pt and can be parameterized by two conserved quantities: which depend only on the conserved quantities (λ, η) via6
the energy-rescaled angular momentum and Carter constant s
r r
pφ p2θ ω+ + ω− − A/3 3 Q P3 Q2
λ=− , η= − a2 cos2 θ + λ2 cot2 θ. (3) z= , ω± = − ± + , (15)
pt p2t 2 2 27 4
B
We are interested in solving for the trajectories xµ (τ) that A = a2 − η − λ2 , = η + (λ − a)2 , C = −a2 η, (16)
connect two spacetime events (ts , rs , θs , φs ) and (to , ro , θo , φo ), 2M  !2 
where the labels ‘s’ and ‘o’ stand for ‘source’ and ‘observer’, A2 A  A  B2
P=− − C, Q = −  − C − , (17)
respectively. We will specialize to distant observers (ro  M) 12 3 6 8
at nonzero inclination θo ∈ (0, π/2) above the equatorial plane
η η + λ2
r !
1
θs = π/2 where we place the source. (The measure-zero cases u± = 4θ ± 4θ + 2 , 4θ =
2
1− . (18)
θo ∈ {0, π/2} technically require separate treatments [5], but a 2 a2
in practice one can simply change the inclination slightly and
We will now restrict our attention to positive spins 0 ≤ a ≤ M.
use the same analytical expressions.) The symmetries of the
Kerr geometry—its stationarity and axisymmetry—allow us
to set ts = 0 and φo = 0 without loss of generality.
B. Kerr critical curve
A photon that reaches an observer with four-momentum pµo
and conserved quantities (λ, η) appears in the sky at Cartesian
position (α, β) given in terms of ±o = sign pθo by Bardeen [15], The radial potential (9) develops a double root at r̃ ≥ r+
(that is, R(r̃) = R0 (r̃) = 0) if and only if [5, 15]
λ
q
α=− , β = ±o η + a2 cos2 θo − λ2 cot2 θo , (4)
sin θo r̃3 4M∆(r̃)
" # " #
r̃ 2∆(r̃)
λ̃ = a + r̃ − , η̃ = 2 − r̃ , (19)
for ro → ∞. If the photon carries a linear polarization, then its a r̃ − M a (r̃ − M)2
observed electric vector polarization angle (EVPA) is [51–54]
where r̃ ∈ [r̃− , r̃+ ] with
νκ1 − βκ2
!
χ = arctan , ν = −(α + a sin θo ),
"  a !#
(5) 2
βκ1 + νκ2 r̃± = 2M 1 + cos arccos ± . (20)
3 M
where κ denotes the complex-valued Penrose–Walker constant
A geodesic with critical conserved quantities λ = λ̃ and η = η̃
κ = κ1 + iκ2 = (PA − iPB )(r − ia cos θ), (6)
asymptotes to an unstably bound orbit at radius r̃ in the Kerr
PA = pt f r − pr f t + a sin2 θ pr f φ − pφ f r ,
   
(7) photon shell. For a distant observer, the Kerr critical curve C
h 
φ θ θ φ
 
t θ θ t
i is the image in the sky of these asymptotically bound orbits:
PB = r + a p f − p f − a p f − p f sin θ, (8)
2 2
n   o
a quantity that is also conserved along null geodesics—thanks C = α̃, β̃ : (λ, η) = λ̃, η̃ . (21)
to the Petrov type D nature of the Kerr metric [55]—and can
be used to algebraically solve the parallel transport problem Though this purely theoretical curve is not in itself observable,
for a linear polarization vector f by evaluating κ at the source. it does play a key role in the study of lensing by a Kerr black
The separability of Kerr geodesic motion allows the r and θ hole [56]. It is traced using Eqs. (4) evaluated on Eqs. (19) for
trajectories to be decoupled [47], These independent motions all the values r̃ ∈ [r̃− , r̃+ ] such that β̃2 ≥ 0. It is always closed,
are then controlled by radial and angular geodesic potentials convex, and reflection-symmetric about the α axis [34].
whose zeros give the turning points of their respective motion:
 2 h i
R(r) = r2 + a2 − aλ − ∆(r) η + (λ − a)2 , (9) √
6 Here, 3 x denotes the real cube root of x if x is real, or else, the principal
Θ(θ) = η + a cos θ − λ cot θ.
2 2 2 2
(10) value of the function x1/3 (that is, the cubic root with maximal real part).
7

C. Analytical backwards ray tracing If a ray lies outside of C, then its motion is always of type
(2), and the definite integrals down to radius rs on the ray are
The character (real or complex) and ordering of the radial
Ik (rs ) = I(2) (2)
k (ro ) ∓ Ik (rs ), (25)
roots {r1 , r2 , r3 , r4 } (when they are real) lead to a classification
of radial motion into four types [40, 57, 58]. We are, however, with sign −/+ before/after reaching the turning point at r = r4 .
only interested in the rays that connect a distant observer to The Mino time τ(rs ) elapsed along a ray inside C is thus
an equatorial source. This excludes two of the motion types, Z ro
dr
leaving only two others corresponding to rays that start from τ− (rs ) = √ ≡ I0 (rs ) = I(2,3) (2,3)
0 (ro ) − I0 (rs ). (26)
infinity at one end point before they either cross the horizon rs R(r)
or return to infinity at the other.7 The boundary between these Hence, the total Mino time elapsed along the full light ray is
two behaviors in the phase space of null geodesics constitutes Z ro
the Kerr photon shell of asymptotically bound orbits, a special dr
τ−max = √ = τ− (r+ ). (27)
locus in phase space with emergent conformal symmetry [59]. r+ R(r)
We now describe how this separation is manifested in the sky. Likewise, the Mino time τ(rs ) elapsed along a ray outside C is
Given a black hole spin and observer inclination (a, θo ), we " Z ro Z rs #
can pick a direction (α, β) in the observer sky and shoot a light dr
τ+ (rs ) = +2w √ = I(2) (2)
0 (ro ) ∓ I0 (rs ), (28)
ray backwards into the geometry in that direction. Such a ray rs r4 R(r)
will cross the equatorial plane a total number of times N(α, β),
which can be determined by computing the total Mino time τ where the sign is − before the turning point at r = r4 (w = 0)
elapsed along the entire development of the trajectory. Rays in and + after the bounce (w = 1). Hence, the total Mino time
the interior of the critical curve all fall into the black hole; that elapsed along the full light ray is
Z ro
is, they encounter no radial turning point and terminate their + dr
motion across the Kerr exterior on the event horizon at r = r+ . τmax = 2 √ = 2I(2)
0 (ro ), (29)
r+ R(r)
Rays in the exterior of the critical curve are all deflected back
to infinity; that is, their radial motion encounters a turning and its radial turn occurs at the “half-way” Mino time
point at r = r4 , whereupon they bounce back toward r → ∞. Z ro
dr τ+max
Thus, the critical curve (21) delineates the boundary between τ4 = √ = I(2)
0 (ro ) = , (30)
r+ R(r) 2
photon capture (its interior) and photon escape (its exterior),
and may be regarded as the cross-sectional area of the hole. We will use τmax to denote the appropriate choice of τ±max .
In-between rays that lie exactly on C are trapped in the photon At last, the total number of equatorial crossings N(α, β) is
shell where they can in principle orbit forever; in practice, this  p 
never occurs because such orbits are unstable (or equivalently,  τmax −u− a2 + sign(β)Fo 
N =    − H(β) + 1, (31)
because C is infinitely thin). 2K
√ √
Let ri j = ri − r j and define A = r32 r42 , B = r31 r41 , and
where H(x) denotes the Heaviside function while K and Fo
r !
2 r − r4 r31 r32 r41 are defined in Eqs. (A4)–(A5); see also App. A of Ref. [10].
I0 = √
(2)
F arcsin , (22) A light ray crosses the equatorial plane for the (n + 1)th time
r31 r42 r − r3 r41 r31 r42
r−r2 B  at Mino time τ = τ(n)
s , where
 (A + B)2 − r21
  2 

1  1 − r−r 1 A 
I0 = √ F  arccos 
(3)
 , (23)
 
   τ(n)
s = G θ ∈ [0, τmax ] ,
(n)
n ∈ {0, . . . , N − 1} , (32)
1 + r−r1 A
r−r2 B  4AB

AB
where F(ϕ|k) is an incomplete elliptic integral of the first kind. with G(n)
θ given in Eq. (A1). This equatorial crossing occurs at
These functions are the antiderivatives (A10) and (A22), and  
rs(n) = rs(2,3) τ(n)
s , (33)
all the radial geodesic integrals Ik are definite integrals that can
be obtained from their respective antiderivative Ik as follows. φ(n)
s = φo − Iφ − λG φ ,
(n) (n)
(34)
If a ray lies inside of C, then its motion is, following the
labeling introduced in Ref. [40], of type (2) when all roots are ts(n) = to − It(n) −a 2
t ,
G(n) (35)
real (in which case r+ > r− > r4 > r3 > r2 > r1 ), or else with rs(2,3) (τ)
given in Eqs. (A20) and (A36) for geodesics of
of type (3) when r3 = r̄4 are complex-conjugate roots with
types (2) and (3), respectively, while the angular integrals G(n)
r+ > r− > r2 > r1 .8 In either case, the definite integrals down t,φ
are given in Eqs. (A2)–(A3). Meanwhile, the radial integrals
to radius rs on the ray are  
(n)
Ik (rs ) = I(2,3) (ro ) − I(2,3) (rs ). (24) It,φ = It,φ rs(n) (36)
k k
decompose via Eqs. (A8)–(A9) into definite integrals that are
to be evaluated via Eq. (24) for rays inside C or via Eq. (25)
7 The angular motion also has two possible behaviors according to whether for rays outside C, using the antiderivatives I(2,3)
k given by
η ≷ 0. We ignore vortical rays with η < 0 as they cannot reach the equator Eqs. (A10)–(A13) for type (2) geodesics or Eqs. (A22)–(A25)
[46]. Such rays always lie within the apparent image of the horizon. for type (3) geodesics. In practice, we take ro  M to be very
8 There exist type (4) rays with both r1 = r̄2 and r3 = r̄4 , but they are vortical. large but not infinite, for reasons described in Eq. (41) below.
8

15 This last observation leads us to a formula that characterizes


the boundary of the nth lensing band: it is the set of points
(α, β) in the image plane for which
10 p
τmax −u− a2 + sign(β)Fo
− H(β) = n, (37)
2K
5
such that N(α, β) jumps by one across the boundary of each
lensing band. This condition can equivalently be written as
β (M )

0
τmax = G(n)
θ , (38)

−5 with G(n)
θ given in Eq. (A1). The n ≥ 1 lensing bands have a
different character than the n = 0 layer—we now focus on the
former and defer a discussion of the latter to the next section.
−10
Each n ≥ 1 lensing band is annular in shape and foliated by
contours of fixed rs(n) , with every source radius mapping onto a
−15 unique closed curve within the band. Moreover, this bijective
−15 −10 −5 0 5 10 15 map is order-preserving: moving radially outwards within a
α (M ) band, one crosses contours of monotonically increasing rs(n) .
Thus, the inner edge of the nth lensing band is the (n + 1)th
FIG. 4. Illustration of lensing bands with different grid resolutions, image of the equatorial event horizon (the contour of fixed
for black hole spin a/M = 50% and observer inclination θo = 60◦ . rs(n) = r+ ), while its outer edge is the (n + 1)th image of the
Orange, purple and magenta marks depict light rays that intersect the equatorial circle at infinity (the contour of fixed rs(n) = ∞).
equatorial plane once (n = 0 band), twice (n = 1 band) and thrice Rays that connect a distant observer to the event horizon
(n = 2 band), respectively. The black dashed line is the critical curve
cannot encounter a radial turning point along their trajectory,
(21). Embedded is a panel showcasing the increasing grid resolution
in higher-n bands. The region inside the lensed n = 0 equatorial
so the inner edge must always appear inside the critical curve.
horizon, shown in black, is always excluded from our calculations. Conversely, rays that connect a distant observer to infinity
have to make a turn along their radial trajectory, so the outer
edge must always appear outside the critical curve. Hence, the
D. Kerr lensing bands
n ≥ 1 lensing bands are annuli that always straddle the critical
curve. Since successive images of a source are demagnified
and appear exponentially closer to the critical curve [5, 7], the
The functions rs(n) , φ(n) (n)
s , and ts defined in Eqs. (33)–(35) lensing bands form a stack of nested annuli, as seen in Fig. 4.
are the “transfer functions” that map the equatorial plane to To summarize, the inner edge of the nth lensing band is
its (n + 1)th lensed image in the observer sky, where the index the solution of Eq. (38) with τmax = τ−max given in Eq. (27),
n ≥ 0 may be thought of as a photon half-orbit number [5, 7]. while its outer edge is the solution with τmax = τ+max given in
To describe the lensing behavior of a Kerr black hole, it is Eq. (29). The two sides of Eq. (38) are continuous functions
helpful to draw contour plots of these function—that is, level G(n)
θ (α, β) and τmax (α, β) of the image-plane position (α, β).
±
sets of fixed rs(n) , φs(n) , and ts(n) within each image layer n. First, In practice, we use the following procedure to determine
however, one must determine the regions of the image plane these edges. We first select a set of polar angles ϕi ∈ [0, 2π]
in which these functions have support. around the image plane, with associated positions (α̃i , β̃i ) on
This is a nontrivial problem because these functions always the critical curve (21)—that is, (α̃i , β̃i ) ∈ C is the point where
evaluate to some value: for instance, rs(n) (α, β) always returns the critical curve is intersected by the ray emanating from the
some source radius, even when the ray shot back from (α, β) origin at polar angle ϕi . Next, we parameterize these rays
does not in fact intersect the equatorial plane n + 1 times. In as (αi (), βi ()) = (α̃i , β̃i ). For each ϕi , we compute G(n)
θ ()
such cases, rs(n) (α, β) may sometimes be obviously unphysical along the corresponding ray over the range  ∈ (0.5, 3). We
(it could, for example, take a negative value), but not always. also compute τ−max () on the range  ∈ (0.5, 1) and τ+max () on
We define the nth lensing band as the image-plane subregion the range  ∈ (1, 3). We then determine the unique parameters
consisting of those rays that cross the equatorial plane at least  ± within each of these ranges such that G(n) θ ( ) = τmax ( ).
± ± ±

n + 1 times after being shot back from the observer, before These values correspond to the points (αi , βi ) =  (α̃i , β̃i )
± ± ±
either terminating on the horizon (if they are shot back from along the ray of constant ϕi where it intersects the lensing
inside the critical curve C) or else returning to infinity (if they band’s outer and inner edges, respectively. Connecting the
are shot back from outside C) after an elapsed Mino time τmax . points (α±i , β±i ) thus obtained at every angle ϕi traces out
By definition, the nth lensing band is the physical domain the edges of the lensing band, which is finally obtained as
of the transfer functions rs(n) , φ(n) (n)
s , and ts . This definition also the annular region between these two closed curves. These
implies that it is the subregion of the image plane in which the boundaries—and hence, the lensing band—can be determined
function N(α, β) defined in Eq. (31) is precisely equal to n + 1. to arbitrarily high precision by sampling sufficiently many ϕi .
9

To check the computation, one can verify that for each i, This layer-by-layer approach naturally allows for different
rs(n) (α−i , β−i ) ≈ r+ and rs(n) (α+i , β+i ) ≈ ∞ up to numerical error.9 resolutions to be used in each image layer. This is important
Further discussion of lensing bands may be found in Ref. [10]. because the exponential-in-n demagnification of successive
images of a source requires the use of exponentially higher
resolutions to resolve the source at the same level of detail in
E. Apparent image of the horizon higher-n layers. AART increases the resolution in successive
lensing bands by roughly the same demagnification factor eγ
We now return to the n = 0 lensing band, which (unlike its that shrinks them (γ is a Lyapunov exponent governing the
higher-n counterparts) is noncompact. This difference can be orbital instability of bound photon orbits [7]) such that every
understood intuitively: if the black hole mass shrinks to zero, image of the source is resolved by roughly the same number
the n ≥ 1 lensing bands must disappear with the hole, whereas of pixels—see Sec. V for a more detailed discussion of the
the n = 0 band must extend to fill the entire image plane. resolution requirements in each layer. This approach enables
Mathematically, the equations in the previous section still ray tracing up to arbitrarily high n and ensures that all images
hold. The inner and outer edges of the n = 0 band are still the of the source are equally well resolved, with a roughly fixed
contours of fixed rs(0) = r+ and fixed rs(0) = ∞, respectively. computational cost per layer.
These curves respectively correspond to the primary image To summarize, AART makes use of a non-uniform adaptive
of the equatorial event horizon (appearing inside C) and of grid composed of multiple layers, each of which it ray traces
the circle of infinite radius (appearing outside C). The inner at a resolution adapted to the lensing band that makes up the
edge can be determined via the same root-finding procedure layer. Since the lensing bands are completely determined by
outlined above (with n = 0), resulting in a closed curve H the Kerr geometry via the analytic formula (38), this form
contained within the critical curve C. On the other hand, the of adaptiveness is analytic, with the ray tracing also carried
outer edge is at infinity in the image plane (as it would be in out analytically using the exact transfer functions (33)–(35).
flat space), so the n = 0 lensing band is no longer an annulus. This adaptiveness is purely geometrical and independent of
Instead, it consists of the entire (unbounded) exterior of H. image features or gradients (unlike, for example, the method
Meanwhile, the interior of H may technically be viewed as in Refs. [36, 37]).
the n = −1 band, since it consists of rays that never reach the In practice, to produce the n = 0 grid points, AART creates
equator before terminating on the horizon. In an equatorial a Cartesian grid covering the desired area of the image plane,
disk model, this region would be completely dark, giving rise but excluding points in the apparent image of the horizon (that
to an “inner shadow” feature [9], and may naturally be viewed is, the interior of H). In each n ≥ 1 layer, the code produces a
as the apparent image of the horizon. In Fig. 4, we display this Cartesian grid of points in an area centered around the critical
image as a shaded black region for the case of a black hole curve and then discards those points which do not lie within
with spin a/M = 50% observed from an inclination θo = 60◦ . the nth lensing band. Numerically, the code only keeps those
points which lie within the concave hull of the points (α+i , β+i )
and outside the concave hull of the points (α−i , β−i ).
F. Analytic grid adaptiveness We caution the reader that the n ≥ 1 bands are not always
convex: at high inclinations and moderate to high spins, their
It has long been known that images of the equatorial plane edges may become non-convex curves, as may be seen for
of a Kerr black hole are lensed into increasingly demagnified, instance in the right column of Fig. 6 for the n = 1 band or
compact regions of the image plane [2, 60]. More recently, in Fig. 9 for the n = 2 band. In such cases, using the convex
this lensing behavior has been exploited to efficiently ray trace hull of the boundary points (α±i , β±i ) to define the lensing band
equatorial disk models in which the n = 0, 1, and 2 image can produce an incorrect grid that overestimates the size of
components are all fully resolved [35]. the band and includes points which do not really belong in it,
AART is designed with this goal in mind and is built around which is why must use the concave hull instead.
the lensing behavior of a Kerr black hole. It ray traces images In Fig. 4, we display an example of an AART grid (with a
of an equatorial source layer-by-layer, using a non-uniform coarse resolution for visualization purposes), which includes
grid adapted to this layered structure. The nth image layer, points in the n = 0 (orange dots), n = 1 (purple crosses) and
which consists of the (n + 1)th image of the source, is only ray n = 2 (magenta dots) lensing bands.
traced on pixels within the nth lensing band, which is precisely
the region of the image plane occupied by this image. As a
result, AART avoids computing unneeded pixels in each layer, G. Visualization of the equatorial transfer functions
substantially improving its efficiency.
Having defined the transfer functions (33)–(35) that map
the equatorial plane of a Kerr black hole onto the image plane
9 In principle, one could also determine the inner and outer boundaries of the of a distant observer, we will now illustrate their behavior with
nth lensing band as those contours of fixed rs(n) (α, β) = r− (inside C) or fixed the help of contour plots. Though we have yet to describe their
rs(n) (α, β) = ∞ (outside C), respectively, which are closest to C. However, range, we have already elucidated their domain of definition:
this method is impractical because rs(n) (α, β) can vary significantly across as described in Sec. II D, the functions rs(n) , φ(n) (n)
s , and ts are
these contours. Yet, it is this very behavior that makes this check effective. th
only physical within the corresponding n lensing band.
10

20

10

−10

−20
β (M )

20

10

−10

−20

−20 −10 0 10 20 −20 −10 0 10 20


α (M )

FIG. 5. Direct (n = 0) images of the equatorial plane θs = π/2 for Kerr black holes with spins a/M ∈ {50%, 99%} and observer inclinations
θo ∈ {17◦ , 80◦ }. The isoradial curves (purple) correspond to rings of constant Boyer-Lindquist radius rs /M ∈ {3, 6, 9, 12, 15, 18} between rs = r+
(the apparent image of the equatorial horizon, in black) and the cutoff at rs = 20M. The colors change every 45◦ across curves of constant φs .

Following Refs. [2, 5], in Figs. 5 and 6, we display contours In each band, rs(n) spans the entire range [r+ , ∞) once and only
of fixed rs(0) and rs(1) , which foliate the n = 0 and n = 1 lensing once, so every equatorial ring produces precisely one image in
bands, respectively. Physically, these “isoradial” curves are each lensing band. Moreover, the map rs(n) is order-preserving:
the direct (n = 0) and first relativistic (n = 1) images of the the source radius grows monotonically in the radial direction.
rings of constant Boyer-Lindquist radius rs in the equatorial
plane θs = π/2. Since the n = 0 lensing band is unbounded, These two figures also illustrate the behavior of φ(0)
s and φs
(1)

rs(0) has noncompact support, while rs(1) maps the entire Kerr by painting the Kerr equatorial plane with a color wheel, as in
equatorial plane into a finite annulus: the n = 1 lensing band. Ref. [61]. This wheel changes colors across “isopolar” curves
of fixed φs , which are drawn at every 45◦ in the source plane.
11

20

10

−10

−20
β (M )

20

10

−10

−20

−20 −10 0 10 20 −20 −10 0 10 20


α (M )

FIG. 6. Same as Fig. 5 for the n = 1 image, which is a demagnified and rotated copy of the direct n = 0 image (shown as a faded background).

Their corresponding images—the curves of fixed φ(n) s —form a As a result, the equatorial plane is unfolded increasingly
swirling pattern that illustrates the effects of frame-dragging. many times in higher lensing bands, which therefore contain
We emphasize that this swirl is a purely geometric effect, and multiple images of a single point source! This surprising and
that gravitational redshift is not yet included at this stage (its intricate lensing behavior is connected to the development of
effects will be described in Sec. III below). caustics [64] and will be explored elsewhere.
Unlike rs(n) , φs(n) has a rather complicated range. While φ(0)
s While the transfer functions rs(n) and φs(n) suffice to describe
spans the entire range [0, 2π) once and only once, so that every the lensing of a static source, one still needs ts(n) to describe a
point source produces precisely one direct image (as in flat time-dependent source and account for its time-delay effects.
spacetime), as n grows large, the maps φ(n) s cover this range an Since light takes forever to reach a distant observer that is truly
increasing (linearly divergent) number of times [45, 62, 63]. at asymptotic infinity, ts(n) is technically infinite when ro → ∞.
12

20

10

−10

−20
0 10 0 10 20
β (M )

−20 −10 −30 −20 −10

20

10

−10

−20
−80 −40 0 10 −100 −50 0 10 20

−20 −10 0 10 20 −20 −10 0 10 20


α (M )

FIG. 7. Same as Fig. 5 for the renormalized time t˜(0) relative to an observer at ro = 10000M [Eq. (41)], with the same isoradial curves (purple).
The solid, dashed, and dotted grey lines respectively denote contours of fixed t˜/M ∈ {−10, 0, 10}. Each panel has its own color scale for clarity.

More precisely, the null geodesic equation in Kerr dictates that Because of these divergences, ts(n) is inherently dependent on
! the observer radius. However, subtracting these (“infrared”)
dt r→∞ 2M 1
≈ 1+ +O 2 . (39) divergences yields a “renormalized” time
dr r r
t˜(n) ≡ ts(n) − ro + 2M log ro ,

Hence, the coordinate time elapsed along a geodesic grows as (41)
ro →∞ which remains finite even as ro → ∞. Moreover, in this limit,
ts(n) ≈ ro + 2M log ro + O (1) , (40)
the dependence on ro drops out, and t˜(n) tends to a constant
with the leading linear divergence arising from the first term in along any outward-propagating light ray. In the limit M → 0,
Eq. (39), and the subleading log-divergence from the second. t˜(0) → ts(0) −ro and t˜s(0) = 0 is the future light cone of the origin.
13

10 At higher spins and inclinations, the range of t˜(0) grows wider


and more skewed toward negative values. Accordingly, we
use a different color scale in each panel. We also display three
−50 −45 −40 −35 −30 −25 −20 −15 −10 contours of fixed time: t˜(0) = −10M (solid), t˜(0) = 0 (dashed),
and t˜(0) = 10M (dotted). These “isochronous” curves may
5
be either open or closed, and in some cases, they may even
be composed of multiple disconnected segments, highlighting
the warped nature of the black hole spacetime.
The range of t˜(n) is always infinite in every lensing band.
β (M )

0 More precisely, it is unbounded below but bounded above: the


reason is that, if the entire equatorial plane emits light forever,
then an observer can see light from arbitrarily far back in the
past (emitted from regions that are either very close to the
−5 horizon or very far behind it) but it does take some minimum
time for any light to reach an observer, even when it is emitted
from the nearest point in the plane. In Fig. 7, we do not see
arbitrarily large negative values of t˜(0) because we have a finite
resolution (and therefore do not have pixels that resolve rays
−10
−10 −5 0 5 10 arbitrarily close to the horizon) and because we cut off the
α (M ) disk at rs = 20M (and therefore do not see rays emitted from
farther behind the black hole).
FIG. 8. Same as Fig. 7 for the n = 1 image. The solid grey line is an The infinite range of t˜(1) is more readily apparent in Fig. 8,
isochronous curve of fixed renormalized time t˜(1) = −10M. The color where we use a different color scale to highlight the divergent
scale is chosen to highlight the rapid variation in t˜(1) at the edges of time elapsed along light rays at the edges of the lensing band:
the lensing band (which correspond to the horizon and infinity). the inner edge consists of rays that asymptote to the horizon
infinitely far back in the past, while the outer edge consists of
10
rays that bounce back towards null infinity, incurring another
time lapse that diverges as Eq. (40). The rapidly changing
colors at the edges of the lensing band illustrate this behavior.
Again, the finite resolution near the inner edge, together with
the cutoff at rs = 20M near the outer edge, preclude us from
5 resolving light rays emitted arbitrarily far back in the past. In
Fig. 8, we chose a color scale spanning 99.9% of the range of
time lapses sampled by the pixels making up the n = 1 grid.
Comparing Figs. 7 and 8, we see that successive images of
β (M )

0 the equatorial plane are delayed by a time of order τ ∼ 15M.


Likewise, comparing the n = 1 image of the equatorial plane
in Fig. 6 to its n = 2 image in Fig. 9 shows that successive
images are not only time-delayed but also demagnified and
−5 rotated. This lensing behavior results in a self-similar photon
ring substructure that is governed by Kerr critical exponents
γ, δ, and τ, which respectively control the demagnification,
rotation, and time delay of successive subring images [5, 7].
−10
−10 −5 0 5 10
α (M )
H. Analytic ray tracing with Beloborodov’s approximation

FIG. 9. Same as Figs. 5 and 6 for the n = 2 image, with the n = 1 Direct (n = 0) light rays experience the smallest deflection.
image shown as a faded background. Despite a significant zoom, the
In Schwarzschild, Beloborodov used an ingenious expansion
extremely narrow n = 2 lensing band is still barely visible.
to derive an excellent analytic expression that approximates
the trajectories of such rays [44]. His small-deflection-angle
expansion is remarkably effective because its leading, linear
In other words, in flat space, a ray emitted radially outwards
term only receives its first subleading correction from a cubic
from rs = 0 at ts = 0 reaches null infinity ro → ∞ with t˜(0) = 0.
term with a small coefficient. To describe this result, we define
We display density plots of t˜(0) and t˜(1) in Figs. 7 and 8,
respectively. In practice, to take the limit ro → ∞, we evaluate q
β tan θo
Eq. (41) at a large value of ro  M. The color scale in Fig. 7 b= α2 + β2 , cos ψ = − p ∈ [−1, 1]. (42)
ensures that t˜(0) = 0 is white and sits in the middle of the scale. b2 + β2 tan2 θo
14

20

10

−10

−20
β (M )

20

10

−10

−20

−20 −10 0 10 20 −20 −10 0 10 20


α (M )

FIG. 10. Apparent positions of source rings of constant Boyer-Lindquist radius rs in the equatorial plane (θs = π/2), for different values of
the spin and inclination within the n = 0 band, using the exact expression [Eq. (33)] and the Beloborodov approximation [Eq. (43)]. The
Beloborodov approximation is still good even for moderate values of the spin and high inclination values.

In terms of these variables, a straightforward manipulation of In practice, the approximation (43) is excellent for rays that
Beloborodov’s formulas leads to the approximate expression never get within ∼ 4M of the black hole—for rays that come
s
!2 closer and experience greater deflection, it remains good along
rs(0) 1 − cos ψ 1 − cos ψ
!
b2 the portion of the ray up to the first equatorial crossing, but
≈ + − , (43) breaks down afterward, once the deflection angle grows large.
M 1 + cos ψ 1 − cos2 ψ 1 + cos ψ
which is highly accurate for low-to-moderate inclinations θo , Intuitively, the reason Eq. (43) holds even for nonzero spins
as shown in Fig. 10. The agreement may seem surprising in is that n = 0 rays are only weakly lensed, and do not spend
the Kerr case, since this formula is derived for Schwarzschild. sufficient time orbiting the black hole to explore its geometry.
15

Typically, n = 0 rays only come close enough to the black hole On the other hand, rays that appear outside the critical curve
to probe the leading, monopole moment of its gravitational do encounter a radial turning point at radius r = r4 , which they
field (the mass), but are largely insensitive to its subleading, reach at Mino time τ = τ4 given in Eq. (30). Hence, for rays
dipole moment (the spin), which contributes small corrections outside of C, ±r depends on whether the Mino time τ(n) s = Gθ
(n)

to the transfer function in an inverse-radius expansion.10 of the (n + 1) equatorial crossing, which is given in Eq. (A1),
th
With a little effort, under the Beloborodov approximation precedes or follows the radial turn at Mino time τ4 ; that is,
for the direct image, one can show that the sign of the quantity  
Outside C : ±r = sign τ4 − G(n) θ . (48)
1
τB = cos ψ − h  i , (44)
3b 1
1− b̃
cos 3 arccos − b̃b When using the Beloborodov approximation from Sec. II H
p √ to ray trace n = 0 images, one merely replaces the exact rs(0)
where b̃ = α̃2 + β̃2 = 3 3M is the Schwarzschild critical in the preceding discussion by its approximation (43), which
impact parameter (apparent image radius of the critical curve), specifies the geodesic momentum at first equatorial crossing
determines whether a ray shot backward from an image-plane up to signs ±r,θ . The angular sign is still ±θ = − sign (cos θo ).
position (α, β) with b > b̃, which must necessarily encounter a For the radial sign ±r , we must again distinguish between rays
radial turning point, does so before (τB > 0) or after (τB < 0) inside and outside C, but this time the relevant critical curve is
first crossing the equatorial plane at radius (43).11 √ of constant radius b = b̃, where
that ofpSchwarzschild: a circle
b = α2 + β2 and b̃ = 3 3M. Rays inside C (with b < b̃)
have ±r = + as always, while rays outside C with (b > b̃) have
I. Photon four-momentum at equatorial crossings
±r = sign τB . (49)
Last, we describe how to compute the source momentum of
a photon that is loaded onto a ray as it crosses the equator. III. EQUATORIAL EMISSION MODEL
The four-momentum p = pµ dxµ of a Kerr photon is

In the previous section, we described how Kerr black holes
!
R(r)
p = E − dt ±r dr ±θ Θ(θ) dθ + λ dφ ,
p
(45) lens light. We will now introduce a simple, phenomenological
∆(r)
model of electromagnetic emission from the equatorial plane
where E = −pt is the photon energy, while R(r) and Θ(θ) and describe how to compute its observational appearance as
are the radial and angular geodesic potentials (9)–(10), which seen by distant observers such as ourselves.
depend on the specific (energy-rescaled) angular momentum
λ and Carter constant η of the photon defined in Eq. (3).
A ray shot backwards from image-plane position (α, β) A. Intensity images
has conserved quantities (λ, η) obtained by inverting Eq. (4).
When such a ray crosses the equatorial plane for the (n + 1)th Our goal is to compute an observed intensity Io at every
time at the radius rs(n) given by Eq. (33), a photon of energy E image-plane position (α, β). For each pixel in the image, we
loaded onto the ray will have a geodesic momentum given by compute Io (α, β) by tracing the corresponding ray backwards
Eq. (45) evaluated at r = rs(n) and θ = π/2 with those values of into the geometry; each time it passes through the emission
(λ, η). This specifies the momentum up to discrete signs ±r,θ . region—in this case, the equatorial plane—we load additional
We now describe how to also determine these two signs. photons onto the ray according to the local source intensity Is .
First, ±θ = sign pθs is trivial to compute, since n = 0 photons Since I/ν3 is the invariant number of photons of frequency ν
must be emitted toward the observer, and so they always have along a ray, it follows that
±θ = − sign (cos θo ). This sign must flip at every crossing, so
N(α,β)
X    
±θ = (−1) n+1
sign (cos θo ) (46) Io (α, β) = ζn g3 rs(n) , α, β Is rs(n) , φ(n)
s , ts
(n)
, (50)
n=0
at the (n + 1)th equatorial crossing. As for ±r = sign prs , two
possibilities arise. Rays that appear inside the critical curve which generalizes to non-stationary and non-axisymmetric
can never encounter a radial turning point and are therefore sources the analogous formula in Ref. [35]. In this expression,
always outgoing; as such, for rays inside of C, we always have ζn is a geometric “fudge” factor and g = νo /νs is the observed
redshift, while the transfer functions rs(n) , φ(n) (n)
s , and ts , which
Inside C : ±r = +. (47) are given in Eqs. (33)–(35), denote the spacetime position of
the ray’s (n + 1)th equatorial crossing, with n ranging from 0
to the total number N of crossings given in Eq. (31).
The factor ζn is meant to account for the (neglected) effects
10 This observation also underlies the “just add one” prescription rs(0) ≈ b− M, of geometrical thickness, which we can mimick with [35]
which gives an excellent approximation to the transfer function for a polar
observer at θo = 0 and for any value of the spin, as first noticed in Ref. [5]

1 n = 0,

ζn =  0 < ζ ≤ 1.

and then derived, along with subleading O (M/b) corrections, in Ref. [45]. (51)
11 Tsupko provides an analytic expression for the shape of n ≥ 2 images [65]. ζ n > 0,

16

Motion type Four-velocity Definition 1. determine the lensing bands as described in Sec. II D;
Circular Keplerian ů App. B 1
Geodesic 2. in each lensing band, define a regular Cartesian grid of
Radial infall ū App. B 2
Sub-Keplerian û App. B 3 appropriate resolution—the resulting grids (indexed by
Non-geodesic n) form the different image layers described in Sec. II F;
General flow ũ App. B 4

TABLE I. Summary of the different accretion flows that we consider. 3. for each pixel in a given layer, trace the corresponding
The circular Keplerian flow ů corresponds to Cunningham’s geodesic ray back into the geometry and compute its equatorial
prescription [67], while ū denotes the radial geodesic inflow. The crossings xs(n) using the transfer functions in Sec. II C;
non-geodesic, sub-Keplerian flow û is the circular motion obtained
by rescaling the Keplerian specific angular momentum `˚ = −ůφ /ůt 4. use the prescribed flow u to compute the redshift (54);
by a factor 0 < ξ < 1. The non-geodesic flow ũ is a general linear
superposition of these purely circular and radial motions. 5. finally, for each pixel in the image, use the equatorial
crossings from 3, the redshift from 4, and the prescribed
source profile Is to compute the observed intensity (50).
The inclusion of this geometric factor significantly improves
the agreement of this simplified equatorial model of emission A key observation is that the source profile Is only enters this
with time-averaged radiative GRMHD simulations [9]. procedure at the very last step. In particular, steps 1 through 3
Finally, to derive the observed redshift g, we must prescribe depend only on the black hole spin and observer inclination,
a four-velocity for the accretion flow of radiating matter. We and can therefore be computed once and for all for each choice
consider flows with a four-velocity of the general form of (a, θo ). The output can then be reused for each new choice
  of accretion flow u or source profile Is , offering a significant
u = ut ∂t − ι ∂r + Ω ∂φ , (52) time advantage (particularly when ray tracing movie frames).
Some comments are in order about step 2. First, while there
where the angular and radial-infall velocities are defined as is no reason to favor regular grids in principle, they do of-
fer computational advantages in practice, since many numer-
uφ ur
Ω= , ι=− , (53) ical algorithms (such as grid interpolation) are optimized for
ut ut such grids. Second, there is no real reason to favor Cartesian
and uµ are the contravariant components of the four-velocity. grids, and AART only relies on them for simplicity—it may be,
In App. B, we introduce two purely circular flows: a geodesic, however, that for extremely high n, regular but non-Cartesian
Keplerian flow ů, together with a non-geodesic, sub-Keplerian grids become more computationally efficient. Third, the exact
flow û obtained by rescaling the Keplerian angular momentum choice of “appropriate resolution” adapted to each grid will be
by a sub-Keplerianity parameter 0 < ξ ≤ 1, such that û → ů as described in detail in Sec. V below.
ξ → 1. We also introduce the four-velocity ū of purely radial
geodesic infall. Then, following Refs. [11, 66], we introduce
a flow ũ given by a linear superposition of purely circular and B. Polarimetric images
purely radial motion. The resulting combined motion, which
we present and derive in detail in App. B, is a non-geodesic In principle, besides the observed intensity (Stokes I), we
family of flows with three parameters: the sub-Keplerianity could also ray trace the observed linear polarization (Stokes
factor 0 < ξ ≤ 1, and two other parameters 0 ≤ βr ≤ 1 and P = Q+iU) to obtain a polarimetric image. In practice, rather
0 ≤ βφ ≤ 1 controlling the radial and angular components of than computing the Stokes parameters Q and U separately, we
the superposition, respectively. For βr = βφ = 0, ũ reduces to will instead simply express the observed linear polarization as
the ξ-independent radial inflow ū. At the other extreme, when
βr = βφ = 1, ũ reduces to the sub-Keplerian flow û, which for Po = mIo e2iχ , (55)
ξ = 1 recovers the circular-equatorial geodesic flow first used
by Cunningham [67]. We summarize these flows in Table I. where 0 ≤ m ≤ 1 is a fractional degree of polarization and χ
Once a specific four-velocity (52) has been prescribed, the denotes the EVPA (5). In a realistic model, m itself would vary
observed redshift is then given by [Eq. (54)] across the image, since the degree of polarization of a photon
usually depends on the angle at which the corresponding light
√ !#−1
ray intersects the local magnetic field at the source.
"
E R(r)
g= = u t
1 ± r ι − λΩ , (54) Here, we will be content to set m to a constant, resulting in
−pµ uµ ∆(r)
sufficiently realistic polarimetric images for our purposes. In
where the conserved quantities (λ, η) of the photon trajectory effect, this amounts to assuming that our astrophysical source
are related to its apparent position (α, β) via Eqs. (4), while emits isotropically, which is a convenient choice that allows us
±r = sign prs denotes the sign of the photon radial momentum to continue treating the emission profile Is as a scalar, rather
at its source whose computation is described in Sec. II I above. than a directed quantity. With this choice (redefining the color
Given a black hole spin a, an observer inclination θo , some scale to absorb the proportionality factor m), the magnitude
prescribed accretion flow u, and an equatorial emission profile |Po | = mIo of the observed linear polarization is identical to
Is , we now have everything needed to compute an observed the image intensity Io , leaving only its direction—the EVPA—
image (50). To summarize, the procedure is the following: as the sole new feature to be displayed in polarimetric images.
17

We indicate the orientation of the plane of polarization using Finally, to prescribe the source polarization fs , we impose
“polarimetric ticks”: these are polarization vectors [53, 68]
  fs · ps = 0, fs · us = 0, fs · Bs = 0. (59)
fo = mIo − sin χ ∂α + cos χ ∂β , (56)
The first condition states that the polarization is perpendicular
where m is chosen to make the ticks legible and χ is computed to the direction of light propagation—this must by definition
from a prescribed source polarization profile (described in the be true everywhere, including at the source. Meanwhile, the
next section) that we parallel transport along light rays using second condition implies that the polarization vector is purely
the conservation of the Penrose-Walker constant κ defined in spatial in the emitter frame, since
Eq. (6).12 The length | fo | = |Po | ∝ Io of the ticks also encodes
f (t) = − f(t) = − fµ eµ(t) = − fs · us , (60)
the magnitude of the polarization, which improves readability.
We present an example of a polarimetric image in Fig. 11. where Latin indices (a) label tensor components in the local
In the remainder of this section, we describe the equatorial orthonormal frame eµ(a) of the emitter, with time leg eµ(t) = uµs ,
emission profiles implemented in AART. The modularity of the while Greek indices µ label spacetime components. Imposing
code makes it easy to include any quantity of interest in the ray this condition does not lead to any loss of generality. Rather,
tracing, but we limit ourselves to Stokes I and P in this paper. it is simply a way to fix gauge: under gauge transformations,
the polarization undergoes a gauge shift fs → fs +cps [53] that
leaves the first condition invariant but shifts fs · us by cps · us .
C. Stationary and axisymmetric source profiles
Thus, setting c = −( fs · us )/( fs · us ) results in a gauge-fixed fs
that obeys the second condition by construction. Lastly, the
State-of-the-art time-averaged GRMHD-simulated images third condition asserts that the polarization is perpendicular to
of realistic sources can be mimicked by ray tracing images the spacetime vector Bs . This is the only physical assumption
of stationary, axisymmetric, equatorial emission profiles [9]. imposed by the relations (59). A suitable choice for modeling
Such models therefore offer a computationally cheap way to synchrotron emission is to let Bs be the local magnetic field.13
study either time-independent or long-time-averaged sources. Together, the three conditions (59) determine three out of
Following [10, 35], we model such sources by setting four components fsµ of the source polarization, which suffices
to fix its spacetime orientation and hence the observed EVPA.
Is (rs , φs , ts ) = J(rs ), (57)
The remaining component essentially controls the magnitude
where we take the radial profile J(r) to be the analytic function of the observed polarization, which in our model is not fixed
at the source, but rather at the observer where we peg it to
r−µ 2
e− 2 [γ+arcsinh( ϑ )]
1
the observed intensity Io . Mathematically, this prescription
JSU (r; µ, ϑ, γ) ≡ q , (58) fixes the Penrose-Walker constant (6) up to an overall scale
(r − µ) + ϑ
2 2 factor that drops out of the formula (5) for the EVPA χ. More
precisely, solving (59) yields, in terms of bµν = Bµs pνs − Bνs pµs ,
which is derived from Johnson’s SU distribution and can be
rapidly computed to arbitrary precision everywhere. The three bθφ − btθ Ω
fr = ft , (61)
parameters µ, ϑ, and γ respectively control the location of the bθφ ι + brφ Ω
profile’s peak, its width, and the profile asymmetry. We invite brφ + btφ ι + btr Ω
the reader to consult Sec. 5 of Ref. [10] for a more in-depth fθ = ft , (62)
bθφ ι + brφ Ω
description of these parameters and their interpretation. brθ + btθ ι
To illustrate the computation of a polarimetric image and fφ = − θφ ft , (63)
the validity of Beloborodov’s approximation in Sec. II H, we b ι + brφ Ω
now consider a black hole of spin a/M = 50% viewed from an where, for clarity, we have temporarily dropped the subscripts
inclination θo = 80◦ . This is the case in Fig. 10 for which the ‘s’. Lowering indices in Eq. (6) and evaluating it at θ = π/2
differences between the exact and approximate rs(0) are largest. then gives the source Penrose-Walker constant κs = κ1 + iκ2 :
We focus on the n = 0 image layer, which we ray trace with    
both the exact transfer function (33) and its approximation rκ1 = a pr fφ − pφ fr − r2 + a2 (pt fr − ft pr ) , (64)
(43). We adopt Cunningham’s prescription u = ů (described  
rκ2 = pθ fφ − pφ fθ − a (pt fθ − pθ ft ) . (65)
in App. B 1) for the accretion flow, resulting in the redshift g̊
given by Eq. (B22) for sources outside the ISCO (B16) and by
Eq. (B29) for sources within. This completes steps 1 through
4 above. In step 5, we set J(rs ) = JSU (rs ) with µ = r− [Eq. (1)], 13 As a technical aside, we emphasize here that the condition fs · us = 0 is also
ϑ = M/4, and γ = −1, so the emission peaks past the horizon. crucial for the following reason. Suppose that we start with a polarization
fs0 which only obeys fs0 · ps = 0, while fs0 · us , 0 and fs0 · Bs , 0. Then
a gauge shift fs0 → fs = fs0 + cps with c = −( fs0 · Bs )/(ps · Bs ) results in a
physically equivalent polarization vector that obeys fs · Bs = 0. Thus, one
12 A slightly more realistic synchrotron emission model would set m ∝ sin2 ζ, can always make the initial polarization perpendicular to the magnetic field
where ζ denotes the angle of emission relative to the local magnetic field, by applying a gauge transformation that leaves the observed polarization
but including this non-isotropy does not produce a large effect [54]. invariant. The gauge-fixing condition in Eq. (59) is therefore essential.
18

10−1
10

5
10−2
β (M )

10−3
−5

−10
Exact Ray Tracing Beloborodov Approximation
10−4
−10 −5 0 5 10 −10 −5 0 5 10
α (M )

FIG. 11. Observed intensity Io (in logarithmic scale) and linear polarization Po corresponding to the stationary and axisymmetric source profile
(58) with parameters µ = r− , ϑ = M/4, and γ = −1. The linear polarization is represented by ticks (56) that are aligned with the plane of
polarization of the light—given by the electric-vector position angle (EVPA) [Eq. (5)]—and whose length encodes the magnitude of Stokes
P. Left: Polarimetric n = 0 image ray traced using the exact analytic transfer functions in Sec. II C. Right: Same image ray traced using the
Beloborodov approximation (43) for the source radius rs(0) . The black hole spin a/M = 50% and observer inclination θo = 80◦ correspond to
the case from Fig. 10 with the largest difference between the exact and approximate isoradial curves, yet the resulting images appear almost
identical. Here, the accretion flow follows the Cunningham prescription [67] corresponding to geodesic circular motion, i.e., u = ů in App. B 1,
and we took the source magnetic field to have purely radial spacetime components, i.e., only Brs is nonzero.

The computation of the photon momentum ps at the source D. Modeling variable sources with Gaussian random fields
is described in Sec. II I, while the parameters (ι, Ω) of the flow
ů are given in App. B 1. The component ft that controls the Variable (non-stationary and non-axisymmetric) sources
magnitude of the polarization remains undetermined, but it can broadly be divided into two classes. On the one hand, one
factors out of both κ1 and κ2 , and therefore drops out of the can consider specific physical phenomena that are governed
ratio in the formula (5) for the EVPA χ, as mentioned above. by deterministic equations. A commonly studied example is
The resulting explicit expressions are implemented in AART. that of a “hot spot”: a localized source of enhanced emissivity
All that is left to specify are the spacetime components Bµs that usually orbits around the black hole and thereby produces
of the local magnetic field. The decomposition of the electro- a characteristic pattern of light echoes [see, e.g., 62, 69–73].
magnetic field strength Fµν into electric and magnetic fields is Adding such a source to our model is straightforward in AART:
frame-dependent. In a realistic synchrotron emission model, one merely needs to insert the deterministic source to the RHS
the polarization would be perpendicular to the local magnetic of Eq. (57) before carrying out the usual ray tracing.
field in the emitter frame, with components Bµs = −?F µν uµ . On the other hand, one can also consider sources subject to
For our example, we simply choose the spacetime field Bs statistical fluctuations. Hot spots are transient and unlikely to
to be purely radial, keeping only a nonzero Brs component be present in any given observation of M87*, but we certainly
whose magnitude also scales out of the EVPA χ. We display expect its surrounding plasma to flare and produce emission
the resulting n = 0 images, ray traced using both the exact ropes (or other photon ring mimickers), which in the absence
and approximate rs(0) , in the left and right panels of Fig. 11, of a definite physical model we can still represent as random
respectively. Even though the corresponding isoradial curves fluctuations in the astrophysical source. A widely used tool to
(plotted in the right column of Fig. 10) differ noticeably, the model such astrophysical noise is the Gaussian random field
ray-traced images in Fig. 11 nonetheless look very similar and (GRF); see, e.g., Ref. [74] for applications to cosmology. A
would be indistinguishable with near-term observations. The random field G is a function on a space X such that G(x) is
Beloborodov approximation is thus excellent for ray tracing a random variable for every x ∈ X. A GRF is a random field
direct images of axisymmetric source configurations. that is completely determined by its mean µ and covariance C,
Having showcased the polarimetric capabilities of AART,
we now leave polarization aside and turn to variable sources. µ(x) = hG(x)i , C(∆x) = hG(x)G(x + ∆x)i ≥ 0. (66)
19

More precisely, we define a random field G(x) with mean µ(x) Similarly, higher-point functions of G̃(k) may also be obtained
and covariance C(∆x) to be a GRF if it satisfies the property by Fourier transforming those of G(x). One finds that G̃(k) is
a GRF obeying Eq. (67) with mean µ̂(k) and covariance (74),
ei `=1 s` G(x` ) = e− 2 `, j=1 C(∆x` j )s` s j +i `=1 µ(x` )s` ,
D Pk E 1 Pk Pk
(67) which is also referred to as the power spectrum P(k) of G(x).
When G(x) is also isotropic, so is P(k) = P(k) with k = |k|.
which states that its joint probability distribution on any set of
Like covariances, power spectra are always non-negative.
k points x` , with ` ∈ {1, . . . , k}, is a k-dimensional multivariate
The most ubiquitous GRF is the white noise process W(x).
Gaussian distribution with mean vector µ = {µ1 , . . . , µk } and
It is the standard, homogeneous and isotropic GRF defined by
covariance matrix C` j = C(∆x` j ), where ∆x` j = x` − x j .
a flat power spectrum with uniform C̃ (k0 ) = 1, so that
The following discussion is technical and readers interested
only in applications may skip down to Sec. III H below. D E
W̃∗ (k)W̃ k0 = (2π)d δ(d) k − k0 ,

(75)
The mean and covariance are also known as the one-point
and two-point (or autocorrelation) function, respectively. As a W(x)W x = δ x − x .
0 (d) 0
(76)
result of the defining relation (67), these low-point correlation
functions determine all higher k-point functions of a GRF via With Eq. (67), this delta-function autocorrelation implies that
Wick’s theorem (also known as the Isserlis theorem), which W(x) consists of independent Gaussian random variables at
may be derived by differentiation with respect to multiple s` . every x ∈ Rd , and likewise in momentum space for W̃ (k).
For instance, letting G` denote G(x` ) for a zero-mean GRF, Thus, creating realizations of white noise is trivial: it suffices
it is immediately seen by differentiating Eq. (67) with respect to draw from independent normal distributions at every point.
to s1 , s2 , and s3 that G(x) must have a vanishing three-point In fact, Eq. (74) provides a straightforward way to produce
function, hG1 G2 G3 i = 0. An additional derivative with respect realizations of any homogeneous GRF G(x): first, one creates
to s4 implies that the four-point function hG1 G2 G3 G4 i is a realization W̃ (k) of white noise; next,pone multiplies it by
p
C̃(k) to obtain a realization of Ĝ(k) = C̃(k)W̃ (k); finally,
hG1 G2 i hG3 G4 i + hG1 G3 i hG2 G4 i + hG1 G4 i hG2 G3 i . (68) inverse Fourier transforming Ĝ(k) yields a realization of G(x).
In general, for a zero-mean GRF, all k-point functions vanish (We note that this procedure is well-defined since C̃(k) ≥ 0.)
for k odd, whereas for k even they are givenD by allE the possible Just as a standard Gaussian distribution has zero mean and
unit variance, we define a standard GRF as a zero-mean GRF
“contractions” of the two-point functions G` G j = C(∆x` j ).
with unit covariance at the origin: µ(x) = 0 and C(0) = 1.
This is Wick’s theorem in a nutshell, and it makes precise the We now consider the Matérn field Fν (x) of order ν, which
idea that a GRF is a “simple” random field whose correlation is another standard, homogeneous, and isotropic GRF that is
structure is fully determined by two functions µ(x) and C(∆x). well-known to statisticians, who use it to model a wide range
This property is very convenient for our modeling purposes, of processes; see, e.g., Ref. [75] for its various applications.
since only two functions need to be specified to prescribe all of This zero-mean GRF obeys the Matérn covariance of order ν,
the statistics of our astrophysical source. For example, Wick’s
theorem also fixes all the even moments of a zero-mean GRF: 1  x ν  x 
Cν (x) = ν−1 Kν , (77)
2 Γ(ν) λ λ
D E D E
G2k (x) = (2k − 1)!! [C(0)]k , G2k+1 (x) = 0, (69)
where the k = 2 case is compatible with Eq. (68) evaluated at where Γ denotes the Gamma function and Kν (x) the modified
coincident points. (Of course, the odd moments all vanish.) Bessel function of the second kind of order ν. Here, λ is a
We now specialize to GRFs in d-dimensional Euclidean correlation length, while ν is a differentiability parameter that
space X = Rd . A GRF is homogeneous if it is invariant under we will always take to be ν = n−d/2 for some positive integer
xλ
translations, so G(x+a) = G(x) for a ∈ Rd , and it is isotropic if n. At short distances, Cν (x) ≈ 1 + cν x2ν for some constant
it looks the same in all directions, so G(x) = G(x) with x = |x|. cν , so it is best (though not strictly necessary) to require ν > 0.
A real-space GRF G(x) defines a (generally complex) GRF The Matérn field Fν (x) is also ubiquitous in physics: though
G̃(k) in momentum space via the Fourier transform its position-space covariance (77) may seem unfamiliar, in
Z Z momentum space it becomes14 (see App. C for a derivation)
G̃(k) ik·x d
G̃(k) = G(x)e−ik·x dd x, G(x) = e d x. (70)  
d Γ ν + 2
d
(2π)d N 2 λd
C̃ν (k) = ν+ d2 , N = (4π)
2 2 , (78)
If G(x) is homogeneous, then we let x0 = x + ∆x and Fourier 1+λ k 2 2 Γ(ν)
transform its autocorrelation to derive the covariance of G̃(k):
E " which for n = 1 (or ν = 1 − d/2), we recognize as the quantum
propagator for a free scalar field of mass m = λ−1 in Euclidean
0 0
D
G̃ (k)G̃ k =
∗ 0
G(x)G x0 eik·x−ik ·x dd x dd x0 (71)

" signature. To the best of our knowledge, this link has not been
=
0 0
C(∆x)ei(k−k )·x e−ik ·∆x dd x dd ∆x (72) stated explicitly before, and we explore it in detail in App. C.
Z Z
i(k−k0 )·x d 0
= e d x C(∆x)e−ik ·∆x dd ∆x (73)
14 Equivalently, C̃ν (k) = σ−2
with σ2 = λ2ν Γ(ν)  as in Ref. [76].
= (2π)d δ(d) k − k0 C̃ k0 .
 
(74) ν+ d2 d
(4π) 2 Γ ν+ 2
d
(k2 +λ−2 )
20

The Matérn field Fν (x) enjoys a special connection to linear However, if the constants λ` are unequal, then the resulting
stochastic partial differential equations (SPDE). Realizations GRF remains homogeneous but becomes anisotropic along a
of its Fourier transform F˜ν (k) are related to white noise via direction set by the unit vectors u` . An example with d = 2 is
q displayed in Fig. 1 of Ref. [43]. Following the steps that led
F˜ν (k) = C̃ν (k)W̃(k). (79) to Eq. (74), the corresponding momentum-space covariance is

Such a relation holds for any homogeneous GRF, but it takes N 2 |Λ|
C̃ν (k) = . (86)
(1 + k · Λk)ν+ 2
d
a particularly simple form for the Matérn covariance (78):
 ν+d d
1 + λ2 k2 2 4 F˜ν (k) = Nλ 2 W̃(k). (80) In this case, we can still use Eq. (79) to obtain realizations of
the anisotropic Matérn field. We can also follow the derivation
Returning to position space gives the linear (fractional) SPDE of Eq. (81) to find that the anisotropic field obeys the SPDE
 2ν + d4 ν
(1 − ∇ · Λ∇) 2 + 4 Fν (x) = N |Λ| 4 W(x).
 d d 1
1 − λ2 ∇2 Fν (x) = Nλ 2 W(x), (81) (87)

which exactly matches Eq. (2) of Ref. [76] since N 2 σ2 = λ2ν . Finally, to include inhomogeneity, we allow the unit vectors
As we show in App. C, for n = 1 and ν = 1−d/2, this SPDE is u` → u` (x) to become functions. Then Λ → Λ(x) becomes
compatible with the identification Fν (x) ≡ Φ(x), where Φ(x) a function too—with the same definition (83)—but the rest of
is a free Euclidean scalar field of mass m = λ−1 , as expected. the previous discussion breaks down. Instead, we must reverse
If instead, n = 2 and ν = 2 − d/2, then the resulting Matérn the logic and define the inhomogeneous, anisotropic Matérn
field, which we will simply denote F (x), obeys a linear SPDE field Fν (x) by the SPDE (87) with variable coefficients Λ(x).
We emphasize that the resulting field Fν (x) no longer obeys
  d
1 − λ2 ∇2 F (x) = Nλ 2 W(x), (82) the covariance (86), which is not a function on momentum
space once Λ(x) is position-dependent, but it is a well-defined
which has no fractional derivative and is thus simpler to solve GRF arising as a solution to the SPDE (87); see also Ref. [77].
numerically, as explained in Ref. [43] below Eq. (3) therein. As explained below Eq. (81), in practice, we will only use
This connection between the Matérn field F (x) and a linear the field F (x) with n = 2 and ν = 2−d/2 as it obeys the SPDE
SPDE is especially useful once we introduce inhomogeneities. 1
Inhomogeneous GRFs are hard to generate via other means [1 − ∇ · Λ(x)∇] F (x) = N |Λ(x)| 4 W(x), (88)
than solving the associated SPDE; in particular, the trick (79)
for generating realizations of F˜ (k) breaks down because the which has no fractional derivatives and is therefore more
Fourier modes are no longer delta-correlated as in Eq. (74). tractable numerically. Since we took our white noise to have
unit variance, this SPDE exactly matches Eq. (5) of Ref. [43]
(after setting σ = 1 therein), which is what inoisy solves.
E. Inhomogeneous and anisotropic Matérn fields One last comment is in order: in the inhomogeneous case,
the Matérn field F (x) that inoisy produces by numerically
Before describing how to include inhomogeneity, we first integrating Eq. (88) does not follow the covariance (85); in
discuss how to model homogeneous anisotropies, following particular, it may not be standard with µ(x) = 0 and C(0) = 1.
closely the excellent treatment in Ref. [43]. To “standardize” it, we will always work with a rescaled field
Still in d dimensions, we introduce d orthonormal vectors
F (x) − hF (x)i
u` , d correlation lengths λ` > 0, with ` ∈ {1, . . . , d}, and define Fˆ (x) = q , (89)
d d F 2 (x) − hF (x)i2
X Y
Λ= λ2` u` uT` , |Λ| ≡ det Λ = λ2` . (83)
`=1 `=1
which also happens to be precisely what inoisy implements.
For a homogeneous field, this makes no statistical difference.
We will also allow the unit vectors u` to not be orthogonal, as
long as |Λ| retains the same form. The matrix Λ is invertible,
and its inverse Λ−1 defines a metric on Rd with line element F. Non-stationary and non-axisymmetric source profiles
d !2
∆x · u`
With the inhomogenous, anisotropic Matérn field Fˆ (x) in
X
ds (∆x) = ∆x · Λ ∆x =
2 −1
. (84)
`=1
λ` hand, we are now in a position to model the fluctuations of an
equatorial accretion disk and produce inoisy simulations.
We now use this to define the generalized Matérn covariance Before doing so, we first describe what we want to achieve
1 intuitively, without being too rigorous. We let xs = (rs , φs , ts ).
Cν (x) = dsν (x)Kν (ds(x)) . (85) As in Ref. [8], we want the total intensity (57) to consist of a
2ν−1 Γ(ν)
background radial profile J(rs ) with some fluctuations ∆J(xs ):
When λ1 = . . . = λn = λ, so that Λ = λ2 I, this reproduces the
homogeneous, isotropic Matérn covariance (77), as desired. Is (xs ) = J(rs ) + ∆J(xs ) = J(rs ) [1 + F(xs )] . (90)
21

Here, F ≡ ∆J/J is the fractional variation in the source, which Evaluating the sum results in
at first we assume to be fractionally small: 0 < |F(xs )|  1.
hJ(xs )i  J(rs )e 2 σ ,
1 2
We also want these fluctuations to wash out under averaging, (98)
hF(xs )i = 0, (91) and we now see why the definition (93) is not what we wanted,
so that after averaging over fluctuations using hJ(rs )i = J(rs ), since it does not satisfy the requirement (92). Instead, we let
we recover the stationary, axisymmetric profile (57):
Is (xs ) = J(xs ) ≡ J(rs )eσF (xs )− 2 σ ,
ˆ 1 2
(99)
hIs (xs )i = hJ(rs )i + hJ(rs )F(xs )i = J(rs ). (92)
which does have the desired property (92), and hence recovers
Eventually, we will also want to consider large fluctuations the background profile (57) after averaging. We note that this
with ∆J ∼ J, or F ∼ O(1), but this could pose a problem: can also be checked directly from the property (67) with k = 1,
if at some point xs , a fluctuation grew so large and negative
that F(xs ) < −1, then we would have ∆J(xs ) < −J(rs ), and 1 2
D E
eisG(x) = e− 2 C(0)s , (100)
hence Is (xs ) < 0 at that point, which is nonsensical because
an intensity is a (necessarily positive) count of photons—we by taking s = −iσ and G = Fˆ with µ = 0 and C(0) = 1. This
must therefore prevent this from happening. also shows that σ2 effectively rescales the covariance C(0). In
Of course, we could demand that the fluctuations always particular, the fluctuations disappear as s → 0 and we recover
remain small, but there is a way to allow them to grow large the pure envelope J(rs ) of the background surface brightness.
while still ensuring they do not result in a negative intensity.
The idea is to note that for small fluctuations 0 < |F(xs )|  1,
1 + F ≈ 1 + F + 2!1 F 2 + 3!1 F 3 + . . . = eF , so we may well replace G. Complete statistics of the variable source
the RHS of Eq. (90) by a (necessarily positive) exponential:
Is (xs )  J(rs )eF(xs ) . (93) We can also use the GRF property to explicitly write down
the complete correlation structure of the random field J(xs ).
For small fluctuations, this seems completely equivalent, since With s` = −iσ, G = Fˆ , µ = 0 and C(0) = 1, Eq. (67) becomes
∆J(xs ) Is (xs ) − J(rs ) |F|1
=  eF(xs ) − 1 ≈ F(xs ), (94) eσ `=1 F (x` ) = e 2 σ `, j=1 C(∆x` j ) .
D Pk ˆ E 1 2 Pk
(101)
J(rs ) J(rs )
reproducing Eq. (90). The advantage of this definition is that it
Since hJ(x1 ) . . . J(xk )i = J(r1 ) . . . J(rk ) eσ `=1 F (x` )− 2 σ ,
D Pk ˆ k 2
E
can now be freely extended to arbitrarily large |F|, even in the
negative direction: a fluctuation F(xs )  −1 would just result  k

in a small but still positive (and hence still physical) intensity, hJ(x1 ) . . . J(xk )i  kσ2 σ2 X 
= exp − + C(∆x` j ) . (102)
rather than an unphysical, negative emissivity. Hence, our new J(r1 ) . . . J(rk ) 2 2 `, j=1
definition (93) for Is (xs ) remains sensible even for large and
negative field excursions, unlike our first attempt (90). This result characterizes all the statistics of the variable source
However, as we are about to see, this new definition suffers J(xs ) in terms of a single covariance function, as advertised.
from a subtle issue: it does not actually satisfy Eq. (92). For Of course, this formula with k = 1 confirms that Eq. (92)
P Meanwhile, for k = 2 the double sum in the exponential
now, we use ‘’ to remind us of this, and proceed naively. holds.
The preceding intuitive discussion hinges on being able to is 2`, j=1 C(∆x` j ) = C(∆x11 ) + C(∆x12 ) + C(∆x21 ) + C(∆x22 )
produce a fluctuation field F(xs ) with some desired properties, so we obtain the two-point function
such as Eq. (91). We now make this mathematically precise.
We replace F(xs ) → σFˆ (xs ), where Fˆ (xs ) is the standard, hJ(x1 )J(x2 )i 2 σ2
= e−σ + 2 [2C(0)+2C(∆x12 )] = eσ C(∆x12 ) ,
2
(103)
zero-mean Matérn field defined in Eq. (89) as the solution to J(r1 )J(r2 )
the SPDE (87). We then define an emissivity field [43]
which is the exponential of the Matérn field’s autocorrelation.
Is (xs ) ≡ J(xs )  J(rs )eσF (xs ) ,
ˆ
(95) Going further, for k = 3 the double sum in the exponential is
`, j=1 C(∆x` j ) = 3C(0) + 2C(∆x12 ) + 2C(∆x13 ) + 2C(∆x23 ) so
P2
where the parameter σ controls the scale of the fluctuations,
since Fˆ (xs ) has unit covariance C(0) = 1. Because Fˆ (xs ) has hJ(x1 )J(x2 )J(x3 )i
= eσ [C(∆x12 )+C(∆x13 )+C(∆x23 )] ,
2
zero mean, Eq. (91) is satisfied. Let us now check Eq. (92): (104)
J(r1 )J(r2 )J(r3 )

X σn Fˆ n (xs )
hJ(xs )i  J(rs ) eσF (xs ) = J(rs )
D ˆ E
, (96) and one can check that J(xs ) is a non-Gaussian random field.
n=0
n! That is, it does not quite obey the GRF property (67), though
its statistics (102) are still governed by a single covariance.
where we series-expanded the exponential and used linearity For some intuition, let us reexamine the fractional variation
of the expectation value. Then by Eq. (69), relabeling k = 2n, F ≡ ∆J/J, whose statistics are encoded in the random field
∞ ∞
hJ(xs )i X σ2k Fˆ 2k (xs ) X (2k − 1)!! 2k J(xs ) − J(rs )
 = σ . (97) Ĵ(xs ) ≡ = eσF (xs )− 2 σ − 1.
ˆ 1 2
(105)
J(rs ) k=0
(2k)! k=0
(2k)! J(rs )
22

Its one-point and two-point functions are Parameter Default value Description
ψ 1.07 BH mass-to-distance ratio
Ĵ(xs )Ĵ x0s = eσ C(∆x) − 1.
2
D E D E
Ĵ(xs ) = 0, (106) a/M 94% BH spin
θo 17◦ Observer inclination
If the fluctuations are very small, then this is approximately ζ 1.5 Geometrical factor
σ 0.4 Fluctuation scale
D E σ2 1 2 ξ
Ĵ(xs )Ĵ x0s ≈ σ C(∆x), (107) 0.95 Sub-Keplerian factor
βr 0.95 Radial velocity factor
which is the autocorrelation of a GRF with covariance σ2 . βφ 0.95 Angular velocity factor
θ∠ 20◦ Anisotropy direction
This is also true for higher-point functions, using Eq. (102),
λ0 2π/Ω Temporal correlation
we can show that the field Ĵ(xs ) is approximately Gaussian at λ1 5rs Spatial correlation in the xs direction
leading order in σ. Thus, small fluctuations are Gaussian, but λ2 0.1λ1 Spatial correlation in the ys direction
larger fluctuations introduce non-Gaussianities. µ r− Controls location of profile peak
ϑ 1/2 Controls profile width
γ −3/2 Controls profile asymmetry
H. Generating realizations of an accretion flow using inoisy
TABLE II. Summary of the fifteen parameters in our non-stationary
To summarize, we model variable sources via the emissivity and non-axisymmetric stochastic source model, together with their
default values used for the majority of our examples. The first three
(99), which is defined in terms of the Matérn field (89) and
pertain to the geometry, while the other twelve prescribe the statistics
displays the correlation structure (102). The zero-mean field of the equatorial surface brightness.
Fˆ (xs ) and its statistics are fully determined by its covariance,
which is controlled by a metric Λ(xs ) of the form (83). Given
such a metric, a realization of Fˆ (xs ) is obtained by solving the We note that u1 (xs ) · u2 (xs ) = 0 are always orthogonal to
linear SPDE (88) with some realization W(xs ) of white noise. each other, but not to u0 (xs ). Nonetheless, the resulting Λ(xs )
This is essentially what inoisy does, though in practice, the still has a determinant of the form required by Eq. (83), since
continuous SPDE (88) is actually discretized and its solution |Λ(xs )| = λ20 λ21 λ22 . Also, the sign of u1 and u2 is arbitrary, as it
is a Gaussian Markov random field approximating Fˆ (xs ) [43]. sets only the spatial correlation; the flow in time is set by u0 .
As for the envelope J(rs ), we use the radial profile (58). To be consistent with our choice of accretion flow (52), we
Therefore, all we need to do to specify our model is to fix a must take the temporal correlations to have the same velocity:
choice of unit vectors u` (xs ) and correlation lengths λ` , where
now ` ∈ {0, 1, 2}. Such a choice will define Λ(xs ) and hence dr dφ
~v ≡ ∂r + ∂φ = −ι ∂r + Ω ∂φ , (112)
our non-stationary, non-axisymetric stochastic source profile. dt dt
Since we want to produce “realistic” images and visibility where Ω and ι respectively denote the angular and radial-infall
amplitudes, we will choose Fˆ (xs ) to be an inhomogeneous, velocities (53). In Cartesian coordinates this velocity becomes
anisotropic Matérn field whose power spectrum is comparable ~v = v x ∂ x + vy ∂y with
to that observed in GRMHD simulations [43], and therefore
xs ys
serves as a good phenomenological model. v x (xs ) = − ι − ys Ω, vy (xs ) = − ι + xs Ω. (113)
Instead of xs = (rs , φs , ts ), we use a regular Cartesian grid rs rs
xs = (ts , xs , ys ) in the Kerr equatorial plane, and inoisy solves As for the spatial correlations, we adopt the prescription [43]
the SPDE (88) on this grid with periodic boundary conditions.
Following Eq. (83), we pick a position-dependent anisotropy θ(xs ) = arctan(ys , −xs ) + θ∠ . (114)

2 This sets the major axis u1 (xs ) of the spatial correlation tensor
to lie at a constant angle θ∠ relative to the equatorial circles of
X
Λ(xs ) = λ2` (rs )u` (xs )uT` (xs ). (108)
`=0 constant radius rs . The resulting flow displays spiral features
with opening angle θ∠ . The choice θ∠ ≈ 20◦ produces spiral
Under this prescription, u0 (xs ) sets the temporal correlation arms broadly consistent with GRMHD simulations [43, 78].
of the flow, with characteristic correlation time λ0 (rs ), while This completes the specification of our stochastic surface
u1 (xs ) and u2 (xs ) determine its spatial structure, which at any brightness. To summarize, the accretion flow model takes as
given time exhibits correlations of characteristic lengths λ1 (rs ) input twelve parameters, each of which is listed in Table II,
and λ2 (rs ), respectively. Following Ref. [43], we take these 3D along with a short description and the default value used for
unit vectors to have (ts , xs , ys ) components the main examples in this paper. To compute images, we must
  also specify three more parameters: the black hole spin a, the
u0 (xs ) = 1, v x (xs ), vy (xs ) , (109) observer inclination θo , and the mass-to-distance ratio
u1 (xs ) = (0, cos θ(xs ), sin θ(xs )) , (110) 1 M
u2 (xs ) = (0, − sin θ(xs ), cos θ(xs )) , (111) ψ= , (115)
3.62 µas ro
where we have yet to specify v x , vy , and θ. In practice, we will which combines the black hole mass M and observer distance
only let these functions depend on the spatial position (xs , ys ). ro , and which is defined relative to a fiducial value for M87*.
23

10

n=0 n=1
5

β (M )
0
ys

−5

−10
n=2

−10 −5 0 5 10 −10 −5 0 5 10
xs α (M )

FIG. 12. Left: Intensity profile (in logarithmic scale) of a single snapshot from an inoisy simulation with the parameters in Table. II. Right:
Ray-traced image corresponding to this equatorial source profile, treating the realization of the random field on the left as static (the “fast-light”
approximation). The inset panels decompose the image into layers: the direct n = 0 image and the first two (n = 1 and n = 2) photon rings.

IV. APPLICATIONS

inoisy Having fully described our stochastic source model, we can


1.0 inoisy + AART now simulate it with inoisy and then use AART to ray trace its
appearance and compute its visibility. We will present a whole
suite of simulations in a follow-up paper, but for now we limit
0.8 ourselves to one complete example from which Figs. 1, 2, and
Normalized Light Curve

3 are derived. We first explain in detail how to obtain Figs. 1


and 2, while Fig. 3 will be the focus of Sec. VI below.
0.6
For this simulation, we ran inoisy on a regular Cartesian
grid (xs , ys , ts ) of size 2048 × 2048 × 512. For each of the
spatial coordinates (xs , ys ), we uniformly placed 2048 pixels
0.4
within the range [−50, 50]M, resulting in a spacing of 0.05M
between pixels, whereas for the time coordinate ts , we placed
512 pixels uniformly within the range [0, 1000]M, resulting
0.2 in a cadence of one frame every 1.95M. We used the default
values listed in Table II for the fifteen parameters in the model.

0.0 The left panel in Fig. 12 shows an example of a snapshot


0 200 400 600 800 1000 extracted from the resulting inoisy run. The accretion flow
Time (M ) appears qualitatively similar to GRMHD-simulated flows. In
particular, we tuned the scale σ of fluctuations to ensure that
FIG. 13. Normalized light curves of the stochastic inoisy source the total observed flux (the blue light curve in Fig. 13) varies
with parameters in Table II. Green: total emitted flux measured from by a factor of ∼ 3 between its minimum and maximum. This
inoisy snapshots. Blue: total observed flux measured from ray- level of time-variability mimics the light curves obtained from
traced images. Relativistic effects (gravitational redshift, Doppler GRMHD simulations: see, for example, Figs. 3 and 4 of
boosting, and light bending) make the observed flux smoother and Ref. [79], Fig. 8 of Ref. [80] or Fig. 5 of Ref. [31]. Ensuring
less variable than the emitted flux. The vertical dotted lines indicate that this agreement also holds quantitatively is an interesting
the times corresponding to the four snapshots displayed in Fig. 1. challenge that we hope to tackle in the future.
24

A. Fast-light: stationary non-axisymmetric source We simply use Eq. (50) and plug in all the transfer functions
defined in Sec. II C. This requires us to use the entire series of
After producing our inoisy simulation, we use AART to ray snapshots generated by inoisy; that is, we must now use the
trace it. A commonly adopted simplification in carrying out full time evolution of the accretion flow.
ray tracing is the so-called “fast-light” approximation. When Slow-light tracing does introduce two new complications.
staring at a variable source, a single frame Io (to , α, β) of its First, since the pixels in a single frame of a movie of the source
observational appearance at a fixed moment in time is formed can depend on a very wide range of emission times ts(n) (α, β),
by photons that were emitted at different times ts(n) (α, β) in the in order to produce a movie of some duration T o , we typically
history of the source. The reason is that photons that appear need to simulate the source over a longer duration T s > T o . In
at different positions (α, β) in the image plane follow different practice, we can estimate how much longer T s needs to be by
paths in the geometry, thereby incurring different time delays examining isochronal curves in the n = 0 layer, such as those
∆t = to − ts(n) on their way from source to observer, as shown displayed in Fig. 7. At low inclinations, the range of ts(0) (α, β)
in Figs. 7 and 8. The fast-light approximation simply ignores over sampled pixels is about 50M, but this range grows much
this variable time delay and maps the image plane (α, β) onto a larger at higher inclinations. In addition, the higher-n images
single snapshot of the equatorial plane, using only the transfer are composed of photons emitted roughly a time nτ earlier, so
functions rs(n) and φ(n) in order to ray trace a movie containing n layers, we expect to
s from Sec. II C and replacing Eq. (50) by
need T s −T o & 50M +nτ ≈ 80M. Some rays may occasionally
N(α,β)
X     require sampling the source even earlier than the start of the
Io (α, β) = ζn g3 rs(n) , α, β Is rs(n) , φ(n)
s . (116) simulation, but this is not really an issue with inoisy thanks
n=0 to its use of periodic boundary conditions (including in time).
Second, because ts(n) (α, β) varies smoothly across the image
Mathematically, this is equivalent to keeping the emission plane, the code must be able to sample the source Is (rs , φs , ts )
time ts fixed while letting the observation time to vary across for continuous values of ts , including at times in between the
the image. Naturally, if the source is stationary, then this is a frames computed by inoisy. This problem can be dealt with
moot distinction since the time-dependence drops out anyway, using interpolation, but only so long as the underlying inoisy
so this amounts to treating each individual inoisy snapshot simulation has sufficiently high resolution to smoothly resolve
as a stationary (though non-axisymmetric) source. Thus, the the flow’s motion. This is the case in our example simulation,
fast-light approximation is exact for a stationary source, and it as the inoisy movie of the equatorial source has high enough
can offer a decent approximation as long as the source varies cadence to be smoothly interpolated in time.
slowly relative to the “fast” light being traced, hence the name. Thus, AART readily produces a movie of the source (whose
As an example, we take the first (ts = 0) snapshot from our parameters are listed in Table II). We display four snapshots
inoisy simulation and ray trace it as described in Sec. II C, from this movie, taken at intervals of 250M, in the top row
assuming a black hole spin of a/M = 94% and an observer of Fig. 1. We also plot its light curves in Fig. 13: the total
inclination of θo = 17◦ . In the n = 0 layer, we ray trace at the observed flux (measured in the image plane) and total emitted
same spatial resolution as the underlying inoisy simulation; flux (measured in the equatorial plane) are shown in blue and
that is, we use a spacing of δx(0) = δxinoisy = 0.05M between green, respectively, with the former appearing smoother and
grid pixels. In the higher-n layers, we adaptively increase the less variable than the latter due to relativistic effects.
resolution by a factor of two in each successive band, and ray As in the stationary (fast-light) case, the snapshot images
trace on grids with spacings of δx(n) = 2n δx(0) between pixels. in Fig. 1 present bright transient features that are qualitatively
The right panel of Fig. 12 displays the resulting image (in a similar to those seen in state-of-the-art snapshots of GRMHD
logarithmic color scale to highlight the accretion flow), which simulations [9, 31]. These transient features can obscure the
looks qualitatively similar to the single snapshots produced photon ring in instantaneous snapshots, but they wash out of
with state-of-the-art GRMHD simulations [9, 31]. The figure the average over 100 snapshots in Fig. 2, leaving the photon
also displays a decomposition of the image into its n ∈ {0, 1, 2} ring as the only prominent feature in the time-averaged image
layers, which are shown individually in the inset panels. Each shown in the left panel. This image can be directly compared
separate layer appears well-resolved, suggesting that our grid the one shown in Fig. 1 of Ref. [7], which was also obtained
resolutions are sufficient to resolve the photon rings. Indeed, by time-averaging over 100 uniformly spaced snapshots taken
we have verified (using a convergence test detailed in Sec. V) from a GRMHD simulation with a time range of 1000M.
that the results derived from the observable of interest (which
in our case is the visibility amplitude, presented in Sec. IV C
below) do not change when the resolutions are doubled. C. Visibility on long baselines

A radio interferometer like the EHT samples a (complex)


B. Slow-light: non-stationary non-axisymmetric source radio visibility V(u). By the van Cittert-Zernike theorem, this
is related to a sky brightness Io (xo ) via a 2D Fourier transform
With the transfer functions (33)–(35) implemented in AART, Z
it is also easy to perform a “slow-light” ray-tracing that takes V(u) = Io (xo )e−2πiu·xo d2 xo . (117)
into account varying photon emission times across the image.
25

Here, xo are dimensionless image-plane coordinates measured to compute Vn (u, ϕ) along slices of fixed polar angle ϕ in the
in radians, such as (α, β)/ro , while the dimensionless vector u Fourier plane. The procedure is as follows. For each angle ϕ,
is a “baseline” projected onto the plane perpendicular to the we first compute the Radon transform along the cut at angle ϕ
line of sight and measured in units of the observation wave- across the image; that is, in each lensing band, we integrate the
length [68]. We use radio-astronomy conventions to make this observed intensity In along lines perpendicular to the slice of
Fourier transform 2π-symmetric, unlike the definition (70). constant ϕ. Then, we interpolate each Radon transform to the
Therefore, to relate simulated images to actual observables, resolution of the highest-order lensing band. Lastly, we sum
we must Fourier transform them to compute their visibility. In up the contributions from all the image layers In , and perform
particular, the narrow image features—like the photon rings— a one-dimensional Fourier transform to finally obtain V(u, ϕ).
that encode precise information about the spacetime dominate The bottom row of Fig. 1 presents the visibility amplitudes
the interferometric signal on very long baselines u = |u|  1 |V(u, ϕ)| along slices of constant ϕ = 0◦ (in red) and ϕ = 90◦
[7, 35], requiring us to compute Fourier transforms up to very (in blue) of the corresponding snapshots in the top row. Since
high frequencies. To do so, we again exploit the layered image these individual snapshots display a strong dependence on the
structure: since the observed intensity (50) decomposes into variable emission profile, the visibility amplitudes also exhibit
N(α,β) a large variability. Nevertheless, as we mentioned previously,
these fluctuations wash out under time-averaging, leaving the
X
Io (α, β) = In (α, β), (118)
n=0
characteristic ringing signature of the photon ring as the main
persistent feature that dominates the visibility.
with In (α, β) denoting the nth image layer, we can likewise use We also display the visibility amplitudes at baseline angles
the linearity of the Fourier transform (117) to decompose the ϕ = 0◦ and ϕ = 90◦ (again, in red and blue, respectively) for
total complex visibility into individual subring components: our 100 snapshots in the right panel of Fig. 2. Their incoherent
X∞ Z time-average is also shown with bold solid lines. “Incoherent”
V(u) = Vn (u), Vn (u) = In (xo )e−2πiu·xo d2 xo . (119) here means that the averaging is performed at the level of the
n=0 visibility amplitudes |V(u, ϕ)|, as in a realistic experiment [35],
rather than at the level of the complex visibilities themselves,
The subring images In (xo ) with n > 0 consist of narrow rings before taking the amplitude—these operations (averaging and
of characteristic widths e−γn (where γ is a Lyapunov exponent) taking amplitudes) do not commute for complex quantities.
and roughly equal intensities. Hence the total flux Vn (0) of the (Experimentally, the latter type of “coherent” time-averaging
nth image layer also scales like e−γn [7]. Moreover, a narrow would require tracking the visibility phase over the course of
ring produces a characteristic, weak u−1/2 power-law falloff in all observations, which is beyond our near-term capabilities.)
its Fourier transform: indeed one can show that in the regime These time-averaged amplitudes can be directly compared to
1 1 those of a radial profile; see, e.g., Figs. 4 and 5 of Ref. [35].
u , (120)
d w
the visibility (117) in polar coordinates u = (u, ϕ) of a ring of
diameter ∼ d and width ∼ w takes the “universal” form [33, 34]
V. RAY TRACING REQUIREMENTS
e−2πiCϕ u h L − iπ iπdϕ u iπ i
V(u) = √ αϕ e 4 e + αRϕ e 4 e−iπdϕ u , (121)
u As described in Sec. II F, we decompose images into layers
labeled by photon half-orbit number n and ray trace each layer
where αL,R
ϕ = αR,L
ϕ+π > 0 encodes the polar intensity profile separately. In principle, one can choose any grid resolution
around the ring, while dϕ and Cϕ are its projected diameter and in each lensing band. In practice, however, these resolutions
centroid displacement at angle ϕ in the image, respectively. must be sufficiently high to resolve the image features present
As first pointed out in Ref. [7], this suggests that the nth in each layer. Otherwise, the output is not a faithful image of
photon ring dominates the interferometric signal in the regime the source, and the resulting visibility is likewise inaccurate.
1 1 In addition, the n = 0 lensing band occupies all of the image
u , (122) plane, whereas the n = 0 grid must of course have a finite size.
wn−1 wn Hence, the direct image of the source may need truncatation,
where wn ≈ e−γn w0 is the width of the nth subring, producing a which could in turn introduce errors in the visibility. In this
characteristic cascading structure of damped oscillations with section, we describe the requirements on grid resolution and
periodicity encoding the diameter of successive subrings; see size that must be met in order to ensure that both the image
also Secs. 2 and 4 of Ref. [10] for a more detailed discussion. and the visibility of a source are correctly computed. We also
Since we compute images layer-by-layer—as described in strive to explain the types of errors encountered when these
Sec. II F—with a different grid and resolution in each layer In , requirements are not met.
it is convenient to also compute the visibility layer-by-layer as We first discuss grid resolution and then grid size. A finite
well, and obtain each Vn as the Fourier transform (119) of In . resolution erases small-scale features and a finite field of view
Instead of taking this 2D Fourier transform directly, as in cuts off large-scale features, but these effects can be remedied
Refs. [10, 35], we make use of the projection-slice theorem with the use of interpolation and extrapolation, respectively.
26

A. Resolution requirements Likewise, in the higher-n layers, we must ray trace on grids
with exponentially fine spacings δx(n) ≈ e−γn δx(0) . The reason
1. Requirements on the underlying simulation of the source is simply that if the smallest feature in the n = 0 image is
of size w0 ≈ w, then its nth mirror image in the nth lensing
When computing images of some source model Is (xs ), we band will be roughly of size wn ≈ e−γn w0 . To avoid missing
must be able to sample the source intensity Is at arbitrary xs these features (or underresolving them), we must demand that
within the range of the simulation. This is because the transfer δx(n)  wn , a condition that is automatically satisfied with this
functions (33)–(35) vary smoothly across the image plane, as adaptive grid scaling tailored to the Kerr lensing behavior.
discussed in Sec. II G and as illustrated in Figs. 5 through 9. Provided that each image layer is ray traced at a sufficiently
For a stationary and axisymmetric source profile of the form high resolution to ensure that no sub-grid structure is omitted,
(57), this requirement is trivial, since we typically specify the we can safely interpolate the image layers In (xo ). In turn, we
radial profile J(rs ) analytically, as we did in Eq. (58) with can evaluate their Radon transforms along any desired slice,
Johnson’s SU distribution (which can be computed anywhere and then take a 1D Fourier transform to obtain their correctly
to arbitrary precision). On the other hand, if the source profile computed visibilities Vn (u), as described in Sec. IV C.
Is (xs ) is not known exactly, then we must define it everywhere Conversely, we note that if the resolution is not increased in
in the range where it is to be sampled via interpolation. each successive layer, then eventually some image layer In (xo )
In that case, it is important that the interpolation be done on will have lensed features smaller than its grid spacing. This
a dense grid, so that we do not miss any features of the source. unresolved sub-grid structure will render the visibility Vn (u),
In practice, this means that if its smallest features are of size and thus the full visibility V(u), inaccurate. Hence, adaptive
∼ w, then before interpolating the source, we must first sample ray tracing is necessary to guarantee accurate visibilities.
it on a grid with pixel spacing δx . w. Otherwise, if the grid
spacing is some δx & w, then we may miss a feature of size w
in between two successive grid points xi and xi + δx, leading 3. Convergence tests
to an error in both the computed image and its visibility.
Moreover, the grid density must be high enough to resolve To summarize, for both the emission model Is (xs ) and its
all the features in the source with many points, so as to ensure ray-traced image layers In (xo ), we must use sufficiently high
that the resulting interpolation is smooth: if some features are resolutions to ensure that we do not truncate any sub-grid
underresolved (for instance, if a bump in a radial profile has structure (that is, structure on length scales smaller than the
only one point underneath it), then interpolation will produce resolution). This then allows us to safely interpolate Is (xs )
spurious sharp features in the image that will have a significant and In (xo ), without missing any features or introducing any
impact on its Fourier transform, and hence the visibility. Thus, spurious new ones in the interpolation procedure.
δx . w is not enough—we must in fact require that δx  w.15 Interpolating Is (xs ) is needed to obtain the layers In (xo ), as
For the GRFs that we introduced in Sec. III to model astro- ray tracing requires us to compute the intensity loaded on rays
physical fluctuations, this means that the spacing between grid that intersect the equatorial source at generic crossing points.
points xs = (ts , xs , ys ) must be smaller than the characteristic Interpolating the image layers In (xo ) is also needed to compute
size of fluctuations, which is set by the correlation lengths λ` their Fourier transforms Vn (u), and hence the visibility (119).
with ` ∈ {0, 1, 2}. That is, we require a grid spacing δx  λ` . Any errors in either of these steps introduce errors in the final
We reiterate that interpolation is unavoidable here because visibility V(u), so we must carefully avoid them by using the
the inoisy simulation grid does not coincide (barring some appropriate resolutions, as explained above.
incredible fine-tuning) with the equatorial crossings of rays In practice, however, these requirements are only heuristic.
traced backwards from points in the image-plane grids. For instance, the Lyapunov exponent is an angle-dependent
function γ(ϕ) around the image plane, so we cannot precisely
realize the exponential scaling δx(n) ∼ e−γn δx(0) on a Cartesian
2. Requirements on the image grids used in ray tracing grid. In addition, eγ(ϕ) can be as large as ∼ 20, but such an
increase in pixel density is not always necessary. In fact, recall
We now describe the resolutions needed in the grids used to that for the example in Sec. IV, we simply set δx(n) = 2n δx(0) ,
ray trace within each lensing band, as described in Sec. II F. a much more modest but still adequate scaling of pixel density.
For the direct n = 0 image, we must ray trace on a grid with Inversely, a realization of an inoisy source realization may
spacing δx(0) . δxmodel for precisely the same reasons as listed occasionally make a field excursion that creates a far smaller
above for the underlying model—provided we do so, we will fluctuation than the minimum size w typically expected from
not miss any image features and they will all be well resolved. its correlation matrix Λ(xs ). While statistically improbable,
such events cannot be precluded from occurring, and could
result in a too-large grid spacing δxmodel & w. Unfortunately,
15 it is intractable to directly check that this does not happen.
This smooth interpolation effectively defines the source down to arbitrarily
small length scales by decreeing that no new features appear below size w, Instead, it is best to numerically check for the absence of
which is the minimal (most conservative) “ultraviolet completion” of the errors using a simple convergence test, in which one doubles
model. Any other choice would modify the visibility on baselines u & 1/w all the resolutions (of the simulated profile and image grids)
and should then be specified as part of the model itself, rather than by fiat. and computes again the observables (images and visibilities).
27

100 100 ϑ = M/2; γ = −3/2


- ϕ = 0◦ - ϕ = 90◦
ϑ = M/4; γ = −1
0.5

10 −1

Normalized Radial Emission Profile


0.4
10−1
Visibility Amplitude (Jy)

(mJy)
10−2
0.3

0.2

10−2 0.1 10−3


425 430 435 440 445

10−4
10 −3

10−5

10−4 10−6
1
∆(%)

0 r− r+ rms
−1 10−7
0 100 200 300 400 500 1 10 100
Baseline Length u (Gλ) rs (M )

FIG. 14. Convergence test of the resolution needed to ray trace an FIG. 15. Radial emission profile JSU (r; µ, ϑ, γ) [Eq. (58)] with µ = r−
inoisy snapshot. We plot its visibility amplitude at baseline angles to ensure that the profile peaks past the outer horizon at r = r+ . The
ϕ = 0◦ (red) and ϕ = 90◦ (blue) computed using the grid resolutions values of the other parameters ϑ and γ are shown in the legend. The
described in the main text. The black dashed lines correspond to the dashed lines depict the profile multiplied by the cutoff function (123)
visibility amplitudes obtained after doubling the nominal resolution. with s¬ = M and r¬ = 15M. From left to right, the vertical lines
The inset panel zooms into the region were the fits were performed indicate the inner and outer horizons r = r± and the ISCO radius rms .
to infer the projected ring diameter as described in Sec. VI. Since the
differences are consistently below 1%, we can validate the resolution.
1. Field of view of the direct image

In Fig. 14, we show an example of a convergence test for the We first focus on the direct image, as it is the most prone to
example from Sec. IV: after halving the grid resolutions, the the Gibbs phenomenon. This is because it occupies the n = 0
changes in the visibility amplitude remain small, confirming lensing band, which has noncompact support: as described in
that the resolution scaling δx(n) = 2n δx(0) is adequate. Sec. II E, it fills the entire image plane (except for the interior
of the apparent equatorial horizon).
If the emission profile has compact support in the equatorial
B. Field of view requirements plane (rs , φs ), however, its direct image I0 (xo ) will also have
compact support within the n = 0 layer, and we can terminate
Being able to completely resolve an emission profile Is (xs ) our FOV just past the edge of this image without introducing
and a (central) part of its image Io (xo ) is a necessary but not any truncation errors or triggering the Gibbs phenomenon.
sufficient condition to correctly compute its visibility V(u). On the other hand, if the emission profile extends infinitely
Another requirement for the accurate computation of a far in the equatorial plane (remaining nonzero as rs → ∞), or
source model’s visibility is related to the size of its image’s else if its support is too large to fully ray trace over in practice,
field of view (FOV), and how fast the emission profile decays then we must necessarily cut off its n = 0 image.16
past it. Even when all the photon rings are fully resolved (that For instance, consider the simple, stationary, axisymmetric
is, the resolution in each lensing band is sufficiently high), it is source (57) with radial profile J(rs ) given by Johnson’s SU
still imperative to check that the FOV is sufficiently large for distribution (58). While this emission profile does decay very
the image to include all of the features with significant flux. fast as rs → ∞, it always remains finite. In Fig. 15, we plot
If the FOV is too small and cuts out a portion of the source, it in logarithmic scale for two sets of values of its parameters.
then this truncation may create a sharp drop in the observed From now on, we focus on the red profile, with parameters
intensity at the edges of the image. Such an artificial transition µ = r− , ϑ = M/2, and γ = −3/2. We first ray trace its direct
would effectively introduce high-frequency components to the n = 0 image on a grid of size 400M, and then we compute the
image, polluting its Fourier transform, and hence its visibility. corresponding visibility amplitude |V0 (u, 0◦ )|, which we plot
This is the well-known Gibbs phenomenon. as a black dashed curve in the top left panel of Fig. 16.
Since we are interested in narrow features of the image, and
in particular the interferometric signature of its n = 2 photon
16
ring that is encoded in the visibility on very long baselines, We saw in Sec. V A that a finite grid resolution introduced an “ultraviolet”
we cannot avoid dealing with this potential effect. cutoff; now we see a finite grid size as the corresponding “infrared” cutoff.
28

100 100
100 100

10−1
10−2 10−1 10−1

10−2
V0 (u)/V0 (0)

V1 (u)/V1 (0)
10−2 10−2
0 5 10 15 0 20 40 60 80 100
10−4

10−3

10−6
FOV = 400 (M ) 10−4 FOV = 400 (M )
FOV = 50 (M ) FOV = 50 (M )
FOV = 50 (M ); rcutoff = 15 (M ) FOV = 50 (M ); rcutoff = 15 (M )
10−8 10−5
100 20
∆ (%)

∆ (%)
0 0
−100 −20
0 100 200 300 400 500 0 100 200 300 400 500
Baseline Length u (Gλ) Baseline Length u (Gλ)

100 100
100 0.8

0.7

0.6
10−1

(mJy)
0.5

0.4

10−1 0.3
10 −2
V2 (u)/V2 (0)

V (u)/V (0)

0.2
0 20 40 60 80 100 260 280 300 320 340
10−1

10−3

FOV = 400 (M ) 10−4 FOV = 400 (M )


FOV = 50 (M ) FOV = 50 (M )
FOV = 50 (M ); rcutoff = 15 (M ) FOV = 50 (M ); rcutoff = 15 (M )
10−2 10−5
1 25
∆ (%)

∆ (%)

0 0
−1 −25
0 100 200 300 400 500 0 100 200 300 400 500
Baseline Length u (Gλ) Baseline Length u (Gλ)

FIG. 16. Normalized visibility amplitudes |Vn (u, 0◦ )| corresponding to the image layers In (α, β) produced by the red radial emission profile in
Fig. 15 (Johnson’s SU distribution with µ = r− , ϑ = M/2 and γ = −3/2), for n = 0 (top left), n = 1 (top right), and n = 2 (bottom left). Bottom
right: Normalized visibility amplitude |V (u, 0◦ ) | corresponding to the full image Io (α, β) obtained by summing over the preceding three layers.
In each quadrant, the main panel shows the visibility amplitudes obtained using different values for the field of view (FOV): dashed black for
a large 400M FOV, blue for a smaller 50M FOV, and red for the 50M FOV with a cutoff applied [Eq. (123) with s¬ = M and r¬ = 15M]. Inset
panels zoom into the shaded regions of each main plot to highlight the differences between the visibility amplitudes, while the smaller panels
beneath the main ones plot percentage differences relative to the amplitude with large 400M FOV. In all these examples, we have assumed that
the flow follows the Cunningham prescription [67], which in our notation corresponds to the Keplerian circular motion described in Sec. B 1.

At the edge of our FOV, J(rs ) has dropped to . 10−11 times We then plot the resulting visibility amplitude as the solid blue
its maximum intensity. As such, the sudden drop introduced curve in the top left panel of Fig. 16. Perhaps surprisingly, it
by our truncation is too small to produce a significant ringing differs quite significantly from the “correct” one! In fact, their
effect on long baselines, and so it cannot meaningfully affect relative difference—which we plot as a blue curve in the small
the signal. We will consider this large FOV and its associated panel below the main—shows a large and growing percentage
visibility amplitude as the “correct” one. Next, we ray trace error that exceeds 100% before even reaching 100 Gλ. Even
the same image with its FOV truncated at rs ≈ 15M, a cutoff though the total flux that we ignored is tiny, the sudden drop in
past which J(rs ) drops to . 10−4 times its maximum intensity. intensity at the FOV’s edge triggered the Gibbs phenomenon.
29

The resulting error can affect the interferometric signature Thus, differences in the visibility amplitude |V1 (u, 0◦ )| only
of the n = 2 photon ring and is therefore an obstacle that we become apparent on baselines that start to resolve its width, as
need to overcome, especially since this very profile is roughly can be seen in the small plot of percentage deviation. On the
consistent with the 2017 EHT observations of M87* [35]. other hand, truncating the FOV of the n = 1 image from 400M
There are two approaches to mitigating this problem. The to 50M has no discernible effect on its visibility amplitude
first is to increase the FOV until the profile decays enough (solid blue curve) and results in a vanishingly small deviation.
at the edges that it “effectively vanishes”—in that case, these Finally, we can repeat this exercise for the n = 2 ring.
spurious ringing effects will not be noticeable. In effect, this Again, windowing the profile has a negligible effect of . 1%,
is what we did when we chose an FOV of 400M. However, it as does the truncation. In fact, the latter deviation in this case
may not always be feasible to extend the FOV this much. For ought to vanish entirely and is only nonzero due to numerical
instance, if the emission profile is an inoisy source simulated error introduced by the discretization of the Fourier transform:
on a grid that stretches out to some maximal rs , then extending extending the FOV of the image to 400M by“zero padding”
the source past this cutoff would require extrapolation. This is changes the sampling of the fast Fourier transform, resulting
already subtle for smooth profiles, as it amounts to introducing in a minute effect.
ad hoc data to the model. We note that these tiny errors are essentially negligible in
The second approach—referred to as “apodization”—has the full visibilities, which we compare in the bottom right
already been used to deal with this problem in Ref. [10]. This panel of Fig. 16, where most of the percentage error can be
method involves the multiplication of the image by a suitable attributed to the direct image, as expected.
window (also known as tapering or apodization function) that
smoothly tapers the intensity to precisely zero before the edge
of the FOV. For our example, we will use the window function VI. MEASURING THE PHOTON RING SHAPE

1 − tanh [s¬ (r − r¬ )]
C(s¬ , r¬ ) = , (123) Using black hole imaging to test general relativity in a
2
regime where gravity is strong and yet non-dynamical poses
with a cutoff value of r¬ = 15M (well within the “small” FOV significant challenges: one must disentangle the astrophysical
of 50M) and s¬ = M, which results in the smooth but rapidly effects of the radiating source from the purely gravitational
truncated profile depicted with a dashed red line in Fig. 15. ones [26, 29]. A promising approach is to focus on the n = 2
Ray tracing the direct image of this profile and computing photon ring, whose interferometric signature is expected to
its visibility amplitude now results in the solid red curve that dominate the time-averaged radio visibility on long baselines.
we plot in the top left panel of Fig. 16. Its difference relative More precisely, Ref. [35] proposed a test of strong-field
to the “correct” amplitude is much smaller, and the percentage GR based on a shape measurement of the n = 2 photon ring
difference remains fairly constant well past 100 Gλ. This can around M87*, which has been recently reviewed in Ref. [10].
therefore be an effective method. Here, we show in the context of our example from Sec. IV
In practice, we prefer to always use emission profiles that how AART may be used to simulate this test using a source that
smoothly terminate within the FOV of our direct image, which includes “realistic” astrophysical fluctuations (generated with
we can easily take to be 50M wide. In case our source profile inoisy, as described in Sec. III). We offer a brief overview of
extends beyond this FOV, we apply the window function (123) the key points in the test, and refer the reader to Refs. [10, 35]
to it and take the resulting profile as our “true” source. for more details.
In the regime (120), a ring of projected diameter dϕ has
a visibility that takes the universal form (121). Therefore, its
2. Field of view of higher layers amplitude can be well approximated in this regime by [33, 34]

Since the n > 0 lensing bands all have compact support, it


r
 2  2  
is straightforward to avoid the Gibbs phenomenon in higher-n |V(u, ϕ)| = αLϕ + αRϕ + 2αLϕ αRϕ sin 2πdϕ u , (125)
image layers and visibilities: we just take a FOV that includes
the totality of the nth photon ring (that is, all the flux in the nth where the functions
image layer).
eupper (u) ± elower (u)
We briefly describe the effects that windowing the emission αL/R
ϕ (u) = (126)
profile as discussed above has on higher visibility amplitudes. 2
The top right panel of Fig. 16 shows that the windowed profile encode the intensity profile around the ring image, while
produces roughly the same visibility (solid red curve) as the eupper/lower (u) are the upper/lower envelopes of the visibility
original one (dashed black curve) up to baselines u ∼ 100 Gλ, amplitude, respectively.
which are necessary to resolve image features of size The width w and diameter d of the n = 1 ring do not always
1011 1 satisfy the condition w/d  1 needed for the universal regime
100 Gλ = ≈ . (124) (120) to exist, but those of the n = 2 ring do. Hence, we can
rad 2 µas
fit Eq. (125) to the visibility amplitudes in the regime (122)
For M87*, whose photon ring has a diameter ∼ 10M ≈ 40 µas, dominated by the n = 2 ring after numerically computing the
this corresponds to a size of 0.5M, or half the n = 1 ring width. functions αL/R ϕ (u) (a simple cubic interpolation is enough).
30

N R0 R1 R2 ϕ0 RMSE VII. FUTURE OUTLOOK


5 37.045 1.289 0.879 1.319◦ 4.5 × 10−4
10 36.250 2.077 1.679 1.308◦ 3.2 × 10−4 Here, we have presented AART: a new, publicly available,
20 36.909 1.419 1.027 0.987◦ 2.4 × 10−4
numerical code designed for precision studies of a black hole’s
100 36.052 2.283 1.894 0.426◦ 1.1 × 10−4
∞ 36.115 2.219 1.830 0.202◦ 4.2 × 10−5 photon rings. AART exploits the integrability of null geodesics
in the Kerr geometry to ray trace images analytically, with no
TABLE III. Best-fit values for the projected diameter dϕ [Eq. (127)] loss of numerical precision even for strongly lensed photons
obtained after averaging N snapshots. The resulting fits are shown in that orbit the black hole multiple times. The code decomposes
Fig. 3. The normalized residuals RMSE are root-mean-square errors the image plane into layers—lensing bands—with increasing
divided by the average value of dϕ . grid resolution adapted to the lensing behavior of the hole.
The modular structure of AART can accommodate any time-
dependent and non-axisymmetric equatorial emission profile,
The inset panel in Fig. 2 shows the visibility amplitude for and the code could be extended to non-equatorial, anisotropic
two baseline angles ϕ = 0◦ and ϕ = 90◦ in the baseline range sources. Also, the components described herein can be used
u ∈ [425, 445] Gλ (corresponding to an Earth-Moon baseline separately: AART is designed to individually export the lensing
with an operation band of 332–344 GHz) with their best fit to bands, critical curve, apparent horizon, and redshift factors,
Eq. (125) overplotted in black dashed lines. which can therefore serve as input for other studies.
According to GR, the shape of the n = 2 photon ring is the As a prime application, we showed how the experimental
sum of a circle, with radius R0 , and an ellipse centered at the forecast for the test of general relativity proposed in Ref. [35]
origin, with radii R1 and R2 [35]. This “circlipse” shape has a could be further refined by including source fluctuations. We
projected diameter given by [34] used inoisy [43] to simulate a stochastic model of equatorial
emission, and we produced high-resolution synthetic images
dϕ q together with their corresponding visibility on long baselines.
= R0 + R21 sin2 (ϕ − ϕ0 ) + R22 cos2 (ϕ − ϕ0 ), (127) We then successfully extracted the GR-predicted shape for the
2
projected diameter of the n = 2 ring from its interferometric
where ϕ0 is an offset angle to account for the orientation of signature. In a follow-up paper, we will use this framework
the ring within the image plane. to carry out a parameter estimation survey that also includes
Using the statistical model described in Sec. III for the instrument response and noise—this will further validate the
emission profile with the parameters shown in Table II, we test proposed in Ref. [35] and bears relevance to SALTUS and
(i) average the visibility amplitudes of N snapshot images; other space-VLBI missions targeting the photon ring [81, 82].
(ii) fit Eq. (125) to the visibility amplitude for 36 angles Although we only studied the Kerr geometry, our approach
ϕ = {0◦ , 5◦ , . . . , 175◦ } to obtain the projected diameters dϕ ; may serve as a useful guide to studying other theories. For
and (iii) fit the general GR prediction given in Eq. (127) instance, the Kerr lensing bands may offer a starting point for
to obtain {R0 , R1 , R2 , ϕ0 }. The best global fits of Eqs. (125) numerically finding those of slightly deformed spacetimes.
and (127) were found using a Markov chain Monte Carlo Finally, recent EHT observations of Sgr A* found that only
(MCMC) method.17 ∼ 3.5% of the EHT GRMHD models passed the light-curve
We report the best-fit values of the parameters in Eq. (127) variability constraint [22]. Given the uncertainty associated
and their normalized root-mean-square error (RMSE) for each with the variability excess, and the possibility of missing
N in Table III. The resulting curves, shown in Fig. 3, follow physical ingredients in current astrophysical models, efforts to
the GR prediction very closely, even when only five snapshots develop astrophysics-agnostic phenomenological approaches
are (incoherently) averaged. The lower panels in Fig. 3 show such as the one presented in this work are just as valuable as
the difference of the best-fit curve for each case with the best- the improvement of accretion disk simulations.
fit curve for the purely radial profile consisting of the envelope
of the emission model (that is, when there are no fluctuations).
The results of this example are very encouraging for future ACKNOWLEDGMENTS
missions targeting measurements of the photon ring shape:
with only a few snapshots, one may already start to check We thank Charles Gammie, Delilah Gates, Samuel Gralla,
whether the data follow the GR prediction. A systematic study Aaron Held, Daeyoung Lee, Hadrien Paugnat, Eliot Quataert,
including instrumental noise is required to simulate a realistic Leo Stein, Frédéric Vincent, Maciek Wielgus, and George
experimental forecast—we leave this for future work. Wong for their useful comments. This work used resources
provided by Princeton Research Computing, a consortium that
includes PICSciE (the Princeton Institute for Computational
Science and Engineering) as well as the Office of Information
17 Technology’s Research Computing division.
Since we are neglecting the phase of the oscillation, Eq. (125) allows for
several values of dϕ separated by ∼ 1/u to provide a good fit [35]. However, A.C.A. and A.L. gratefully acknowledge support from Will
the global maximum is not always the actual diameter dϕ of the ring as and Kacie Snellings. A.C.A also acknowledges support from
measured directly in the image domain. As explained in Ref. [10], in those the Simons Foundation. A.L. also thanks Pam Davis Kivelson
cases, one must infer the diameter by fitting at multiple angles. for her black hole art inspired from this AART.
31

Appendix A: Explicit form of null geodesics in the Kerr exterior Their precise form depends on the nature of the radial roots
{r1 , r2 , r3 , r4 }: there are four different cases for null geodesics
This appendix lists all the formulas needed to compute the in the Kerr exterior, but only two of these—labeled type (2)
observational appearance of Kerr equatorial sources in the sky and type (3)—arise for light rays that connect the equatorial
of a distant observer. Readers are referred to Refs. [5, 40] for plane to a distant observer; meanwhile, type (4) geodesics can
derivations and further details. Throughout, F(ϕ|k), E(ϕ|k), reach distant observers but never the equator, since they are all
and Π(n; ϕ|k) denote the incomplete elliptic integrals of the vortical, whereas type (1) geodesics are all bound to the black
first, second and third kinds, respectively, defined according to hole and cannot reach a distant observer at infinity [40].
the conventions listed in App. A of Ref. [46]. K(k) ≡ F(π/2|k)
denotes the complete integral of the first
 kind and we also let a. Type (2)
E 0 (ϕ|k) ≡ ∂k E(ϕ|k) = E(ϕ|k) − F(ϕ|k) /(2k).

In this case, all the roots are real and the motion is restricted
1. Angular geodesic integrals to r ≥ r4 > r3 > r2 > r1 .
The antiderivatives of the radial geodesic integrals take the
manifestly real and smooth forms [40]
The angular motion of a Kerr photon can display two
qualitatively different behaviors depending on the sign of its I0 = F (2) (r), (A10)
energy-rescaled Carter constant η. Since we only consider
I1 = r3 F (r) +
(2)
r43 Π(2)
1 (r), (A11)
sources in the Kerr equatorial plane, we may ignore vortical √
motion with η < 0, which can never reach the equator [46]. R(r) r1 r4 + r2 r3 (2)
I2 = − F (r) − E (2) (r), (A12)
We thus restrict our attention to ordinary motion with η > 0, r − r3 2
in which case the angular geodesic integrals in Sec. II are [5] F (2) (r)
I± = −Π(2)
± (r) − , (A13)
1 r±3
G(n) = √ [2m(n)K ∓o Fo ] , (A1)
θ
a −u− with (recall that ri j = ri − r j )
1 2  
G(n)
φ = √ [2m(n)Π ∓o Πo ] , (A2) F (2) (r) = √ F arcsin x2 (r) k2 ≥ 0, (A14)
a −u− r31 r42
2u+  √  
G(n) =− √ 2m(n)E 0 ∓o Eo0 , E (2) (r) = r31 r42 E arcsin x2 (r) k2 ≥ 0,

t (A3) (A15)
a −u−
!
2 r41
where ±o = sign β, m(n) ≡ n + H(β) counts the number of Π(2)
1 (r) = √ Π ; arcsin x2 (r) k 2 ≥ 0, (A16)
r31 r42 r31
angular turning points encountered along the trajectory, H(x) !
denotes the Heaviside function, and we also introduced 2r43 r±3 r41
Π(2)
± (r) = √ Π ; arcsin x 2 (r) k2 , (A17)
r±3 r±4 r31 r42 r±4 r31
π u+
! !
u+
K=K =F , (A4) where the auxiliary function x2 (r) is defined as
u− 2 u− r " r #
cos θo u+ r − r4 r31 r31
! !
Fo = F arcsin √ , (A5) x2 (r) = ∈ 0, ⊂ [0, 1), (A18)
u+ u− r − r3 r41 r41
cos θo u+
! !
while the elliptic modulus is
Eo0 = E 0 arcsin √ , (A6)
u+ u− r32 r41
k2 = ∈ (0, 1). (A19)
cos θo u+
! !
r31 r42
Πo = Π u+ ; arcsin √ . (A7)
u+ u− Finally, the source radius is given in terms of the Jacobi
elliptic sine function sn(ϕ|k) by
 
2. Radial geodesic integrals r4 r31 − r3 r41 sn2 X2 (τ) k2
rs (τ) =
(2)
  , (A20)
r31 − r41 sn2 X2 (τ) k2
The radial integrals can be decomposed into [40]
1√  
X2 (τ) = r31 r42 τ − F arcsin x2 (ro ) k2 . (A21)
2Ma  aλ   aλ   2
Iφ = r+ − I+ − r− − I− , (A8)
r+ − r− 2M 2M
(2M)2   aλ   aλ   b. Type (3)
It = r+ r+ − I+ − r− r− − I−
r+ − r− 2M 2M
+ (2M)2 I0 + (2M)I1 + I2 , (A9) In this case, only r1 and r2 are real roots while r3 = r̄4
are complex-conjugate roots, and the range of radial motion is
with the integrals I0 , I1 , I2 and I± reducible to Legendre form. r ≥ r+ > r− > r2 > r1 .
32

The antiderivatives of the radial geodesic integrals take the Finally, the source radius is given in terms of the Jacobi
manifestly real and smooth forms [40] elliptic cosine function cn(ϕ|k) by
 
I0 = F (3) (r), (A22) (Ar1 − Br2 ) − (Ar1 + Br2 ) cn X3 (τ) k3
 Ar + Br  rs (τ) =
(3)
  , (A36)
I1 =
1 2
F (3) (r) + Π(3) (A − B) − (A + B) cn X3 (τ) k3
A+B 1 (r), (A23)
√  
 Ar + Br 2
1 2
 Ar + Br 
1 2
X3 (τ) = ABτ − F arccos x3 (ro ) k3 . (A37)
I2 = F (3) (r) + 2 Π(3)
1 (r)
A+B A+B

+ ABΠ(3) 2 (r), (A24)
Appendix B: Accretion flow four-velocities
A+B
I± = − F (3) (r)
Ar±1 + Br±2
√ Let us consider an equatorial four-velocity
2r21 AB  
+ R 1 α ± ; arccos x3 (r) k 3 , (A25) u = ut ∂t + ur ∂r + uφ ∂φ = ut dt + ur dr + uφ dφ, (B1)
(Ar±1 )2 − (Br±2 )2
with uµ and uµ the contravariant and covariant components,
with (recall that ri j = ri − r j )
respectively. The angular and radial-infall velocities are
1 uφ
 
F (3) (r) = √ F arccos x3 (r) k3 , (A26) ur
Ω= , ι=− . (B2)
AB ut ut
 √ ` 
 2r21 AB  
Meanwhile, the energy and specific angular momentum are
Π` (r) =  2
(3)
 R ` α 0 ; arccos x 3 (r) k3 , (A27)
B − A2


1
"
α2
! # E = −ut , `= . (B3)
E
 
R1 α; ϕ j = Π ; ϕ j − α f 1 , (A28)
1 − α2 α2 − 1
   For geodesic motion, these quantities are conserved, whereas
α2 E ϕ j 

1   
R2 (α; ϕ| j) = 2 F ϕ j − 
ν=
ur
(B4)
α −1 j + (1 − j) α2  ut
q
α3 sin ϕ 1 − j sin2 ϕ is not. These variables parameterize the four-velocity as
+ 2
α − 1 j + (1 − j) α2 (1 + α cos ϕ)
   
u = ut ∂t − ι ∂r + Ω ∂φ = E (− dt − ν dr + ` dφ) . (B5)
! R α; ϕ j

α2 1
+ 2j − 2 , (A29) Any two components of u determine it entirely, as the third
α − 1 j + (1 − j) α2
q can be recovered from the normalization condition u · u = −1.
p1 1 − j sin2 ϕ + sin ϕ We will parameterize the four-velocity using either the pair
p1
f1 = log q , (A30) (ι, Ω), the pair (ν, `), or the pair (E, `), as follows. First, define
2
p1 1 − j sin2 ϕ − sin ϕ  2
s Π(r) = r2 gφφ θ=π/2 = r2 + a2 − a2 ∆(r). (B6)
α2 − 1
p1 = , (A31) Given (ι, Ω), unit-normalization fixes
j + (1 − j) α2
#−1/2
r2 2
"   2M
where the auxiliary function x3 (r) is defined as ut = 1 − r2 + a2 Ω2 − (1 − aΩ)2 − ι , (B7)
r ∆(r)
!
A (r − r1 ) − B (r − r2 ) 1 which is physically admissible provided the quantity in square
x3 (r) = ∈ − , 1 ⊂ (−1, 1), (A32)
A (r − r1 ) + B (r − r2 ) α0 brackets is positive. We took the positive square root to ensure
ut > 0 is future-directed. Lowering the index of uµ yields
while the elliptic modulus is
r2 ι Π(r)Ω − 2aMr
E = χut , ν= , `= , (B8a)
(A + B) − 2 2
r21 ∆(r) χ r2 χ
k3 = ∈ (0, 1), (A33) 2M
4AB χ=1− (1 − aΩ) . (B8b)
r
and the other parameters are
Conversely, given (ν, `), unit-normalization fixes the energy to
B+A Br±2 + Ar±1 1
α0 = > 1, α± = =− , (A34) ∆(r)
v
t
B−A Br±2 − Ar±1 x3 (r± ) E= i2 , (B9)
√ √ Π(r)
 
2 − ∆(r) ν
h
A = r32 r42 > 0, B = r31 r41 > 0. (A35) r 2 − 4aM`
r − 1 − 2M
r ` r
33

which is physically admissible provided the quantity under the a. Outside the ISCO radius
square root is positive. Raising the index of uµ then yields
For r ≥ rms , we can set ι̊ = 0 and the Kerr geodesic equation
E ∆(r) χ determines the Keplerian angular velocity (Eq. (2.16) of [56]):
u = ,
t
ι = 2 χν, Ω = 2 (` + aH) , (B10a)
χ r r √
M
1 2Mr − a` Ω̊ = √ . (B17)
χ= , H= . (B10b) r +a M
3/2
1 + 2M
r (1 + H)
∆(r)
The contravariant four-velocity is then [Eqs. (B5) and (B9)]
It is often convenient to express Ω and χ in terms of ` only as √
r3/2 + a M
ůt = q , (B18a)

r − 3Mr + 2a Mr
 
a+ 1− 2M
r
(` − a) ∆(r)
3 2 3/2
Ω= , χ= . (B11)
Π(r)
− 2aM` Π(r)
− 2aM` ůr = 0, (B18b)
r2 r r2 r √
M
Lastly, given (E, `), the full four-velocity can be recovered by ůφ = q , (B18c)

solving the normalization condition u·u = −1 [or equivalently, r3 − 3Mr2 + 2a Mr3/2
Eq. (B9)] for
and the Keplerian conserved quantities are [Eqs. (B8)]
√ √
r (r − 2M) + a M
s
Π(r) 4aM` 2M 2 ∆(r)
!
ν=
r
− − 1− ` − 2 . (B12) E̊ = q , (B19a)

∆(r) r 2 r r E r3 − 3Mr2 + 2a Mr3/2
√  2 √ 
Finally, the observed redshift g(ι, Ω) is M r + a2 − 2a Mr
`˚ = √ √ . (B19b)
r (r − 2M) + a M
E 1
g= µ
=  √ . (B13) Likewise [Eqs. (B8)],
−pµ u R(r)
ut 1 ±r ∆(r) ι − λΩ √
2M r
ν̊ = 0, χ̊ = 1 − √ . (B20)
r3/2 + a M
Alternatively, the redshift can also be expressed in terms of
(E, ν, `) as [Eq. (B10)] Hence, the covariant four-velocity is [Eq. (B5)]
√ √
r (r − 2M) + a M
∆(r) ůt = − q , (B21a)
g= √ i, (B14a) √
E G ±r R(r) ∆(r) r3 − 3Mr2 + 2a Mr3/2
h
r2
ν
Π(r) 2M
!
2aM ůr = 0, (B21b)
G = 2 − 1− `λ − (` + λ) . (B14b) √  √ 
r r r M r2 − 2a Mr + a2
ůφ = q . (B21c)

r3 − 3Mr2 + 2a Mr3/2
1. Keplerian circular orbits (geodesic motion) Finally, the observed redshift is [Eq. (B13)]
q √
We let u = ů denote the four-velocity of timelike, circular- r3 − 3Mr2 + 2a Mr3/2
equatorial geodesics. These orbits define Keplerian motion in g̊ = √ . (B22)
r3/2 + M (a − λ)
the Kerr spacetime; they are only stable if (Eq. (2.20) of [56])
All of the above√quantities are physically admissible provided
√ r3 − 3Mr2 + 2a Mr3/2 > 0, which is the case for all r ≥ rms .
r2 − 6Mr + 8a Mr − 3a2 ≥ 0, (B15)

or equivalently, if r ≥ rms , where rms denotes the radius of the b. Inside the ISCO radius
(marginally stable) innermost stable circular orbit (ISCO),
h p i For r ∈ [r+ , rms ], we follow Cunningham’s prescription [67]
rms = M 3 + Z2 − (3 − Z1 ) (3 + Z1 + 2Z2 ) , (B16a) and smoothly extend the flow past the ISCO using geodesic
inspirals with the conserved quantities of the ISCO:
q h p3 p3 i
3
Z1 = 1 + 1 − a2∗ 1 + a∗ + 1 − a∗ , (B16b) r
2 M
a Ems = E̊ r=r = 1 − ,
q
Z2 = 3a2∗ + Z12 , a∗ = . (B16c) (B23a)
M
ms 3 rms
√  √ 
M rms2
− 2a Mrms + a2
Inside the ISCO, we must use an inspiraling geodesic motion, `ms = ` r=r = √
˚ √ . (B23b)
as orbits are unstable and particles must fall into the horizon.
ms
rms (rms − 2M) + a M
34

This choice results in a nonzero radial component [Eq. (B12)] As particles fall in, they pick up a radial velocity [Eq. (B12)]
r
2Mr r2 + a2
p
r2 3/2 
2 M  rms
ν̊ = −1 > 0, (B24) ν̄ = > 0, (B31)
∆(r)E̊ms 3 rms r ∆(r)

where we used the identity rms 2
− 6Mrms + 8a Mrms − 3a2 = 0 and due to frame-dragging (the off-diagonal term gtφ , 0),
[Eq. (B15)] to simplify the expressions for both Ems and ν̊. they also acquire an angular velocity Ω̄ , 0 [Eqs. (B10)]:
It then follows that [Eqs. (B10)] ∆(r)
q
2aMr
ῑ = 2Mr r2 + a2 , Ω̄ = ,

(B32a)
 3/2 Π(r) Π(r)
χ̊ rms
r −1 χ̊  
ι̊ = q , Ω̊ = 2 `ms + aH̊ , (B25a) r2 ∆(r) 2Mr
3 rms r χ̄ = , H̄ = . (B32b)
2 M − 1 Π(r) ∆(r)
1 2Mr − a`ms Explicitly, the covariant four-velocity is then [Eq. (B5)]
χ̊ = , H̊ = , (B25b)
∆(r)

1+ 2M
1 + H̊
2Mr r2 + a2
r
p 
ūt = −1, ūr = − , ūφ = 0, (B33)
so that the covariant four-velocity is explicitly [Eq. (B5)] ∆(r)
r
2 M while the contravariant four-velocity is [Eqs. (B5) and (B10)]
ůt = − 1 − , (B26a)
3 rms
Π(r)
r
r2 2 M  rms 3/2
ūt = ,
ůr = − −1 , (B26b) r2 ∆(r)
(B34a)
∆(r) 3 rs r
√ √ 2Mr r2 + a2
  p 
M rms2
− 2a Mrms + a2 ū = −
r
, (B34b)
ůφ = q . (B26c) r2
√ 3/2
3
rms − 3Mrms 2
+ 2a Mrms 2aM
ūφ = . (B34c)
r∆(r)
The contravariant four-velocity is then [Eqs. (B5) and (B10)]
Finally, the observed redshift is [Eq. (B13)]
2 M Π(r) − 2aM`ms r
r
ů = 1 −
t
, (B27a) r2 ∆(r)
3 rms r2 ∆(r) ḡ = √ . (B35)
Π(r) − 2aMrλ ±r R(r) 2Mr r2 + a2
r p
2 M  rms 3/2
ůr = − −1 , (B27b)
3 rms r
 
2 M a + 1 − r (`ms − a)
r 2M
3. Sub-Keplerian orbits (non-geodesic motion)
φ
ů = 1 − , (B27c)
3 rms ∆(r)
We now define a four-velocity u = û for timelike, equatorial
The contravariant components may also be recast in the form sub-Keplerian orbits. Such trajectories cannot be geodesic;
Ems Ems  following [83], we introduce a “sub-Keplerianity” parameter
ůφ =

ůt = , ` ms + a H̊ , (B28a) ξ ∈ (0, 1) and instead prescribe a non-geodesic motion via
χ̊ r2
√  2 √ 
which matches Eqs. (A12) of [67] since (γe , λe ) = (Ems , `ms ). M r + a2 − 2a Mr
Finally, the observed redshift is [Eq. (B14)] `ˆ = ξ`˚ = ξ √ √ , (B36)
r (r − 2M) + a M
∆(r)
g̊ = q " #, (B29a) where `˚ denotes the Keplerian specific angular momentum
√ (√rmsr −1)3/2
1−2 M
G̊ ± r R(r) (B19b). As for Keplerian motion (with ξ = 1), we must treat
3 rms 3 rms
2 M −1 radii outside and inside of the ISCO radius rms separately.
Π(r)
!
2M 2aM
G̊ = 2 − 1 − `ms λ − (`ms + λ) . (B29b)
r r r
a. Outside the ISCO radius

2. Radial infall (geodesic motion) For r ≥ rms , we demand that, as for Keplerian orbits,

We let u = ū denote the four-velocity of timelike, radially ν̂ = 0. (B37)


infalling equatorial geodesics. One class of such trajectories This fixes the sub-Keplerian orbital energy to be [Eq. (B9)]
consists of particles that fall in from spatial infinity with zero
∆(r)
s
initial velocity and vanishing (conserved) angular momentum:
Ê = Π(r)
 > 0, (B38)
4aM `ˆ

Ē = 1, `¯ = 0. (B30) 2 −
r
− 1 − 2M `ˆ2
r r
35

which is manifestly real for all ξ ∈ (0, 1), since the quantity in The contravariant four-velocity is then [Eqs. (B5) and (B10)]
the square root is strictly greater than when ξ = 1.
It then follows that [Eqs. (B10)–(B11)] Êms Π(r) 2aM `ˆms
" #
û =
t
− , (B48a)
   ∆(r) r2 r
a + 1 − 2M `ˆ − a
ι̂ = 0, Ω̂ =
r
, (B39) ∆(r)
Π(r) `ˆ ûr = − 2 ν̂Êms , (B48b)
r2
− 2aM
r r
" ! #
Ê 2M  ˆ
ûφ =
ms
so that the covariant four-velocity is explicitly [Eq. (B5)] a+ 1− `ms − a . (B48c)
∆(r) r
ût = −Ê, ûr = 0, ûφ = `ˆÊ. (B40)
Finally, the observed redshift is [Eq. (B14)]
The contravariant four-velocity is then [Eqs. (B5) and (B10)]
∆(r)
Ê Π(r) 2aM `ˆ ĝ = i, (B49a)
" #

ût = , Êms Ĝ ±r R(r) ∆(r)
h
− (B41a) ν̂
∆(r) r2 r r2
Π(r)
!
ûr = 0, 2M ˆ 2aM  ˆ 
(B41b) Ĝ = 2 − 1 − `ms λ − `ms + λ . (B49b)
" ! # r r r
Ê 2M ˆ
ûφ =

a+ 1− `−a . (B41c)
∆(r) r
Finally, the observed redshift is [Eq. (B14)] 4. General flow (non-geodesic motion)

∆(r)
ĝ = h Π(r)    i . (B42) We can linearly superpose the preceding four-velocities to
Ê r2
− 1− 2M
r `λ
ˆ − 2aM
r `ˆ + λ obtain a general flow u = ũ combining both circular motion
and radial inflow. Following [11, 66], we do so by defining

b. Inside the ISCO radius ũr = ûr + (1 − βr ) (ūr − ûr ) , (B50)


  
Ω̃ = Ω̂ + 1 − βφ Ω̄ − Ω̂ , (B51)
For r ∈ [r+ , rms ], we do not expect orbits to remain circular.
Instead, we use Cunningham’s prescription (B23) to smoothly
extend the sub-Keplerian condition (B36) past the ISCO using where û is the four-velocity of the sub-Keplerian flow, given
the conserved quantities of the sub-Keplerian ISCO orbit, by Eq. (B41) for r ≥ rms (outside the ISCO) and by Eq. (B48)
for r+ < r < rms (inside the ISCO), while ū is the radial inflow
Êms = Ê r=r , `ˆms = `ˆ r=r = ξ`ms , (B43) four-velocity (B34). Likewise, Ω̂ = ûφ /ût is the sub-Keplerian
ms ms
angular velocity, given by Eq. (B46) outside the ISCO and by
or more explicitly, Eq. (B46) inside the ISCO, while Ω̄ = ūφ /ūt is the angular
v velocity (B32) of radial inflow.
∆(rms ) Given ũr and Ω̃ = ũφ /ũt , we can solve the normalization
t
Êms = , (B44a)
Π(rms )
 4aM `ˆms
 condition u · u = −1 for ũt , the only missing component of the
rms `ms
− 2 − 1 − 2M ˆ2
rms rms four-velocity, resulting in a simple modification of Eq. (B7):
√  2 √ 
M rms − 2a Mrms + a2
`ˆms =ξ √ √ . (B44b)
v
1 + ∆(r)
u 2
r
(ũr )2
u
rms (rms − 2M) + a M
t
ũ =t
 2 , (B52)
1 − r2 + a2 Ω̃2 − 2M Ω̃

This choice results in a nonzero radial component [Eq. (B12)] r 1 − a
s
r Π(r) 4aM `ˆms
!
2M ˆ2 ∆(r) This fully specifies the general four-velocity ũ.
ν̂ = − − 1− `ms − . (B45) For generic values of the three parameters (ξ, βr , βφ ), the
∆(r) r2 r r Êms
flow ũ is not geodesic. However, when βr = βφ = 0, ũ reduces
It then follows that [Eqs. (B10)] to the radial inflow ū, which is ξ-independent and geodesic.
At the opposite end, when βr = βφ = 1, ũ reduces to the
[∆(r)]2 ν̂ χ̂   sub-Keplerian flow ũ(ξ), which in turn reduces when ξ = 1 to
ι̂ = , Ω̂ = 2 `ˆms + aĤ , (B46a) geodesic Keplerian motion ů = ũ(1).
Π(r) − 2aM `ˆms r r
1 2Mr − a`ms
χ̂ =  , Ĥ = , (B46b)
∆(r)

1 + r 1 + Ĥ
2M
Appendix C: More on the Matérn covariance
so that the covariant four-velocity is explicitly [Eq. (B5)]
This appendix offers a field-theoretic interpretation of the
ût = −Êms , ûr = −ν̂Êms , ûφ = `ˆms Êms . (B47) Gaussian random field with Matérn covariance.
36

1. Classical field 3. Higher-derivative fields

A free, massive scalar field Φ(x) with mass m in Euclidean Consider now the higher-derivative theory
spacetime Rd is described by the Lagrangian
1  n
L = − Φ ∇2 − m2 Φ, (C9)
1 1 2
L= ∂ µ Φ ∂ µ Φ + m2 Φ 2 , (C1)
2 2
where n is an integer power. This is a free (Gaussian) theory
and therefore obeys the Euler-Lagrange field equation because it remains quadratic in the field for any n. Classically,
  the field obeys the higher-derivative wave equation
∂L  ∂L 
= ∂µ     ,
n
(C2)

∂Φ ∇2 − m2 Φ(x) = 0, (C10)
∂ ∂µ Φ
whose Green function ∆d,n (x) is defined by
which in this case is the classical Klein-Gordon wave equation
 n
  ∇2 − m2 ∆d,n (x) = −δ(d) (x). (C11)
∇2 − m2 Φ(x) = 0, (C3)
 n
µ
where ∇ = ∂µ ∂ is the scalar Laplacian. The Green function
2 In momentum space, this reads k2 + m2 ∆˜ d,n (k) = 1, so
∆d (x) for this equation is defined by
1
∆˜ d,n (k) =  , (C12)
k2 + m2 n
 
∇2 − m2 ∆d (x) = −δ(d) (x). (C4)
  and the inverse Fourier transform (70) yields
In momentum space, this reads k2 + m2 ∆˜ d (k) = 1, so
eik·x dd k
Z
d
1 ∆d,n (x) = n ∝ (mx)n− 2 Kn− d (mx). (C13)
∆˜ d (k) = 2 , (C5) k2 + m (2π)d
2 2

k + m2
This nontrivial identity, including the proportionality factor, is
where k2 = k · k. The inverse Fourier transform (70) yields derived in Eq. (C34) below.
Letting ν = n − d/2 and λ = 1/m, the propagator becomes
eik·x dd k md−2
Z
d
∆d (x) = = (mx)1− 2 K1− d (mx), (C6)  x ν  x 
k + m (2π)
2 2 d d
(2π) 2 2
∆d,n (x) = C Kν , (C14)
λ λ
where x2 = x · x. We present a full derivation of this nontrivial
which is (proportional to) a Matérn covariance.
identity in Eq. (C34) below.
We conclude that a d-dimensional Gaussian random field
with Matérn covariance of order ν = n − d/2 and correlation
length λ is equivalent to a (Euclidean) quantum scalar field of
2. Quantum field
Compton wavelength λ (and therefore mass m = 1/λ) with a
kinetic term including 2n derivatives.
Integrating the Lagrangian (C1) by parts recasts it as In Lorentzian signature, such a kinetic term would lead to
problems with causality, but in Euclidean signature this seems
1  
L = − Φ ∇2 − m2 Φ. (C7) like an acceptable statistical field.
2
The path integral over this Lagrangian defines the Euclidean
quantum field theory of a free, massive scalar field Φ(x) with 4. Connection with the associated SPDE
mass m in Rd , which is fully characterized by its two-point
function: the Feynman propagator (C6). A d-dimensional Gaussian random field Φ(x) with Matérn
Letting ν = 1 − d/2 and λ = 1/m, the propagator becomes covariance of order ν = n−d/2 and correlation length λ = 1/m
obeys the stochastic PDE
md−2  x ν  x 
∆d (x) = Kν , (C8)  n
(2π) 2 λ
d
λ m2 − ∇2 2 Φ(x) = W(x), (C15)

which is (proportional to) a Matérn covariance. where the pseudodifferential operator on the LHS is defined
We conclude that a d-dimensional Gaussian random field via its spectral properties [76], and the Gaussian random field
with Matérn covariance of order ν = 1 − d/2 and correlation W(x) is the standard white noise process defined in Sec. III D.
length λ is equivalent to a (Euclidean) quantum scalar field of We now wish to reconcile this statement with more familiar
Compton wavelength λ (and therefore mass m = 1/λ). facts from quantum field theory, particularly in the case n = 1.
37

When n = 1, the Gaussian random field Φ(x) is a Euclidean In spherical coordinates,


scalar field. Yet, Eq. (C15) implies that it also obeys
π
Ωd−1 ∞
kd−1
Z Z

m2 − ∇2 Φ(x) = W(x), (C16) ∆d,n (x) = eikx cos θ sind−2 θ dθ dk,
(2π)d k 2 + m2 n

0 0
(C25)
which seems to impose yet another condition on the field. We
now show how this statement is consistent with the propagator
where Ωn denotes the solid angle on the sphere S n , which is
(C6), which fully characterizes the behavior of the field.
The key is to multiply two copies of this equation inserted at n

two different points x1 and x2 , keeping track of which position 2π 2


Ωn =  . (C26)
the derivatives in the Laplacian act upon: Γ 2n
q q
m2 − ∇21 Φ(x1 ) m2 − ∇22 Φ(x2 ) = W(x1 )W(x2 ). (C17) We now need two identities from Gradshteyn and Rhyzik [84]:
by 3.915-5 (p492), we have for Re ν > −1/2,
We now massage the LHS as follows. First, we are free to !ν
move the derivatives to the left, since ∇2i acts only on Φ(xi ):
Z π √
!
2 1
eiβ cos θ sin2ν θ dθ = π Γ ν+ Jν (β), (C27)
q q 0 β 2
m2 − ∇21 m2 − ∇22 Φ(x1 )Φ(x2 ) = W(x1 )W(x2 ). (C18)
and by 6.565-4 (p678), we have for −1 < Re ν < Re(2µ+2/3),
Next, we take expectation values of the fields on both sides: r > 0, and m > 0,

kν+1 Jν (kx) mν−µ xµ
q   Z
m2 − ∇21 m2 − ∇22 hΦ(x1 )Φ(x2 )i = hW(x1 )W(x2 )i . dk = Kν−µ (mx). (C28)
(C19) 0 k2 + m2 µ+1
 2µ Γ(µ + 1)

By definition, the two-point function on the LHS is of course Using the first formula with β = kx and ν = d/2 − 1, we find
the Euclidean quantum propagator
∞ ! d2 −1
kd−1
Z
2
hΦ(x1 )Φ(x2 )i = ∆d (x1 − x2 ). (C20) ∆d,n (x) = X J d −1 (kx) dk, (C29)
k + m2 n kx
2
 2
0
As for the two-point function on the RHS, it is none other than
the autocorrelation function of Gaussian white noise, where the prefactor is

hW(x1 )W(x2 )i = δ(d) (x1 − x2 ), (C21) Ωd−1 √


d−1 d
21− 2
!
d−1 2π 2 √
X= πΓ = π = d
. (C30)
(2π)d 2 (2π)d (2π) 2
which is by definition a delta function. We thus have
We can thus rewrite
q  
m2 − ∇21 m2 − ∇22 ∆d (x1 − x2 ) = δ(d) (x1 − x2 ). (C22)
d ∞ d
x1− 2
Z
k2
Since both two-point functions depend only on x = x1 − x2 , ∆d,n (x) =  J d (kx) dk. (C31)
we have ∇21 = ∇22 = ∇2 , and thus we are left with (2π) 2
d
0 k + m2 n 2 −1
2

The second identity with ν = d/2 − 1 and µ = n − 1 then yields


 
m2 − ∇2 ∆d (x) = δ(d) (x). (C23)
d d
This is exactly Eq. (C4), proving that it is consistent for the x1− 2 m 2 −n xn−1
∆d,n (x) = K d −n (mx), (C32)
Euclidean quantum field Φ(x) to also obey the SPDE (C16). (2π) 2 2n−1 Γ(n) 2
d

This is the quantum-statistical analog of the classical field


equation (C3): the presence of the white noise source term so we finally obtain
W(x) is what lends the Euclidean field its statistical nature.
Finally, when n = 2, the Gaussian random field Φ(x) obeys 21−n  m  d2 −n
∆d,n (x) = d
K d −n (mx). (C33)
(2π) (n − 1)! x
 
m2 − ∇2 Φ(x) = W(x),
2
(C24) 2

which is a linear SPDE and therefore straightforward to solve. Since K−ν (x) = Kν (x), this exactly reproduces Eq. (C13):
This is the reason why it is implemented in inoisy [43].
5. Derivation of the scalar propagator 21−n md−2n d
∆d,n (x) = d
(mx)n− 2 Kn− d (mx), (C34)
2
(2π) (n − 1)!2

Here, we derive Eqs. (C6) and (C13), which are equivalent


to the position-space formula (77) for the Matérn covariance. and setting n = 1 recovers Eq. (C6).
38

[1] C. Darwin, Proceedings of the Royal Society of London Series [29] A. Cárdenas-Avendaño, S. Nampalliwar, and N. Yunes, Classi-
A 249, 180 (1959). cal and Quantum Gravity 37, 135008 (2020), arXiv:1912.08062
[2] J. P. Luminet, A&A 75, 228 (1979). [gr-qc].
[3] H. C. Ohanian, American Journal of Physics 55, 428 (1987). [30] D. Ayzenberg and C. Bambi, arXiv e-prints , arXiv:2111.13918
[4] V. Bozza, General Relativity and Gravitation 42, 2269 (2010), (2021), arXiv:2111.13918 [astro-ph.HE].
arXiv:0911.2187 [gr-qc]. [31] G. N. Wong, B. S. Prather, V. Dhruv, B. R. Ryan,
[5] S. E. Gralla and A. Lupsasca, Phys. Rev. D 101, 044031 (2020), M. Mościbrodzka, C.-k. Chan, A. V. Joshi, R. Yarza, A. Ricarte,
arXiv:1910.12873 [gr-qc]. H. Shiokawa, J. C. Dolence, S. C. Noble, J. C. McKinney, and
[6] S. E. Gralla, D. E. Holz, and R. M. Wald, Phys. Rev. D 100, C. F. Gammie, ApJS 259, 64 (2022), arXiv:2202.11721 [astro-
024018 (2019), arXiv:1906.00873 [astro-ph.HE]. ph.HE].
[7] M. D. Johnson, A. Lupsasca, A. Strominger, G. N. Wong, [32] G. N. Wong, V. Dhruv, and B. S. Prather, “Characteristics of
S. Hadar, D. Kapec, R. Narayan, A. Chael, C. F. Gammie, Thermal Synchrotron Emission in Radiatively Inefficient Ac-
P. Galison, D. C. M. Palumbo, S. S. Doeleman, L. Blackburn, cretion Flows,” (2022), to appear.
M. Wielgus, D. W. Pesce, J. R. Farah, and J. M. Moran, Sci- [33] S. E. Gralla, Phys. Rev. D 102, 044017 (2020),
ence Advances 6, eaaz1310 (2020), arXiv:1907.04329 [astro- arXiv:2005.03856 [astro-ph.HE].
ph.IM]. [34] S. E. Gralla and A. Lupsasca, Phys. Rev. D 102, 124003 (2020),
[8] S. Hadar, M. D. Johnson, A. Lupsasca, and G. N. Wong, arXiv:2007.10336 [gr-qc].
Phys. Rev. D 103, 104038 (2021), arXiv:2010.03683 [gr-qc]. [35] S. E. Gralla, A. Lupsasca, and D. P. Marrone, Phys. Rev. D
[9] A. Chael, M. D. Johnson, and A. Lupsasca, ApJ 918, 6 (2021), 102, 124004 (2020), arXiv:2008.03879 [gr-qc].
arXiv:2106.00683 [astro-ph.HE]. [36] G. N. Wong, ApJ 909, 217 (2021), arXiv:2009.06641 [astro-
[10] H. Paugnat, A. Lupsasca, F. Vincent, and M. Wielgus, arXiv ph.HE].
e-prints , arXiv:2206.02781 (2022), arXiv:2206.02781 [astro- [37] Z. Gelles, B. S. Prather, D. C. M. Palumbo, M. D. John-
ph.HE]. son, G. N. Wong, and B. Georgiev, ApJ 912, 39 (2021),
[11] F. H. Vincent, S. Gralla, A. Lupsasca, and M. Wielgus, arXiv arXiv:2103.07417 [astro-ph.HE].
e-prints , arXiv:2206.12066 (2022), arXiv:2206.12066 [astro- [38] C. J. White, arXiv e-prints , arXiv:2203.15963 (2022),
ph.HE]. arXiv:2203.15963 [astro-ph.HE].
[12] H. Falcke, F. Melia, and E. Agol, ApJ 528, L13 (2000), [39] E. Teo, General Relativity and Gravitation 53, 10 (2021),
arXiv:astro-ph/9912263 [astro-ph]. arXiv:2007.04022 [gr-qc].
[13] R. Narayan, M. D. Johnson, and C. F. Gammie, ApJ 885, L33 [40] S. E. Gralla and A. Lupsasca, Phys. Rev. D 101, 044032 (2020),
(2019), arXiv:1910.02957 [astro-ph.HE]. arXiv:1910.12881 [gr-qc].
[14] V. Perlick and O. Y. Tsupko, Phys. Rep. 947, 1 (2022), [41] J. Dexter and E. Agol, ApJ 696, 1616 (2009), arXiv:0903.0620
arXiv:2105.07101 [gr-qc]. [astro-ph.HE].
[15] J. M. Bardeen, in Black Holes (Les Astres Occlus), edited by [42] X.-L. Yang and J.-C. Wang, A&A 561, A127 (2014),
C. Dewitt and B. S. Dewitt (Gordon and Breach Science Pub- arXiv:1311.4436 [gr-qc].
lishers, 1973) pp. 215–239. [43] D. Lee and C. F. Gammie, ApJ 906, 39 (2021),
[16] K. Akiyama et al. (Event Horizon Telescope Collaboration), arXiv:2011.07151 [astro-ph.IM].
ApJ 875, L1 (2019), arXiv:1906.11238 [astro-ph.GA]. [44] A. M. Beloborodov, ApJ 566, L85 (2002), arXiv:astro-
[17] K. Akiyama et al. (Event Horizon Telescope Collaboration), ph/0201117 [astro-ph].
ApJ 930, L12 (2022). [45] D. E. A. Gates, S. Hadar, and A. Lupsasca, Phys. Rev. D 102,
[18] K. Akiyama et al. (Event Horizon Telescope Collaboration), 104041 (2020), arXiv:2009.03310 [gr-qc].
ApJ 875, L5 (2019), arXiv:1906.11242 [astro-ph.GA]. [46] D. Kapec and A. Lupsasca, Classical and Quantum Gravity 37,
[19] P. Arras, P. Frank, P. Haim, J. Knollmüller, R. Leike, M. Rei- 015006 (2020), arXiv:1905.11406 [hep-th].
necke, and T. Enßlin, Nature Astronomy 6, 259 (2022). [47] B. Carter, Physical Review 174, 1559 (1968).
[20] C. L. Carilli and N. Thyagarajan, ApJ 924, 125 (2022), [48] K. P. Rauch and R. D. Blandford, ApJ 421, 46 (1994).
arXiv:2111.11626 [astro-ph.CO]. [49] S. E. Vázquez and E. P. Esteban, Nuovo Cimento B Serie 119,
[21] W. Lockhart and S. E. Gralla, MNRAS 509, 3643 (2022), 489 (2004), arXiv:gr-qc/0308023 [gr-qc].
arXiv:2107.06948 [astro-ph.HE]. [50] L.-X. Li, E. R. Zimmerman, R. Narayan, and J. E. McClintock,
[22] K. Akiyama et al. (Event Horizon Telescope Collaboration), ApJS 157, 335 (2005), arXiv:astro-ph/0411583 [astro-ph].
ApJ 930, L16 (2022). [51] L.-X. Li, R. Narayan, and J. E. McClintock, ApJ 691, 847
[23] W. Lockhart and S. E. Gralla, MNRAS 517, 2462 (2022), (2009), arXiv:0809.0866 [astro-ph].
arXiv:2208.09989 [astro-ph.HE]. [52] A. Lupsasca, D. Kapec, Y. Shi, D. E. A. Gates, and A. Stro-
[24] S. E. Gralla, Phys. Rev. D 103, 024023 (2021), minger, Proceedings of the Royal Society of London Series A
arXiv:2010.08557 [astro-ph.HE]. 476, 20190618 (2020).
[25] G. Lara, S. H. Völkel, and E. Barausse, Phys. Rev. D 104, [53] E. Himwich, M. D. Johnson, A. Lupsasca, and A. Strominger,
124041 (2021), arXiv:2110.00026 [gr-qc]. Phys. Rev. D 101, 084020 (2020), arXiv:2001.08750 [gr-qc].
[26] A. M. Bauer, A. Cárdenas-Avendaño, C. F. Gammie, and [54] R. Narayan et al. (Event Horizon Telescope Collaboration), ApJ
N. Yunes, ApJ 925, 119 (2022), arXiv:2111.02178 [gr-qc]. 912, 35 (2021), arXiv:2105.01804 [astro-ph.HE].
[27] K. Akiyama et al. (Event Horizon Telescope Collaboration), [55] S. Chandrasekhar, The mathematical theory of black holes (Ox-
ApJ 875, L6 (2019), arXiv:1906.11243 [astro-ph.GA]. ford University Press, 1983).
[28] K. Akiyama et al. (Event Horizon Telescope Collaboration), [56] J. M. Bardeen, W. H. Press, and S. A. Teukolsky, ApJ 178, 347
ApJ 930, L17 (2022). (1972).
39

[57] E. Hackmann, Geodesic equations in black hole space-times ries B (Statistical Methodology) 73, 423 (2011),
with cosmological constant, Ph.D. thesis, Bremen University https://rss.onlinelibrary.wiley.com/doi/pdf/10.1111/j.1467-
(2010). 9868.2011.00777.x.
[58] G. Compère, Y. Liu, and J. Long, Phys. Rev. D 105, 024075 [77] G.-A. Fuglstad, F. Lindgren, D. Simpson, and H. Rue, Statistica
(2022), arXiv:2106.03141 [gr-qc]. Sinica 25, 115 (2015), arXiv:1304.6949 [stat.ME].
[59] S. Hadar, D. Kapec, A. Lupsasca, and A. Strominger, Classi- [78] X. Guan, C. F. Gammie, J. B. Simon, and B. M. Johnson, ApJ
cal and Quantum Gravity 39, 215001 (2022), arXiv:2205.05064 694, 1010 (2009), arXiv:0901.0273 [astro-ph.GA].
[gr-qc]. [79] S. C. Noble and J. H. Krolik, ApJ 703, 964 (2009),
[60] K. Beckwith and C. Done, MNRAS 359, 1217 (2005), arXiv:0907.1655 [astro-ph.HE].
arXiv:astro-ph/0411339 [astro-ph]. [80] K. Chatterjee, Z. Younsi, M. Liska, A. Tchekhovskoy, S. B.
[61] O. James, E. v. Tunzelmann, P. Franklin, and K. S. Markoff, D. Yoon, D. van Eijnatten, C. Hesp, A. Ingram,
Thorne, Classical and Quantum Gravity 32, 065001 (2015), and M. B. M. van der Klis, MNRAS 499, 362 (2020),
arXiv:1502.03808 [gr-qc]. arXiv:2002.08386 [astro-ph.GA].
[62] S. E. Gralla, A. Lupsasca, and A. Strominger, MNRAS 475, [81] L. I. Gurvits, Z. Paragi, R. I. Amils, I. van Bemmel, P. Boven,
3829 (2018), arXiv:1710.11112 [astro-ph.HE]. V. Casasola, J. Conway, J. Davelaar, M. C. Dı́ez-González,
[63] D. E. A. Gates, S. Hadar, and A. Lupsasca, Phys. Rev. D 103, H. Falcke, R. Fender, S. Frey, C. M. Fromm, J. D. Gallego-
044050 (2021), arXiv:2010.07330 [gr-qc]. Puyol, C. Garcı́a-Miró, M. A. Garrett, M. Giroletti, C. Goddi,
[64] V. Bozza, Phys. Rev. D 78, 063014 (2008), arXiv:0806.4102 J. L. Gómez, J. van der Gucht, J. C. Guirado, Z. Haiman,
[gr-qc]. F. Helmich, B. Hudson, E. Humphreys, V. Impellizzeri,
[65] O. Y. Tsupko, Phys. Rev. D 106, 064033 (2022), M. Janssen, M. D. Johnson, Y. Y. Kovalev, M. Kramer,
arXiv:2208.02084 [gr-qc]. M. Lindqvist, H. Linz, E. Liuzzo, A. P. Lobanov, I. López-
[66] H.-Y. Pu, K. Akiyama, and K. Asada, ApJ 831, 4 (2016), Fernández, I. Malo-Gómez, K. Masania, Y. Mizuno, A. V.
arXiv:1608.03035 [astro-ph.HE]. Plavin, R. T. Rajan, L. Rezzolla, F. Roelofs, E. Ros, K. L. J.
[67] C. T. Cunningham, ApJ 202, 788 (1975). Rygl, T. Savolainen, K. Schuster, T. Venturi, M. Verkouter,
[68] D. H. Roberts, J. F. C. Wardle, and L. F. Brown, ApJ 427, 718 P. de Vicente, P. N. A. M. Visser, M. C. Wiedner, M. Wielgus,
(1994). K. Wiik, and J. A. Zensus, Acta Astronautica 196, 314 (2022),
[69] C. T. Cunningham and J. M. Bardeen, ApJ 183, 237 (1973). arXiv:2204.09144 [astro-ph.IM].
[70] A. E. Broderick and A. Loeb, MNRAS 363, 353 (2005), [82] P. Kurczynski, M. D. Johnson, S. S. Doeleman, K. Haworth,
arXiv:astro-ph/0506433 [astro-ph]. E. Peretz, T. K. Sridharan, B. Bilyeu, L. Blackburn, D. Boroson,
[71] A. E. Broderick and A. Loeb, MNRAS 367, 905 (2006), A. Brosius, R. Butler, D. Caplan, K. Chatterjee, P. Cheimets,
arXiv:astro-ph/0509237 [astro-ph]. D. D’Orazio, T. Essinger-Hileman, P. Galison, R. Gamble,
[72] N. Hamaus, T. Paumard, T. Müller, S. Gillessen, F. Eisen- S. Hadar, T. Hoerbelt, H. Jiao, J. Kauffmann, R. Lafon, C.-P.
hauer, S. Trippe, and R. Genzel, ApJ 692, 902 (2009), Ma, G. Melnick, N. R. Newbury, S. Noble, D. Palumbo, L. Par-
arXiv:0810.4947 [astro-ph]. itsky, D. Pesce, L. Petrov, J. Piepmeier, C. J. Roberts, B. Robin-
[73] M. Zamaninasab, A. Eckart, G. Witzel, M. Dovciak, V. Karas, son, C. Shieler, J. Small, N. Spellmeyer, P. Tiede, J. Verniero,
R. Schödel, R. Gießübel, M. Bremer, M. Garcı́a-Marı́n, D. Kun- J. Wang, M. Wielgus, E. Wollack, G. N. Wong, and G. Yang, in
neriath, K. Mužić, S. Nishiyama, N. Sabha, C. Straubmeier, Space Telescopes and Instrumentation 2022: Optical, Infrared,
and A. Zensus, A&A 510, A3 (2010), arXiv:0911.4659 [astro- and Millimeter Wave, Society of Photo-Optical Instrumenta-
ph.GA]. tion Engineers (SPIE) Conference Series, Vol. 12180, edited by
[74] J. M. Bardeen, J. R. Bond, N. Kaiser, and A. S. Szalay, ApJ L. E. Coyle, S. Matsuura, and M. D. Perrin (2022) p. 121800M.
304, 15 (1986). [83] R. F. Penna, A. Kulkarni, and R. Narayan, A&A 559, A116
[75] P. Guttorp and T. Gneiting, Biometrika 93, 989 (2013), arXiv:1309.3680 [astro-ph.HE].
(2006), https://academic.oup.com/biomet/article- [84] I. S. Gradshteyn, I. M. Ryzhik, A. Jeffrey, and D. Zwill-
pdf/93/4/989/593552/934989.pdf. inger, Table of Integrals, Series, and Products (Academic Press,
[76] F. Lindgren, H. v. Rue, and J. Lindström, 2007).
Journal of the Royal Statistical Society: Se-

You might also like