Vous êtes sur la page 1sur 22

Tidal evolution of dark matter annihilation rates in subhalos

M. Sten Delos∗
Department of Physics and Astronomy, University of North Carolina at Chapel Hill,
Phillips Hall CB3255, Chapel Hill, North Carolina 27599, USA

Dark matter halos grow by hierarchical clustering as they merge together to produce ever larger
structures. During these merger processes, the smaller halo can potentially survive as a subhalo
of the larger halo, so a galaxy-scale halo today likely possesses a rich abundance of substructure.
This substructure can greatly boost the rate of dark matter annihilation within the host halo,
but the precise magnitude of this boost is clouded by uncertainty about the survival prospects
of these subhalos. In particular, tidal forces gradually strip material from the subhalos, reducing
their annihilation signals and potentially destroying them. In this work, we use high-resolution
idealized N -body simulations to develop and tune a model that can predict the impact of this tidal
arXiv:1906.10690v1 [astro-ph.CO] 25 Jun 2019

evolution on the annihilation rates within subhalos. This model predicts the time-evolution of a
subhalo’s annihilation rate as a function of three physically motivated parameters of the host-subhalo
system: the energy injected into subhalo particles per orbit about the host, the ratio of stretching to
compressive tidal forces, and the radial distribution of tidal heating within the subhalo. Our model
will improve the accuracy of predictions of the magnitude and morphology of annihilation signals
from dark matter substructure. Additionally, our parametrization can describe the time-evolution of
other subhalo properties, so it has implications for understanding aspects of subhalo tidal evolution
beyond the annihilation rate.

I. INTRODUCTION through hierarchical merging, it accretes smaller halos


that survive as subhalos within the larger host. These
Despite overwhelming evidence for the existence of subhalos gradually lose mass due to the influence of tidal
dark matter (e.g. [1–3]), its microphysical details remain forces from the host (e.g. [14, 15]), surviving until either
unknown. Numerous models have been considered, but they are completely stripped or dynamical friction causes
none have been experimentally confirmed (see Refs. [4–8] them to sink into the host’s center [16]. However, for suf-
for reviews). However, to explain the present abundance ficiently small subhalos, dynamical friction is inefficient
of dark matter, a large class of models, including the pop- [15]. Moreover, numerous analyses have found that if
ular weakly interacting massive particle [9], propose that the subhalos possess divergent central density, as in the
dark matter was pair-produced from the thermal plasma NFW profile, then tidal forces may never fully strip them
in the hot early Universe. In this scenario, the dark mat- [17–20].
ter can annihilate back into standard-model particles to- Thus, a galactic-scale dark matter halo is likely to pos-
day, leading to prospects for the detection of high-energy sess a multitude of subhalos. This substructure can sig-
gamma rays or other annihilation products (e.g. [10]). nificantly boost the rate of dark matter annihilation, de-
The rate of dark matter annihilation scales as the pending on the scale of the smallest halos and when they
square of the dark matter density, so it is strongly sen- form [20–34] (see Ref. [35] for a recent review). When the
sitive to the spatial distribution of the dark matter. At smallest halos are microhalos of roughly earth mass, anni-
galactic scales and above, this spatial distribution is well hilation rates may be boosted by a factor of ∼ 10 [31, 34]
understood. Initially overdense patches in the Universe relative to those expected in the absence of substructure,
collapse into gravitationally bound dark matter halos, assuming that these halos arise from primordial density
which thereafter merge to produce successively larger fluctuations comparable to the large-scale fluctuations in-
structures. Numerical simulations demonstrate that the ferred from the cosmic microwave background (e.g. [3]).
spherically averaged mass distributions of the resulting Moreover, this boost can be raised by orders of magni-
dark matter halos are well described by the NFW den- tude by cosmological scenarios that amplify small-scale
sity profile [11, 12], density fluctuations and thereby lead to earlier and more
ρs abundant microhalo formation. Such scenarios include
ρ(r) = , (1) a period of domination by a heavy species [36–39] or a
(r/rs )(1 + r/rs )2 fast-rolling scalar field [40] prior to nucleosynthesis, along
which has scale parameters rs and ρs . Baryonic effects with a variety of inflationary models [41–63]. In these
may subsequently alter this density profile (e.g. [13]). cases, the high density within these microhalos causes
However, at subgalactic scales, the spatial distribution them to completely dominate any dark matter annihila-
of the dark matter is less clear. As a halo is built up tion signal (e.g. [64–66]).
Unfortunately, all estimates of substructure’s boost to
annihilation rates are subject to uncertainties about the
impact of tidal effects on subhalos. Cosmological simula-
∗ Electronic address: delos@unc.edu tions cannot resolve subhalos that are much smaller than
2

the host, and those that are resolved are prone to artifi- hilation rate, characterized by its J-factor1
cial destruction [67–69]. Numerous semianalytic models Z
have been developed to describe the dynamical evolu- J ≡ ρ2 dV, (2)
tion of subhalos; Refs. [70–77] model a subhalo’s loss of
mass due to tidal stripping, and Refs. [18, 78] predict
as a function of its orbit about the host. To assist the
the impact of this mass loss on the halo’s density pro-
application of our model, we supply convenient fitting
file. However, these models are typically tuned to the
functions.
results of cosmological simulations, so their predictions
This article is organized as follows. In Section II, we
are affected by the artificial subhalo disruption occurring
detail how we carry out our N -body simulations. Sec-
therein. They cannot fully reproduce the results of ide-
tion III qualitatively discusses the trajectory of a sub-
alized simulations [68, 79].
halo’s J-factor, interpreting simulation trends physically
Meanwhile, calculations of the dark matter annihi-
and motivating our model. In Section IV, we develop
lation rate in substructure have employed a number
our predictive model for the evolution of a subhalo’s J-
of different treatments of tidal evolution. Some, like
factor, and Section V summarizes the model and dis-
Refs. [33, 34], employ a combination of the models above
cusses limitations and extensions. In Section VI, we com-
to predict the time-evolution of subhalo density profiles.
pare the model’s predictions to those of previous semian-
Others, like Refs. [30–32, 39], employ simpler models, of-
alytic models. Section VII concludes, after which we sup-
ten either truncating subhalos at a characteristic tidal
ply a variety of appendices. Appendix A supplies further
radius or formulating a destruction condition for subha-
details about our simulations. Appendix B presents fit-
los and assuming the survivors are unaltered. Still oth-
ting formulas and other computational details that aid in
ers, like Refs. [25–28, 64–66] neglect tidal disruption of
applying our model. Appendix C tests our model against
substructure altogether.
a publicly available simulation library [79]. Finally, in
Our work is motivated by this context. Since cos- Appendix D, we observe that our model can be adapted
mological simulations cannot resolve the smallest sub- to describe the evolution of subhalo properties beyond
structures, we follow Refs. [18, 69, 78–81] in using ideal- the J-factor.
ized simulations of an N -body subhalo inside an analytic
galactic potential. However, unlike these works, we focus
on understanding the impact of tides on the subhalo’s II. SIMULATIONS
annihilation rate, a goal that requires significantly bet-
ter resolution than has been attained in previous studies.
Owing to the difference in scales between a host and
Moreover, previous works have focused on understanding
its smallest subhalos, the computational challenge in
the evolution of subhalos of scales resolvable in cosmo-
simulating subhalo evolution in cosmological context is
logical simulations, such as halos associated with dwarf
formidable. A number of previous works have addressed
galaxies within a galactic halo. Accordingly, they probe
this problem by simulating an N -body subhalo inside an
only the subhalo properties and orbits that are found
analytic host potential [18, 69, 78–81]; our approach is
in such simulations. For instance, Ref. [79] only studies
similar but differs in one key step. Instead of placing a
subhalos orbiting above the host’s scale radius rs . In con-
subhalo in orbit about the host potential, we subject the
trast, we seek to probe the full range of subhalos down to
subhalo directly to the time-dependent tidal force field
the smallest microhalos, which span a far broader range
experienced by an analytic orbit about the host. This
of properties and orbits.
procedure minimizes the impact of numerical precision
Using the results of 51 high-resolution N -body sim-
errors that can result from differences in scale between
ulations, we develop a physically motivated model that
the subhalo’s orbital and internal dynamics. In this sec-
can predict the time-evolution of a subhalo’s annihilation
tion, we detail that procedure and present qualitative
rate due to tidal effects. In the process, we isolate three
results.
physical variables that determine this evolution:
We assume that both the host and the subhalo possess
1. The energy injected by tidal forces into subhalo the NFW density profile given by Eq. (1). While there
particles over the course of each orbit about the is evidence that many galactic halos possess constant-
host, in units of the particle’s binding energy to density cores instead of the NFW profile’s cusp [82], at
the subhalo; least some galactic halos appear to be cuspy [83]. Ad-
ditionally, while microhalos are expected to form with
2. The ratio of stretching (radial) tidal forces to com- ρ ∝ r−3/2 inner profiles [24–26, 65, 84–88], it is likely that
pressive (tangential) tidal forces; mergers will drive their inner cusps toward the ρ ∝ r−1
of the NFW profile [87–89].
3. The range of radii in the subhalo across which ma-
terial is heated by tidal forces, which is set by the
shape of the subhalo’s orbit.
1 We assume the dark matter annihilation cross-section is velocity-
This model predicts the suppression of a subhalo’s anni- independent in the nonrelativistic limit.
3

To model the host’s tidal field, we begin with an apocenter, t = 0.00 Gyr pericenter, t = 0.02 Gyr
analytically-computed orbit, described by the time-
dependent vector R(t) pointing from host center to sub-
halo center. The tidal acceleration at position r relative
to the subhalo center is2
216 pc
dF r − (r · R̂)R̂
Ftidal (r) = − (r · R̂)R̂ − F (R) (3)
dR R 48 pc

at linear order in r/R, where F (R) is the force profile


of the host, R = |R|, and R̂ = R/R. We modified the
Gadget-2 N-body simulation code [90, 91] to include
this tidal acceleration.
We prepare the initial N -body subhalo with an NFW 48 pc
profile by drawing particles from an isotropic distribu-
tion function computed using the fitting form in Ref. [92]. 215 pc
Additionally, we sample the subhalo’s central region at
increased resolution: particles whose orbital pericenters
are below rs /3, where rs is the subhalo scale radius,
have 1/64 the mass and 64 times the number density
of the other particles. We cut off the density profile at apocenter, t = 0.70 Gyr pericenter, t = 0.73 Gyr
r = 500rs ; subhalo particles this far out are stripped im-
mediately, so as long as the cutoff radius is much larger FIG. 1. The projected density field of a microhalo simulated
than rs , the precise choice makes no difference. We rep- in orbit about a galactic halo. The width of each frame is
resent the subhalo using a total of 8 million particles, and 0.03 pc, and the arrow indicates the direction and distance to
roughly 70% of them, carrying roughly 4% of the total the host center. The density is computed using a k-nearest-
mass, are high-resolution particles. All of our subhalos neighbor density estimate with k = 50 and is plotted with a
logarithmic color scale (lighter is denser).
have rs ' 10−6 Rs , where Rs is the scale radius of the
host, but as we will soon discuss, the precise choice of rs
has no impact on dynamics.
For our simulations, we consider a variety of orbits
1011
ρ (M kpc−3 )

about the host. An orbit in a spherically symmetric po-


tential is characterized by two parameters: energy E and
angular momentum L, or equivalently, a scale parameter 109 t=0
and a shape parameter. For convenience, we use the cir-
cular orbit radius3 Rc , defined as the radius of the circu- 107 t = 0.9 Gyr
lar orbit with energy E, and the “circularity” η = L/Lc ,
where Lc is the angular momentum of the circular orbit 105
with the same energy. In each simulation, the subhalo 10−7 10−6 10−5
begins at its orbital apocenter.
r (kpc)
Figure 1 illustrates a simulation executed through this
arrangement. The host has scale radius Rs = 0.8 kpc
FIG. 2. The density profile evolution of the halo depicted in
and scale density Ps = 5 × 107 M /kpc3 , while the N -
Fig. 1. 91% of the initial mass of the microhalo is stripped
body subhalo is initially a microhalo with scale radius by t = 0.9 Gyr, but the central density profile is largely un-
rs = 6 × 10−7 kpc that has ρs /Ps = 1285 times the scale affected.
density of the host. The subhalo orbit has Rc = 0.15
kpc and η = 0.5. The simulation runs through 18 orbits
about the host, and Fig. 2 plots the density profile of the
subhalo at each apocenter. 91% of the subhalo’s mass is Our goal in this work is to understand how the annihi-
stripped by simulation termination, but its central den- lation signal decays due to tidal effects. For this purpose,
sity profile is largely unscathed. we consider the J-factor, Eq. (2), integrated over the sub-
halo mass distribution. This J-factor is the factor in the
annihilation rate that depends on mass distribution, and
2 We experimented with using the full tidal force Ftidal (r) = Appendix A discusses our procedure to extract it from
F (R + r) − F (R), but because r  R in our simulations, it the simulations. We also show in Appendix A that the
offers no advantage; moreover, it is less numerically stable due resulting J-factor trajectories are converged with respect
to the subtraction of two close numbers. to simulation parameters.
3 Note that Rc is roughly the time-averaged radius: for a power-
law potential φ(R) ∝ Rn , Rc = hRn i1/n . See Appendix B for a Finally, we conclude this section by discussing the ap-
more precise relationship for NFW profiles. plicability of our simulation results. The linearized tidal
4

force in Eq. (3) is valid for r  R, or


rs  Rc . (4) Rc ρs
100 = 2.37
R s Ps
If this condition is satisfied, then the impact of tides is η = 0.5
independent of rs , and our simulation results are appli-

J/Jinit
cable. Additionally, since we do not simulate the host
halo’s dynamics, we cannot account for dynamical fric- Rc /Rs ρs /Ps
tion. Reference [79] found that dynamical friction4 has 0.0018 1285
minimal impact on a subhalo’s mass-evolution for host- 0.00018 12848
−1
10 0.015 153
to-subhalo mass ratios M/m > ∼ 100. However, since
subhalos relevant to dark matter annihilation may or-
bit over significantly longer time scales than considered 100 101
n = t/T
in Ref. [79], it is also useful to have an analytic estimate
for when dynamical friction can be neglected. It follows
from the analysis in Ref. [15] that for a subhalo that ac- FIG. 3. Trajectories of the J-factor for different systems
Rc ρ s
cretes onto a host at redshift z, dynamical friction can with the same x̃ = R s Ps
and η. Scaled to the orbital period,
these systems all have the same trajectory.
safely be neglected as long as the host-to-subhalo mass
ratio M/m satisfies
M/m > A. Trends in the simulations
10(1 + z)3/2 . (5)
ln(M/m) ∼
We first inspect the results of selected simulations in
Our results may be considered applicable as long as Eqs. order to find trends in the behavior of J. As a further
(4) and (5) are satisfied, but we remark that if one is simplification, we focus on the Rc  Rs regime. The
satisfied, then the other likely is too. host potential is self-similar in this regime, reducing the
tidal evolution problem in two additional ways.
III. TRENDS IN THE TIDAL EVOLUTION 1. The orbital radius Rc is degenerate with properties
of the host and subhalo. For instance, reducing
In this section, we explore trends in the evolution of the orbital radius is equivalent to making the host
J as a function of system and time and attempt to ex- denser.
plain them physically. Our goal is to find the J-factor as
2. Orbits with the same circularity η have the same
a function of time t, orbital parameters Rc and η, sub-
shape: they are rescaled versions of one another.
halo parameters rs and ρs , and host parameters Rs and
Ps . The dimensionality of this space is large, but some The first simplification further reduces the parameter
immediate simplifications are evident. space so that in this self-similar regime, there are only
1. As long as the subhalo is much smaller than its two parameters:5
orbit, or rs  Rc , the value of rs has no impact Rc ρs
on dynamics. All of our simulated subhalos have x̃ ≡ (6)
R s Ps
rs ' 10−6 Rs , which leads to rs  Rc for all orbits
we consider. and η. Figure 3 shows the success of this parameter re-
duction: different systems with the same x̃ and η follow
2. If instead of time t we use the orbit count n = precisely the same J(n) trajectories. Meanwhile, the sec-
t/T , where T is the orbital period, then the overall ond simplification allows us to isolate the impact of these
density scale has no impact on dynamics, and only two parameters: we can vary the “reduced orbital radius”
the ratio ρs /Ps enters. x̃ without altering the orbit’s shape.
3. The overall size of the host-subhalo system is irrel- We first investigate the impact of orbit shape. Figure 4
evant, so only the ratio Rc /Rs affects dynamics. shows the trajectory of the J-factor for several values of
η. We see immediately that the J-factor oscillates with
We have verified that all of these simplifications are borne the orbital period with larger amplitude for more eccen-
out in our simulations. Hence, if Jinit is the initial J- tric orbits. This trend is explained by noting that all
factor, then J/Jinit is now a function of time n = t/T tidal forces are compressive in the self-similar regime, so
and just three system parameters: ρs /Ps , Rc /Rs , and η. the subhalo becomes most compact near the orbital peri-
center. But more interesting trends arise in the broader
4 Specifically, Ref. [79] studied dynamical self-friction, or the dy-
namical friction that results from the subhalo’s own tidal tail.
This friction can be considerably more efficient than that result- 5 We reserve x (without the tilde) for later use as a modified version
ing from the host’s material alone [93, 94]. of x̃.
5

η = 0.997 ρs /Ps = 1285


1.0 η = 0.96 Rc /Rs = 0.018
η = 0.5 1.0
η = 0.1 η = 0.1
η = 0.5

−d log J/d log t


η = 0.96
η = 0.997
J/Jinit

0.5 0.5

ρs /Ps = 1285 0.2


Rc /Rs = 0.018
0.2
100 101 100 101
n = t/T n = t/T

FIG. 4. Trajectories of the J-factor for different orbital FIG. 5. The impact of orbit shape on the J-factor trajectory.
shapes. The J-factor oscillates with the orbital period; the This figure plots the logarithmic slope of the orbital period-
dashed lines show fits using Eq. (7) (for |d ln J/d ln t| < 1). averaged trajectory to make the trends clearer. Notably, the
slope runs more rapidly for more eccentric orbits. The points
show the simulation results, while the lines correspond to fits
using Eq. (7) (for |d ln J/d ln t| < 1).
time-evolution. Figure 5 plots the running power-law in-
dex d ln J/d ln t of the J-factor with time, and the equa-
tion B. Physical interpretation
d ln J
= −bn1−c (7) Behavior similar to that of Eq. (9) can be reproduced
d ln t
in a toy model. Suppose the subhalo has potential
describes the evolution of this index reasonably well as φ(r) ∝ rγ up to an additive constant; for instance, an
long as |d ln J/d ln t| < O(1). Here, b > 0 and c > 0 are NFW profile would have γ = 1 for r  rs . Now discretize
constant parameters, and c is smaller for more eccentric time, perhaps as a count of orbits, letting rn be the radius
orbits. Figures 4 and 5 also show fits to the J-factor of the subhalo’s outer boundary at time tn . Any material
trajectories using this form, which determines J(n) up outside rn at time tn is free, fixing the additive constant
to a constant multiple. The exponent in Eq. (7) is so in the potential: φn (r) ∝ rγ − rnγ (where φ ≥ 0 implies
defined because it leads to the more evocative expression freedom). Now suppose that at each time step, particles
in the subhalo experience an injection of energy ∆E ∝ rα
1 dJ due to tidal forces, and any radius with ∆E + φ > 0 no
= −bn−c . (8) longer belongs to the halo. For instance, if the time step
J dn is constant, then α = 2 since the energy injection is pro-
portional to the square of the tidal force, which is in turn
If c = 0, this equation tells us that the J-factor would
proportional to radius. This tidal heating rule leads to
decay by the same factor e−b over each orbit. The param-
the evolution equation φn (rn+1 ) + ∆E(rn+1 ) = 0, or
eter c, when c > 0, accommodates some physical process
by which tidal effects lose efficiency over time. γ
rn+1 − rnγ + f rn+1
α
= 0, (10)
We show the impact of orbital radius in Fig. 6, which
plots the trajectories of J and d ln J/d ln t for a variety where f incorporates all of the proportionality constants.
of reduced orbital radii x̃ ranging from 0.7 to 230. Ev- For simplicity, we may assume r0 = 1, absorbing its
idently, x̃ affects the initial decay rate of the J-factor, dimensionful value into f . For α > γ, Eq. (10) obeys
described by the parameter b in Eq. (7), without altering
(
the rate at which the decay slows over time. This figure ∆ ln r −f n/γ, fn  1
also shows more clearly that there is a steepness limit to ' (11)
∆ ln n −1/(α − γ), f n  1
the decay of the J-factor:

 where ∆ denotes the discrete difference across time steps.


d ln J
' − min bn1−c , B , (9) If J ∝ rβ , then
d ln t
∆ ln J
where B ∼ O(1). ' − min {bn, B} (12)
∆ ln n
6

combination of these two processes—the increasing den-


1.0 ρs /Ps = 1285 sity of the subhalo as its radius drops and the J-factor
η = 0.5 of its tidal stream—may explain the behavior in Figs.
5 and 6 wherein |d ln J/d ln t| initially shallows toward
0.6 some value larger than 1 before subsequently returning
J/Jinit

Rc /Rs back to 1. Note, however, that the precise evolution of


0.4 0.18 the J-factor in the |d ln J/d ln t| ∼ 1 regime is of little
0.055 consequence. By this point, the subhalo has already lost
0.018
0.0055 most of its J-factor and contributes little to annihilation
0.0018 signals.
0.00055
2.0 The physical explanation for the c > 0 behavior ob-
ρs /Ps = 1285 served in the simulations remains unclear. However, it is
η = 0.5 necessarily connected to how the shape of the subhalo’s
density profile changes in response to tidal effects (see
1.0
Rc /Rs Fig. 2), which the toy model does not account for. In
a more complete picture, the rate dJ/dn of tidal evolu-
−d log J/d log t

0.00055
0.0018 tion should be sensitive only to the instantaneous host-
0.5 0.0055 subhalo system with no explicit dependence on the time
0.018 n. Hence, it should be possible to replace the factor n−c
0.055
0.18 in Eq. (8) with a function of the subhalo’s density profile
(and other properties of the system). However, in the
0.2 |d ln J/d ln t| < 1 regime, the total change in J is much
smaller than J itself, and if we neglect changes in the
shape of the density profile, then any parameter of the
0.1 density profile (e.g. ρs or rs ) must experience a similarly
small change. Since the factor n−c can change by an or-
100 101 der of magnitude in the same regime, it is not possible,
n = t/T except in a very contrived way, to replace this factor with
a function of the density profile.
Thus, the c > 0 behavior must follow from changes in
FIG. 6. The impact of orbital radius on the J-factor trajec- the density profile’s shape. As a result of these changes,
tory. Top: The J-factor trajectory as in Fig. 4 (solid lines);
the density profile picks up new parameters that can po-
the dashed lines show fits using Eq. (7). Bottom: The loga-
rithmic slope of the orbital period-averaged trajectory, as in tentially vary wildly without significantly altering J, and
Fig. 5; the points show the simulation results, while the lines the factor n−c can be replaced with a function of those
correspond to the fits. We only fit the trajectories that do parameters. For instance, by introducing a new parame-
not pass |d ln J/d ln t| = 1. ter q, we can write
1 dJ 1 dq
= −bq, = −cq 1/c . (13)
J dn q dn
with b = βf /γ and B = β/(α − γ). The two separate
regimes arise physically because when f n  1, the total This system no longer has explicit time-dependence, but
radius change |rn − r0 |  r0 . Since the radius does not if q = 1 when n = 1, then it is equivalent to Eq. (8).
change appreciably, the efficiency of tidal heating does
not change, so r and J drop by the same fraction in each
orbit. However, when f n  1, |rn − r0 | ∼ r0 . In this IV. MODELING THE TIDAL EVOLUTION
case, the radius is decreasing significantly, which implies
that the density at the halo’s shrinking outer boundary Motivated by the results of the previous section, we
is increasing and hence that the halo is becoming more seek a model of the form
difficult to strip.  
J 1 
This toy model has reproduced Eq. (7) with c = 0, ln =b a− n1−c − 1 (14)
Jinit 1−c
successfully explaining the apparent upper limit in
|d ln J/d ln t|. We remark, however, that there is an- for the case where |d ln J/d ln t| < 1. The parameters b
other, completely different physical reason to expect a and c follow immediately from Eq. (7), and we have in-
upper limit in |d ln J/d ln t|: an unbound tidal stream serted another parameter a to fix the overall normaliza-
grows in length L as L ∝ t. Hence, its volume grows tion. Our goal is now to relate a, b, and c to the param-
as V ∝ t, so its J-factor drops as J ∝ M 2 /V ∝ t−1 . eters of the host-subhalo system. For this purpose, we
Once a subhalo has been stripped to the point that its use the results of 51 idealized N -body simulations that
own J-factor is dwarfed by that of its tidal stream, the we carried out as described in Section II. The parameter
J-factor of the subhalo remnant decays as J ∝ t−1 . The space covered by these simulations is depicted in Fig. 7.
7

1.0 Rc /Rs < 0.3


0.3
0.99 1
10
0.96

0.90
100 0.1

b
Rc /Rs
0.8
η

η
10−1 0.0 0.2 0.4 0.6 0.8 1.0
0.6 0.90 0.96 0.99
0.03
0.4
10−2
0.2 30 100 300
0.0 x̃ = (Rc ρs )/(Rs Ps )
100 101 102
x̃ = (Rc /Rs )(ρs /Ps )
Rc /Rs < 0.3
0.3

100

101 0.1

b
b = 0.59x−0.58
100

y = R̄/Rs
z = ra /rt

10−1 0.03
−1
10
3 10 30 100
x = |Eb |/∆Eimp
10−2
−2
10
FIG. 8. The dependence of the trajectory parameter b on
101 2
10
system parameters in the self-similar regime (Rc /Rs  1).
x = |Eb |/∆Eimp Top: There is a trend between b and x̃, but it is polluted by
residual sensitivity to the orbit shape parameter η. Bottom: b
is a power-law in x with little residual sensitivity to the orbit
FIG. 7. These figures summarize the 51 simulations we use shape, and the best fit is plotted as a solid line. The color
to tune our model. Top: The simulations distributed in the scale is the same for both panels. Each marker is a simulation,
host-subhalo system parameters. Bottom: The simulations and the marker radius is proportional to the number of orbital
distributed in the reduced parameters x, y, and z (see text). periods, which ranges from 7 to 20 for this sample.
Simulations with x < ∼ 1 are not included in this sample be-
cause they lead to |d ln J/d ln t| ≥ 1 too quickly. The radius of
each marker is proportional to the number of orbital periods, at n = 1.5, whose corresponding J-factor averages from
which ranges from 5 to 20. n = 1 to n = 2). This restriction is intended to remove
the influence of any transient effects associated with sud-
denly turning on the tidal field. Additionally, the av-
For each simulation, we obtain the trajectory of eraging procedure does not completely smooth out the
the subhalo’s J-factor, stopping when |d ln J/d ln t| ≥ 1 periodic oscillations, so to minimize any biasing effect,
or at simulation termination. The number of radial we end the fit at an integer number of orbits (so for in-
(apocenter-to-apocenter) orbits6 spanned by this trajec- stance, we might end at n = 15.5, corresponding to the
tory is represented in Fig. 7 as the marker size; this num- average J from n = 15 to n = 16).
ber is a proxy for how much information that simulation
provides7 . At each time step, we compute the running
logarithmic average of J over the surrounding radial pe- A. Parameter b: the initial J-factor decay rate
riod. This procedure smooths out the oscillatory behav-
ior observed in Section III. Next, we fit Eq. (14) to this
From Fig. 6, we anticipate that b should depend
smoothed trajectory of J, but we only employ times af-
strongly on the orbital radius. For simplicity, we first
ter the end of the first radial period (so the first point is
study the self-similar regime, Rc  Rs . The upper panel
of Fig. 8 plots b against the reduced orbital radius x̃ for
the 35 of our simulations that satisfy Rc /Rs < 0.3. While
6 b is strongly sensitive to x̃, there is also significant sensi-
For circular orbits, the radial orbit period is ill-defined, and we
substitute its limit as the orbit approaches circular, as obtained tivity to the orbital shape, parametrized by η. However,
using the fitting form in Appendix B. it turns out that we can eliminate the shape-dependence
7 When
√ performing fits, we weight a simulation spanning n orbits of b by defining the reduced orbital radius more carefully.
by n. We first remark that x̃ ∼ |Eb |/∆E, where Eb is the
8

binding energy of a particle at the subhalo’s initial scale


radius rs , and ∆E is the energy injected into that particle 0.3
by tidal forces over the subhalo’s orbital period. To see b=
this, observe that the particle’s binding energy is 0.5
9x − 0.5
8
0.1

b
Eb = −4π(ln 2)Gρs rs2 (15)
y = R̄/Rs
(per mass). Meanwhile, the tidal acceleration on this 10−2 10−1 100 101
particle is roughly (rs /R)F , where R is the subhalo’s 0.03
orbital radius and F is the host force (per mass) at radius
R. In the self-similar regime, F ∼ GPs Rs . The total 3 10 30 100
velocity injected into the particle is ∆vp∼ F T , where T is x = |Eb |/∆Eimp
the subhalo orbital period. Since T ∼ R/F , the energy
injection (per mass) is ∆E ∼ (∆v)2 ∼ GPs Rs rs2 /R, and 3.0 b
= 1 + 1.27f (y)


since R ∼ Rc , this leads to |Eb |/∆E ∼ x̃. 0.59x−0.58

b/ 0.59x−0.58
With this motivation, we define 2.5
100
2.0
x ≡ |Eb |/∆Eimp (16) 30

x
1.5 10
as a more exact version of x̃. Here, ∆Eimp is the energy
injection per orbit on a particle at rs computed using the 1.0 3
impulse approximation as in Ref. [95]. For convenience,
we supply a fitting formula for ∆Eimp in Appendix B. 10−2 10−1 100 101
In the bottom panel of Fig. 8, we plot b against x for y = R̄/Rs
the self-similar regime. Evidently, the more exact energy
calculation in x captures most or all of the sensitivity of FIG. 9. The dependence of b on x and y. Top: At each radius
the parameter b to the orbit shape, and y = R̄/Rs , b appears to follow a similar power law in x with
a different normalization. The solid line is duplicated from
b = b0 x−b1 , if Rc  Rs , (17) Fig. 8. Bottom: Scaling of the normalization of b with radius
y. The best fit is plotted, with f (y) defined in Eq. (20). Each
with b0 = 0.59 and b1 = 0.58. marker is a simulation, and the marker radius is proportional
To complete our understanding of the parameter b, we to the number of orbital periods, which ranges from 5 to 20.
must move beyond the self-similar regime. In the upper
panel of Fig. 9, we plot b against x for all of our simulated
subhalos. The color scale indicates the time-averaged or- the host force profile F (R) is self-similar, the ratio be-
bital radius R̄ in units of Rs . It appears that the effect tween the stretching and compressive forces is fixed8 . Be-
of leaving the self-similar regime is to alter the normal- yond the self-similar regime, however, the ratio between
ization of b while keeping the power-law sensitivity to x these forces can change. With this motivation, we define
unchanged. In particular, we may write f (R/Rs ) ≡ (dF/dR)/(F/R) as this ratio. For the NFW
profile, this definition implies that
b = b0 x−b1 [1 + b2 f (y)] , (18)

for some function f (y) and parameter b2 , where we define 2 ln(1 + y) − y(2 + 3y)/(1 + y)2
f (y) = . (20)
ln(1 + y) − y/(1 + y)
y ≡ R̄/Rs . (19)
In the bottom panel of Fig. (9), we plot b/(b0 x−b1 )
For convenience, we supply a fitting formula for R̄ in against y for the purpose of tuning the parameter b2
Appendix B. While we could use the circular orbit radius in Eq. (18). We find that this equation9 works reason-
Rc instead, we favor R̄ because its physical significance ably well, and we obtain b2 = 1.27. The introduction
is clearer. of stretching tidal forces increases the efficiency of tidal
To define the function f (y), we consider the physi- effects, which is reflected as an increase in the decay rate
cal impact of leaving the self-similar regime. The mag- b of the J-factor.
nitudes of the tidal forces are altered, but this effect
should be accounted for by the definition of x. How-
ever, the directions of the tidal forces also change. In
particular, Eq. (3) implies that there are stretching tidal
forces proportional to dF/dR along the radial axis from 8 In fact, for an NFW profile, dF/dR = 0 when R  Rs .
9
the host, and there are compressive tidal forces propor- The force-ratio argument motivates any expression of the form
tional to F/R along the perpendicular directions. When [1 + b2 f (y)α ]β , but we assume for simplicity that α = β = 1.
9

1 initi
al ρ

rt = ra
a = 0.45 − 1.33f (y) (r)

rt
ρr (M kpc−2 )

ra
0 104
a

t = 4T
t = 18T
−1 x
3 10 30 100 eccentric orbit
circular orbit
103
10−2 10−1 100 101 10−8 10−7 10−6
y = R̄/Rs r (kpc)

FIG. 10. The dependence of the trajectory parameter a FIG. 11. The influence of the shape of a subhalo’s orbit
on the orbital radius parameter y. The best-fitting curve is on its density profile after tidal evolution. One subhalo is
plotted as a solid line using the definition of f (y) in Eq. (20). on a highly eccentric orbit (η = 0.1) while the other its on
Each marker is a simulation, and the marker radius is propor- a circular orbit, but both have roughly the same x (4.4 and
tional to the number of orbital periods, which ranges from 5 4.3 respectively). We plot the density profiles after 4 and
to 20. 18 orbits. The key difference is that the circular orbit strips
material primarily from the outskirts, while the eccentric orbit
strips much more material from the interior. This difference
can be understood in terms of the adiabatic shielding radius
B. Parameter a: the J-factor normalization
ra and its comparison to the tidal radius rt (see text), shown
as dotted lines for both orbits.
We next handle the overall normalization of J/Jinit .
According to Eq. (14), the J-factor changes by the fac-
tor eab after the first orbit, which is sensitive to a second
parameter: a. Due to the way we defined this parameter,
subhalo density profile. The connection to orbital shape
it turns that a is almost wholly sensitive to y = R̄/Rs
is that circular orbits tidally heat material more predom-
alone. Figure 10 plots a against y for all of our simula-
inantly in the outskirts of the subhalo, while eccentric or-
tions, and we find that with only moderate scatter,
bits can alter the density profile further inward. This ten-
a = a0 − a1 f (y) (21) dency is illustrated in Fig. 11, which depicts the tidally
altered density profiles of two subhalos with roughly the
with a0 = 0.45 and a1 = 1.33. same x ' 4.4 but different orbit shapes. The subhalo on
In some sense, the parameter a describes the initial the more eccentric orbit experiences more tidal stripping,
behavior of the subhalo as it equilibrates—to the extent but in particular, its interior is stripped much more than
that this is possible—into the tidal field generated by the that of the subhalo on a circular orbit.
host. When y  1, all tidal forces are compressive, so the
Differences in the radii at which material is heated can
J-factor is initially slightly boosted (a > 0). However,
be understood in terms of adiabatic shielding (e.g. [96–
when y >∼ 1, the stretching tidal forces cause the J-factor 99]). Deep within the subhalo, the internal dynamical
to be initially suppressed (a < 0).
time scale is much shorter than the time scale over which
the external tidal field changes. In this case, the conser-
vation of adiabatic invariants prevents any energy injec-
C. Parameter c: the loss of tidal efficiency
tion by tidal forces: these radii are adiabatically shielded.
Meanwhile, adiabatic shielding is connected to the shape
Finally, we address the parameter c that characterizes of the subhalo’s orbit. The time scale over which tidal
the drop in the efficiency of tidal effects over time. As we forces change is related to the time scale of pericenter
found in Section II, c is sensitive to the orbit shape: more passage, which can be very short for highly eccentric or-
eccentric orbits yield smaller values of c, while more cir- bits.
cular orbits yield larger values. In the self-similar regime,
we could write c as a function of η, since η completely Up to factors of order unity, the subhalo’s internal dy-
describes the orbit shape. However, beyond this regime, namical time scale is tdyn ∼ (Gm(r)/r3 )−1/2 , where m(r)
orbits with the same η could have different shapes. Thus, is the subhalo mass profile [14]. Meanwhile, the pericen-
to accurately describe the sensitivity of the parameter c ter passage time scale is tp ∼ Rp /Vp , where Rp and Vp
to the host-subhalo system, it is necessary to find the are the radius and velocity at pericenter, respectively. To
correct orbit-shape parametrization. make precise the connection between orbit shape and the
We argued in Section II that the loss of tidal efficiency radii at which tidal heating is efficient, we define the adia-
encoded in c is related to changes in the shape of the batic shielding radius ra as the radius at which tdyn = tp .
10

This definition motivates a characteristic density scale 1.0


c = 0.74z 0.21
Vp2 M (Rc )Rc
ρa ≡ = η2 , (22)
GRp2 Rp4
0.5

c
so that ra is the radius at which m(ra )/ra3
= ρa . Note
that we used the definitions of the circular orbit radius x
Rc and orbit circularity η to eliminate Vp from Eq. (22); 3 10 30 100
M (R) is the host mass profile at radius R.
To quantify changes in the shape of the subhalo density 0.2
profile, we can compare the radius ra below which mate- 10−2 10−1 100
rial is shielded to the tidal radius rt above which all mate- z = ra /rt
rial is stripped. The tidal radius is the radius above which
the tidal force from the host exceeds the gravitational FIG. 12. The dependence of the trajectory parameter c on
force from the subhalo. There are several definitions of the system parameter z = ra /rt . With moderate scatter, c
the tidal radius in the literature, but they are all re- follows a power law in z, shown as the solid line. Each marker
lated to the expression [100, 101] rt = R[m(rt )/M (R)]1/3 is a simulation, and the marker radius is proportional to the
by (possibly nonconstant) factors of order unity (see e.g. number of orbital periods, which ranges from 5 to 20.
Ref. [68]). The tidal radius is only well-defined for cir-
cular orbits, but it is common to apply the concept to as the J-factor decays slower than |d ln J/d ln t| = 1, its
eccentric orbits as well [68]. In particular, if we seek the trajectory is well-fit by the expression
radius above which all material is stripped, we can define  
the tidal radius rt using the orbital apocenter radius Ra . J 1 1−c

In this case, there is a characteristic density scale ln =b a− n −1 . (14)
Jinit 1−c

ρt ≡ M (Ra )/Ra3 , (23) Here, n = t/T is the number of subhalo orbits, and a,
b, and c are parameters that depend on the host-subhalo
and rt is the solution to m(rt )/rt3 = ρt . reduced system parameters x, y, and z through
Anticipating that the drop in the efficiency of tidal
a = a0 − a1 f (y) (21)
effects encoded in the parameter c is a consequence of
−b1
changes to the shape of the subhalo density profile, we b = b0 x [1 + b2 f (y)] (18)
may hypothesize that c is sensitive to the ratio c = c0 z c1 . (25)

z ≡ ra /rt . (24) with a0 = 0.45, a1 = 1.33, b0 = 0.59, b1 = 0.58,


b2 = 1.27, c0 = 0.74, and c1 = 0.21. In Appendix B, we
For simplicity, in defining z, we employ the subhalo’s detail how to compute x, y, z, and T from the subhalo
initial NFW mass profile [see Eq. (B1)]. Figure 12 shows parameters rs and ρs , host parameters Rs and Ps , and
the relationship between c and z: there is some scatter, orbital parameters Rc and η.
but the trend is that Equation (14) applies only when |d ln J/d ln t| < 1.
The J-factor’s precise behavior when |d ln J/d ln t| >
∼ 1 is
c = c0 z c 1 (25) of little consequence, as the subhalos in this regime con-
tribute only minimally to the annihilation signal. Never-
with c0 = 0.74 and c1 = 0.21. Note that z is not solely theless, it is useful to have an approximate treatment
a function of the subhalo’s orbit. Because it depends on in this regime. As we discussed in Section III, when
the subhalo mass profile m(r), it is also sensitive to the |d ln J/d ln t| ≥ 1, it is a reasonable approximation to
density ratio ρs /Ps . We explored using ρt /ρa , a purely enforce −d ln J/d ln t = 1, i.e. J ∝ n−1 . We define
orbital parameter, instead of ra /rt . This parameter ex-
hibited a similar power-law relationship with c, but it left n1 = b1/(c−1) (26)
significant residual sensitivity to the parameter x, which
as the orbit count at which −d ln J/d ln t = 1. Addi-
is related to ρs /Ps . Using z = ra /rt mostly eliminates
tionally, when b > 1 (so n1 < 1), we cannot expect our
that sensitivity.
treatment of the normalization of J (Section IV B) to be
accurate. To handle these issues, we can write
 n h io
V. MODEL SUMMARY AND DISCUSSION 
 exp b a − 1
n 1−c
−1 , if n ≤ n1 , b < 1
J  n h 1−c io

= exp b a − 1 1 − 1 n1
In the last section, we developed a model for the evolu- Jinit  1−c b n , if n > n1 , b < 1

 a
tion of a subhalo’s J-factor due to tidal effects as a func- e /n, if b ≥ 1,
tion of parameters of the host-subhalo system. As long (27)
11

where the last case follows from continuity considera- discussion in Section IV B. However, it turns out that
tions. We further note that this equation is only valid while this treatment works reasonably well for a portion
when n ≥ 1. of the a0 , b0 , c0 , a, b, c parameter space, it does not accu-
One can now use our model to understand the emis- rately predict every scenario: the parameter q in Eq. (13)
sion from a host halo due to dark matter annihilation in is not a function of subhalo alone. We leave a detailed
subhalos. In particular, one can sample subhalos from investigation of this problem to future work.
an orbital distribution in Rc and η (e.g. [67, 73, 102– As another caveat, the long-term accuracy of the tra-
105]). Accounting for tidal evolution, each subhalo’s jectory in Eq. (14) relies on the assumption that the effi-
contribution to the dark matter annihilation signal is ciency of tidal effects follows precisely the power law n−c ,
then scaled by the orbit-dependent function given by as described by Eq. (8). While such a power law is a
Eq. (27). In Ref. [106] (in preparation), we will use this natural assumption [e.g Eq. (13)] and is borne out in our
model to study the annihilation signature arising from simulations, it does not have a direct physical motivation:
the extreme-density microhalos that result from certain tidal heating models considered in Section III B and else-
early-Universe scenarios. In this case, the orbital distri- where [76] can only reproduce c = 0. Without such moti-
bution of subhalos is the same as that of particles, and vation, it is unclear that this power-law behavior should
one may employ the host halo’s distribution function (e.g. extend beyond the n = 20 orbits of our longest simula-
[92]) to sample subhalo orbits. tions. Also, the system parameter z = ra /rt that sets the
Our model does not include the periodic oscillations power-law index c is defined based on two concepts that
in the J-factor observed in Section III. These oscillations are not themselves entirely well-defined: the adiabatic
do not affect the overall annihilation rate in subhalos, shielding radius ra and the tidal radius rt . Moreover,
but they still introduce a systematic biasing effect where since our model does not predict the larger evolution of
subhalos at smaller radii have larger J, and this effect the subhalo density profile, we use the subhalo’s initial
can alter the morphology of an annihilation signal. How- density profile to define ra and rt even though the density
ever, we remark that these oscillations only have signif- profile quickly begins to change. For these reasons, we
icant amplitude in the Rc  Rs regime, when all tidal anticipate that it is possible to find a better-motivated
forces are compressive, and at small x < ∼ 10. Because of
parameter to replace z = ra /rt .
these restrictions, we anticipate that their impact is mi- Nevertheless, this model describes the results of our
nor. However, in forthcoming work [106], we will quantify simulations with remarkable success. As further valida-
the impact of these oscillations. tion, we consider the library of idealized subhalo sim-
We also address another potential limitation to our ulations, called DASH (for Dynamical Aspects of Sub-
model. The differential equation driving it, Eq. (8), has Haloes), published by Ref. [79]. These simulations have
explicit time-dependence in the factor n−c , so the result- lower resolution than ours, but because of the extraor-
ing tidal evolution is not completely determined by the dinary volume of this library, it still supplies a valuable
system’s instantaneous state. Physically, we view n−c as test for our model. In Appendix C, we verify that modulo
a proxy for unknown physical variables [e.g. Eq. (13)], substantial scatter and certain systematic effects associ-
and as long as the host halo’s density profile and the sub- ated with their lower resolution, the DASH simulations
halo’s orbit are static, this formulation poses no difficulty. are consistent with our model.
Since halos grow from the inside outward, subsequent ac-
cretion is not expected to significantly alter the density
profile of a host halo at the radii of already-present sub- VI. COMPARISON TO PREVIOUS WORK
halos, so the host halo is generally expected to remain
static. Moreover, if dynamical friction is negligible [see Numerous prior works have endeavored to model the
Eq. (5)], the subhalo’s orbit is also static. However, there impact of tidal effects on a subhalo’s dynamical evolution
is a scenario where a subhalo’s host is expected to change [18, 70–78]. In this section, we explore how our results
dramatically. If the host is itself a satellite of a larger host compare to those of previous studies. Motivated primar-
halo, then the subhalo may be tidally stripped from its ily by simulations, our model is based on the notion that
host, becoming itself a satellite of the superhost. In this a subhalo’s J-factor evolution is determined by
scenario, it is not obvious how to continue the subhalo’s
tidal evolution. 1 dJ
If the initial host-subhalo system yields trajectory pa- = −bn−c , (8)
J dn
rameters a0 , b0 , and c0 , and the new host-subhalo system
yields parameters a, b, and c, then a self-consistent way where b and c are functions of the host-subhalo system
to treat this problem 0is to substitute the factor n−c in and n counts the number of orbits. In contrast to our
Eq. (8) with (n + n0c /c )−c and integrate the resulting focus on the J-factor, previous works have largely fo-
expression. This treatment follows from the assumption cused on the evolution of a subhalo’s total bound mass
that the parameter q = n−c in Eq. (13) is a function mbound and of its maximum circular velocity vmax and
of the subhalo alone. 0 0
Additionally, the J-factor should corresponding radius rmax . However, the general form of
be rescaled by eba−b a , a consideration motivated by the our model is not specific to the J-factor, and we show
12

in Appendix D that it can also describe the evolution of


1.0
vmax , rmax , and mbound .
Despite the broad applicability of our model suggested
by Appendix D, no prior work (to our knowledge) has
proposed tidal evolution of the form given in Eq. (8). 0.5

J/Jinit
Broadly, prior models of tidal evolution fall into two main
categories, although a given work may employ more than (252, 1, 0.5)
one: (69, 11, 0.09)
1. Tidal stripping models, where material outside the 0.2 (31, 5.2, 0.4)
characteristic tidal radius [e.g. Eq. (23)] is assumed (13, 1.1, 0.1)
to be stripped over some time period; (21, 11, 0.07)
simulation (this
n= work)
t/T
2. Tidal heating models, where energy injected by 1.0 model (this work)
tidal forces heats subhalo material, causing it to P14 model

−d log J/d log t


rise and possibly become freed from the subhalo.
We found in Section IV A that the parameter b in Eq. (8),
0.3
which characterizes the rate of tidal evolution, is tightly
sensitive to the energy injected by tidal forces (see Fig. 8).
Additionally, we observed in Section III that the tidal
evolution in our simulations closely resembles that pre- 0.1
dicted by a toy model of tidal heating. Consequently, we
anticipate that of the two classes of model, tidal heating
models should yield results most similar to those of our 100 101
model. We will first compare the results of our model to n = t/T
those of the tidal heating model developed by Ref. [76],
hereafter P14.
However, prior treatments of dark matter annihilation FIG. 13. A comparison between our tidal evolution model
within subhalos predominantly treat the impact of the (solid lines) and the analytic tidal heating model developed in
host halo’s tidal forces using models based on tidal strip- Ref. [76] (P14, dashed lines). This figure shows the J-factor
trajectory and its logarithmic derivative for different host-
ping [30–35, 39]. Tidal stripping models cannot prescribe
subhalo parameters (x, y, z), listed on the figure. We also
how to change a subhalo’s density profile below the tidal show our simulation results (as circles) for these parameters.
radius, but it is possible to apply a simulation-tuned pre- The P14 predictions exhibit the correct trends, but they are
scription for how the density profile responds to mass only reasonably accurate for a small range of host-subhalo
loss [18, 78]. We will subsequently compare the results system parameters.
of our model to those of a tidal stripping model devel-
oped by Ref. [77] (hereafter J16), using the prescription
of Ref. [18] (hereafter P10) to predict the subhalo’s den-
pute the evolution of a subhalo’s density profile, subse-
sity profile. This pair of models has been employed by
quently integrating it to obtain the J-factor. Figure 13
Refs. [33, 34] to predict dark matter annihilation rates in
shows a sample of the resulting J-factor trajectories, and
subhalos.
we compare those trajectories to our model’s predictions
and to the results of our simulations. Generally, we find
A. Comparison to a tidal heating model
that for a model constructed from first principles, the
P14 model is remarkably accurate. However, it does not
fully capture the sensitivity of tidal evolution to system
We first compare our model’s predictions to those of parameters, a matter we explore next.
the analytic tidal heating model given in P14. In the tidal
The quantity Q = ∆E/r2 employed by P14 is re-
heating picture, energy injected by tidal forces causes
lated to host-subhalo system parameters by Q/(GPs ) =
subhalo material to move to higher radii, and P14 em-
4π(ln 2)N/x, where N is the number of orbits over which
ployed the assumption of virial equilibrium to predict this
energy injection is taken, and x is the system parame-
change in radius and consequently the subhalo’s new den-
ter (see Section IV). For our comparison, we take N = 1
sity profile. We follow the prescription in P1410 to com-
and iterate the calculation, assuming the halo revirializes
during each orbit. Since the density profile evolution in
P14 is only sensitive to the ratio Q/(GPs ), we see imme-
10 For simplicity, we compute the energy injection ∆E directly us- diately that this model’s predictions are only sensitive
ing the impulse approximation (Appendix B), neglecting addi- to the system parameter x and are insensitive to y and
tional corrections suggested in P14; these corrections will not z. Additionally, the J-factor evolution predicted by P14
qualitatively alter the results. turns out to be only sensitive to x in the combination
13

100

J/Jinit (orbit-averaged)
n/x, where n = t/T is the number of orbits, so every this work
system follows the same trajectory rescaled in time. In J16+P10
this respect, the P14 model is similar to the toy model we
explored in Section III B, which was only sensitive to the
combination f n of system parameters f and orbit count
n. In fact, the P14 model approximately obeys the toy
model solution Eq. (12) with b ' 3.2/x and B = 1, but 10−1 m/M, csub , chost
it can potentially transition between the n/x  1 and 10−6 , 51, 6
n/x  1 regimes extremely slowly, and all behavior seen
10−3 , 24, 6
in Fig. 13 is in the intermediate regime.
The combination of its single time-rescaled trajectory 100 101
and its insensitivity to y and z leaves the P14 model t/tdyn
unable to accurately predict tidal evolution in the full
host-subhalo parameter space. We see evidence of this FIG. 14. A comparison between our tidal evolution model
deficiency in Fig. 13, but we further note that we did not and the semianalytic model developed in Refs. [77] (J16)
plot any subhalos in the y  1 regime. In this regime, and [18] (P10). This figure shows the orbit-averaged J-
the P14 model dramatically overestimates the impact of factor trajectory of subhalos of mass m and concentration csub
tidal stripping, since it does not account for the directions within a host halo of mass M and concentration chost . Com-
of tidal forces, which are encapsulated in the parameter pared to our model, the semianalytic model underestimates
y. While the P14 model yields reasonably accurate pre- the impact of tidal stripping early on while overestimating its
dictions over a small range of host-subhalo system pa- impact at late times.
rameters, our model can accurately predict the evolution
of a much broader variety of systems.
Ref. [73]. The circular orbit radius Rc is taken to be uni-
formly distributed between 0.6Rvir and Rvir , where Rvir
B. Comparison to a tidal stripping model is the host’s virial radius. Meanwhile, the circularity η is
distributed proportionally to sin πη, and we assume that
the distributions of Rc and η are independent. By draw-
Finally, we compare our model to a semianalytic model
ing subhalo orbits from this distribution, we are able to
of tidal stripping that has been employed in previous cal-
compute the orbit-averaged11 value of J/Jinit using our
culations of annihilation rates in substructure [33, 34].
model given in Eq. (27). In Fig. 14, we plot the result-
This semianalytic model uses the tidal stripping model
ing orbit-averaged J-factor trajectories along with those
in J16 to characterize a subhalo’s mass loss, subsequently
predicted by the semianalytic model of J16 and P10. We
using the results of P10 to connect this mass loss to the
consider two different host-subhalo systems, listed on the
subhalo’s density profile and hence annihilation signal.
figure, and since the semianalytic model is sensitive to
In J16, the rate of mass loss for subhalos of mass m in-
the total virial masses of the host and subhalo, we employ
side a host halo of mass M , averaged over subhalo orbits,
the concentration parameter c ≡ rvir /rs to describe these
is modeled using
systems: chost is the host halo’s concentration, while csub
dm m  m ζ is the subhalo’s concentration when it is accreted.
= −A . (28) Compared to our model, Fig. 14 shows that the semi-
dt tdyn M
analytic model underestimates the impact of tidal forces
Here, A and ζ are simulation-tuned parameters, and tdyn early on while overestimating their impact at late times.
is the host’s dynamical time scale at its virial radius (e.g. These discrepancies arise from several sources. As we
[14]). For this comparison, we adopt J16’s central values show in Appendix D, the structural parameter vmax is not
A = 0.86 and ζ = 0.07. uniquely determined by the subhalo’s remaining mass,
In P10, it is shown that the subhalo’s maximum cir- contrary to the finding in P10, and moreover, the subhalo
cular velocity vmax and the radius rmax at which it is does not retain an NFW profile. However, this source
attained are uniquely related to the fraction m/macc of of error is relatively minor. The main differences arise
the subhalo’s mass that remains gravitationally bound, from the model in J16 given by Eq. (28). Since this
where macc is the subhalo’s virial mass at accretion. model does not account for the subhalo’s density pro-
If we assume that the subhalos possess NFW profiles, file, it takes too long to strip the subhalo’s weakly bound
then m/macc thereby determines the subhalo’s J-factor. outskirts (beyond rs ) that contribute little to annihila-
While we find in Appendix D that this correspondence is
not exact, we will adopt it for this comparison, using the
relations in P10 to connect the subhalo’s density profile,
and hence J-factor, to its mass. 11 Specifically, we take the median J/Jinit at each time, but using
To compare our model, we employ the same subhalo or- the mean or the logarithmic mean instead does not significantly
bital distribution considered in J16, which is drawn from alter the results.
14

tion rates12 . This behavior partially explains why the Our model has limitations. As presented, it is re-
semianalytic model underestimates the early impact of stricted to host-subhalo systems in which both halos
tidal effects. Meanwhile, for small ζ  1, Eq. (28) de- possess NFW density profiles. The NFW profile (possi-
scribes nearly exponential decay, analogous to our model, bly with minor corrections, e.g. [107]) arises generically
Eq. (8), with c = 0. Without the braking behavior con- in dark matter simulations of halos built by hierarchi-
tributed by c > 0 (and attributed to changes in the shape cal clustering [11, 12]. However, the smallest subhalos,
of the subhalo’s density profile; see Section III) along forming by direct collapse, exhibit steeper density pro-
with the limiting |d ln J/d ln t| ∼ 1 behavior, the semi- files [24–26, 65, 84–88]. Additionally, the density profiles
analytic model overestimates the impact of tidal effects of many galactic halos (but not all [83]) are inferred to
at late times. For these reasons, our simulation-tuned be shallower than the NFW profile, an observation that
model supplies significantly more accurate predictions of may be explained by baryonic effects or unknown dark
subhalo annihilation rates. matter properties (see Ref. [13] for a review). Despite be-
ing developed using NFW profiles, we anticipate that the
physical manner in which we defined the model parame-
VII. CONCLUSION ters x, y, and z implies that our model can be adapted to
accommodate different host or subhalo density profiles.
In this work, we used 51 idealized N -body simulations Also, our model only accounts for tidal forces from the
to develop a model that can predict the impact of a host host halo. Subhalos can also be disrupted by encounters
halo’s tidal forces on the rates of dark matter annihilation with other subhalos, but the results of Ref. [68] suggest
within its subhalos. Our model is given by Eq. (27) and that this effect is subdominant. More importantly, sub-
summarized in Section V, and it predicts the evolution of halos can be affected by baryonic content residing within
the subhalo’s J-factor, the factor in the annihilation rate the host, such as stars (e.g. [17, 21, 24, 108–113]) or
that depends on mass distribution, as a function of the a disk (e.g. [31, 70, 111–117]). These effects are not
subhalo’s orbit and other properties of the host-subhalo included in our model and must be accounted for sep-
system. These properties are distilled into three phys- arately. However, we remark that many of the dwarf
ically motivated variables x, y, and z that characterize spheroidal galaxies, already some of the most promising
the energy injected by tidal forces, the ratio of stretching targets for dark matter annihilation searches [10], have
to compressive tidal forces, and the radial distribution of such little baryonic content (e.g. [118]) that stellar and
tidally heated material, respectively. Appendix B details disk effects can likely be neglected.
how to compute these variables from standard properties Despite these limitations, we anticipate that our model
of the host-subhalo system. will prove useful in understanding the annihilation sig-
Our model is based on the notion that for sufficiently nals of dark matter substructure. In a subsequent paper
small changes in J, the J-factor evolves according to [106], we will explore the consequences of our model by
using it to study microhalo-dominated annihilation sig-
1 dJ nals in nearby dwarf galaxies. Such signals are expected
= −bn−c , (8)
J dn to arise from certain cosmological scenarios, such as an
where n = t/T is the time in units of the subhalo’s orbital early matter-dominated era prior to nucleosynthesis, and
period, and b and c are parameters that depend on the our model enables precise characterization of the magni-
system. If c = 0, Eq. (8) states that the subhalo loses tude and morphology of these signals.
a fixed fraction e−b of its J-factor in each orbit. The
parameter c ≥ 0 is motivated by simulation results and
ACKNOWLEDGMENTS
adds a braking mechanism to the J-factor’s decay. To
our knowledge, a model of this form has not previously
been put forward, even though we find that it can also The simulations for this work were carried out on the
describe other structural properties of the subhalo. We KillDevil and Dogwood computing clusters at the Univer-
also find that our model predicts significantly different sity of North Carolina at Chapel Hill. This work was sup-
J-factor trajectories than prior semianalytic models. We ported by Fermi Guest Investigator Cycle 10 Award No.
further validate our model by testing it against the pub- 80NSSC17K0751 (PI A. Erickcek). The author thanks
licly available DASH library of subhalo simulations [79], Adrienne Erickcek and Tim Linden for helpful discus-
finding reasonable agreement. sions. Key figures in this work employ the cube-helix
color scheme developed by Ref. [119].

12 For the two cases shown in Fig. 14, it takes respectively 6 and 3 Appendix A: Simulation details
dynamical times for the J16 model to bring the subhalo’s bound
mass below its initial mmax , the mass enclosed within the radius
rmax at which the maximum circular velocity is attained. As we In this appendix, we supply additional details about
find in Appendix D (see Fig. 18), the drop in J-factor is minimal our simulations. First, we describe how subhalos’ J-
above this mass threshold. factors are extracted from the simulations, and sec-
15

ond, we detail the choices of simulation parameters and Eq. (A2) inaccurate. Thus, we also compute the J-factor
demonstrate that the simulations are converged with re- as the sum over simulation particles
spect to these parameters. X
J= ρi mi , (A3)
i
1. Density profiles and J-factors where mi is the mass of particle i, and ρi is its local
density. The density ρi is estimated as
We obtain each subhalo’s density profile by binning it
N
X
in factors of 1.1 in the radius. At small radii, there is a
resolution limit driven by three effects: force softening, ρi = mj W (rij , hi ), (A4)
j=1
Poisson noise, and artificial relaxation. Each effect is as-
sociated with a minimum resolved radius below which over the N = 50 nearest particles j, where rij is the
the density profile artificially flattens. For force soften- distance to particle j, hi is the distance to the N th par-
ing, that radius is the distance rsoft = 2.8, where  is ticle, and W (r, h) is the cubic spline kernel defined as in
Gadget-2’s force-softening parameter, at which forces Ref. [91].
become non-Newtonian. For Poisson noise, we take it to Equation (A3) underestimates the J-factor contribu-
be the radius r100 enclosing 100 particles. To estimate tion at r < rmin due to artificial flattening of the density
the radius rrel at which artificial relaxation becomes sig- profile. To accommodate the extrapolation procedure in
nificant, we compute the relaxation time [14] Eq. (A2) that addresses this problem, an additional step
is required. We find the bound remnant of the subhalo
N r
trelax = p (A1) using a procedure similar to that in Ref. [68]. Beginning
8 ln Λ GM/r with the assumption that all particles are bound, we iter-
atively compute the gravitational potential of each par-
at each radius r, where M and N are the mass and parti- ticle due to all other bound particles using a Barnes-Hut
cle count interior to r, and Λ = max{N, r/}. If αtrelax at octree [120] with θ = 0.7 and the same softening length
radius r, with an appropriate proportionality constant α, as the simulation. Subsequently, we mark each particle
is shorter than the system age, then r < rrel . The pro- as unbound if its total energy is positive. At each step,
portionality constant α is tuned to predict the correct we find the center-of-mass position and velocity of the
rmin in a simulation of the subhalo without a host: in 100 most bound particles and re-center the full system
that case, any change to the density profile is artificial to be relative to this center of mass. All particles are
since the halo was built from an equilibrium distribution. initially marked as bound, and the procedure terminates
From this calibration, we use α = 5. when the count of bound particles converges13 .
The minimum resolved radius of the density profile is By assuming that the bound remnant is spherically
rmin = max{rsoft , r100 , rrel }. For the purpose of accu- symmetric, we can estimate the J-factor both including
rately computing J-factors, we extrapolate the density spherical asymmetry and compensating for the flattening
profile below rmin as ρ = Ar−1 , where A is the average of the density profile below rmin . If Jfull is the J-factor of
of ρr in the three smallest radial bins above rmin , so that the full system computed using Eq. (A3), and Jbd,rad and
Z ∞ Jbd are the J-factors of the bound remnant computed
2
J = 4πA rmin + ρ(r)2 4πr2 dr. (A2) using Eqs. (A2) and (A3) respectively, then
rmin
J = Jfull − Jbd + Jbd,rad . (A5)
Effectively, this procedure produces a lower bound on
J under the assumption that larger radii are always
stripped more than smaller radii. Below rmin , we simply 2. Numerical convergence
assume all radii are stripped equally. We can also com-
pute an upper bound on J by assuming that radii below In our simulations, we set Gadget-2’s force-softening
rmin are completely unaffected (so A = ρs rs ), and this length to be  = 0.003rs . This small value is intended
allows us to estimate the uncertainty in our J-factors. to evade the artificial subhalo disruption observed by
We find that by the termination time of each subhalo’s Ref. [69]. Meanwhile, the subhalo’s high-resolution parti-
J-factor trajectory (as defined in Section IV), the uncer- cles (see Section II) have mass 4.3 × 10−7 ρs rs3 . To check
tainty in the J-factor, taken as Jupper /Jlower − 1, is 31% that numerical artifacts in our simulations are under con-
for one simulation (parameters x = 31, y = 11, z = 0.07; trol, we test the impact of changing the softening length
see Section IV), smaller than 17% for the remaining 50 and the particle resolution. Additionally, we test the im-
simulations, and smaller than 10% for 43 of them. pact of altering the (adaptive) integration time steps in
To understand the J-factors in the cases where
|d ln J/d ln t| >
∼ 1, another step is necessary. In this
regime, the elongated tidal stream can contribute sig-
13 This halting condition is stricter than the one in Ref. [68].
nificantly to the J-factor, making the spherical integral
16

1.0 1. Computing x = |Eb |/∆Eimp


reference
mass ×0.42
softening ×2.5 The binding energy Eb of a particle at the subhalo’s
0.6 time step ×0.45 scale radius rs is given by Eq. (15). Meanwhile, the to-
J/Jinit

tal energy ∆Eimp injected into a particle at radius r by


0.4 tidal forces over the course of a subhalo orbit is com-
puted using the impulse approximation, as described in
Ref. [95]. This energy depends on the particle’s full three-
dimensional position within the subhalo, but we simplify
0.2 the picture by averaging this energy over the sphere at
2 4 6 8 10 12 14 radius r. Dimensionally, ∆Eimp /r2 ∼ F (Rc )/Rc , and we
n = t/T can approximate
∆Eimp n h io F (R )
P3 (yc ) c
FIG. 15. Simulations of the same tidal evolution scenario = P 1 (y c ) exp P 2 (yc ) 1 − η ,
r2 Rc
carried out with different simulation parameters. For each (B3)
simulation, two J-factor trajectories are plotted correspond- where P1 (yc ) is defined
ing to the lower and upper limits discussed in Appendix A
(the lower limit is the value we use throughout this work). A(1 + B ln(1 + yc ) − Cyc /(D + yc ))
The upper and lower limits of each simulation overlap, imply- P1 (yc ) = ,
1 + E(ln(1 + yc ) − 2yc /(2 + yc ))
ing numerical convergence.
A = 3.327, B = 0.6463, C = 0.8837, D = 0.8809,
E = 0.2156, (B4)
order to ensure there are no artifacts arising from the P2 (yc ) is defined
application of the host’s tidal field over these discrete in-
tervals. In Fig. 15, we plot the J-factor trajectory in P2 (yc ) = A(1 + (yc /c)a )b ,
a reference simulation (with system parameters x = 34, A = 3.005, a = 3.641, b = 0.08513, c = 0.5703, (B5)
y = 0.018, and z = 0.15; see Section IV) along with three
simulations of the same system with different particle res- and P3 (yc ) is defined
olution, force softening, and integration time steps. We
plot the upper and lower limits of the J-factor trajectory A(1 + (yc /c1 )a1 )b1
P3 (yc ) = ,
as discussed above. These limits overlap for all simulation (1 + (yc /c2 )a2 )b2 (1 + (yc /c3 )a3 )b3
parameters, suggesting that the simulation is converged. A = 0.2150, a1 = 1.017, b1 = 0.8650, c1 = 0.5057,
a2 = 2.774, b2 = 0.2426, c2 = 0.6415,
a3 = 0.7663, b3 = 0.6508, c3 = 18.84. (B6)
Appendix B: Computational details
For η > 0.04, this expression is accurate to within 3% for
yc < 10 and within 14% for yc < 103 .
In this appendix, we present practical ways to compute
the reduced variables x, y, and z, along with the orbital
period T . For convenience, we include fitting formulas 2. Computing y = R̄/Rs
to approximate the necessary integrals. In what follows,
the host is assumed to possess an NFW profile with scale The time-averaged radius R̄ of the orbit is approxi-
radius Rs and scale density Ps : its mass profile is mately Rc , and in fact, R̄/Rc → 1 as Rc /Rs → 0. More
broadly, the expression
   
R R/Rs
M (R) = 4πPs Rs3 ln 1 + − , (B1) R̄ 1+B(1−F η G ) ln(1+yc )−Cyc /[D(1−Hη I )+yc ]
Rs 1 + R/Rs = ,
Rc 1 + E(ln(1 + yc ) − 2yc /(2 + yc ))
B = 0.3777, C = 0.4892, D = 2.412, E = 0.2426,
its force profile is F (R) = GM (R)/R2 , and its potential
profile is F = 0.3556, G = 1.860, H = 0.1665 (B7)

is accurate to within 0.3% for η > 0.04 and yc < 103 .


ln(1+ R/Rs )
Φ(R) = −4πGPs Rs2 . (B2)
R/Rs
3. Computing z = ra /rt
Meanwhile, the subhalo’s orbit about the host is
parametrized by the circular orbit radius Rc and circu- We define the adiabatic shielding radius ra and the
larity η, and as shorthand, we define yc ≡ Rc /Rs . tidal radius rt as the solutions to m(ra )/ra3 = ρa and
17

0.10 resolved (r > rmin ) this work simulations. This library includes the results of 2177
r < rmin simulations, with different system parameters, of an N -
ρ2 r3 /(ρ2s rs3 )
x = 69.3

n
extrapolated

=
y = 11.0 body subhalo orbiting an analytic host potential. These

0
0.05 z = 0.087 simulations resolve significantly less of the subhalo den-

n
sity profile than do ours: as shown in Fig. 16, they can

=
leave large fractions of the J-factor unresolved. Also, the

5
DASH library covers a smaller parameter range in x, y,
0.00
0.10 resolved (r > rmin ) DASH and z. Nevertheless, because of the extraordinary volume
r < rmin
ρ2 r3 /(ρ2s rs3 )

extrapolated x = 69.0 of this library, it can serve as a test for our model.

n
y = 12.3 We use the procedure in Appendix A to find the J-

=
0
0.05 z = 0.061 factor trajectory of each DASH simulation14 , imposing
an additional constraint that the trajectory halt when

n
=
the maximum uncertainty in the J-factor is larger than

5
0.00
a factor of 3. Next, we fit the parameters a, b, and c to
this trajectory as in Section IV. Figure 17 repeats our
10−2 10−1 100 101
plots in Figs. 9, 10, and 12 of the trajectory parame-
r/rs ters a, b, and c against the system parameters x, y, and
z. Superposed are our model predictions, as solid lines,
FIG. 16. A resolution comparison between our simulations using the parameters obtained in Section IV.
(top) and those of the DASH library (bottom). The (log-
We first remark that all DASH simulations have y > 2,
space) integrand for the J-factor, ρ2 r3 , is plotted for an ex-
ample subhalo from each catalogue with similar system pa- so we cannot directly test Eq. (17) describing the behav-
rameters (x, y, and z; see Section IV) at n = 0 and n = 5 ior of b in the y  1 self-similar regime. Nevertheless,
orbits. The J-factor is the area under the curve. Below the the upper-left panel of Fig. 17 shows that the DASH sim-
resolution limit rmin , we plot a pessimistic extrapolation of ulations exhibit the same power-law behavior b ∝ ∼x
−0.58

the density profile; see Appendix A. predicted by Eq. (18) (the offset between our curve and
the simulations here is not a discrepancy). The lower
panels show the sensitivity of a and b to y. Because
m(rt )/rt3 = ρt , respectively, where m(r) is the subhalo’s the DASH simulations only cover a small range of y, we
initial NFW mass profile [see Eq. (B1)]. Here, ρa and ρt cannot verify the functional form of each parameter in
are functions of the subhalo’s orbit; in particular, y. Also, there is substantial scatter, especially at large
Vp2 M (Rc )Rc M (Ra ) x. Nevertheless, our model predicts roughly the correct
ρa ≡ 2
= η2 4
and ρt ≡ , (B8) values of a and b for these simulations, although there
GRp Rp Ra3 is a tendency for the simulations to have smaller values
where Rp and Ra are the orbital pericenter and apocenter of a and larger values of b. Finally, although the scatter
radii, which may be obtained as the two solutions R to in c is quite large, the relationship between c and z is
  approximately borne out in the DASH simulations.
R2 GM (Rc ) The tendency for the DASH simulations to yield small
Φ(Rc ) − Φ(R) + 1 − η 2 c2 = 0. (B9)
R 2Rc a and large b can be understood as a resolution artifact.
Below the resolution limit, we extrapolate the density
profile in a way that always underestimates the J-factor
4. Computing T (see Appendix A). This underestimation both increases
the immediate loss of J-factor, reducing a, and increases
Dimensionally,
p the radial orbit period T ∼ t0 , where the rate at which J decays (since artificial relaxation
t0 ≡ Rc /F (Rc ). More precisely, the expression worsens the resolution over time), raising b.
Also, there is a tendency for systems at the large-x end
T A(1 + F η G )[1 + B ln(1 + yc ) − Cyc /(D + yc )] to exhibit large scatter in a, b, and c as well as a precip-
= ,
t0 1 + E(1 + Hη I )(ln(1 + yc ) − 2yc /(2 + yc )) itous drop in b (sometimes even to b < 0). This trend
A = 3.460, B = 0.6076, C = 0.8831, D = 2.312, is also an unphysical artifact. In our simulations, we ob-
served the same trend when x > ∼ 200, which is why our
E = 0.3325, F = 0.04827, G = 1.261,
simulation sample in Section IV only includes x < 200.
H = 0.03606, I = 1.288 (B10) For the lower-resolution DASH simulations, the trend be-
gins at x >∼ 50. The numerical difficulty with large x is
is accurate to within 0.2% for η > 0.04 and yc < 103 .
unclear, but it is likely connected to the the fact that

Appendix C: Comparison to the DASH library


14 We use a larger α = 20 to find rrel for the DASH simulations,
Reference [79] published a library called DASH (for obtained by recalibrating for these simulations. Note that larger
Dynamical Aspects of SubHaloes) of idealized subhalo α implies more optimism about simulation resolution.
18

0.3 1.0

predic
tion f
b

c
0.1 or y  0.5
y 1 only
x
10−1 100 101 10 30 100
0.03
0.2
10 30 100 10−1 100
x = |Eb |/∆Eimp z = ra /rt
4
0


3
b/ 0.59x−0.58

−1
2

a
1 x −2 x
10 30 100 10 30 100

0 −3
10−1 100 101 10−1 100 101
y = R̄/Rs y = R̄/Rs

FIG. 17. A test of our model against the DASH simulations. This figure plots the J-factor trajectory parameters a, b, and
c for the DASH simulations against the system parameters x, y, and z. The solid curves are our model predictions: they are
the same curves shown in Figs. 9, 10, and 12. Note that the offset between the solid line and the simulations in the upper-left
panel is not a discrepancy: the solid line is only valid for y  1. The DASH simulations exhibit significant scatter but broadly
support our model, with some systematic discrepancies discussed in Appendix C. The radius of each marker is proportional to
the number of orbital periods, which ranges from 5 to 11.

large x implies that the subhalo’s internal forces are much discussed in Section IV. Additionally, we halt the rmax
stronger than the external tides. The vast difference in and vmax trajectory when rmax becomes smaller than the
the scales of these forces could lead to issues in numeri- resolution limit (see Appendix A), and we only include
cal precision when the tiny tidal forces are added to the simulations whose trajectories cover at least five orbits
large internal forces. about the host. This restriction reduces our simulation
count to 40, 32 of which are in the self-similar regime
(Rc /Rs < 0.3).
Appendix D: The broader density profile
Following H03 and P10, we first explore the rela-
tionship between a subhalo’s structural parameters and
Refs. [78] (hereafter H03) and [18] (hereafter P10) its total mass loss. These prior works parametrize the
studied the tidal evolution of a subhalo’s density profile, mass loss using the ratio mbound /macc , where mbound
focusing on the structural parameters vmax , the maxi- is the mass that remains bound to the subhalo, and
mum circular velocity within the subhalo, and rmax , the macc is the subhalo’s virial mass at accretion. How-
radius at which this velocity is attained. Prior treat- ever, this parametrization implies that the impact of tides
ments of the annihilation rate in subhalos (e.g. [33, 34]) is strongly sensitive to the subhalo’s initial concentra-
have employed these works’ predictions of rmax and vmax , tion, and we propose that this sensitivity is unphysical
along with the assumption that subhalos retain NFW since the outer layers may be stripped almost immedi-
profiles, to predict subhalo J-factors. To understand the ately upon accretion onto the host. To evade this prob-
connection between our work and these prior works, this lem, we instead parametrize the mass loss using the ratio
appendix investigates the evolution of rmax and vmax in m̃ ≡ mbound /mmax,init of mbound to the mass initially en-
our simulations. closed within rmax ; this ratio is initially larger than unity.
At each snapshot of our simulations, p we find rmax as We compute mbound using the procedure in Appendix A,
the radius r < rt that maximizes vcirc = Gm(r)/r, and and Fig. 18 shows these relationships. We find that while
vmax is the corresponding maximum. We only consider rmax is cleanly related to mbound , the relationship be-
snapshots up to the point where |d ln J/d ln t| = 1, as tween vmax and mbound exhibits nontrivial scatter that is
19

1.0 1.0
rmax
rmax /rmax,init
= 0.52m̃0.46 W
rmax,init NF

J/Jinit
0.5 0.5
x
0.3 3 10 30 100
x
3 10 30 100
1.0
P10 0.2
vmax /vmax,init

H03 0.3 1.0


0.8
G−2 vmax
4 −1
rmax /Jinit
0.6 vmax −0.24 FIG. 19. The relationship between a subhalo’s J-factor and
= 1 + m̃−1.56 its structural parameters rmax and vmax ; for an NFW profile,
vmax,init
J = 1.23G−2 vmax4 −1
rmax (solid line). In our simulations, the J-
1.0 factor lies consistently below the value that would be expected
assuming an NFW profile. Each point represents the average
J/Jinit

over a single orbit in our simulations.


0.5

0.3 J −0.54
= 1 + m̃−1.59 where n = t/T is the number of orbits [c.f. Eq. (14)].
Jinit Note that the total bound mass, mbd , behaves similarly:
its trajectory follows from Eq. (D1) by inverting the equa-
0.3 1 3
tion in the top panel of Fig. 18. As shown in Fig. 20, br ,
m̃ ≡ mbound /mmax,init cr , bv , and cv appear to depend on the system parame-
ters x, y, and z in the same way that b and c did (albeit
FIG. 18. The relationship of the subhalo properties rmax with more noise in the cases of cr and cv ):
(top), vmax (middle), and J (bottom) to its bound mass
mbound after tidal stripping. rmax is cleanly related to mbound , br = 0.49x−0.39 [1 + 0.72f (y)] (D3)
but the scatter is larger for vmax and still larger for J. Each 0.13
point represents the average over a single orbit in our sim- cr = 0.92z (D4)
ulations, and the solid lines represent the displayed fitting −0.58
bv = 0.28x [1 + 1.41f (y)] (D5)
functions. The dashed and dotted lines correspond to the 0.23
predictions of P10 and H03, respectively, assuming that the cv = 0.77z . (D6)
initial mass is 4.5mmax,init . The P10 prediction in the last
panel additionally assumes an NFW profile. However, the middle panels of Fig. 20 show that unlike
a, the parameters ar and av depend not only on y but
also on x. Moreover, they are only sensitive to x in the
correlated with the system parameter x = |Eb |/∆Eimp . self-similar regime (Rc /Rs < 0.3). We fit the equation
For comparison, we also plot the predictions of H03 and ar = ar0 ln(x/ar1 ), and likewise for av , in the self-similar
P10 assuming macc = 4.5mmax,init , which corresponds to regime. Next, we fit ar − ar0 ln(x/ar1 )[1 − f (y)/2] =
subhalo concentration c ' 20 at accretion. −ar2 f (y), and likewise for av , using all simulations. The
The bottom panel of Fig. 18 shows that there is even function f (y) asymptotes at 2 for large y, so the combi-
more scatter in the relationship between J and mbound . nation [1 − f (y)/2] suppresses the x-dependent part of ar
This scatter partially results from the scatter in vmax , and av at large r. Hence, we obtain
4
since J ∝ vmax , but it also reflects that tidally altered
ar = 0.52 ln(x/86) [1 − f (y)/2] − 1.24f (y) (D7)
density profiles differ significantly from NFW. Figure 19
investigates this effect further and shows that a subhalo’s av = 0.38 ln(x/12) [1 − f (y)/2] − 1.19f (y), (D8)
actual J-factor is generically smaller than what would be
predicted from rmax and vmax assuming an NFW profile. as depicted in Fig. 20.
We can also predict the evolution of the subhalo’s
structural parameters more explicitly. Subjected to tidal
forces, rmax and vmax follow qualitatively similar trajec-
tories to the J-factor:
 
rmax 1 
ln = br ar − n1−cr − 1 , (D1)
rmax,init 1 − cr
 
vmax 1 
ln = bv av − n1−cv − 1 , (D2)
vmax,init 1 − cv
20

0.3

0.3
br = 0.1 bv =
0.49 − 0.28 −
x 0.39 x 0.58
br

bv
0.1
y 0.03 y
10−2 10−1 100 101 10−2 10−1 100 101

3 10 30 100 3 10 30 100
x = |Eb |/∆Eimp x = |Eb |/∆Eimp
br bv
= 1 + 0.72f (y) = 1 + 1.41f (y)



3
0.49x−0.39 0.28x−0.58
br / 0.49x−0.39

2.0

bv / 0.28x−0.58
1.5 2

x x
1.0 3 10 30 100 3 10 30 100
1

10−2 10−1 100 101 10−2 10−1 100 101


y = R̄/Rs y = R̄/Rs

1
0 −2 −1 0 1
10−2 10−1 100 101
10 10 10 10 y
y
0
av

−1
ar

−1
−2
ar = 0.52 ln (x/86) av = 0.38 ln (x/12)

101 102
101 102
av −0.38 ln(x/12) [1−f (y)/2]

x = |Eb |/∆Eimp
ar −0.52 ln(x/86) [1−f (y)/2]

x = |Eb |/∆Eimp
0.5
0 3 10 30 100
0.0 3 10 30 100 x
x
−0.5
−1
−1.0

−1.5 −2
ar −0.52 ln(x/86) [1−f (y)/2] av −0.38 ln(x/12) [1−f (y)/2]
−2.0 = −1.24f (y) = −1.19f (y)
10−2 10−1 100 101 10−2 10−1 100 101
y = R̄/Rs y = R̄/Rs

1.0
1.0 cr = 0.92z 0.13
cv = 0.77z 0.23
cv
cr

100
0.3 30
x
x

3 10 30 100 10
3
0.3
10−2 10−1 100 10−2 10−1 100
z = ra /rt z = ra /rt

FIG. 20. The rmax trajectory parameters ar , br , and cr (left panels) and the vmax trajectory parameters av , bv , and cv (right
panels) plotted against the system parameters x, y, and z (c.f. Figs. 9, 10, and 12); see Eqs. (D1) and (D2). The radius of
each marker is proportional to the number of orbital periods, which ranges from 5 to 15 for this sample.
21

[1] F. Zwicky, Helvetica Physica Acta 6, 110 (1933). A. Helmi, Nature 456, 73 (2008), 0809.0894.
[2] D. Clowe, A. Gonzalez, and M. Markevitch, Astrophys. [31] M. Stref and J. Lavalle, Phys. Rev. D 95, 063003 (2017),
J. 604, 596 (2004), astro-ph/0312273. 1610.02233.
[3] P. A. R. Ade et al. (Planck Collaboration), Astron. As- [32] M. Stref, T. Lacroix, and J. Lavallle, 1905.02008.
trophys. 594, A13 (2016), 1502.01589. [33] R. Bartels and S. Ando, Phys. Rev. D 92, 123508 (2015),
[4] G. Bertone and T. M. Tait, Nature 562, 51 (2018), 1507.08656.
1810.01668. [34] N. Hiroshima, S. Ando, and T. Ishiyama, Phys. Rev.
[5] J. L. Feng, Ann. Rev. Astron. Astrophys. 48, 495 D 97, 123002 (2018), 1803.07691.
(2010), 1003.0904. [35] S. Ando, T. Ishiyama, and N. Hiroshima, 1903.11427.
[6] G. Bertone, Particle dark matter: Observations, models [36] A. L. Erickcek and K. Sigurdson, Phys. Rev. D 84,
and searches (Cambridge University Press, 2010). 083503 (2011), 1106.0536.
[7] G. Bertone, D. Hooper, and J. Silk, Phys. Rep. 405, [37] G. Barenboim and J. Rasero, J. High Energy Phys.
279 (2005), hep-ph/0404175. 2014, 138 (2014), 1311.4034.
[8] L. Bergström, Rep. Prog. Phys. 63, 793 (2000), hep- [38] J. J. Fan, O. Özsoy, and S. Watson, Phys. Rev. D 90,
ph/0002126. 043536 (2014), 1405.7373.
[9] G. Jungman, M. Kamionkowski, and K. Griest, Phys. [39] A. L. Erickcek, Phys. Rev. D 92, 103505 (2015),
Rep. 267, 195 (1996), hep-ph/9506380. 1504.03335.
[10] L. E. Strigari, S. M. Koushiappas, J. S. Bullock, and [40] K. Redmond, A. Trezza, and A. L. Erickcek, Phys. Rev.
M. Kaplinghat, Phys. Rev. D 75, 083526 (2007), astro- D 98, 063504 (2018), 1807.01327.
ph/0611925. [41] J. Silk and M. S. Turner, Phys. Rev. D 35, 419 (1987).
[11] J. F. Navarro, C. S. Frenk, and S. D. M. White, Astro- [42] D. S. Salopek, J. R. Bond, and J. M. Bardeen, Phys.
phys. J. 462, 563 (1996), astro-ph/9508025. Rev. D 40, 1753 (1989).
[12] J. F. Navarro, C. S. Frenk, and S. D. M. White, Astro- [43] A. A. Starobinskij, JETP Lett. 55, 489 (1992).
phys. J. 490, 493 (1997), astro-ph/9611107. [44] P. Ivanov, P. Naselsky, and I. Novikov, Phys. Rev. D
[13] A. Brooks, Ann. Phys. (Berlin) 526, 294 (2014), 50, 7173 (1994).
1407.7544. [45] L. Randall, M. Soljačić, and A. H. Guth, Nucl. Phys.
[14] J. Binney and S. Tremaine, Galactic Dynamics (Prince- B472, 377 (1996), hep-ph/9512439.
ton University Press, Princeton, NJ, 1987). [46] E. D. Stewart, Phys. Rev. D 56, 2019 (1997), hep-
[15] H. Mo, F. Van den Bosch, and S. White, Galaxy forma- ph/9703232.
tion and evolution (Cambridge University Press, 2010). [47] J. A. Adams, G. G. Ross, and S. Sarkar, Nucl. Phys.
[16] S. Chandrasekhar, Astrophys. J. 97, 255 (1943). B503, 405 (1997), hep-ph/9704286.
[17] T. Goerdt, O. Y. Gnedin, B. Moore, J. Diemand, and [48] A. A. Starobinsky, Grav. Cosmol. 4, 489 (1998), astro-
J. Stadel, Mon. Not. R. Astron. Soc. 375, 191 (2007), ph/9811360.
astro-ph/0608495. [49] L. Covi and D. H. Lyth, Phys. Rev. D 59, 063515 (1999),
[18] J. Penarrubia, A. J. Benson, M. G. Walker, G. Gilmore, hep-ph/9809562.
A. W. McConnachie, and L. Mayer, Mon. Not. R. As- [50] J. Martin, A. Riazuelo, and M. Sakellariadou, Phys.
tron. Soc. 406, 1290 (2010), 1002.3376. Rev. D 61, 083518 (2000), astro-ph/9904167.
[19] R. Errani and J. Peñarrubia, 1906.01642. [51] D. J. H. Chung, E. W. Kolb, A. Riotto, and
[20] V. Berezinsky, V. Dokuchaev, and Y. Eroshenko, Phys. I. I. Tkachev, Phys. Rev. D 62, 043508 (2000), hep-
Rev. D 77, 083519 (2008), 0712.3499. ph/9910437.
[21] V. Berezinsky, V. Dokuchaev, and Y. Eroshenko, Phys. [52] J. Martin and R. H. Brandenberger, Phys. Rev. D 63,
Rev. D 68, 103003 (2003), astro-ph/0301551. 123501 (2001), hep-th/0005209.
[22] J. Diemand, B. Moore, and J. Stadel, Nature 433, 389 [53] M. Joy, V. Sahni, and A. A. Starobinsky, Phys. Rev. D
(2005), astro-ph/0501589. 77, 023514 (2008), 0711.1585.
[23] L. Pieri, G. Bertone, and E. Branchini, Mon. Not. R. [54] N. Barnaby and Z. Huang, Phys. Rev. D 80, 126018
Astron. Soc. 384, 1627 (2008), 0706.2101. (2009), 0909.0751.
[24] T. Ishiyama, J. Makino, and T. Ebisuzaki, Astrophys. [55] N. Barnaby, Phys. Rev. D 82, 106009 (2010), 1006.4615.
J. Lett. 723, L195 (2010), 1006.3392. [56] I. Ben-Dayan and R. Brustein, J. Cosmol. Astropart.
[25] D. Anderhalden and J. Diemand, J. Cosmol. Astropart. Phys. 09, 007 (2010), 0907.2384.
Phys. 04, 009 (2013), 1302.0003; 08, E02 (2013). [57] J.-O. Gong and M. Sasaki, J. Cosmol. Astropart. Phys.
[26] T. Ishiyama, Astrophys. J. 788, 27 (2014), 1404.1650. 03, 028 (2011), 1010.3405.
[27] M. A. Sánchez-Conde and F. Prada, Mon. Not. R. As- [58] D. H. Lyth, J. Cosmol. Astropart. Phys. 07, 035 (2011),
tron. Soc. 442, 2271 (2014), 1312.1729. 1012.4617.
[28] B. Anderson, S. Zimmer, J. Conrad, M. Gustafsson, [59] E. Bugaev and P. Klimai, J. Cosmol. Astropart. Phys.
M. Sánchez-Conde, and R. Caputo, J. Cosmol. As- 11, 028 (2011), 1107.3754.
tropart. Phys. 2016, 026 (2016), 1511.00014. [60] N. Barnaby and M. Peloso, Phys. Rev. Lett. 106,
[29] L. Gao, C. Frenk, A. Jenkins, V. Springel, and 181301 (2011), 1011.1500.
S. White, Mon. Not. R. Astron. Soc. 419, 1721 (2011), [61] A. Achúcarro, J.-O. Gong, S. Hardeman, G. A. Palma,
1107.1916. and S. P. Patil, J. Cosmol. Astropart. Phys. 01, 030
[30] V. Springel, S. D. White, C. S. Frenk, J. F. Navarro, (2011), 1010.3693.
A. Jenkins, M. Vogelsberger, J. Wang, A. Ludlow, and [62] S. Cespedes, V. Atal, and G. A. Palma, J. Cosmol.
22

Astropart. Phys. 05, 008 (2012), 1201.4848. [91] V. Springel, Mon. Not. R. Astron. Soc. 364, 1105
[63] N. Barnaby, E. Pajer, and M. Peloso, Phys. Rev. D 85, (2005), astro-ph/0505010.
023525 (2012), 1110.3327. [92] L. M. Widrow, Astrophys. J. Suppl. Ser. 131, 39 (2000).
[64] T. Bringmann, P. Scott, and Y. Akrami, Phys. Rev. D [93] M. Fujii, Y. Funato, and J. Makino, Pub. Astron. Soc.
85, 125027 (2012), 1110.2484. Japan 58, 743 (2006), astro-ph/0511651.
[65] M. S. Delos, A. L. Erickcek, A. P. Bailey, and M. A. [94] M. Fellhauer and D. Lin, Mon. Not. R. Astron. Soc.
Alvarez, Phys. Rev. D 98, 063527 (2018), 1806.07389. 375, 604 (2007), astro-ph/0611557.
[66] C. Blanco, M. S. Delos, A. L. Erickcek, and D. Hooper, [95] O. Y. Gnedin, L. Hernquist, and J. P. Ostriker, Astro-
1906.00010. phys. J. 514, 109 (1999), astro-ph/9709161.
[67] F. C. van den Bosch, Mon. Not. R. Astron. Soc. 468, [96] L. S. Spitzer Jr, Dynamical evolution of globular clusters
885 (2017), 1611.02657. (Princeton University Press, 1987).
[68] F. C. van den Bosch, G. Ogiya, O. Hahn, and A. Burk- [97] M. D. Weinberg, Astron. J. 108, 1398 (1994), astro-
ert, Mon. Not. R. Astron. Soc. 474, 3043 (2017), ph/9404015.
1711.05276. [98] M. D. Weinberg, Astron. J. 108, 1403 (1994), astro-
[69] F. C. van den Bosch and G. Ogiya, Mon. Not. R. Astron. ph/9404016.
Soc. 475, 4066 (2018), 1801.05427. [99] O. Y. Gnedin and J. P. Ostriker, Astrophys. J. 513, 626
[70] J. E. Taylor and A. Babul, Astrophys. J. 559, 716 (1999), astro-ph/9902326.
(2001), astro-ph/0012305. [100] A. Klypin, S. Gottlöber, A. V. Kravtsov, and
[71] J. Peñarrubia and A. J. Benson, Mon. Not. R. Astron. A. M. Khokhlov, Astrophys. J. 516, 530 (1999), astro-
Soc. 364, 977 (2005), astro-ph/0412370. ph/9708191.
[72] F. C. Van Den Bosch, G. Tormen, and C. Giocoli, Mon. [101] A. Klypin, F. Prada, G. Yepes, S. Heß, and
Not. R. Astron. Soc. 359, 1029 (2005), arXiv:astro- S. Gottlöber, Mon. Not. R. Astron. Soc. 447, 3693
ph/0409201. (2015), 1310.3740.
[73] A. R. Zentner, A. A. Berlind, J. S. Bullock, A. V. [102] G. Tormen, Mon. Not. R. Astron. Soc. 290, 411 (1997),
Kravtsov, and R. H. Wechsler, Astrophys. J. 624, 505 astro-ph/9611078.
(2005), astro-ph/0411586. [103] S. Khochfar and A. Burkert, Astron. Astrophys. 445,
[74] M. Kampakoglou and A. J. Benson, Mon. Not. R. As- 403 (2006), astro-ph/0309611.
tron. Soc. 374, 775 (2006), astro-ph/0607024. [104] A. R. Wetzel, Mon. Not. R. Astron. Soc. 412, 49 (2011),
[75] J. Gan, X. Kang, F. C. Van Den Bosch, and J. Hou, 1001.4792.
Mon. Not. R. Astron. Soc. 408, 2201 (2010), 1007.0023. [105] L. Jiang, S. Cole, T. Sawala, and C. S. Frenk, Mon.
[76] A. R. Pullen, A. J. Benson, and L. A. Moustakas, As- Not. R. Astron. Soc. 448, 1674 (2015), 1409.1179.
trophys. J. 792, 24 (2014), arXiv:1407.8189. [106] M. S. Delos, A. L. Erickcek, and T. Linden, (in prepa-
[77] F. Jiang and F. C. van den Bosch, Mon. Not. R. Astron. ration).
Soc. 458, 2848 (2016), 1403.6827. [107] J. F. Navarro, A. Ludlow, V. Springel, J. Wang, M. Vo-
[78] E. Hayashi, J. F. Navarro, J. E. Taylor, J. Stadel, gelsberger, S. D. M. White, A. Jenkins, C. S. Frenk, and
and T. Quinn, Astrophys. J. 584, 541 (2003), astro- A. Helmi, Mon. Not. R. Astron. Soc. 402, 21 (2010).
ph/0203004. [108] V. Berezinsky, V. Dokuchaev, and Y. Eroshenko, Phys.
[79] G. Ogiya, F. C. van den Bosch, O. Hahn, S. B. Green, Rev. D 73, 063504 (2006), astro-ph/0511494.
T. B. Miller, and A. Burkert, Mon. Not. R. Astron. [109] H. Zhao, D. Hooper, G. W. Angus, J. E. Taylor,
Soc. (2019), 1901.08601. and J. Silk, Astrophys. J. 654, 697 (2007), astro-
[80] S. Kazantzidis, L. Mayer, C. Mastropietro, J. Diemand, ph/0508215.
J. Stadel, and B. Moore, Astrophys. J. 608, 663 (2004), [110] A. M. Green and S. P. Goodwin, Mon. Not. R. Astron.
astro-ph/0312194. Soc. 375, 1111 (2007), astro-ph/0604142.
[81] J. I. Read, M. Wilkinson, N. Evans, G. Gilmore, and [111] A. Schneider, L. Krauss, and B. Moore, Phys. Rev. D
J. T. Kleyna, Mon. Not. R. Astron. Soc. 366, 429 82, 063525 (2010), 1004.5432.
(2006), astro-ph/0506687. [112] G. Angus and H. Zhao, Mon. Not. R. Astron. Soc. 375,
[82] B. Moore, Nature 370, 629 (1994). 1146 (2007), astro-ph/0608580.
[83] J. I. Read, M. G. Walker, and P. Steger, Mon. Not. R. [113] V. S. Berezinsky, V. I. Dokuchaev, and Y. N.
Astron. Soc. 481, 860 (2018), 1805.06934. Eroshenko, Physics-Uspekhi 57, 1 (2014), 1405.2204.
[84] E. Polisensky and M. Ricotti, Mon. Not. R. Astron. Soc. [114] E. D’Onghia, V. Springel, L. Hernquist, and D. Keres,
450, 2172 (2015), 1504.02126. Astrophys. J. 709, 1138 (2010), 0907.3482.
[85] G. Ogiya and O. Hahn, Mon. Not. R. Astron. Soc. 473, [115] Q. Zhu, F. Marinacci, M. Maji, Y. Li, V. Springel,
4339 (2018), 1707.07693. and L. Hernquist, Mon. Not. R. Astron. Soc. 458, 1559
[86] M. S. Delos, A. L. Erickcek, A. P. Bailey, and M. A. (2016), 1506.05537.
Alvarez, Phys. Rev. D 97, 041303 (2018), 1712.05421. [116] T. Kelley, J. S. Bullock, S. Garrison-Kimmel,
[87] R. E. Angulo, O. Hahn, A. D. Ludlow, and M. Boylan-Kolchin, M. S. Pawlowski, and A. S. Graus,
S. Bonoli, Mon. Not. R. Astron. Soc. 471, 4687 (2017), (2018), 1811.12413.
1604.03131. [117] M. Hütten, M. Stref, C. Combet, D. Maurin, and
[88] M. S. Delos, M. Bruff, and A. L. Erickcek, 1905.05766. J. Lavalle, Galaxies 7, 60 (2019), 1904.10935.
[89] G. Ogiya, D. Nagai, and T. Ishiyama, Mon. Not. R. [118] A. W. McConnachie, Astron. J. 144, 4 (2012),
Astron. Soc. 461, 3385 (2016), 1604.02866. 1204.1562.
[90] V. Springel, N. Yoshida, and S. D. M. White, New [119] D. A. Green, Bull. Astron. Soc. India 39, 289 (2011),
Astron. 6, 79 (2001), astro-ph/0003162. 1108.5083.
[120] J. Barnes and P. Hut, Nature 324, 446 (1986).

Vous aimerez peut-être aussi