ArticlePDF Available

Cluster-counterpart Voids: Void Identification from Galaxy Density Field

Authors:

Abstract and Figures

We identify cosmic voids from galaxy density fields under the theory of void–cluster correspondence. We extend the previous novel void-identification method developed for the matter density field to the galaxy density field for practical applications. From cosmological N -body simulations, we construct galaxy number- and mass-weighted density fields to identify cosmic voids that are counterparts of galaxy clusters of a specific mass. The parameters for the cluster-counterpart void identification such as Gaussian smoothing scale, density threshold, and core volume fraction are found for galaxy density fields. We achieve about 60%–67% of completeness and reliability for identifying the voids of corresponding cluster mass above 3 × 10 ¹⁴ h ⁻¹ M ⊙ from a galaxy sample with the mean number density, n ¯ = 4.4 × 10 − 3 ( h − 1 Mpc ) − 3 . When the mean density is increased to n ¯ = 10 − 2 ( h − 1 Mpc ) − 3 , the detection rate is enhanced by ∼2%–7% depending on the mass scale of voids. We find that the detectability is insensitive to the density weighting scheme applied to generate the density field. Our result demonstrates that we can apply this method to the galaxy redshift survey data to identify cosmic voids corresponding statistically to the galaxy clusters in a given mass range.
Content may be subject to copyright.
Cluster-counterpart Voids: Void Identication from Galaxy Density Field
Junsup Shim
1,2
, Changbom Park
1
, Juhan Kim
3
, and Sungwook E. Hong ()
4,5
1
School of Physics, Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea
2
Institute of Astronomy and Astrophysics, Academia Sinica, No.1, Sec. 4, Roosevelt Rd, Taipei 10617, Taiwan
3
Center for Advanced Computation, Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea; kjhan@kias.re.kr
4
Korea Astronomy and Space Science Institute, 776 Daedeokdae-ro, Yuseong-gu, Daejeon 34055, Republic of Korea
5
Astronomy Campus, University of Science & Technology, 776 Daedeok-daero, Yuseong-gu, Daejeon 34055, Republic of Korea
Received 2022 October 6; revised 2023 April 29; accepted 2023 May 15; published 2023 July 18
Abstract
We identify cosmic voids from galaxy density elds under the theory of voidcluster correspondence. We extend
the previous novel void-identication method developed for the matter density eld to the galaxy density eld for
practical applications. From cosmological N-body simulations, we construct galaxy number- and mass-weighted
density elds to identify cosmic voids that are counterparts of galaxy clusters of a specic mass. The parameters for
the cluster-counterpart void identication such as Gaussian smoothing scale, density threshold, and core volume
fraction are found for galaxy density elds. We achieve about 60%67% of completeness and reliability for
identifying the voids of corresponding cluster mass above 3 ×10
14
h
1
M
e
from a galaxy sample with the mean
number density,
n
h4.4 10 Mpc
31 3
¯()
-- -
. When the mean density is increased to
n
h10 Mpc
21 3
¯()=-- -
, the
detection rate is enhanced by 2%7% depending on the mass scale of voids. We nd that the detectability is
insensitive to the density weighting scheme applied to generate the density eld. Our result demonstrates that we
can apply this method to the galaxy redshift survey data to identify cosmic voids corresponding statistically to the
galaxy clusters in a given mass range.
Unied Astronomy Thesaurus concepts: Large-scale structure of the universe (902);Voids (1779)
1. Introduction
Cosmic voids are the largest volume component of the large-
scale structures in the universe (Joeveer & Einasto 1978;
Einasto et al. 1980; de Lapparent et al. 1986; Gott et al. 1986;
Vogeley et al. 1994)but contain a relatively small amount of
mass. Because of their underdense nature, voids have been
utilized as a probe for dark energy (Lee & Park 2009; Lavaux
& Wandelt 2010;Li2011; Pisani et al. 2015; Sutter et al. 2015;
Achitouv 2017), gravitation (Nusser et al. 2005; Li et al. 2012;
Cai et al. 2015; Achitouv 2019), initial conditions (Kim &
Park 1998), as well as a laboratory for studying the
environmental effect on galaxy formation and evolution
(Ceccarelli et al. 2008; Park & Lee 2009; Kreckel et al.
2011; Beygu et al. 2013; Shim et al. 2015; Ceccarelli et al.
2022).
For void identication from simulation and obser-
vation, various void-nding methods have been developed
(Kauffmann & Fairall 1991; El-Ad et al. 1996; Colberg et al.
2005; Patiri et al. 2006b; Hahn et al. 2007; Platen et al. 2007;
Neyrinck 2008; Forero-Romero et al. 2009; Lavaux &
Wandelt 2010; Sutter et al. 2015; Shim et al. 2021). With
reasonable void-nding parameter values, a void-nder may
identify voids that are aspherical and hierarchical. However,
neither the asphericity nor hierarchical structure of voids
are taken into account in the spherical void formation
model (Fillmore & Goldreich 1984; Suto et al. 1984;
Bertschinger 1985). Thus, it is not unnatural to nd some
discrepancy between voids in data and the spherical void
model. For example, the shell-crossing density threshold
derived from the statistics of voids in data is inconsistent with
the prediction based on the spherical void approximation (Chan
et al. 2014; Achitouv et al. 2015; Nadathur & Hotchkiss 2015).
This motivates a search for a new denition and identication
method for voids that mitigate the inconsistency between voids
in theory and data.
Recently, a new concept of dening voids has been
presented (Pontzen et al. 2016; Shim et al. 2021; Stopyra
et al. 2021; Desmond et al. 2022). The key idea of void
identication in these studies is that a void of a certain size can
be related with a dark matter halo of a particular mass because a
halo region would have become a void if the initial overdensity
eld were inverted. The correspondence between voids and
dark matter halos in the inverted eld is more reliable on the
scale of massive galaxy clusters. This argument leads us to
dene cosmic voids as a counterpart to galaxy clusters, and we
will call the void of a certain size, identied and related with
the cluster of the corresponding mass, the cluster-counterpart
void (CCV). Other interesting features of CCVs include (1)
almost universal density proles (Shim et al. 2021)and (2)
their mean density contrast being close to the prediction from
the spherical expansion model (Stopyra et al. 2021; Desmond
et al. 2022). Interestingly, (3)they are still in the quasi-linear
regime, and information on the initial conditions is better
preserved relative to clusters (Kimzx & Park 1998; Stopyra
et al. 2021).
In order to identify CCVs from a given density eld, two
different approaches have been suggested. The major differ-
ence between the two approaches is whether the void
identication method requires the reconstruction of the initial
density elds. Without utilizing the initial density eld, Shim
et al. (2021)established a simple CCV-identication method
adopting free parameters for smoothing scale, density, and
volume thresholds. On the other hand, reversely evolving a
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 https://doi.org/10.3847/1538-4357/acd852
© 2023. The Author(s). Published by the American Astronomical Society.
Original content from this work may be used under the terms
of the Creative Commons Attribution 4.0 licence. Any further
distribution of this work must maintain attribution to the author(s)and the title
of the work, journal citation and DOI.
1
given density eld to the initial state and then forwardly
evolving its sign-inverted density uctuation eld to the present
epoch to identify CCVs was suggested by Stopyra et al. (2021)
and implemented by Desmond et al. (2022). In this study, we
extend and apply the former approach adopted in Shim et al.
(2021)to galaxy density elds.
This paper is organized as follows. We describe our
simulations and the scheme for assigning galaxies to halos in
Section 2. In Section 3, we briey review the identication of
CCVs from dark matter density elds and describe how we
apply this approach to galaxy density elds in Section 4.We
discuss our results and conclude in Section 5.
2. Simulation and Galaxy Mocks
We use a pair of N-body simulations with inverted initial
overdensity elds to study the CCVs and their corresponding
clusters. The Mirror simulation used in this work is one of the
Multiverse simulations introduced by Park et al. (2019),
Tonegawa et al. (2020). The Reference simulation starts with
the initial density uctuations that have the same amplitude as in
the Mirror simulation but have the opposite sign. The simulations
adopt the GOTPM (Dubinski et al. 2004)code and WMAP 5 yr
ΛCDM cosmology (Dunkley et al. 2009)with the matter, baryon,
dark energy density parameters set to 0.26, 0.044, and 0.74,
respectively. Each simulation evolves N
p
=2048
3
dark matter
particles, with mass m
p
;9×10
9
h
1
M
e
, in a periodic cubic box
of side length L
box
=1024 h
1
Mpc.
We identify dark matter halos with 30 or more particles
using the friend-of-friend algorithm with the linking length
ll0.2
link p
=, where lpis the mean particle separation. We then
populate the identied halos with mock galaxies based on the
most bound member particle (MBP)galaxy correspondence
model (Hong et al. 2016). In this galaxy assignment scheme, all
MBPs marked in halo merger trees are the proxies for galaxies
(De Lucia et al. 2004; Faltenbacher & Diemand 2006). The
position and velocity of a galaxy are taken from those of an
MBP, whereas galaxy luminosity is determined by the
abundance matching between the mass function of mock
galaxies (see Hong et al. 2016, for the denition of galaxy
mass)and the luminosity function of the SDSS main galaxies
(Choi et al. 2007). The survival time of a satellite halo is given
by the merger timescale modeled with a modied version of the
tting formula described in Jiang et al. (2008). We set the
power-law index of the host-to-satellite mass ratio to 1.5. This
is to yield a good match for the projected two-point correlation
functions between the mock galaxies in our simulations and
SDSS main galaxies down to scales below 1 h
1
Mpc (Zehavi
et al. 2011).
We construct two galaxy samples with different mean number
densities to test the effect of sample density on the CCV
identication. The mean number densities of the sparse and
dense samples are
n
4.4 10
gal 3
¯
-and 1.0 ×10
2
h
3
Mpc
3
,
respectively. We chose the sparse sample density as such
assuming that the sparse galaxy sample mimics a volume-limited
sample from the SDSS main galaxy sample with the largest
survey volume (Choi et al. 2010). For the dense sample, its mean
density is set close to the highest achievable density from the
simulations. The dense sample-like data will be available from
upcoming surveys such as Dark Energy Spectroscopic Instru-
ment (DESI Collaboration et al. 2016)and SPHEREx (Doré
et al. 2014)given that their expected number densities of
observed galaxies up to z0.2 are higher than our dense sample
density.
3. Identifying CCVs with Dark Matter
We briey summarize how CCVs are dened using paired
simulations in Section 3.1 and are recovered in dark matter
density elds without using the inverted simulation in
Section 3.2.
3.1. Cluster-counterpart Voids
In Figure 1, we illustrate how CCVs are identied using the
Reference and Mirror simulations. We note that the cluster-
scale dark matter halos in the Mirror simulation become large
voids in the Reference simulation. Therefore, the dark matter
particles belonging to each cluster-scale halo in one simulation
can be used to dene the void region in the other one. Thus,
each cluster in the Mirror simulation uniquely denes a CCV in
the Reference simulation.
We generate a catalog containing the position, effective
radius, and mean and central densities of CCVs in the
Reference simulation that correspond to 422,818 halos with
halo mass M
h
10
13
h
1
M
e
,atz=0 in the Mirror simulation.
The volume of a CCV is dened as the region occupied by the
void member particles in the Reference simulation. A CCV
associated with a more massive dark matter halo tends to have
a larger volume (Shim et al. 2021).
3.2. CCVs from Dark Matter Density Fields
In this subsection, we describe a method developed by Shim
et al. (2021)for nding the CCVs in the Reference simulation
without resorting to the Mirror simulation. First, we nd the
most underdense regions in the smoothed density eld below a
certain threshold density, and then select those whose volume
is larger than a certain fraction of the typical void volume. The
expected typical or mean void volume is directly related with
the corresponding cluster mass. Namely, a choice of cluster
mass determines the corresponding CCV size, which is used to
select the void core regions. Our method is based on the nding
that a larger CCV tends to develop a larger void core, the
central region with density below the specic threshold value.
There are three parameters used in this method: Gaussian
smoothing scale (R
s
), core density threshold (δ
c
), and core
volume fraction (f
c
)with respect to the CCV volume at that
mass scale. The optimal values of these parameters change
depending on the mean void size or corresponding cluster
mass. The matter density eld is smoothed with a Gaussian
lter over R
s
, and the void core regions with overdensity below
δ
c
are found. Among them, those with volume larger than f
c
times the typical volume of the CCVs under interest are
selected. Because void cores only cover the innermost volume
of CCVs, we then grow the cores by repeatedly attaching
neighboring higher-density volume elements. This growing
process stops when the total volume of the recovered voids
reaches that of the CCVs of the interested mass scale in the
Reference simulation.
4. Identifying CCVs from Galaxy Density Fields
We test whether or not the CCV-identication method
originally developed for dark matter density elds could be also
applied to galaxy density elds. The success of the extension of
2
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
the prescription depends on how accurately the galaxy density
eld follows the underlying dark matter density eld, especially
in underdense regions. To be more specic, the rank order of
pixel density should be preserved; a region with a higher dark
matter density should have a higher galaxy density value if we
aim to identify the same voids in the galaxy and dark matter
density elds using this method. Any reversal in the rank order
will disturb the correspondence. Thus, we rst examine the
relation between galaxy and dark matter density elds in
Section 4.1 and determine the optimal values of free parameters
for galaxy density elds in Section 4.2.
4.1. Dark Matter versus Galaxy Density Fields
Before we develop a void-nding algorithm for a galaxy
sample, we rst compare between the number- and mass-
weighted galaxy density elds to study which one has the
tighter relation with the underlying matter distribution. The
weighted density is calculated at the center of each pixel with a
volume h2Mpc
13
(
)
-using the cloud-in-cell assignment
scheme. We then smooth these density elds with Gaussian
smoothing kernel. In Figure 2, we compare between smoothed
dark matter and galaxy density elds. As can be seen, the
overdensities of both galaxy density elds are enhanced
compared to that of the dark matter density eld, which
reects galaxies being the biased tracers of the underlying
matter eld (Kaiser 1984; Bardeen et al. 1986; Desjacques et al.
2018). Interestingly, the nonlinear relation between dark matter
and galaxy density elds is well modeled by the second-order
polynomials of logarithmic densities as discovered in Jee et al.
(2012). Note that this relation reduces to a linear bias model
when |δ|=1.
We nd that the mass-weighted galaxy density is more
tightly correlated with the dark matter density than the number-
weighted case. The standard deviation of the logarithmic dark
matter overdensity for a given logarithmic galaxy overdensity
bin is on average 10% smaller in the mass-weighted galaxy
eld than that in the number-weighted one. This is consistent
Figure 1. A schematic diagram showing how CCVs are identied under the paradigm of the voidcluster correspondence theory using the Reference and Mirror
simulations. Top panels: initial density uctuations of the same regions within the Reference (left)and Mirror (right)simulations. The initial density troughs (blue
regions)in the Reference simulation correspond to the initial density peaks (red regions)in the Mirror simulation because the initial overdensity elds of the two
simulations are designed to be the sign-inverted version of the other. Bottom panels: distributions of dark matter particles forming voids (left)and their counterpart
clusters (right)through gravitational evolution from their initial particle distributions representing initial underdensities (top left)and overdensities (top right),
respectively. Note that we illustrate only the density troughs and voids for the Reference simulation, and their corresponding density peaks and clusters for the Mirror
simulation, for simplicity.
3
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
with the previous results showing that mass weighting reduces
the scatter between dark matter and halo density eld (Park
et al. 2007,2010; Seljak et al. 2009). However, the standard
deviation for the mass-weighted galaxy density eld becomes
comparable to that for the number-weighted case at an
extremely low-density range. For this reason, we consider
both the number- and mass-weighted galaxy density elds even
though the scatter is smaller overall in the mass-weighted case.
The scatter in the relation between the galaxy and dark matter
elds reects the stochastic nature of bias (Pen 1998; Dekel &
Lahav 1999; Matsubara 1999)and the shot-noise effect on
density measurement. Because of these two effects, the rank
order of pixel values in the density array changes when we
move from the matter density eld to galaxy density eld.
In Figure 3, we show how much the degree of randomization
in the rank order depends on the density-weighting scheme and
smoothing scale. We compute the volume overlap fraction
(f
overlap
)of the lowest-density regions between galaxy and dark
matter density elds as
fV
V,1
overlap
overlap
low
()=
where V
low
is the volume with density lower than the input
threshold, and V
overlap
is the overlap volume between those
lowest-density regions in the two density elds. And the fraction
of volume occupied by the lowest-density regions is given as
fV
V,2
low
low
sim
()=
where
V
sim is the entire simulation volume. For example,
f
overlap
=0, at f
low
=0.02, means that the most underdense 2%
volumes in galaxy and dark matter density eld have no
overlap between them.
We nd the decreasing trend of the volume overlap fraction
toward a lower f
low
, which becomes steeper for a smaller
smoothing scale. This implies that the order shufing between
two density elds is severe at the center of voids, or in the most
underdense regions. It is also shown that the overlap fraction is
higher for a larger smoothing scale implying that the shufing
becomes less signicant when galaxy density elds are
smoothed on relatively larger smoothing scales. However, the
density order at the extreme low-density range is still changed
by 10% for the largest smoothing scale.
4.2. Optimal Parameter Values for CCV Identication
We need to ne-tune the free parameters for CCV
identication in the galaxy density eld. This is because the
galaxy density eld is nonlinearly biased with respect to the
underlying dark matter density eld, and hence the ne-tuned
parameters obtained for the dark matter eld cannot be applied
to the galaxy density eld without adjustment. We thus follow
the approach described in Shim et al. (2021). The optimal
Figure 2. Relation between dark matter and galaxy overdensities. The upper
(lower)panel shows the case for the number- (mass-)weighted galaxy density
eld. The Gaussian smoothing scale applied is R
s
=7.3 h
1
Mpc. Solid lines
are the second-order polynomial ts to the relations (Jee et al. 2012), whereas
dash lines represent the identity relation, δ
DM
=δ
gal
.
Figure 3. Volume overlap fraction (f
overlap
)of underdense regions between
dark matter and galaxy density elds as a function of the volume fraction (f
low
)
of the low-density regions. The galaxy density elds are constructed using the
dense galaxy sample. Solid (dashed)lines represent number- (mass-)weighted
galaxy density elds. Different colors indicate various Gaussian smoothing
scales.
4
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
values of the free parameters are determined so that complete-
ness and reliability are maximized. We calculate the complete-
ness and reliability of void nding given by
NN,3
sv
()º
and
NN,4
sc
()º
respectively. Here, N
v
,N
c
, and N
s
represent the number of the
CCVs of a target mass scale, all the identied void cores, and
successfully reproduced CCVs, respectively. From here on, the
mass scale of a CCV refers to the corresponding cluster mass.
The completeness represents the successfully recovered
fraction among the CCVs with the corresponding cluster mass
above Mmi
n
. On the other hand, the reliability is the fraction of
recovered voids with the corresponding MM0.8 min
(to allow
for a buffer in mass)among all the identied voids by this
method. Here, it is counted as a successful detection when the
nearest CCV from the identied void core satises the mass
criterion or when the most massive CCV within 1.5r
c
from the
void core satises the mass criterion. Here, r
c
is the effective
radius of the void core and is measured from the void core
volume V
c
using the following relation
rV3
4.5
cc
1
3
()
p
=
We repeat calculating the detection completeness and
reliability in three-dimensional parameter space of R
s
,δ
c
, and
f
c
, and provide their optimal values in Table 1. Note that we
limit the smoothing scales to be the reciprocals of integers
multiplied by the effective radius of the CCVs of the target
mass scale. The optimal length scales for smoothing galaxy
density elds are R
s
=R
v
/3 in most cases, where R
v
is the
effective radius of a CCV of a particular mass scale. This
implies that the optimized smoothing scale of a given galaxy
density eld is determined by the target mass scale, or
equivalently the target size, of voids. Interestingly, the relation
between the smoothing scale and void size is consistent with
the case for the dark matter density elds (Shim et al. 2021).
In Figure 4, we show the detection completeness and
reliability for different minimum target mass scales of CCVs. It
is shown that the detection completeness, and equivalently the
reliability, does not depend much on the density-weighting
Table 1
Optimal Values of the Free Parameters for the CCV Identication from Galaxy Density Fields Constructed Using the Dense and Sparse Galaxy Samples
ngal
¯
M
min R
v
R
s
δ
c
δ
upper
f
c
((h
1
Mpc )
3
)(10
14
h
1
M
e
)(h
1
Mpc)(R
v
)
number-weighted density eld 1.0 ×10
2
3 14.5 1/40.981 0.785 0.02
5 17.2 1/30.915 0.738 0.02
7 19.3 1/30.923 0.725 0.01
10 21.7 1/30.890 0.727 0.05
4.4 ×10
3
3 14.5 1/40.995 0.853 0.01
5 17.2 1/40.981 0.858 0.05
7 19.3 1/40.967 0.861 0.10
10 21.7 1/20.835 0.619 0.01
mass-weighted density eld 1.0 ×10
2
3 14.5 1/40.987 0.835 0.02
5 17.2 1/30.928 0.797 0.03
7 19.3 1/30.942 0.779 0.01
10 21.7 1/30.921 0.771 0.04
4.4 ×10
3
3 14.5 1/40.996 0.894 0.01
5 17.2 1/30.957 0.829 0.02
7 19.3 1/30.950 0.824 0.03
10 21.7 1/30.937 0.816 0.05
Note. The upper (lower)half of the table corresponds to the number- (mass-)weighted galaxy density elds. Listed in columns are the mean galaxy number density
n
gal
¯, minimum target mass scale of a CCV
M
min, effective radius of the CCV R
v
, smoothing scale R
s
, core density threshold δ
c
, upper density threshold δ
upper
when
core growing stops, and minimum core fraction f
c
.
Figure 4. Completeness and reliability of identifying CCVs from the mass-
weighted (dashed)and number-weighted (solid)density elds constructed
using the dense (red)and sparse (blue)galaxy sample. The shaded regions are
the detection rate uncertainties calculated as the standard deviations of the
completeness and reliability measured from bootstrap resampled galaxy density
elds. The detection rates and uncertainties are calculated when the optimal
parameter values in Table 1are adopted. Note that the completeness and
reliability are identical for the optimal parameter values.
5
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
scheme. On the other hand, the mean galaxy number density
signicantly affects the detection quality. This is just because
of the unavoidable nature of the cosmic voids whose
identication is sensitive to the tracer number density. For
the sparse sample, the detection rates increase from
0.60
== to 0.66 with the void mass scale before
MhM710
min 14 1
´-
, and then they drop to 0.62 at
MhM10
min 15 1
=-
. When using the dense sample, the
detection rates are enhanced approximately by 2%7%. This
is because the dense sample includes low-mass galaxies that are
more likely to be in voids than in other environments (Alonso
et al. 2015). Thus,increasing sample density decreases the
shot-noise effect, and hence the density-order shufing in
underdense regions. The detection rate enhancement is most
noticeable at the largest mass scale, MhM10
min 15 1
=-
.
Because the halo abundance in a large void is lower than that
in a small void (Patiri et al. 2006a), increasing the sample
density by adding lower-mass galaxies has more signicant
effect on decreasing the shot-noise for larger voids. Conse-
quently, the density-order shufing diminishes, leading to the
rapid increase in the detection rate at the largest mass scale. In
line with this interpretation, we nd that the uncertainty of the
detection rate, which increases with the void mass scale, is the
largest for the largest scale voids. We compute the uncertainty
as the standard deviation of the detection rate measured from
1000 realizations of bootstrap resampled galaxy density elds.
More importantly,increasing sample density reduces the
uncertainties of the detection rates approximately by 26% and
7% at the largest void mass scale, whereas they only decreased
by 7.6% and 3.7% on average on smaller mass scales for the
number- and mass-weighted cases, respectively. Thus, the
benet of increasing galaxy sample density is most signicant
in the largest CCV detection.
Finally, we show in Figure 5the spatial correspondence
between the recovered voids (black contours)using this method
and the CCVs (colored regions)dened under the voidcluster
Figure 5. Recovered voids (black contours)from the number-weighted galaxy density elds and CCVs (nonblack)of various mass scales in a 2 h
1
Mpc thick slice.
The optimal parameter values listed in Table 1are adopted. All panels show the identical eld. Recovered voids corresponding to galaxy cluster mass scale above
1×10
15
,7×10
14
,5×10
14
, and 3 ×10
14
h
1
M
e
are shown in clockwise direction from the upper left panel to bottom left.
6
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
correspondence theory. The voids are recovered from the
number-weighted galaxy density elds constructed using the
dense galaxy sample. It is shown that the recovered void
regions well overlap with the CCVs of the target mass scales. A
recovered void typically corresponds to a single CCV unless
the corresponding CCV has neighboring CCVs. This nearly
one-to-one correspondence between a recovered void and a
CCV becomes relatively weaker for a smaller minimum target
void mass. Thus, the trend of a recovered void encompassing
multiple CCVs is most noticeable for M
void
>3×10
14
h
1
M
e
.
However, in principle, by comparing the recovered void
regions for two different target mass scales, one can decompose
such void complexes into several smaller voids that respec-
tively correspond to single CCVs.
5. Summary and Discussion
We identify cosmic voids from galaxy density elds under
the voidcluster correspondence theory. The correspondence
between voids and clusters of the same mass scale is
established using paired cosmological simulations with sign-
opposite initial overdensity elds. CCVs in one simulation are
dened as the regions occupied by the same dark matter
particles that form cluster-scale halos in its inverted simulation.
Extending the CCV-identication method (Shim et al. 2021)
developed for matter density elds, we nd the optimal values
of the free parameters for the void identication in galaxy
density elds: Gaussian smoothing scale, density threshold,
and core volume fraction. The optimal length scale for
smoothing galaxy density elds is determined by the target
mass scale of voids, which is related to the target void radius as
R
s
=R
v
/3. The optimal parameter values for the number- and
mass-weighted density elds constructed using the sparse and
dense galaxy samples are listed in Table 1. Using the dense
galaxy sample, we can achieve about 64%69% of CCV
detection completeness and reliability from the number-
weighted galaxy density elds. The detection rates decrease
by 2%7% when using the sparse galaxy sample due to the
increasing effects of shot-noise on density.
Our result demonstrates that we can identify CCVs from
galaxy density elds in real space. In actual galaxy surveys, on
the other hand, one needs to consider the redshift-space
distortion (RSD)when constructing density elds from the
observed galaxy distribution in redshift space. Directly
identifying CCVs in redshift space without considering the
RSD effect will be less reliable than in real space. This is
because the RSD further introduces the density-rank order
randomization in addition to those induced by the shot-noise
and stochasticity of bias. Thus, it is better to adopt
reconstruction techniques and identify CCVs in real space
than in redshift space. The ngers of God (Jackson 1972)due
to the internal motion of galaxies within clusters can be
effectively removed by forcing the line-of-sight elongation of
clusters equal to their size perpendicular to the line of sight
(Tegmark et al. 2004; Park et al. 2012; Tully 2015; Hwang
et al. 2016). On larger scales, the Kaiser effect (Kaiser 1987)
due to coherent ow induced by large-scale overdensities and/
or underdensities can be corrected using Zeldovich approx-
imation (ZelDovich 1970)or second-order Lagrangian
perturbation theory (Scoccimarro 1998). Using these
approaches, one can place galaxies back to their real space
positions by subtracting the displacement due to their peculiar
motions (Wang et al. 2009; Kitaura et al. 2012; Bos et al.
2019).
Identifying CCVs from observational data could help us
perform a more precise cosmological and astrophysical analysis
using voids. Because CCVs have a universal mass density
distribution (Shim et al. 2021)and common central and average
densities (Pontzen et al. 2016; Stopyra et al. 2021; Desmond
et al. 2022), they represent a relatively more homogeneous void
population than that of arbitrary underdense regions extracted
from density elds. Consequently, when using CCVs, it is
possible to decrease the uncertainties arising due to the scatter
in void properties. This will be the subject of a future study.
One of the possible applications of CCVs in the cosmological
context is to measure the linear growth rate using CCVs. In
addition to the homogeneous properties of the CCVs, their
internal dynamics remain near linear (Stopyra et al. 2021), and
the linear bias relation holds at around CCVs (Pollina et al.
2017). Because the linearity in and around the CCV
environment allows a simple linear modeling for voidgalaxy
cross-correlation, an accurate measurement of the linear growth
rate using voids (Achitouv 2019; Hamaus et al. 2022;
Woodnden et al. 2022)can be made using CCVs. We leave
such cosmological analysis using CCVs from galaxy surveys to
future work.
Acknowledgments
We thank an anonymous referee for helpful comments that
helped improve the original manuscript. J.S. was supported by
Academia Sinica Institute of Astronomy and Astrophysics and
KIAS individual grant PG071202 at Korea Institute for
Advanced Study. C.B.P. was supported by KIAS individual
grant PG016903 at Korea Institute for Advanced Study. J.K.
was supported by a KIAS individual grant (KG039603)via the
Center for Advanced Computation at Korea Institute for
Advanced Study. S.E.H. was supported by the project
(Understanding
the Dark Universe Using Large Scale Structure of the
Universe), funded by the Ministry of Science. The computing
resources were kindly provided by the Center for Advanced
Computation at Korea Institute for Advanced Study.
ORCID iDs
Junsup Shim https://orcid.org/0000-0001-7352-6175
Changbom Park https://orcid.org/0000-0001-9521-6397
Juhan Kim https://orcid.org/0000-0002-4391-2275
Sungwook E. Hong ()https://orcid.org/0000-0003-
4923-8485
References
Achitouv, I. 2017, PhRvD,96, 083506
Achitouv, I. 2019, PhRvD,100, 123513
Achitouv, I., Neyrinck, M., & Paranjape, A. 2015, MNRAS,451, 3964
Alonso, D., Eardley, E., & Peacock, J. A. 2015, MNRAS,447, 2683
Bardeen, J. M., Bond, J. R., Kaiser, N., & Szalay, A. S. 1986, ApJ,304, 15
Bertschinger, E. 1985, ApJS,58, 1
Beygu, B., Kreckel, K., van de Weygaert, R., van der Hulst, J. M., &
van Gorkom, J. H. 2013, AJ,145, 120
Bos, E. G. P., Kitaura, F.-S., & van de Weygaert, R. 2019, MNRAS,488, 2573
Cai, Y.-C., Padilla, N., & Li, B. 2015, MNRAS,451, 1036
Ceccarelli, L., Duplancic, F., & Garcia Lambas, D. 2022, MNRAS,509, 1805
Ceccarelli, L., Padilla, N., & Lambas, D. G. 2008, MNRAS,390, L9
Chan, K. C., Hamaus, N., & Desjacques, V. 2014, PhRvD,90, 103521
Choi, Y.-Y., Park, C., Kim, J., et al. 2010, ApJS,190, 181
7
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
Choi, Y.-Y., Park, C., & Vogeley, M. S. 2007, ApJ,658, 884
Colberg, J. M., Sheth, R. K., Diaferio, A., Gao, L., & Yoshida, N. 2005,
MNRAS,360, 216
Dekel, A., & Lahav, O. 1999, ApJ,520, 24
de Lapparent, V., Geller, M. J., & Huchra, J. P. 1986, ApJL,302, L1
De Lucia, G., Kauffmann, G., & White, S. D. M. 2004, MNRAS,349, 1101
DESI Collaboration, Aghamousa, A., Aguilar, J., et al. 2016, arXiv:1611.
00036
Desjacques, V., Jeong, D., & Schmidt, F. 2018, PhR,733, 1
Desmond, H., Hutt, M. L., Devriendt, J., & Slyz, A. 2022, MNRAS,511, L45
Doré, O., Bock, J., Ashby, M., et al. 2014, arXiv:1412.4872
Dubinski, J., Kim, J., Park, C., & Humble, R. 2004, NewA,9, 111
Dunkley, J., Komatsu, E., Nolta, M. R., et al. 2009, ApJS,180, 306
Einasto, J., Joeveer, M., & Saar, E. 1980, MNRAS,193, 353
El-Ad, H., Piran, T., & da Costa, L. N. 1996, ApJL,462, L13
Faltenbacher, A., & Diemand, J. 2006, MNRAS,369, 1698
Fillmore, J. A., & Goldreich, P. 1984, ApJ,281, 9
Forero-Romero, J. E., Hoffman, Y., Gottlöber, S., Klypin, A., & Yepes, G.
2009, MNRAS,396, 1815
Gott, J., Richard, I., Melott, A. L., & Dickinson, M. 1986, ApJ,306, 341
Hahn, O., Porciani, C., Carollo, C. M., & Dekel, A. 2007, MNRAS,375, 489
Hamaus, N., Aubert, M., Pisani, A., et al. 2022, A&A,658, A20
Hong, S. E., Park, C., & Kim, J. 2016, ApJ,823, 103
Hwang, H. S., Geller, M. J., Park, C., et al. 2016, ApJ,818, 173
Jackson, J. C. 1972, MNRAS,156, 1P
Jee, I., Park, C., Kim, J., Choi, Y.-Y., & Kim, S. S. 2012, ApJ,753, 11
Jiang, C. Y., Jing, Y. P., Faltenbacher, A., Lin, W. P., & Li, C. 2008, ApJ,
675, 1095
Joeveer, M., & Einasto, J. 1978, in IAU Symp. 79, Large Scale Structures in
the Universe, ed. M. S. Longair & J. Einasto (Dordrecht: Reidel),241
Kaiser, N. 1984, ApJL,284, L9
Kaiser, N. 1987, MNRAS,227, 1
Kauffmann, G., & Fairall, A. P. 1991, MNRAS,248, 313
Kim, M., & Park, C. 1998, JKAS, 31, 109
Kitaura, F.-S., Angulo, R. E., Hoffman, Y., & Gottlöber, S. 2012, MNRAS,
425, 2422
Kreckel, K., Platen, E., Aragón-Calvo, M. A., et al. 2011, AJ,141, 4
Lavaux, G., & Wandelt, B. D. 2010, MNRAS,403, 1392
Lee, J., & Park, D. 2009, ApJL,696, L10
Li, B. 2011, MNRAS,411, 2615
Li, B., Zhao, G.-B., & Koyama, K. 2012, MNRAS,421, 3481
Matsubara, T. 1999, ApJ,525, 543
Nadathur, S., & Hotchkiss, S. 2015, MNRAS,454, 2228
Neyrinck, M. C. 2008, MNRAS,386, 2101
Nusser, A., Gubser, S. S., & Peebles, P. J. 2005, PhRvD,71, 083505
Park, C., Choi, Y.-Y., Kim, J., et al. 2012, ApJL,759, L7
Park, C., Choi, Y.-Y., Vogeley, M. S., et al. 2007, ApJ,658, 898
Park, D., & Lee, J. 2009, MNRAS,400, 1105
Park, H., Kim, J., & Park, C. 2010, ApJ,714, 207
Park, H., Park, C., Sabiu, C. G., et al. 2019, ApJ,881, 146
Patiri, S. G., Betancort-Rijo, J., & Prada, F. 2006a, MNRAS,368, 1132
Patiri, S. G., Betancort-Rijo, J. E., Prada, F., Klypin, A., & Gottlöber, S. 2006b,
MNRAS,369, 335
Pen, U.-L. 1998, ApJ,504, 601
Pisani, A., Sutter, P. M., Hamaus, N., et al. 2015, PhRvD,92, 083531
Platen, E., van de Weygaert, R., & Jones, B. J. T. 2007, MNRAS,380, 551
Pollina, G., Hamaus, N., Dolag, K., et al. 2017, MNRAS,469, 787
Pontzen, A., Slosar, A., Roth, N., & Peiris, H. V. 2016, PhRvD,93, 103519
Scoccimarro, R. 1998, MNRAS,299, 1097
Seljak, U., Hamaus, N., & Desjacques, V. 2009, PhRvL,103, 091303
Shim, J., Lee, J., & Hoyle, F. 2015, ApJ,815, 107
Shim, J., Park, C., Kim, J., & Hwang, H. S. 2021, ApJ,908, 211
Stopyra, S., Peiris, H. V., & Pontzen, A. 2021, MNRAS,500, 4173
Suto, Y., Sato, K., & Sato, H. 1984, PThPh,71, 938
Sutter, P. M., Lavaux, G., Hamaus, N., et al. 2015, A&C,9, 1
Tegmark, M., Blanton, M. R., Strauss, M. A., et al. 2004, ApJ,606, 702
Tonegawa, M., Park, C., Zheng, Y., et al. 2020, ApJ,897, 17
Tully, R. B. 2015, AJ,149, 171
Vogeley, M. S., Park, C., Geller, M. J., Huchra, J. P., & Gott, J. R. I. 1994,
ApJ,420, 525
Wang, H., Mo, H. J., Jing, Y. P., et al. 2009, MNRAS,394, 398
Woodnden, A., Nadathur, S., Percival, W. J., et al. 2022, MNRAS,516, 4307
Zehavi, I., Zheng, Z., Weinberg, D. H., et al. 2011, ApJ,736, 59
ZelDovich, Y. B. 1970, A&A, 5, 84
8
The Astrophysical Journal, 952:59 (8pp), 2023 July 20 Shim et al.
... As void regions are the direct opposite of nodes within the context of local density (Shim et al. 2023 ), we expect to see far less frequent MNRAS 531, 3858-3875 (2024) interaction between galaxies in this environment (Jian, Lin & Chiueh 2012 ). Due to fewer interactions with other (sub)structures in the environment, one would expect less SFR enhancement, leading to less chemically enriched gas released by SNe and stellar winds into the ISM, and hence lower average gas metallicity values within the galaxy compared to denser environments. ...
Article
Metallicity offers a unique window into the baryonic history of the cosmos, being instrumental in probing evolutionary processes in galaxies between different cosmic environments. We aim to quantify the contribution of these environments to the scatter in the mass-metallicity relation (MZR) of galaxies. By analysing the galaxy distribution within the cosmic skeleton of the Horizon Run 5 cosmological hydrodynamical simulation at redshift z = 0.625, computed using a careful calibration of the T-ReX filament finder, we identify galaxies within three main environments: nodes, filaments and voids. We also classify galaxies based on the dynamical state of the clusters and the length of the filaments in which they reside. We find that the cosmic environment significantly contributes to the scatter in the MZR; in particular, both the gas metallicity and its average relative standard deviation increase when considering denser large-scale environments. The difference in the average metallicity between galaxies within relaxed and unrelaxed clusters is ≈0.1dex, with both populations displaying positive residuals, δZg, from the averaged MZR. Moreover, the difference in metallicity between node and void galaxies accounts for ≈0.14 dex in the scatter of the MZR at stellar mass M⋆ ≈ 109.35 M⊙. Finally, both the average [O/Fe] in the gas and the galaxy gas fraction decrease when moving to higher large-scale densities in the simulation, suggesting that the cores of cosmic environments host – on average – older and more massive galaxies, whose enrichment is affected by a larger number of Type Ia Supernova events.
Preprint
Full-text available
Cosmic filaments are the main transport channels of matter in the Megaparsec universe, and represent the most prominent structural feature in the matter and galaxy distribution. Here we describe and define the physical and dynamical nature of cosmic filaments. It is based on the realization that the complex spatial pattern and connectivity of the cosmic web are already visible in the primordial random Gaussian density field, in the spatial pattern of the primordial tidal and deformation eigenvalue field. The filaments and other structural features in the cosmic web emerging from this are multistream features and structural singularities in phase-space. The caustic skeleton formalism allows a fully analytical classification, identification, and treatment of the nonlinear cosmic web. The caustic conditions yield the mathematical specification of weblike structures in terms of the primordial deformation tensor eigenvalue and eigenvector fields, in which filaments are identified -- in 2D -- with the so-called cusp caustics. These are centered around points that are maximally stretched as a result of the tidal force field. The resulting mathematical conditions represent a complete characterization of filaments in terms of their formation history, dynamics, and orientation. We illustrate the workings of the formalism on the basis of a set of constrained $N$-body simulations of protofilament realizations. These realizations are analyzed in terms of spatial structure, density profiles, and multistream structure and compared to simpler density or potential field saddle point specifications. The presented formalism, and its 3D generalization, will facilitate the mining of the rich cosmological information contained in the observed weblike galaxy distribution, and be of key significance for the analysis of cosmological surveys such as SDSS, DESI, and Euclid.
Article
Full-text available
The cosmological principle is fundamental to the standard cosmological model. It assumes that the Universe is homogeneous and isotropic on very large scales. As the basic assumption, it must stand the test of various observations. In this work, we investigated the properties of the Pantheon+ sample, including redshift distribution and position distribution, and we give its constraint on the flat Lambda CDM model: $ m $ = 0.36pm 0.02 and $H_ $ = 72.83pm 0.23 km s$^ $ Mpc$^ $. Then, using the region fitting (RF) method, we mapped the all-sky distribution of cosmological parameters ($ m $ and $H_ $) and find that the distribution significantly deviates from isotropy. A local matter underdensity region exists toward ($ circ circ $) as well as a preferred direction of the cosmic anisotropy ($ circ circ $) in galactic coordinates. Similar directions may imply that local matter density might be responsible for the anisotropy of the accelerated expansion of the Universe. Results of statistical isotropy analyses including Isotropy and Isotropy with real-data positions (RP) show high confidence levels. For the local matter underdensity, the statistical significances are 2.78sigma (isotropy) and 2.34sigma (isotropy RP). For the cosmic anisotropy, the statistical significances are 3.96sigma (isotropy) and 3.15sigma (isotropy RP). The comparison of these two kinds of statistical isotropy analyses suggests that inhomogeneous spatial distribution of real sample can increase the deviation from isotropy. The similar results and findings are also found from reanalyses of the low-redshift sample (lp+) and the lower screening angle ( = 60degr ), but with a slight decrease in statistical significance. Overall, our results provide clear indications for a possible cosmic anisotropy. This possibility must be taken seriously. Further testing is needed to better understand this signal.
Article
Full-text available
We present measurements of the anisotropic cross-correlation of galaxies and cosmic voids in data from the Sloan Digital Sky Survey Main Galaxy Sample (MGS), Baryon Oscillation Spectroscopic Survey (BOSS) and extended BOSS (eBOSS) luminous red galaxy catalogues from SDSS Data Releases 7, 12 and 16, covering the redshift range 0.07 < z < 1.0. As in our previous work analysing voids in subsets of these data, we use a reconstruction method applied to the galaxy data before void-finding in order to remove selection biases when constructing the void samples. We report results of a joint fit to the multipole moments of the measured cross-correlation for the growth rate of structure, fσ8(z), and the ratio DM(z)/DH(z) of the comoving angular diameter distance to the Hubble distance, in six redshift bins. For DM/DH, we are able to achieve a significantly higher precision than that obtained from analyses of the baryon acoustic oscillations (BAO) and galaxy clustering in the same datasets. Our growth rate measurements are of lower precision but still comparable with galaxy clustering results. For both quantities, the results agree well with the expectations for a ΛCDM model. Assuming a flat Universe, our results correspond to a measurement of the matter density parameter $\Omega _\mathrm{m}=0.337^{+0.026}_{-0.029}$. For more general models the degeneracy directions obtained are consistent with and complementary to those from other cosmological probes. These results consolidate void-galaxy cross-correlation measurements as a pillar of modern observational cosmology.
Article
Full-text available
The redshift-space distortion (RSD) in the observed distribution of galaxies is known as a powerful probe of cosmology. Observations of large-scale RSD, caused by the coherent gravitational infall of galaxies, have given tight constraints on the linear growth rate of the large-scale structures in the universe. On the other hand, the small-scale RSD, caused by galaxy–random motions inside clusters, has not been much used in cosmology, but it also has cosmological information because universes with different cosmological parameters have different halo mass functions and virialized velocities. We focus on the projected correlation function w ( r p ) and the multipole moments ξ l on small scales (1.4–30 h ⁻¹ Mpc). Using simulated galaxy samples generated from a physically motivated most bound particle (MBP)–galaxy correspondence scheme in the Multiverse Simulation, we examine the dependence of the small-scale RSD on the cosmological matter density parameter Ω m ; the satellite velocity bias with respect to MBPs, ; and the merger timescale parameter α . We find that α = 1.5 gives an excellent fit to the w ( r p ) and ξ l measured from the Sloan Digital Sky Survey–Korea Institute for Advanced Study value-added galaxy catalog. We also define the “strength” of the Fingers of God as the ratio of the parallel and perpendicular size of the contour in the two-point correlation function set by a specific threshold value and show that the strength parameter helps constrain by breaking the degeneracy among them. The resulting parameter values from all measurements are , indicating a slight reduction of satellite galaxy velocity relative to the MBP. However, considering that the average MBP speed inside halos is 0.94 times the dark matter velocity dispersion, the main drivers behind the galaxy velocity bias are gravitational interactions, rather than baryonic effects.
Article
Full-text available
We develop an improved Alcock–Paczynski (AP) test method that uses the redshift-space two-point correlation function (2pCF) of galaxies. Cosmological constraints can be obtained by examining the redshift dependence of the normalized 2pCF, which should not change apart from the expected small nonlinear evolution. An incorrect choice of cosmology used to convert redshift to comoving distance will manifest itself as redshift-dependent 2pCF. Our method decomposes the redshift difference of the two-dimensional correlation function into the Legendre polynomials whose amplitudes are modeled by radial fitting functions. Our likelihood analysis with this 2D fitting scheme tightens the constraints on Ω m and w by ∼40% compared to the method of Li et al. that uses one-dimensional angular dependence only. We also find that the correction for the nonlinear evolution in the 2pCF has a non-negligible cosmology dependence, which has been neglected in previous similar studies by Li et al. With an accurate accounting for the nonlinear systematics and use of full two-dimensional shape information of the 2pCF down to scales as small as 5 h ⁻¹ Mpc it is expected that the AP test with redshift-space galaxy clustering anisotropy can be a powerful method to constraining the expansion history of the universe.
Article
A recently-proposed algorithm identifies voids in simulations as the regions associated with halos when the initial overdensity field is negated. We apply this method to the real Universe by running a suite of constrained simulations of the 2M++ volume with initial conditions inferred by the BORG algorithm, along with the corresponding inverted set. Our 101 inverted and uninverted simulations, spanning the BORG posterior, each identify ∼150,000 “voids as antihalos” with mass exceeding 4.38 × 1011 M⊙ (100 particles) at z = 0 in a full-sky sphere of radius 155 Mpc/h around the Milky Way. We calculate the size function, volume filling fraction, ellipticity, central density, specific angular momentum, clustering and stacked density profile of the voids, and cross-correlate them with those produced by VIDE on the same simulations. We make our antihalo and VIDE catalogues publicly available.
Article
Euclid is poised to survey galaxies across a cosmological volume of unprecedented size, providing observations of more than a billion objects distributed over a third of the full sky. Approximately 20 million of these galaxies will have their spectroscopy available, allowing us to map the three-dimensional large-scale structure of the Universe in great detail. This paper investigates prospects for the detection of cosmic voids therein and the unique benefit they provide for cosmological studies. In particular, we study the imprints of dynamic (redshift-space) and geometric (Alcock–Paczynski) distortions of average void shapes and their constraining power on the growth of structure and cosmological distance ratios. To this end, we made use of the Flagship mock catalog, a state-of-the-art simulation of the data expected to be observed with Euclid . We arranged the data into four adjacent redshift bins, each of which contains about 11 000 voids and we estimated the stacked void-galaxy cross-correlation function in every bin. Fitting a linear-theory model to the data, we obtained constraints on f / b and D M H , where f is the linear growth rate of density fluctuations, b the galaxy bias, D M the comoving angular diameter distance, and H the Hubble rate. In addition, we marginalized over two nuisance parameters included in our model to account for unknown systematic effects in the analysis. With this approach, Euclid will be able to reach a relative precision of about 4% on measurements of f / b and 0.5% on D M H in each redshift bin. Better modeling or calibration of the nuisance parameters may further increase this precision to 1% and 0.4%, respectively. Our results show that the exploitation of cosmic voids in Euclid will provide competitive constraints on cosmology even as a stand-alone probe. For example, the equation-of-state parameter, w , for dark energy will be measured with a precision of about 10%, consistent with previous more approximate forecasts.
Article
We study the population of active galaxies in void environment in the SDSS. We use optical spectroscopic information to analyse characteristics of the emission lines of galaxies, accomplished by WHAN and BPT diagrams. Also, we study WISE mid-IR colours to assess AGN activity. We investigate these different AGN classification schemes, both optical and mid-IR, and their dependence on the spatial location with respect to the void centres. To this end, we define three regions: void, the spherical region defined by voidcentric distance relative to void radius (distance/rvoid) smaller than 0.8, comprising overdensities lesser than -0.9, an intermediate/transition shell region (namely void–wall) 0.8 < distance/rvoid < 1.2, and a region sufficiently distant from voids, the field: distance/rvoid > 2. We find statistical evidence for a larger fraction of AGN and star–forming galaxies in the void region, regardless of the classification scheme addressed (either BPT, WHAN or WISE). Moreover, we obtain a significantly stronger nuclear activity in voids compared to the field. We find an unusually large fraction of the most massive black holes undergoing strong accretion when their host galaxies reside in voids. Our results suggest a strong influence of the void environment on AGN mechanisms associated with galaxy evolution.
Article
We develop a method to identify cosmic voids from the matter density field by adopting a physically motivated concept that voids are the counterpart of massive clusters. To prove the concept we use a pair of ΛCDM simulations, a reference and its initial density-inverted mirror simulation, and study the relation between the effective size of voids and the mass of corresponding clusters. Galaxy cluster-scale dark matter halos are identified in the Mirror simulation at z = 0 by linking dark matter particles. The void corresponding to each cluster is defined in the Reference simulation as the region occupied by the member particles of the cluster. We study the voids corresponding to the halos more massive than 10 ¹³ h ⁻¹ M ⊙ . We find a power-law scaling relation between the void size and the corresponding cluster mass. Voids with a corresponding cluster mass above 10 ¹⁵ h ⁻¹ M ⊙ occupy ∼1% of the total simulated volume, whereas this fraction increases to ∼54% for voids with a corresponding cluster mass above 10 ¹³ h ⁻¹ M ⊙ . It is also found that the density profile of the identified voids follows a universal functional form. Based on these findings, we propose a method to identify cluster-counterpart voids directly from the matter density field without their mirror information by utilizing three parameters such as the smoothing scale, density threshold, and minimum core fraction. We recover voids corresponding to clusters more massive than 3 × 10 ¹⁴ h ⁻¹ M ⊙ at a 70%–74% level of completeness and reliability. Our results suggest that we are able to identify voids in a way to associate them with clusters of a particular mass scale.
Article
Cosmic voids provide a powerful probe of the origin and evolution of structures in the Universe because their dynamics can remain near-linear to the present day. As a result, they have the potential to connect large-scale structure at late times to early Universe physics. Existing ‘watershed’-based algorithms, however, define voids in terms of their morphological properties at low redshift. The degree to which the resulting regions exhibit linear dynamics is consequently uncertain, and there is no direct connection to their evolution from the initial density field. A recent void definition addresses these issues by considering ‘anti-haloes’. This approach consists of inverting the initial conditions of an N-body simulation to swap overdensities and underdensities. After evolving the pair of initial conditions, anti-haloes are defined by the particles within the inverted simulation that are inside haloes in the original (uninverted) simulation. In this work, we quantify the degree of non-linearity of both anti-haloes and watershed voids using the Zel’dovich approximation. We find that non-linearities are introduced by voids with radii less than $5\, \mathrm{Mpc}\, h^{-1}$, and that both anti-haloes and watershed voids can be made into highly linear sets by removing these voids.
Article
We present a new analysis of the inferred growth rate of cosmic structure measured around voids, using the LOWZ and the CMASS samples in the 12th data release (DR12) of SDSS. Using a simple multipole analysis we recover a value consistent with ΛCDM for the inferred linear growth rate normalized by the linear bias: the β parameter. We find β=0.33±0.11 for the LOWZ sample and β=0.36±0.05 for the CMASS sample. This work demonstrates that we can expect redshift-space distortions around voids to provide unbiased and accurate constraints on the growth rate, complementary to galaxy clustering, using simple linear modeling.
Article
We present a self-consistent Bayesian formalism to sample the primordial density fields compatible with a set of dark matter density tracers after a cosmic evolution observed in redshift space. Previous works on density reconstruction did not self-consistently consider redshift space distortions or included an additional iterative distortion correction step. We present here the analytic solution of coherent flows within a Hamiltonian Monte Carlo posterior sampling of the primordial density field. We test our method within the Zel'dovich approximation, presenting also an analytic solution including tidal fields and spherical collapse on small scales. Our resulting reconstructed fields are isotropic and their power spectra are unbiased compared to the true field defined by our mock observations. Novel algorithmic implementations are introduced regarding the mass assignment kernels when defining the dark matter density field and optimization of the time-step in the Hamiltonian equations of motions. Our algorithm, dubbed barcode, promises to be specially suited for analysis of the dark matter cosmic web down to scales of a few megaparsecs. This large-scale structure is implied by the observed spatial distribution of galaxy clusters - such as obtained from X-ray, Sunyaev-Zel'dovich, or weak lensing surveys - as well as that of the intergalactic medium sampled by the Ly α forest or perhaps even by deep hydrogen intensity mapping. In these cases, virialized motions are negligible, and the tracers cannot be modelled as point-like objects. It could be used in all of these contexts as a baryon acoustic oscillation reconstruction algorithm.