Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

MNRAS 000, 000–000 (0000) Preprint November 27, 2019 Compiled using MNRAS LATEX style file v3.

Fisher for complements: Extracting cosmology and neutrino mass


from the counts-in-cells PDF

Cora Uhlemann1,2 , Oliver Friedrich3,4 , Francisco Villaescusa-Navarro5,6 ,


Arka Banerjee7,8,9 , Sandrine Codis10
1 Centre for Theoretical Cosmology, DAMTP, University of Cambridge, CB3 0WA Cambridge, United Kingdom
2 Fitzwilliam College, University of Cambridge, CB3 0DG Cambridge, United Kingdom
arXiv:1911.11158v1 [astro-ph.CO] 25 Nov 2019

3 Kavli Institute for Cosmology, University of Cambridge, CB3 0HA Cambridge, United Kingdom
4 Churchill College, University of Cambridge, CB3 0DS Cambridge, United Kingdom
5 Department of Astrophysical Sciences, Princeton University, Peyton Hall, Princeton NJ 08544-0010, USA
6 Center for Computational Astrophysics, Flatiron Institute, 162 5th Avenue, 10010, New York, NY, USA
7 Kavli Institute for Particle Astrophysics and Cosmology, Stanford University, 452 Lomita Mall, Stanford, CA 94305, USA
8 Department of Physics, Stanford University, 382 Via Pueblo Mall, Stanford, CA 94305, USA
9 SLAC National Accelerator Laboratory, 2575 Sand Hill Road, Menlo Park, CA 94025, USA
10 CNRS & Sorbonne Université, UMR 7095, Institut d’Astrophysique de Paris, 75014, Paris, France

November 27, 2019

ABSTRACT
We comprehensively analyse the cosmology dependence of counts-in-cell statistics. We focus
on the shape of the one-point probability distribution function (PDF) of the matter density field
at mildly nonlinear scales. Based on large-deviation statistics, we parametrise the cosmology
dependence of the matter PDF in terms of the linear power spectrum, the growth factor, the
spherical collapse dynamics, and the nonlinear variance. We extend our formalism to include
massive neutrinos, finding that the total matter PDF is highly sensitive to the total neutrino
mass Mν and can disentangle it from the clustering amplitude σ8 .
Using more than a million PDFs extracted from the Quijote simulations, we determine the
response of the matter PDF to changing parameters in the νΛCDM model and successfully
cross-validate the theoretical model and the simulation measurements. We present the first
νΛCDM Fisher forecast for the matter PDF at multiple scales and redshifts, and its combina-
tion with the matter power spectrum. We establish that the matter PDF and the matter power
spectrum are highly complementary at mildly nonlinear scales. The matter PDF is particu-
larly powerful for constraining the matter density Ωm , clustering amplitude σ8 and the total
neutrino mass Mν . Adding the mildly nonlinear matter PDF to the mildly nonlinear matter
power spectrum improves constraints on Ωm by a factor of 5 and σ8 by a factor of 2 when con-
sidering the three lowest redshifts. In our joint analysis of the matter PDF and matter power
spectrum at three redshifts, the total neutrino mass is constrained to better than 0.01 eV with
a total volume of 6 (Gpc/h)3 . We discuss how density-split statistics can be used to translate
those encouraging results for the matter PDF into realistic observables in galaxy surveys.
Key words: cosmology: theory — large-scale structure of Universe — methods: analytical,
numerical

1 INTRODUCTION parameters representing fundamental physics quantities, e.g. the to-


tal neutrino mass, the equation-of-state of dark energy, as well as
The ΛCDM model of a universe filled with a cosmological constant the amplitudes and shapes of primordial non-Gaussianity. Within
and cold dark matter has proved to be an extraordinarily successful current observational limits, the latter two extensions are consis-
paradigm. This concordance model is capable of explaining a large tent with a cosmological constant and Gaussian initial conditions,
variety of cosmological observations, from the anisotropies of the respectively. By contrast, we already know that at least two neu-
cosmic microwave background (Planck Collaboration et al. 2018) trino families must have a nonzero mass in order to explain the
to the spatial distribution of galaxies at low redshift. This “Stan- observations of neutrino oscillations (Becker-Szendy et al. 1992;
dard Model” of cosmology can be extended by a few additional Fukuda et al. 1998; Ahmed et al. 2004). However, both the total

© 0000 The Authors


2 C. Uhlemann et al.
mass of the three neutrino mass eigenstates, as well as the hierar- σ8 Ωm Ωb ns h Mν [eV ]
chy of these mass states are still unknown, and can provide crucial fid 0.834 0.3175 0.049 0.9624 0.6811 0
hints to physics beyond the Standard Model of particle physics. The ∆ 0.015 0.01 0.002 0.02 0.02 0.1, 0.2, 0.4
minimal total neutrino
P mass in the presence of known mass split- Table 1. Cosmological model parameters for the set of νΛCDM Quijote
tings is Mν = mν > 0.056 eV for a normal hierarchy and simulations under consideration in this paper.
Mν > 0.095 eV for an inverted hierarchy (Lesgourgues & Pas-
tor 2006). Cosmological observations from the cosmic microwave
background (CMB) already provide upper bounds on the sum of cosmology dependence can be accurately predicted. Recently, this
neutrino masses. Upcoming Stage-IV CMB polarization experi- idea has been applied to surveys like DES (The Dark Energy Survey
ments and large-scale structure surveys like Euclid (Laureijs et al. Collaboration 2005) and KiDS (de Jong et al. 2013) using galaxy
2011), LSST (Ivezić et al. 2019), DESI (Levi et al. 2013), PFS troughs and ridges (Gruen et al. 2016; Brouwer et al. 2018), mo-
(Takada et al. 2014) and WFIRST (Spergel et al. 2015) will seek ments of galaxy density and lensing convergence (Bel & Marinoni
to detect the signature of total neutrino mass in galaxy clustering 2014; Petri et al. 2015; Clerkin et al. 2016; Salvador et al. 2018;
and weak lensing statistics conclusively (Abazajian et al. 2015). Gatti et al. 2019) and density-split statistics (Friedrich et al. 2018;
Constraining the value of all the fundamental physical param- Gruen et al. 2018). In particular, Gruen et al. (2018) have shown
eters mentioned above is one of the most important goals of mod- that density-split statistics from joint counts- and lensing-in-cells
ern precision cosmology. For that reason, many different cosmo- yields cosmological constraints competitive with two-point func-
logical missions are going to survey the sky to collect data that tion measurements. At the same time, density-split statistics re-
allow to constrain the value of the cosmological parameters as ac- cover additional information about higher-order moments of the
curately and precisely as possible. Unfortunately, a large fraction of density field and the relation between galaxy and matter density.
the raw data available from these surveys cannot be converted into Their combined use of galaxy counts and lensing allowed them to
tighter constraints on cosmological parameters, because the infor- connect models of the matter density PDF to photometric data of
mation is embedded on small scales and in non-Gaussian observ- the galaxy density field, demonstrating that the methodology pre-
ables for which accurate theory predictions are challenging (Scoc- sented here can be carried over to real data analyses.
cimarro et al. 1999; Cooray & Sheth 2002; Rimes & Hamilton In this work, we combine insights from an analytical model for
2005; Neyrinck et al. 2006; Nishimichi et al. 2016). These scales the matter PDF based on large-deviation statistics and spherical col-
are typically in the mildly to fully nonlinear regime due to grav- lapse (Bernardeau et al. 2014; Uhlemann et al. 2016) with measure-
itational collapse, and for this reason, analytical predictions based ments from the large suite of the Quijote simulations (Villaescusa-
on perturbation theory are invalid. At high redshifts, the matter den- Navarro et al. 2019). We quantify, for the first time, the amount
sity fluctuations are close to a Gaussian random field, which is fully of cosmological information encoded in the matter density PDF at
characterised by its power spectrum or two-point correlation func- multiple scales and redshifts, and compare it to the one from the
tion. However, as non-linear gravitational clustering proceeds, the matter power spectrum. In our analysis, we take into account the
density field becomes non-Gaussian. The information that initially full covariance between density PDFs measured at different scales
was contained in the power spectrum, leaks into higher-order mo- and its cross-covariance with the power spectrum. To perform the
ments of the density field (Peebles 1980; Bernardeau et al. 2002). full analysis we extracted more than a million matter density PDFs
Thus, if the analysis of large-scale structure survey data is limited to from the Quijote suite for different cosmologies and made them
the power spectrum, a significant amount of information is unused. publicly available.1 We consider the νΛCDM model, which ex-
In the non-linear regime, it is unknown what fraction of the infor- tends ΛCDM by including the sum of neutrino masses, Mν , as pa-
mation is contained in each statistic. In this paper, we focus our rameter. In particular, we use what we call derivative simulations,
attention on one of the simplest statistics of a three-dimensional which vary exactly one parameter in the νΛCDM model compared
field: the probability distribution function (hereafter PDF) of the to the fiducial model with magnitudes given in Table 1.
matter density field smoothed on a given scale. Our paper is structured as follows: in Section 2, we describe
a theoretical model for the PDF of matter densities in spheres and
Empirically, it has been found that one-point matter density
discuss the physical ingredients that determine the resulting shape
PDFs are close to lognormal (Coles & Jones 1991; Kayo et al.
of the PDF. In Section 3, we generalise the PDF model to account
2001), with further improvements by skewed lognormal models
for the presence of massive neutrinos. We then cross-validate our
(Colombi 1994; Repp & Szapudi 2018) or generalised normal dis-
theoretical predictions with measurements from the Quijote simu-
tributions (Shin et al. 2017). While the lognormal model is only a
lation suite in Section 4. In Section 5 we present a Fisher analy-
crude approximation, it highlights that by limiting the cosmolog-
sis that demonstrates the constraining power of the matter PDF for
ical analysis to two-point statistics one inevitably misses a large
ΛCDM parameters and total neutrino mass. Section 6 provides a
amount of information encoded in a non-Gaussian field. This idea
conclusion and an outlook to further work.
has been formalised by considering the power spectrum of log-
densities (Neyrinck et al. 2009; Seo et al. 2011; Wolk et al. 2015)
and more generally sufficient statistics (Carron & Szapudi 2014).
In order to unlock additional information in upcoming large- 2 PHYSICAL INGREDIENTS FOR THE PDF OF
scale surveys like Euclid, we need accurate predictions for non- MATTER DENSITIES IN SPHERES
Gaussian statistics and their dependence on cosmology. Having Large deviation statistics provides a means to compute the proba-
multiple complementary large-scale structure probes is particularly bility distribution function (PDF) of nonlinear matter densities in
important for breaking degeneracies when jointly constraining fun- spheres (that is to say density smoothed with a top-hat kernel). In
damental physics such as neutrino masses, modified gravity and
dynamical dark energy (Font-Ribera et al. 2014; Sahlén 2019).
Counts-in-cells statistics like density PDFs are ideal candidates for 1 The PDFs can be accessed as part of the public data release of the Quijote
this purpose, as they can be easily measured in surveys and their simulations, see github.com/ franciscovillaescusa/Quijote-simulations.

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 3
the present paper, we limit ourselves to Gaussian initial conditions, (i) the scale-dependent linear variance and the linear growth,
but primordial non-Gaussianities can also be implemented in the (ii) the mapping between initial and final densities in spheres,
formalism (Uhlemann et al. 2018b, Friedrich et al. 2019). For Gaus- (iii) and the nonlinear variance .
sian initial conditions, the PDF P(δL ) of the linear matter density
contrast δL in a sphere of radius r is fully specified by the linear We will discuss each of these ingredients in the following three sub-
variance at that scale sections. Then we generalise the formalism to include the impact
s of massive neutrinos in Section 3.
2
 
1 δL While in general, the transformation from the decay-rate of
Prini (δL ) = 2
exp − . (1)
2πσL (r) 2σL2 (r) the PDF (5) to the PDF itself has to be evaluated numerically, one
can find an excellent analytical approximation using a saddle-point
The linear variance at scale r is obtained from an integral over the technique. As shown in Uhlemann et al. (2016), the inverse Laplace
linear power spectrum with spherical top-hat filter in coordinate transform can be evaluated with a saddle-point approximation for
space the log-density µ = ln ρ, the expression reads
Z
dk
σL2 (r) = PL (k)k2 W3D
2 r
(kr) , (2) Ψ00R (ρ) + Ψ0R (ρ)/ρ
2π 2 PR (ρ) = exp (−ΨR (ρ)) . (6a)

where W3D (k) is the the Fourier transform of the 3D spherical top-
hat kernel Because of the use of the logarithmic variable, one has to ensure
r the correct mean density hρi = 1 by specifying the mean of the
π J3/2 (k) log-density hln ρi. This can be implemented by properly rescaling
W3D (k) = 3 , (3)
2 k3/2 the ‘raw’ PDF (6a)
and J3/2 (k) is the Bessel function of the first kind of order 3/2. 
hρ̃i

hρ̃i
To describe the impact of nonlinear gravitational dynamics on P̂R (ρ) = PR ρ · · , (6b)
h1i h1i2
the shape of the initially Gaussian matter PDF, it is informative to R
look at the exponential decay of the PDF with increasing density where hf (ρ̃)i = dρ̃ , f (ρ̃)P(ρ̃). Note that the normalization is
contrast. The decay-rate function is the negative argument of the only necessary because the inverse Laplace transform is not com-
exponent in equation (1) and reads puted explicitly, which would automatically preserve the normali-
2 sation and ensure a correct mean. Since the saddle-point approxi-
δL
Ψini
r (δL ) = 2
. (4) mation makes use of the log-density, the nonlinear variance that en-
2σL (r)
ters the decay-rate function (5) is the one of the logarithmic density
According to the contraction principle of large deviation µ = ln ρ.2 If one computes the PDF numerically using an inverse
statistics (Bernardeau & Reimberg 2016), the exponential decay of Laplace transform (as was for instance performed in Bernardeau
the PDF of final densities (at scale R and redshift z) can be obtained et al. (2015), the nonlinear variance of the density enters directly.
from the initial one by inserting the most likely mapping between Since the final matter density PDFs obtained from both approaches
linear and nonlinear densities in spheres and their radii into agree very well (Uhlemann et al. 2016), the shape of the density
PDF allows to translate the variances to each other.
σL2 (R) δL (ρ)2
ΨR (ρ) = 2
, (5)
σNL (z, R) 2σL2 (Rρ1/3 )
where σNL is the nonlinear variance of the density at scale R and 2.1 The scale-dependence of the linear variance
redshift z. Thanks to the symmetry of the statistics and statistical
isotropy, the most probable evolution of densities in spheres δL (ρ) The amplitude of the linear density fluctuation at different scales
can be accurately approximated by spherical collapse and the initial are calculated from the CAMB (Lewis et al. 2000) linear power
and final radii are related by mass conservation r = Rρ1/3 . This spectrum at z = 0. To obtain good agreement with the finite reso-
argument can be made more precise by writing the final PDF as lution simulations, the integral for the linear variance should be cut
a path integral over all possible histories relating final densities to at the Nyquist frequency kNy = π · Nmesh /Lbox , which is around
linear densities. As shown in Bernardeau (1994a); Valageas (2002), 1.6h/Mpc for the box size L = 1000 Mpc/h and Nmesh = 512.
the dominant contribution to this integral comes from spherical col- In Figure 1 we show how different cosmologies with fixed σ8 im-
lapse, which is a saddle point of the corresponding functional inte- pact the scale-dependence of the linear variance. While changes
grals. Recently, non-perturbative effects not captured by the saddle in σ8 simply modify the overall amplitude, changes in Ωm , Ωb ,
point have been analysed analytically in one dimension (Pajer & ns and h modify the variance in a scale-dependent way. As ex-
van der Woude 2018) and estimated using a path-integral approach pected, a change in the primordial spectral index ns results in a
2
based on perturbation theory with a renormalisation of small-scale constant shift of the logarithmic derivative d log σL (R)/d log R.
physics (Ivanov et al. 2019). In practice, non-perturbative effects In contrast, changes in the matter and baryon densities, Ωm and
mostly renormalise the nonlinear variance entering equation (5) and Ωb , as well as the Hubble parameter h induce an additional scale-
higher order reduced cumulants are still reliably predicted by spher- dependent running of the spectral index.
ical collapse. While it is possible to numerically determine the precise
From the decay-rate function in equation (5) one can compute ΛCDM parameter dependence of the linear power spectrum from
the cumulant generating function via a Legendre transform. Then, CAMB (Lewis et al. 2000) or CLASS (Blas et al. 2011), a closed-
one obtains the final PDF from the cumulant generating function form expression for forecasts and data analysis is desirable. In
via an inverse Laplace transform that can be computed numerically
(Bernardeau et al. 2014, 2015; Friedrich et al. 2018). For a stan-
dard ΛCDM universe, there are only three ingredients that enter 2 Note that the relevant variable is the logarithm of the smoothed density ρ
this theoretical model for the matter PDF, in a sphere of radius R, not the smoothed logarithm of the density.

MNRAS 000, 000–000 (0000)


4 C. Uhlemann et al.
Figure A1 in Appendix A we show that the linear variance com-
0.02 ΔΩm Δns
puted from the Eisenstein-Hu formula for the linear power spec-

σ2L,+/- (R)/σ2L, d (R)-1


trum (Eisenstein & Hu 1998) is accurate at about 0.5% for the fidu- 0.01 ΔΩb Δh
cial model and the derivative simulations.
0.00

Impact on the matter PDF and its cumulants - 0.01


The scale-dependence of the linear variance determines the expo-
- 0.02
nential decay of the PDF according to equation (5). With different
densities ρ, one scans the linear variance at scales Rρ1/3 in a range 5 10 20
of values around the radius R. One can also understand this be- 0.02
R [Mpc/h]

S3,+/- (R)-S3,fid (R)


haviour from the tree order perturbation theory prediction for the
0.01
reduced skewness, S3 , of the density at scale R. In an EdS uni-
verse, this quantity is determined by the first logarithmic derivative
0.00
of the linear variance (Bernardeau 1994b)
hδ 3 (R)i 34 d log σL2
(R) - 0.01
S3 (R) = = + . (7)
hδ 2 (R)i2 7 d log R
- 0.02
Note that differences between the reduced skewness amplitude be-
5 10 20
fore smoothing in EdS, which predicts 34/7 and the ΛCDM spher-
ical collapse prediction are below 0.0017 for the changes in Ωm R [Mpc/h]
considered here. Hence, the main change in the reduced skewness 2 (R) as computed
is indeed caused by a difference in the scale-dependent variance. Figure 1. Comparison between the linear variance σL
from CAMB at z = 0 for ΛCDM cosmologies with varying Ωm (red), Ωb
Note that the clustering amplitude σ8 cancels in the logarithmic
(blue), ns (yellow) and h (green) with positive sign (solid lines) and neg-
derivative and does not change the reduced skewness. We show how ative sign (dashed lines) as indicated in Table 1 with fixed σ8 . The lower
the other ΛCDM parameters impact the ratio of linear variances panel shows the differences in the predicted reduced skewness S3 for the
and hence the differences in the reduced skewness in Figure 1. The different cosmologies according to equation (7).
lower panel demonstrates how changing cosmological parameters
offsets the reduced skewness. When focusing on a single radius
R, one can only detect the overall offset, but not distinguish be- that for the matter PDF at a single nonzero redshift z > 0, the
tween the cosmological parameters. In particular, the skewness at linear growth function controlled by Ωm becomes degenerate with
one scale cannot distinguish a constant tilt caused by ns from a run- the overall amplitude σ8 , which is not the case at z = 0, because
ning of the tilt induced by Ωm , Ωb and h. This degeneracy is par- D(z = 0) = 1. This degeneracy can be broken by performing an
tially broken by considering the full PDF, whose shape is also sen- analysis in multiple redshift slices, which also helps to disentangle
sitive to a combination of the reduced kurtosis S4 and higher order the matter density Ωm from the spectral index ns , as we demon-
cumulants, which depend on higher order logarithmic derivatives. strate in a Fisher forecast shown in Figure 15.
In the presence of irreducible noise such as cosmic variance, the In Appendix A1, we discuss the impact of a dark energy equa-
amount of this additional information is however limited. This is tion of state beyond a cosmological constant. In Figure 19 in Sec-
why it is important to jointly consider matter density PDFs at mul- tion 6, we compare the dependence of the growth-rate on the matter
tiple radii. In this study, we limit ourselves to modelling the PDFs density Ωm for a flat ΛCDM universe to a change in the dark energy
at different radii individually, but include their cross-covariance, equation of state parameter w0 and to the modifications induced by
which captures some part of the joint one-point PDF. the presence of massive neutrinos.

Growth of density fluctuations 2.2 Spherical collapse dynamics

In the linear regime and for a ΛCDM model, the amplitude of den- In an Einstein-de Sitter (EdS) universe, there is a parametric solu-
sity fluctuations grows with the growth rate such that tion for spherical collapse dynamics (Peebles 1980), which relates
the linear density contrast δL to the nonlinear density ρNL
σ 2 (z, R) = D2 (z)σL
2
(R) . (8)  2
 ρNL = 9 (θ − sin θ)

For a flat ΛCDM universe, the linear growth of structure depends 2 (1 − cos θ)3


only on the matter density Ωm and a closed form is known (Mat- δL > 0 : (10a)
 δL = 3 [6(θ − sin θ)]2/3

subara 1995)


h i 20
−1+Ωm
p
1 − Ωm + Ωm (1 + z)3 2 F1 65 , 23 , 11 ,
6 Ωm (1+z) 3 ) and
D(z) = ,
9 (sinh η − η)2
h i 
(1 + z)5/2 2 F1 56 , 32 , 11 , −1+Ω m
 ρNL = 2 (cosh η − 1)3
6 Ωm



(9) δL < 0 : , (10b)
3

where 2 F1 is the hypergeometric function. At redshift z = 1, a
 2/3
δL = − [6(sinh η − η)]


change in the matter density Ωm of ± 3% around the fiducial value 20
leads to ∓ 1% difference in the square of the growth function. From where θ ∈ [0, 2π] is the development angle and η its counter-
the overall amplitude of density fluctuations in equation (8), we see part for an open universe and both parameters can be eliminated

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 5

δL,app (ρNL )/δL,LCDM (ρNL )-1 2.3 The nonlinear variance


In full analogy to its linear counterpart, the nonlinear variance is
0.04 EdS ν=1.5 defined in terms of the nonlinear power spectrum
0.02 ν=21/13 ν=1.686
d3 k
Z
2
σNL (z, R) = PNL (z, k)W3D (kR)2 . (12)
0.00 (2π)3

- 0.02 While the reduced cumulants are well predicted by spherical col-
lapse dynamics, the nonlinear variance cannot be inferred accu-
- 0.04 rately enough. In the following we discuss how one can efficiently
parameterise the cosmology dependence of this quantity, and how
0.1 0.5 1 5 10 accurately this quantity can be predicted using the nonlinear mat-
ρNL ter power spectrum from fitting functions (HALOFIT from Peacock
& Smith (2014) and RESPRESSO from Nishimichi et al. (2017)) or
Figure 2. Comparison between the parametric spherical collapse dynamics perturbative techniques (2-loop SPT and RegPT from Taruya et al.
for EdS given by equation (10) (black) and the ν-parametrization given (2012); Osato et al. (2019)).
by equation (11) for the parameters ν = 1.5 (red), ν = 21/13 (blue)
and ν = 1.686 (green) compared to the numerical solution for spherical
collapse in a ΛCDM universe. 2.3.1 Cosmology dependence
For lognormal fields with unit mean, the variance of the logarithm
of the smoothed density is related to the density variance as (Coles
from the relation. Note that these formulae can be extended to any & Jones 1991)
background with zero cosmological constant (see Appendix A of 2 2
 
σln ρ (z, R) = ln 1 + σρ (R, z) . (13)
Bernardeau et al. 2002). Let us note that the initial and final radii
of the sphere are related by mass conservation r3 = R3 ρ. To estimate the impact of changing cosmology, we will use the lin-
For simplicity, one can rely on an approximate explicit param- ear density variance in this relation.3 In the absence of massive neu-
eterisation for spherical collapse in an EdS universe, given by trinos, we use a factorisation of the linear variance into the growth
function and the scale-dependent linear variance extrapolated to to-
 −ν day, σL (R, z) = D(z)σL (R).
δL
ρNL,ν (δL ) = 1− , (11) This simple relation proves useful for parameterising the cos-
ν
mology dependence of the nonlinear log-variance, which is induced
where δL is the linear density at redshift zero. The parameter ν by changes in the linear variance (computed from the linear power
controls the amplitude of the skewness before smoothing 3(1 + spectrum) and the growth rate D(z). We use this relation to predict
1/ν) and can be matched to the prediction in equation (7), yielding the scaling of the nonlinear variance for cosmologies with changed
ν = 21/13. Originally, this parametric form has been suggested in cosmological parameters from the measured variance at the fiducial
Bernardeau (1994a), with approximately ν = 1.5, which becomes cosmology
exact for Λ=0 in the limit of Ωm → 0 and drives the shape of  2
ln 1 + σL,cos

(R, z) 2
2
the PDF in low density regions. In excursion-set inspired models, σln ρ,cos (z, R) = h i σln ρ,fid (z, R) . (14)
2
usually the critical linear density for collapse is used, setting ν = ln 1 + σL,fid (R, z)
δc = 1.686 (Lam & Sheth 2008).
We have checked that this yields residuals smaller than 0.1% for
In Figure 2, we compare the parametric EdS form (10) (black
radii R = 10, 15, 20 Mpc/h at all redshifts and will use the approx-
line) and the ν-parameterisation (11) with different parameters
imation (14) for predicting the change of nonlinear log-variances
(blue, green and red lines) to the numerical solution of the ΛCDM
with cosmology.
spherical collapse dynamics, described in Friedrich et al. (2018, see
Appendix A). We find that the parametric solution for EdS approx-
imates the numerical ΛCDM solution extremely well, with sub- 2.3.2 Calibration for fiducial cosmology
percent residuals in the range of relevant densities ρ ∈ [0.1, 10].
For the fiducial cosmology (Ωm = 0.3175), the deviations are less The nonlinear variance can be predicted from a given nonlinear
than 0.2%. As expected, the agreement improves with increasing power spectrum according to equation (12). We compare the re-
Ωm . But even for a matter density of Ωm ' 0.21, differences stay sult from fitting functions (halofit and respresso) and perturbative
below 0.3%. Therefore, for the purpose of constraining cosmol- techniques (SPT and RegPT) with the variance obtained from the
ogy with the bulk of the matter PDF, the cosmology dependence of measured nonlinear power spectrum. With this comparison, we cir-
spherical collapse can be neglected. In what follows, we rely on the cumvent potential convergence issues that might affect the nonlin-
parametric EdS spherical collapse solution (10) for the theoretically ear variance measured from the PDF at small scales, as discussed
predicted PDF. in Section 4.2. We compare the nonlinear power spectra using the
Note that spherical collapse is potentially affected more se-
riously by dynamical dark energy (Mota & van de Bruck 2004; 3 As discussed in Repp & Szapudi (2017), one could introduce a free pa-
Abramo et al. 2007; Pace et al. 2010; Mead 2017) or modified grav- rameter in this simple approximation in order to improve the matching on
ity (Schäfer & Koyama 2008; Barreira et al. 2013; Kopp et al. 2013; small scales. Note that Repp & Szapudi (2017) use this approach for obtain-
Cataneo et al. 2016), which are beyond the scope of this work. The ing expression the power spectrum of the log-density. Consequently, their
impact of massive neutrinos on spherical collapse is discussed in relationships are formulated for the smoothed log-density rather than the
Section 3. logarithm of the smoothed density that we consider here.

MNRAS 000, 000–000 (0000)


6 C. Uhlemann et al.
fid, HR R [Mpc/h] z=0 z=0.5 z=1 z=2 z=3 matter density and can be approximated as
10 0.619 0.498 0.404 0.285 0.218
σµ 15 0.486 0.384 0.308 0.215 0.164 Mν Ων
Ων = , fν = . (15)
20 0.388 0.304 0.242 0.169 0.129 93.14h2 eV Ωm
10 0.743 0.560 0.436 0.296 0.223 For the simulations considered here with Mν = 0.1, 0.2, 0.4 eV,
σρ 15 0.532 0.406 0.319 0.219 0.166 we have fν = 0.0075, 0.015, 0.030. Due to their large thermal
20 0.408 0.313 0.247 0.171 0.13 velocities, neutrinos do not cluster on scales below their (physi-
Table 2. Variances of the density ρ and the log-density µ = ln ρ for dif- cal) free-streaming length, which is defined in analogy to the Jeans
ferent radii R [Mpc/h], redshifts z and cosmological models (see Table 1) length and depends on their mass according to
as measured from the mean of the 100 realisations of the high resolution
7.7(1 + z) 1eV
fiducial simulation. λfs (z) = p Mpc/h , (16)
ΩΛ + Ωm (1 + z)3 mν
where mν is the mass of the considered neutrino species. The cor-
different approaches in Figure A2 in Appendix A1. At low redshifts responding free-streaming wavenumber is kfs = 2πa/λfs . In this
z = 0, 0.5, 1, the halofit power spectrum is accurate at a few per- study, we consider degenerate neutrino masses, and therefore, we
cent level in the whole range up to k ' 0.4, which is relevant for have mν = Mν /3. For total neutrino masses of Mν = 0.1, 0.2, 0.4
obtaining the variance down to R = 10 Mpc/h. When using the eV, the free streaming lengths are λfs (z = 0) = 231, 115, 57
halofit fitting function for the nonlinear power spectrum to predict Mpc/h. While the free streaming scale sets the scale below which
the nonlinear variance at R = 15 Mpc/h, we find about 1% and neutrinos are not clumping under the influence of gravity at some
2% disagreement at redshifts z = 0.5 and z = 0. At the smaller redshift z, there is an additional scale of interest in the problem.
scale R = 10 Mpc/h, we obtain residuals of 1.5% and 2.5% for The maximum free streaming scale in comoving units achieved at
z = 0.5 and z = 0, respectively. The cosmology dependence of any redshift is much larger, as it is related to the time when neutri-
the nonlinear variance, measured by the residuals of the ratio of nos become non-relativistic
variances in different cosmologies, is predicted at sub-percent level  m 1/2
ν
for all derivative simulations with changed ΛCDM parameters ac- knr ' 0.018 Ω1/2
m h/Mpc . (17)
cording to Table 1, where the largest deviations of are found for 1eV
Ωm and ns . The associated comoving length scale λnr = 2π/knr determines
The nonlinear power spectrum generated from the response above which scales massive neutrinos behave like an ordinary clus-
function approach (Nishimichi et al. 2016, 2017)4 is extremely tering dark matter component. Below this scale, the neutrino power
close to the measured power spectrum, having sub-percent resid- spectrum is damped compared to the cold dark matter one. For
uals throughout. The response function approach is aided by a few the total neutrino masses considered here, those comoving length
sets of simulations, one of which is for a Planck 2015 cosmology scales are beyond the size of the simulation box.
with a very similar set of parameters as chosen in Quijote. While In the presence of massive neutrinos, we need to extend the
this could explain the spectacular agreement for fiducial cosmol- list of physical ingredients for the matter PDF from Section 2 by
ogy, also the predictions for the dependence on changed ΛCDM an additional element. In order to predict the total matter PDF, we
parameter according to Table 1 are sub-percent throughout. need to specify the nonlinear matter density of the clustering com-
Obtaining accurate nonlinear variances from perturbative re- ponent. After discussing this key change below, we will describe
sults for the nonlinear power spectrum is difficult, since a broad the imprint of massive neutrinos on the other standard ingredients,
range of scales up to k ' 0.4h/Mpc enters the integration from namely spherical collapse and the variances.
equation (12) when decreasing radius towards R = 10 Mpc/h,
as we demonstrate in Figure A3 in Appendix A1. RegPT and SPT
3.2 Matter density of clustering component
at 2-loop order 5 have residuals typically larger than 4% at red-
shift z = 0, with improvements at higher redshifts and larger Let us denote normalised densities by ρ = 1 + δ and physical
radii. While RegPT gives a more accurate result than SPT out densities by ρ̃, such that ρ = ρ̃/ρ̄. The normalized total density ρm
to wavenumbers of about k ' 0.2h/Mpc, its exponential cutoff can be expressed in terms of the normalised matter density in cold
causes predictions for the nonlinear variance to seem worse than dark matter and baryons, ρcb , and the neutrino density ρν , and the
SPT. Note that within the framework of the effective field theory of relative abundances of the two species
large-scale structure (Baumann et al. 2012; Carrasco et al. 2012), ρ̃cb+ν ρ̃cb + ρ̃ν Ωcb Ων
the nonlinear variance gets renormalised in order to account for the ρm = ' = ρcb + ρν (18a)
ρ̄cb+ν ρ̄cb + ρ̄ν Ωm Ωm
short-scale effects (Ivanov et al. 2019).
Ωcb Ων
δm = δcb + δν . (18b)
Ωm Ωm
A first estimate for the effect of neutrinos masses can be ob-
3 THE MATTER PDF WITH MASSIVE NEUTRINOS tained by considering a uniform background density constituted by
the massive neutrinos, thus setting ρν = 1. We illustrate this most
3.1 Basic effects of massive neutrinos
simplistic relationship between the total normalized matter density
Before we start to discuss the effect of massive neutrinos on the and the normalized matter density in CDM and baryons as dotted
matter PDF, let us review the relevant basics following Lesgour- line in Figure 3. We see that the effect is biggest for underdensi-
gues & Pastor (2006). The neutrino abundance is related to the total ties, where the presence of massive neutrinos causes one to probe
CDM+baryon densities that are effectively even lower than the total
matter density, hence rarer and less probable. To take into account
4 computed using the publicly available Respresso Python package that neutrinos do cluster on scales larger than their free-streaming
5 computed using the publicly available Eclairs code length, we assume that a portion of massive neutrinos cluster like

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 7

ρcb (ρm )/ρm -1, R=10 Mpc/h Mν [eV] σµ,cb σµ,m σρ,cb σρ,m
0.05 0 0.6144 0.7399
0.1 0.6183 0.6124 0.7468 0.7414
0.2 0.6222 0.6113 0.7541 0.7438
0 0.4 0.6297 0.6107 0.7686 0.7501
Mν [eV]
- 0.05 z= 0 0.1 Table 3. Measured variances for the log-density µ = log ρ and the density
ρ using all matter (m) or the cold dark matter and baryon component (cb)
z= 1 0.2 at z = 0 for a sphere of radius R = 10 Mpc/h.
- 0.1
bg 0.4
- 0.15
0.1 0.2 0.5 1 2 5 10
3.3 Spherical collapse
ρm
For realistic total neutrino masses that are in agreement with cur-
rent bounds, the effect of massive neutrinos on spherical collapse is
Figure 3. Effect of massive neutrinos on the ratio of the normalised den- typically sub-percent. LoVerde (2014) demonstrated that the main
sities of CDM plus baryons (cb) and total matter (m) in a given massive net effect of massive neutrinos with total mass Mν < 0.5eV is to
neutrino cosmology. We show predictions from equation (18a) considering
increase the collapse threshold by at most 1%. In our theoretical
only the neutrino background (dotted lines) or also the scale-dependent neu-
trino clustering according to equations (18) for radius R = 10 Mpc/h at
model for spherical collapse, the impact of this change can be es-
redshift z = 1 (dashed lines) and z = 0 (solid lines) for total neutrino mass timated by changing the parameter ν ∝ δc entering the spherical
Mν = 0.1eV (blue), 0.2eV (green) and 0.4eV (red). collapse approximation from equation 11. We have checked that
this effect remains below 1% in the entire 2-sigma region around
the mean in logarithmic scale.
When focusing on nonlinear objects like halos and voids that
the cold dark matter and baryon component. We approximate the intrinsically live in the tails of the PDF, the impact of massive
clustering portion using the ratio between the linear variances com- neutrinos on their formation can be more significant. On the one
puted from the CAMB power spectra for massive neutrinos and hand, massive neutrinos lower the abundance of massive dark mat-
cold dark matter and baryons ter halos that host galaxy clusters (Ichiki & Takada 2012; LoVerde
σL,ν (R, z) 2014; Villaescusa-Navarro et al. 2011, 2013) by delaying the col-
δν ' δcb . (18c) lapse time, which is only slightly counteracted by the nonlinear
σL,cb (R, z)
clustering of neutrinos with the halo. On the other hand, neutrinos
This corresponds to considering a simplified scale-dependent bias do not evacuate voids as efficiently as CDM due to their thermal
between neutrinos and cold dark matter plus baryons, and a tight velocities (Massara et al. 2015). This is why the scale-dependent
correlation between the linear fields. Hence, the total matter density bias for voids defined in the total matter field in massive neutrino
is a biased version of the cold dark matter and baryon density and cosmologies (Banerjee & Dalal 2016) is stronger than the scale-
the impact of this bias on the PDF is given by a change of variables dependent halo bias at fixed neutrino mass and rarity of the fluctu-
ation. Unfortunately, this signal is difficult to use currently since it
is challenging to robustly detect voids in total matter through lens-
dρcb
Pm (ρm ) = Pcb (ρcb (ρm )) . (19) ing measurements. Additionally, if one attempts to look for these
dρm
effects by defining voids using tracer populations like galaxies or
In Figure 3, we show the impact of this neutrino clustering on massive clusters, it has been shown that the massive neutrino effects
the ratio between the cold dark matter and baryon component to the sensitively depends on the choice of tracers (Kreisch et al. 2019).
total matter for two redshifts, z = 1 (dashed lines) and z = 0 (solid
lines). We see that the presence of massive neutrinos lowers the
clustering density in underdense regions. This shifts the underdense
tail of the total matter PDF to slightly higher densities, as seen in
3.4 Scale-dependent linear variance and nonlinear variance
the upper panel of Figure 10. Additionally, there is a few percent
effect on the shape around the peak, that is not visible in a log-log In Figure 4 we show the scale-dependence in the linear variance
plot but shown in the lower panels of Figure 10. For the prediction, induced by the presence of massive neutrinos, for the total mat-
we inferred the impact of massive neutrinos on the nonlinear log- ter component (solid lines) and the clustering matter component
variance of cold dark matter plus baryons using equation (14). (dashed lines) at redshift z = 0. For the theoretical predictions that
Since the weak lensing convergence is a projected version of follow, we use the linear variance computed for the clustering mat-
the total matter density contrast (18b), we expect that the scale- ter component and approximate the impact of massive neutrinos on
dependent neutrino clustering according to equation (18c) is di- the nonlinear log-variance according to equation (14). Note that the
rectly related to the imprint of massive neutrinos in the weak lens- nonlinear variance could also be predicted from halofit (Peacock &
ing PDF recently measured in simulations (Liu & Madhavacheril Smith 2014), its extensions (Bird et al. 2012) or perturbative mod-
2019). In fact, the residuals between the PDFs with and without els including massive neutrinos (Saito et al. 2009).
massive neutrinos shown in Figure B2 look qualitatively very sim- In Figure 19 in Section 6, we demonstrate the impact of mas-
ilar. However, we note that part of the observed signature could sive neutrino on the growth of structure at the scales of interest here
be due to a change in the nonlinear variance that is driven by σ8 , R & 10 Mpc/h. For the case of massive neutrinos, the density of
which is not fixed in the MassiveNu simulations used in Liu & the clustering matter component (CDM+baryons) can be estimated
Madhavacheril (2019). from the density of the total matter according to equation (18a).

MNRAS 000, 000–000 (0000)


8 C. Uhlemann et al.

σ2Mν (R)/σ2fid (R)-1 Np = 256 (low resolution), Np = 512 (fiducial resolution) and
0.12 Np = 1024 (high resolution).
Mν=0.4eV cb In this section, we focus on the PDF of the matter field,
0.10
smoothed with a spherical top-hat at different scales, computed
0.08 Mν=0.2eV m
from the Quijote simulations. The PDFs have been computed from
0.06 Mν=0.1eV the Quijote simulations as follows. First, particle positions and
0.04 masses are assigned to a regular grid with Nm (typically 5123 ) cells
using the Cloud-in-Cell (CIC) mass assignment scheme. Next, the
0.02
value of the normalised density field ρ = 1 + δ in each grid cell
0.00 is computed by dividing the mass of each cell by the average cell
- 0.02 mass. We then smooth the density field with a top-hat filter of ra-
5 10 20 dius R. This procedure is done in Fourier-space, by first computing
R [Mpc/h] the Fourier transform of the density field and then multiplying it by
the Fourier transform of the filter on the regular grid itself, to avoid
numerical artifacts on small scales. Finally, the smoothed field is
Figure 4. Comparison between the linear variance σL 2 (R) computed from
estimated by computing the inverse Fourier transform of the pre-
the linear power spectrum of total matter (solid lines) and cold dark mat- vious quantity. The PDF is measured in 99 logarithmically spaced
ter plus baryons (dashed lines) for cosmologies with total neutrino mass bins of normalised density between 10−2 and 102 by calculating
Mν = 0.1eV (blue), 0.2eV (green) and 0.4eV (red) with fixed σ8 . Note
the fraction of cells that lie in a given bin and dividing by the bin
that the offset of the curves for cold dark matter and baryons is because
width.
their clustering amplitude needs to be enhanced to compensate for the lack
of neutrino clustering on large scales to yield the desired σ8 for total matter.
4.2 Matter PDF resolution effects
3.5 Summarised recipe and result We tested the convergence of the PDF measurements with respect
to two resolution parameters — the number of particles Np , and
Combining the recipes outlined above, one can compute the to-
the number of mesh cells Nm . The fiducial resolution is Np =
tal matter density PDF in the presence of massive neutrinos from
Nm = 5123 , giving typical initial inter-particle and grid spac-
the linear power spectrum for cold dark matter plus baryon, their
ing of about 2 Mpc/h. The high-resolution PDF is obtained from
nonlinear log-variance, standard spherical collapse and a mapping
Np = Nm = 10243 , where the initial inter-particle and grid spac-
between the total clustering density and the cold dark matter plus
ing are about 1 Mpc/h. To disentangle the impact of the particle
baryon density according to equations (18).
number and mesh resolution effects, we compared all combinations
In Figure 10 we show the predicted and measured effect of
of Np , Nm ∈ {5123 , 10243 }.
massive neutrinos (with fixed matter clustering amplitude σ8 ). The
In the upper panel of Figure 5 we show a convergence check
impact of massive neutrinos at scales around 10 Mpc/h is largest in
for the PDF, comparing the measured fiducial cosmology PDF in
the underdense regions, where it is at the 10% level (ρ ' 0.2) and
the standard resolution simulation with the high resolution simula-
at the few percent level for overdense (ρ ' 5) regions. This effect
tion. From the plot we see that the resolutions affects the PDF of
is significantly stronger than cosmic variance, as we highlight in a
densities in spheres of radii R = 5, 10, 15 Mpc/h at 6%, 3%, 1%
corresponding residual plot shown in Figure B2 in Appendix B.
level even around the peak. This effect depends only weakly on
redshift. We found that in order to obtain accurate matter PDFs,
the final mesh resolution has to be about a tenth of the radius of
4 VALIDATING THEORY AND SIMULATIONS the spheres. For a quantitative comparison of the standard devia-
tion and the skewness, we therefore focus on radii R = 10, 15, 20
4.1 Measurements from the Quijote simulations
Mpc/h and redshifts z = 0, 0.5, 1. The standard deviation σ is un-
The Quijote simulations are a large suite of full N-body simu- derestimated by 2.5 − 3% for radius R = 10 Mpc/h in the lower
lations designed for two main purposes: 1) to quantify the in- resolution, which is mostly due to the smaller mesh resolution. We
formation content of cosmological observables and 2) to provide find that the lower resolution overestimates the reduced skewness
enough data to train machine learning algorithms. They contain S3 , defined in equation (7), relatively independent of scale and red-
43100 simulations spanning more than 7000 cosmological mod- shift by about 1.5 − 2.5%, mostly caused by the lower particle
els in the {Ωm , Ωb , h, ns , σ8 , Mν , w} hyperplane. At a single red- number (and therefore more subject to rare events).
shift, the total number of particles in the simulations excess 8.5 The finite resolution effects on the matter PDF measured from
trillions, over a combined volume of 43100 (Gpc/h)3 . The simula- simulations highlight the importance of having reliable theoreti-
tions follow the gravitational evolution of Np3 particles (2 × Np3 for cal predictions for the PDF available, allowing to cross-validate
simulations with massive neutrinos) over a comoving volume of 1 the simulation measurements and the theory. We mitigate the im-
(Gpc/h)3 starting from redshift z = 127. Initial conditions for the pact of finite resolution effects by discarding the smallest radius
pure ΛCDM simulations (without massive neutrinos) are generated R = 5 Mpc/h and cutting a percentage of rare density spheres
from 2LPT, while the simulations to assess the impact of massive in the Fisher analysis, essentially limiting the range of logarithmic
neutrinos are run using initial conditions from the Zeldovich ap- densities to the 1.5σ region around the peak.
proximation taking into account the scale-dependent growth factor In the lower panel of Figure 5 we show residuals between the
and growth rate present in these models. Simulations were run us- matter PDF extracted from the fiducial simulations run either from
ing the SPH-T REE PM G ADGET-III CODE. We refer the reader to Zeldovich approximation or 2LPT initial conditions. We can see
Villaescusa-Navarro et al. (2019) for further details on the Quijote that the impact of the initial conditions is sub-percent in the region
simulations. Different values of particle number Np are provided: around the peak, but increases in the tails. Overall, we observe that

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 9

z=0,0.5,1 (ρ), R=10 Mpc/h



2.0
0.10


●●
● ●●

●● ●

z
●● ● ●

0.05

●● ● ●●
●●●● ●● ●●● ●●● ● ●
dHR (ρ)-1

● ● ●●
●●
●● ● ●● ●

3
● ● ●
●● ● ●● ● ● ●
● ● ● ●

1.5
●● ●● ● ● ●●● ●●●● ● ● ●
● ●● ● ●
● ●● ●
●● ● ●● ●
●●●
● ● ● ● ●● ●

0.00
● ●● ● ● ●
●● ●●●●



●● ●●● ●● ● ●
● ● ●● ● ●● ●●●●
●● ●● ●●●●●●● ● ● ● ●●●●
●● ● ● ●● ● ●●
● ●● ●● ●
● ●● ●
●● ●● ●
●●●● ●●●
●●● ●
● ●● ●●●●●●●●●● ●●

● ●● ● ●● ●● ●●●● ●

2
●●● ● ● ●
● ● ●
● ●● ● ● ● ●●●● ● ●
● ● ●● ● ● ● ●● ●
● ● ●● ● ● ●
- 0.05
●● ●● ● ●
● ● ●● ●● ● ●● ● ●
● ● ● ●
● ● ●● ●●

1
● ● ●

R [Mpc/h]
●●

1.0



● ● ● ●● ● ● ●
● ●
●● ● ●● ●
d (ρ)/

- 0.10
● ● ●
● ●
●● ●
● ● ● ●
● ● ●

5 15
●● ●●
●●

0.5
●● ● ●● ● ● ●
● ●●●●
● ● ● ● ● ●
●●

- 0.15
●● ●
● ● ●● ● ● ●● ● ●
● ● ● ● ●

10 20
● ● ● ● ●●
● ●●

0

0.5
● ● ●● ●
● ● ●
● ● ●● ●

- 0.20
● ●

●● ● ● ●
● ●● ●

● ● ● ●
● ●
● ● ● ● ●

-2 0 2 4

0.10 ●

R [Mpc/h]
● ●

0.0

0.0 0.5 1.0 1.5 2.0


fidZA (ρ)/fid (ρ)-1

5 15

● ●
● ●

● ●

0.05
● ●
●●

10 20
● ● ●

● ●

● ●●

● ●
ρ

(ln ρ), R=10 Mpc/h


● ●
● ●● ● ● ● ●
● ● ●
●● ● ● ●● ●●
● ●●

2.0
●● ● ● ● ●● ●

0.00
●● ● ●
● ● ● ●●●● ●
● ●
●●●●●● ●
● ● ●● ●●●
● ●
●●●● ●

●●● ●●●●


●● ●
●●●

●●●
●●
● ● ●●●
●●●●●●●●
● ● ●●
●● ●●
●●●●●●●●● ● ●

●● ●
●●● ●
●●●● ● ● ●
● ●●● ●●
●●●●
●●●●●●
●●
●●

●●
●●
●●
●●
● ● ● ●

● ●●●
● ●
●●●●●

●● ●● ● ●● ● ● ●

●● ●●●●●
●● ●●

●●●

●●●


●●
●●●
●●●

●●●

●●●● ●

● ●●●●● ● ●●
●●●
● ●●●
●● ● ●●●● ●●●
● ●●●●●●
● ● ●●
● ●● ●●●
●●
●●●●●● ●●● ● ● ● ● ●● ● ●

z
●●● ●
● ● ●
●● ●●●●●● ● ●
● ● ●
● ● ● ●● ● ● ● ●● ● ●● ●

● ● ●
●●●● ● ●● ●●●● ● ● ●●
● ● ● ● ●●
● ●● ● ●
●● ●● ● ● ● ●

3
●● ● ●
● ● ●● ● ●
● ● ●● ●

1.5
●● ●

- 0.05
● ● ● ●
● ●● ●
●●

2

-2 0 2 4 ●

(ln ρ-<ln ρ>fid )/σln ρ,fid 1.0 ● 1


mean ● 0.5
Figure 5. Residual between the fiducial matter PDF in spheres of ra- 0.5 ● 0
dius R = 5, 10, 15, 20 Mpc/h (red, yellow, green, blue) at redshifts
z = 0, 0.5, 1 (slightly blue-shifted in color towards larger z), measured
from the means of different simulation runs. The dashed and solid grey 0.0
lines indicate 1% and 2% accuracy, respectively. (Top panel) Residual be- -3 -2 -1 0 1 2 3
tween the mean over 100 simulations with standard resolution and high- (ln ρ-<ln ρ>)/σln ρ
resolution. (Bottom panel) Residual between 500 realisations with initial
conditions using the Zeldovich approximation and 2LPT.
Figure 6. Matter density PDF for spheres of radius R = 10 Mpc/h at red-
shifts z = 0, 0.5, 1, 2, 3 (from red to blue, as indicated in the legend) for the
the impact of initial conditions is smaller than the finite resolution high-resolution run of the fiducial model as measured from the mean over
effects and negligible in the 1.5σ region around the peak, on which 100 realisations (data points) compared to the theoretical prediction (lines).
we will focus in our analysis. We show the density PDF (upper panel), and the PDF of the logarithmic
density, with a shifted and rescaled x-axis to align the peak positions and
unify the widths (lower panel). Lognormal PDFs are shown for comparison
as thin dashed lines, which clearly deviate from the measurements towards
4.3 Shape of the matter PDF for fiducial cosmology
lower redshifts.
In Figure 6 we show a comparison of the theoretical prediction
for the PDF and the measurement from the mean over 100 reali-
sations of the high-resolution simulation for the fiducial cosmol- imberg 2016) to get the correct PDF shape by accounting for a
ogy. We find that the theoretical prediction for the PDF, with the complete (and therefore meaningful) hierarchy of cumulants. This
measured nonlinear variance as an input, performs very well. To has be be contrasted with an Edgeworth-like expansion which trun-
fairly compare the performance of the prediction at different radii cates this cumulant hierarchy and therefore necessarily fails to re-
and redshift, we plot residuals in Figure 7 as a function of the de- produce the tails of the distribution. In fact, the inclusion of higher-
viation of the log-density from its mean in units of the standard order cumulants is essential even to capture the full shape of the
deviation. Data points with error bars indicate the mean and stan- PDF around the peak, as we demonstrate in an accompanying pa-
dard deviation of the PDF bin measurements across the 100 realisa- per where we compare the constraining power of the moments to
tions. Residuals are at a few percent level in the 2-σ region around the central region of the PDF (Friedrich et al. 2019).
the mean for all scales and redshifts at which the nonlinear vari-
2
ance, listed in Table 3, is sufficiently below unity (σNL . 0.5).
4.4 Change of PDF shape with cosmological parameters
As expected, the agreement between the theory and the simula-
tion improves with decreasing the nonlinear variance, which can In the following, we compare the differences of the matter PDFs
be achieved by increasing either redshift or the smoothing radius. for changes in ΛCDM parameters and total neutrino mass, as pre-
Let us note that when going from high (blue curves) to low dicted by our theoretical model and as measured in the simulation.
(red) redshifts, the PDF is skewed towards underdensities as ex- We show differences rather than ratios to highlight deviations in
pected since voids occupy more volume while overdensities be- the shape of the PDF around the peak, where the signal to noise
come more concentrated. While this evolution is mainly driven is highest and most constraining power is located. Even when ex-
by the growth of the skewness, one needs to implement a large- cluding the tails, the full shape of the PDF still carries significant
deviation argument (Bernardeau 1992, 1994a; Bernardeau & Re- non-Gaussian information, as we will show in Figure 9 below. Ad-

MNRAS 000, 000–000 (0000)


10 C. Uhlemann et al.

sim,HR (ρ)/ th (ρ)-1 σ8 + (ln ρ)-σ8 - (ln ρ), z=0


0.2 0.03
z R [Mpc/h]
0 0.5 1 2 3 0.02
● ● ● ● ●
10
0.1

0.01
R=10 Mpc/h

0.00 ● 15
0.0 - 0.01 ● 20
- 0.02
- 0.1 - 0.03
- 0.04
- 0.2 -3 -2 -1 0 1 2 3
0.2- 3 -2 -1 0 1 2 3
z (ln ρ-<ln ρ>fid )/σln ρ,fid
● 0 ● 0.5 ● 1 ● 2 ● 3
0.1
R=15 Mpc/h

Figure 8. Differences between the PDFs with changed σ8 at redshift z = 0


0.0 with radii R = 10, 15, 20 Mpc/h as predicted from our formalism (solid
lines), the lognormal approximation (thin dashed lines) and measured in the
derivative simulations with standard resolution (data points). The gray ver-
- 0.1 tical lines indicate the region that is used for the Fisher analysis in Section 5.

- 0.2
ln ρ)- ln ρ)
0.2- 3 -2 -1 0 1 2 3 +( -(
z
ΔΩm Δns Δh ΔΩb
0 0.5 1 2 3
● ● ● ●
● ● ● ● ●
0.1 0.005
z=0, R=10 Mpc/h
R=20 Mpc/h

0.0
0.000

- 0.1 full pred.


- 0.005
σNL only
- 0.2
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
0.010
(ln ρ-<ln ρ>)/σln ρ
z=0, R=15 Mpc/h

0.005
Figure 7. Residuals between the measured and predicted matter PDF in
0.000
spheres of radius R = 10, 15, 20 Mpc/h (top to bottom) for redshifts z =
0, 0.5, 1, 2, 3 (red to blue, as indicated in the legend).
- 0.005

- 0.010
ditionally, the differences between PDFs for different parameters
determine the response of the PDF to changing cosmology that en- 0.020- 3 -2 -1 0 1 2 3
ter in the Fisher analysis presented in Section 5. Data points with 0.015
z=0, R=20 Mpc/h

error bars indicate the mean and standard deviations of the differ- 0.010
ences in the PDFs measured across 500 realisations. 0.005
0.000
-0.005
4.4.1 ΛCDM cosmological parameter dependence -0.010
-0.015
In Figures 8 and 9 we show differences between the matter PDFs -0.020
for changes in ΛCDM parameters. For the predictions shown as
-3 -2 -1 0 1 2 3
solid lines, we use the theoretical model from equations (6) along
with the predicted cosmology dependence of the log-variance from (ln ρ-(<ln ρ>)fid )/σln ρ, fid
equation (14), normalised with the measured variance for the fidu-
Figure 9. Differences between the matter PDFs for changes in Ωm (red),
cial simulations. Increasing the clustering amplitude σ8 leads to an
Ωb (blue), ns (orange) and h (green) as predicted (solid lines) and measured
increase in the nonlinear variance, which broadens the PDF and
in the simulations (data points) at redshift z = 0 with radii R = 10, 15, 20
hence decreases the peak height, as can be seen by the dip around Mpc/h (from top to bottom). To highlight the impact of cosmology on the
the origin in Figure 8. Since the PDF is normalised, this dip around non-Gaussian shape on small scales, we add predictions that only account
the peak is compensated by an increase of the PDF in regions that for the change in the nonlinear variance (thin dashed lines). The gray verti-
are more significantly underdense or overdense. For comparison, cal lines indicate the region that is used for the Fisher analysis in Section 5.
we also show the expected differences assuming a lognormal matter
PDF (thin dashed lines), which disagree with the simulation mea-
surements in particular in the underdense regions.

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 11
At small scales R ∼ 10 Mpc/h, changing cosmological pa- log10 (ln ρ), R=10 Mpc/h
rameters other than σ8 mostly cause an additional skewness, which 0 ● ● ▲▲
● ● ●▲
● ▲▲
● ●▲
●▲
●▲

▲▲▲▲▲ ● ●▲
●▲●●
●▲
manifests in an asymmetry between densities on different sides ●▲
● ▲▲

●▲

●▲

●● ●▲
▲ ●●
●▲
▲ ▲
●▲●

●●
▲▲

● ▲ ● ●●

of the peak of the PDF. This is in line with the expectation that ▲
● ● ●▲

-1
▲ ●
▲ ▲
● ▲
●▲●
● ● ●▲
▲ ●
▲ ●▲
▲ ●
changes in the amplitude of the variance are small around R = 8 ●
▲ ▲
● ●
▲ ▲

● ▲ ●
● ▲ ●
▲ ●
▲ ▲
● ▲●

Mpc/h due to fixed σ8 , and the main effect is the change of the ●
▲ ▲ ▲ ▲
-2

● ▲
● ▲

▲ ▲ ▲

scale-dependence shown in Figure 1 that modifies the skewness ac- ●


▲ ●




● ▲

▲ ▲
cording to equation (7). At larger scales, the asymmetry disappears ● ● ▲
z= 0 M =0eV


-3
● ▲
● ▲


▲ ν
because the main effect is a few percent change in the variance, re- ●



● ▲





sembling the impact of changing σ8 shown in Figure 8. To highlight
z=0.5 M =0.2eV
▲ ▲

▲ ▲
-4
● ● ▲
● ν ●
the cosmological information in the PDF shape, we contrast the full ▲ ▲

●▲

theoretical prediction including the impact of the cosmological pa- ● ●

-5
●▲ ▲ ●
rameters on the scale-dependence of the variance (solid lines) to a
mere change in the nonlinear variance (thin dashed lines) in Fig- -2 -1 0 1 2
ure 9. This demonstrates the importance of the reduced skewness, ln ρ
which responds to a change in cosmological parameters accord-
Mν (ln ρ)- dZA (ln ρ)
ing to equation (7). Indeed, the full hierarchy of cumulants affects
the PDF shape even in the central region excluding the tails, as we 0.015 0.4eV
0.010 0.2eV
z=0, R=10 Mpc/h
discuss in an accompanying paper focused on the impact of primor-
dial non-Gaussianity (Friedrich et al. 2019). For completeness, we 0.005
show the corresponding ratios of the PDFs when varying ΛCDM 0.1eV
parameters other than σ8 in Figure B1 in Appendix B. 0.000
- 0.005
- 0.010
- 0.015

0.015 - 3 -2 -1 0 1 2 3
0.010
z=0, R=15 Mpc/h

4.4.2 Total neutrino mass dependence 0.005


In Figure 10 we show differences in the total matter PDF for the 0.000
massive neutrino models compared to the fiducial model. Note that - 0.005
both simulations have been run using initial conditions generated
- 0.010
from the Zeldovich approximation. We find that massive neutri-
nos affect the shape of the PDF in a distinct way, that is well pre- - 0.015
dicted by our model (solid lines) and not degenerate with a change 0.020 - 3 -2 -1 0 1 2 3
in σ8 , for which differences are displayed in Figure 8. To highlight 0.015
z=0, R=20 Mpc/h

this, we contrast our model with naive predictions only account- 0.010
ing for the change in the nonlinear variance as thin dashed lines. 0.005
For the smallest radius R = 10 Mpc/h, we can see a significant 0.000
suppression in underdense regions, as expected from Figure 3. Ad- -0.005
ditionally, the skewness is enhanced by the presence of massive -0.010
neutrinos due to a combination of the enhancement by the scale- -0.015
dependent variance demonstrated in Figure 4 and the change of -0.020
variables (18). This leads to a characteristic signature that can be
-3 -2 -1 0 1 2 3
even distinguished by eye from the shapes induced by changing
other ΛCDM parameters shown in Figure 9. For larger radii, this (ln ρ-<ln ρ>fidZA )/σln ρ, fidZA
signature gets concealed, because the range of probed densities be-
Figure 10. (Upper panel) Total matter PDF in spheres of radius R = 10
comes smaller and the differences in the variances grow. The scale-
Mpc/h for redshifts z = 0, 1 for the fiducial model without massive neutri-
dependence of the impact of massive neutrinos on the matter PDF nos and with Mν = 0.2 eV, as indicated in the legend. (Lower panels) The
has a significant advantage over the mildly nonlinear matter power fractional difference of the PDF as measured (data points) and predicted
spectrum, where the clustering amplitude σ8 and the total neutrino (solid lines) for massive neutrinos with mν = 0.1eV (blue), mν = 0.2eV
mass Mν are largely degenerate (Villaescusa-Navarro et al. 2018), (green) and mν = 0.4eV (red) and the fiducial model (with equal σ8 ) as a
as we show in Figure 17. The main effect in the matter PDF is function of density at redshift z = 0 and radii R = 10, 15, 20 Mpc/h (top
caused by the presence of a massive neutrino background in under- to bottom). To highlight that the imprint of massive neutrinos in the matter
dense regions and a partial clustering of massive neutrinos encoded PDF is distinct from a change in σ8 , we add predictions that only account
in the scale-dependent bias from equations (18)). for the change in the nonlinear variance (thin dashed lines). The gray verti-
cal lines indicate the region used for the Fisher analysis in Section 5.
For completeness, we provide a residual plot between the to-
tal matter PDF in the presence of massive neutrinos and the fidu-
cial model in Figure B2. This plot demonstrates that our theoretical
model for massive neutrinos achieves a similar accuracy as the pre-
dictions for changes in the ΛCDM parameters.

MNRAS 000, 000–000 (0000)


12 C. Uhlemann et al.
5 FISHER FORECAST FOR νΛCDM COSMOLOGY Mν = 0.1, 0.2, 0.4eV. For the Fisher analysis, we assume a total
effective cosmic volume of (6 Gpc/h)3 spread equally across the 3
In this section, we quantify the information content of the matter
lowest redshifts in the simulation z = 0, 0.5, 1. The total volume
PDF on the full set of ΛCDM cosmological parameters and the
roughly corresponds to the effective volume of the BOSS galaxy
total neutrino mass using a Fisher matrix formalism together with
survey (Cuesta et al. 2016) and about one tenth of Euclid.
the large suite of Quijote simulations.
Note that such a Fisher analysis has a number of limitations:
After briefly reviewing the basis of the Fisher analysis in Sec-
it only yields realistic error bars if measurements of the consid-
tion 5.1, we explain our data vector in Section 5.2. We discuss the
ered data vectors have Gaussian noise and if the responses of these
covariance matrix in Section 5.3 and the derivatives with respect
data vectors to changing cosmological parameters are close to lin-
to cosmological parameters in Section 5.4. In Section 5.5, we de-
ear. We checked that the distribution of individual bins of the PDFs
termine a suitable combination of smoothing radii and redshifts for
measured in the Quijote sims are sufficiently close to a Gaussian
the matter density PDF, and establish the complementarity between
distribution to expect a small impact on the total width of our fore-
the matter PDF and the matter power spectrum on mildly nonlinear
casted contours. Additionally, realistic data analyses might have to
scales. The final constraints on the full set of νΛCDM parameters
account for systematic effects by marginalising over nuisance pa-
are presented in Figure 18.
rameters. However, the main focus of this study is to explore the
complementarity between the mildly nonlinear power spectrum and
5.1 Basics for the Fisher analysis the matter density PDF as cosmological probes. Hence, we expect
these limitations to have limited impact on our findings.
~ given a
The Fisher matrix on a set of cosmological parameters, θ,
~
(combination of) statistics S is defined as
X ∂Sα −1 ∂Sβ 5.2 Data vector for matter PDFs and power spectra
Fij = C , (20)
∂θi αβ ∂θj In our case the data vector S ~ is built from the values of the matter
α,β
PDF in density bins. First, we consider a single redshift z and ra-
~ and C is the covariance
where Si is the element i of the statistic S dius R. Then, we combine multiple redshifts and radii. In all cases,
matrix, defined as we only use those bins for the Fisher analysis, where the cumu-
Cαβ = h(Sα − S̄α )(Sβ − S̄β )i , S̄α = hSα i . (21) lative probability distribution function (CDF) is between 0.03 and
0.9. This amounts to removing 3% of the lowest densities and 10%
We multiply the inverse of the covariance matrix measured in the of the highest densities. We choose this approach in order to limit
simulation by the Kaufman-Hartlap factor (Kaufman 1967; Hartlap the impact of finite resolution effects that are most severe for rare
et al. 2006), h = (Nsim − 2 − NS )/(Nsim − 1), to correct for a events, as shown in Figure 5, while still capturing the PDF shape
potential bias for the inverse of the maximum-likelihood estimator around the peak, which is located in underdense regions. We chose
of the covariance depending on the ratio of the length of the data an asymmetric cut in the CDF, because the PDF rises more steeply
vector NS to the number of simulations Nsim . Since in our case towards the peak in underdense regions (see Figure 6). Addition-
the number of simulations for covariance estimation is very large ally, it takes into account that the theoretical modelling of the mat-
(15,000) compared to the maximal length of the data vector (218 ter density PDF and the impact of tracer bias or projections will
for our three-redshift analysis of the PDF at three scales and the be more challenging in the tails (particularly for overdensities) and
mildly nonlinear power spectrum), this factor will be close to unity correspondingly degrade the signal to noise from those regions. Fi-
throughout. Additionally, we mimic a BOSS-like effective survey nally, we will include the matter power spectrum in Fourier bins
volume by multiplying the covariance with the ratio of the consid- up to a given kmax in the data vector. The matter power spectrum
ered survey volume V and the simulation volume Vsim . is linearly binned in k-space in steps of the fundamental frequency
The Fisher matrix allows us to determine the error contours kF = 2π/Lbox and the bin center is determined by averaging over
on a set of cosmological parameters under the assumption that the all modes in the interval, exactly as in Villaescusa-Navarro et al.
likelihood is Gaussian. The inverse of the Fisher matrix gives the (2019).
parameter covariance. The error on the parameter θi , marginalised In the next section, we determine the covariance matrix of the
over all other parameters, is given by matter PDF at different scales and its cross-covariance with the
q matter power spectrum. The final ingredient are the partial deriva-
δθi > (F −1 )ii . (22) tives of the summary statistics with respect to the cosmological pa-
rameters, which are discussed in section 5.4.
The Fisher analysis relies on three ingredients,
(i) the chosen summary statistics that enter the data vector,
(ii) their covariance matrix, and 5.3 Covariance of matter PDF bins
(iii) their derivatives with respect to cosmological parameters.
We estimate the covariance using 15,000 realisations of the Quijote
As discussed in Villaescusa-Navarro et al. (2019), the Quijote sim- simulation with fiducial cosmology. While the covariance matrix
ulations are designed to numerically evaluate those three pieces for defined in equation (21) is the quantity that enters the Fisher anal-
different summary statistics, including matter power spectra and ysis, for visualisation purposes it is useful to normalise this matrix
matter density PDFs. There are 15000 simulations at fiducial cos- on the diagonal. For that purpose, we consider the correlation ma-
mology available to estimate covariances, along with 500 simula- trix, defined as
tions each for increasing/decreasing every single ΛCDM parame- Cij
ter. To assess the impact of massive neutrinos, there are 500 simu- Corrij = p , (23)
Cii Cjj
lations run from Zeldovich approximation (instead of 2LPT) initial
conditions for fiducial cosmology, and for total neutrino masses of where C is the covariance matrix.

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 13
In Fig. 11 we show the correlation matrix of the matter den- Correlation matrix of density PDF bins
sity PDF at z = 0 and R = 10 Mpc/h showing the correlation 1.00
between different density bins used for the Fisher analysis. We see 0.4
that, as expected, neighbouring bins are positively correlated, while 0.75
intermediate underdense and overdense bins are anticorrelated with 0.6
each other. Note that the tails of the PDF, which are excluded in 0.50
0.8
the plot, are strongly correlated with each other and anti-correlated
0.25
with the peak. Note that this is completely in line with the corre-
1.0
lation matrix predicted by the large-deviation formalism (see Ap-
0.00


pendix C in Codis et al. 2016b), which can be decomposed into a 1.2
shot noise contribution, a cosmic variance term due to the finite vol- 0.25
ume of the survey and a term describing the spatial correlation of 1.4
spheres. This last contribution dominates when enough spheres are 0.50
considered and is proportional to the product of the sphere bias of 1.6
the respective density bins, multiplied by the average dark matter 0.75
correlation function at the typical separation of the spheres. Be- 1.8
cause the sphere bias is negative for underdensities and positive for 1.00
0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
overdensities, this product is positive when the cross-correlations

of overdensities (or underdensities) is considered and negative for
the cross-correlations of underdensities and overdensities.
In an accompanying paper (Friedrich et al. 2019), we show
Figure 11. Correlation matrix for the bins of the matter density PDF at
that the correlation matrix measured in zero-mean shifted lognor- radius R = 10 Mpc/h and redshift z = 0.
mal realisations (Hilbert et al. 2011; Xavier et al. 2016) closely
resembles the simulation result. This is an encouraging result, as it
provides a simple way to estimate covariances and is used in cur-
Correlation matrix PDFs and P(k) bins
rent analyses (Friedrich et al. 2018; Gruen et al. 2018). In particular, 0 1.00
this could be used to estimate the impact of super-sample covari-
ance (Takada & Hu 2013; Chan et al. 2018; Barreira et al. 2018), 0.75
10
which is not captured in our analysis using the full periodic simu-
lation boxes. 0.50
Note that, when assuming a diagonal covariance matrix for the 20
matter PDF at one scale and radius, corresponding parameter errors 0.25
are significantly underestimated. For constraints on the clustering 30
amplitude σ8 and matter density Ωm , a single PDF with diagonal 0.00
covariance underestimates the contour area by a factor of 5. On 40
the other hand, when combining the PDF at two different radii and 0.25
assuming a block-diagonal covariance matrix, parameter errors are
50 0.50
overestimated. Combining two PDFs as if they were independent
using a block-diagonal covariance leads to a wrong orientation of
the error ellipse and an overestimation of its area by a factor of 60 0.75
about 2. This highlights the importance of an accurate covariance
matrix and potentially valuable information in cross-correlations of 1.00
0 10 20 30 40 50 60
summary statistics.
Figure 12 shows the cross-correlation matrix between the mat-
ter PDF at two different scales and the mildly nonlinear power spec- Figure 12. Correlation matrix for the matter PDF PR (ρ) at radii R =
trum. First, we observe that PDFs at different scales are relatively 10, 15 Mpc/h and the mildly nonlinear power spectrum P (k), both at red-
strongly correlated with each other. This is expected, as the matter shift z = 0. The density PDF bins correspond to a range of densities as
density fields smoothed at radii of R = 10 and R = 15 Mpc/h shown in Figure 11. The power spectrum is shown in 31 bins of the funda-
are qualitatively similar, as about of a third of the mass in a sphere mental frequency kf ' 0.0063h/Mpc up to kmax = 0.2h/Mpc.
of radius R = 15 Mpc/h comes from a sub-sphere of R = 10
Mpc/h. The cross-correlations between PDF bins of different radii
look very similar to the bin correlations for the individual PDFs, be- bins varies very mildly with the considered k-bin, because it re-
cause the clustering of spheres encoded in the sphere bias changes flects the correlation of the variance with the PDF bins. As can be
mildly with radius (see Figure 5 in Uhlemann et al. 2017). The vari- inferred from Figure 8, the variance is negatively correlated with
ance of the PDF at the smaller scale probes a slightly wider range the overall height of the PDF, but positively correlated with the
in the nonlinear power spectrum, as we demonstrate in Figure A3. tails. Due to our conservative cut in overdense regions (chosen to
In contrast, there is hardly any cross-correlation between the in- mitigate the impact of resolution effects and nonlinear tracer bias),
dividual k-bins of the matter power spectrum since we focus on the positively correlated overdense tail does not appear in the cor-
mildly nonlinear scales. The correlation between the PDFs and the relation matrix for the Fisher analysis.
power spectrum is overall small, suggesting that the two probes are For our multi-redshift analysis, we assume no cross-
complementary and their combination can increase the constraining correlation between different redshift slices and build a block-
power. The correlation between the power spectrum and the PDF diagonal covariance matrix. We adopt this approach as it is com-

MNRAS 000, 000–000 (0000)


14 C. Uhlemann et al.
mon practice in cosmological analysis of galaxy clustering in cur-
z=0, V=2 (Gpc/h)3
rent surveys like DES (Krause et al. 2017), and because correla- PDF, R=20 Mpc/h
tions between different redshift bins are expected to be negligible PDF, R=15 Mpc/h
for non-neighbouring bins of size ∆z = 0.1 as intended for Eu- PDF, R=10 Mpc/h
clid (Bailoni et al. 2017). Additionally, estimating those correla- PDF, all R
tions from our simulation suite would not lead to realistic results,
as the snapshots are extracted from identical runs and hence highly
correlated.

5
85
0.
0
8
84
0.
5.4 Derivatives with respect to cosmological parameters

5
82
0.
The third ingredient of the Fisher matrix (20) are the derivatives of

0
81
the matter PDFs with respect to cosmological parameters. They can

0.
be obtained from the differences between the matter PDFs with one

6
1.
cosmological parameter varied each, which have been discussed in

2
Section 4.4. For ΛCDM parameters, θ ∈ {Ωm , Ωb , h, ns , σ8 }), we

1.
ns
compute partial derivatives from two-point finite differences

8
0.
~
∂S ~ + dθ) − S(θ
S(θ ~ − dθ)

4
0.
' . (24)
∂θ 2dθ

15

30

45

60
0

6
81

82

84

85

0.

0.

1.

1.
0.

0.

0.

0.
ns

0.

0.

0.

0.
For variations in the total neutrino mass Mν , we use finite dif- ⌦m 8
ference formulas to estimate the derivative at the left endpoint
Mν = 0 using two, three and four points
~ ~ ν ) − S(M~ ν = 0) Figure 13. Fisher forecast constraints on {σ8 , Ωm , ns } (with fixed Ωb and
∂S S(M
' , h) from the matter PDF at redshift z = 0 using a single radius R = 20
∂Mν Mν Mpc/h (yellow), R = 15 Mpc/h (blue), R = 10 Mpc/h (green) or three
~
∂S ~
−S(2M ~ ~
ν ) + 4S(Mν ) − 3S(Mν = 0) radii combined (red dashed).
' , (25)
∂Mν 2Mν
~
∂S ~
S(4M ~ ~ ~
ν ) − 12S(2Mν ) + 32S(Mν ) − 21S(Mν = 0)
' .
∂Mν 12Mν
z=0, V = 2 (Gpc/h)3,
By default, we rely on the 4-point derivative and use the other for- R = 10, 15 Mpc/h
mulas for consistency checks. While for technical reasons, the sim- P (k ≤ 0.2h/Mpc)
ulations have been run with zero fiducial total neutrino mass, we σρ,NL
know already that there is a lower limit to the total neutrino mass PDF
of about Mν & 0.056eV (Lesgourgues & Pastor 2006). Hence, we
will quote constraints in terms of ∆Mν and avoid to fold in a hard
prior to enforce a positive neutrino mass.

5.5 Constraining ΛCDM and massive neutrinos


0
84
0.

In the following we present constraints on key ΛCDM parameters


6
83

and the total neutrino mass using the matter PDF at different red-
σ8
0.

shifts and radii. We compare its constraining power with results


2

from the matter power spectrum and combine the two large-scale
83
0.

structure probes.
8

We have verified the convergence of our results, as constraints


82
0.

do not change if the covariance and derivatives are computed from a


4

0
30

31

32

32

82

83

83

84

smaller number of realisations, or the massive neutrino derivatives


0.

0.

0.

0.

0.

0.

0.

0.

are evaluated using lower order approximations. Ωm σ8

5.5.1 Understanding constraints from the PDF at redshift zero Figure 14. Fisher forecast for constraints on {σ8 , Ωm } at redshift z =
A Fisher forecast for constraints on the three ΛCDM parame- 0 from the matter PDF (green) or the nonlinear variance (red) using two
radii R = 10 and R = 15 Mpc/h, and the matter power spectrum up to
ters {Ωm , σ8 , ns } using one redshift z = 0 and different radii
kmax = 0.2h/Mpc (blue). By considering the shape of the PDF the area
R = 10, 15, 20 Mpc/h, as well as their combination is presented of the contours shrinks by a factor of about 2.5 compared to the variance
in Figure 13. The degeneracy between the matter density Ωm and and by more than a factor of 3.5 compared to the mildly nonlinear power
the spectral index ns arises because the shape of the PDF around spectrum.
its peak is sensitive to the overall ‘tilt’ of the scale-dependent lin-
ear variance, as explained in Section 2.1. Since the magnitude of the
induced additional tilt depends on scale, the degeneracy direction
between Ωm and ns slightly rotates when increasing the radius.

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 15

V =6 (Gpc/h)3: R=10,15,20 Mpc/h z=0, 0.5, 1, Vtot=6 (Gpc/h)3


PDF, z =1 P (k), kmax =0.2 h/Mpc
PDF, z =0.5 PDF, R=10, 15, 20 Mpc/h
PDF, z =0 joint P (k)+PDF
PDF, all z
84 .90 .96 .02 .08 810 825 840 855

00 25 50 75
0.

83 .83 .83 .83


0
8

8
0.

0
0.

0. 0
0 0 1 1 0.

20
1.
05
ns

ns
1.
90
0.
75
0.
0.

27
30
33
36

5
84
90
96
02
08

4
2
0
8

0. 6
0. 0
0. 5
0. 0
75

75

90

05

20
81

82

84

85

30
31
32
32
33

0
2
5
0.
0.
0.
0.

0.
0.
0.
1.
1.

83
83
83
83

0.

0.

1.

1.
0.

0.

0.

0.

0.
0.
0.
0.
0.
⌦m 8
ns ⌦m ns
8

∆Ωm ∆σ8 ∆ns


Figure 15. Fisher forecast for PDF constraints on {σ8 , Ωm , ns } (with fixed P (k) ±0.00503 ±0.00153 ±0.0386
Ωb and h) from the matter PDF with three radii R = 10, 15, 20 Mpc/h PDF ±0.00243 ±0.000895 ±0.081
at redshifts z = 0 (green) 0.5 (blue) and 1 (yellow) Mpc/h, each for one joint ±0.00225 ±0.000740 ±0.0205
third of the volume, and combined to the total volume (red dashed). Note
Figure 16. Fisher forecast for marginalised constraints on {σ8 , Ωm , ns }
that the green contours here correspond to the red contours in Figure 13
from an analysis at three redshifts z = 0, 0.5, 1 for the matter PDF at three
radii R = 10, 15, 20 Mpc/h (green), the matter power spectrum up to
kmax = 0.2h/Mpc (blue) and both probes combined (red dashed).
When increasing the radius of the spheres, most of the change in
the PDF shape is due to a change in the nonlinear variance, which
induces a degeneracy between σ8 and {Ωm , ns }, as expected from Figure 15 we show that measuring the matter PDF at different red-
Figure 9. When combining the matter PDF at different radii, those shifts breaks this degeneracy between σ8 and Ωm . A split in red-
degeneracies are broken and the constraining power on Ωm and ns shift slices also helps to disentangle a change in matter density Ωm ,
is significantly enhanced. that changes both the scale-dependence and the growth of the linear
In Figure 14 we demonstrate that the shape of the PDF in variance, from a change in the spectral index ns . We consider three
the central region, even excluding the tails, contains more infor- redshifts z = 0, 0.5, 1, which are of particular interest for upcom-
mation than the nonlinear variance of the smoothed density field. ing galaxy surveys like Euclid and LSST. For simplicity, we split
We use two radii for this analysis, because the variance measured the total volume V evenly across the supposedly independent red-
at N scales can only constrain N parameters. We compare the shift slices, such that the constraints shown for the three individual
Fisher forecasts at redshift z = 0 for the two ΛCDM parameters redshifts use only one third of the total volume.
{σ8 , Ωm }, with all other parameters fixed, using the matter PDF
(green) and the nonlinear variance6 (red). By considering the shape 5.5.3 Combining the matter PDF and power spectrum
of the PDF rather than just the variance, the area of the contours
shrinks by a factor of 2.5. For comparison, we show Fisher con- Using the matter power spectrum and the matter PDF at mildly non-
straints for the matter power spectrum on mildly nonlinear scales linear scales as complementary probes, one can enhance the con-
up to kmax = 0.2h/Mpc (blue), whose constraints are weaker by straining power by combining both observables. To demonstrate
a factor of 3.5. This demonstrates that the matter PDF measured at this, we first focus on the five ΛCDM parameters. In Figure 16 we
two scales contains more information on Ωm and σ8 than both the show constraints on {Ωm , σ8 , ns } (marginalised over Ωb and h)
nonlinear variance and the matter power spectrum, when all other obtained from the matter PDF at three radii and the mildly nonlin-
parameters are fixed. ear matter power spectrum, both analysed at three redshifts. The
1σ constraints quoted in the Table are obtained by marginalising
over all other parameters. We find that the matter PDF is strong at
5.5.2 Increasing constraining power with multiple redshifts constraining the clustering amplitude σ8 and matter density Ωm , as
From theoretical grounds, we expect that the PDF at nonzero red- expected from Figure 14. Constraints on those two parameters im-
shift will have a degeneracy between the clustering amplitude prove only by about 10-15% when adding the matter power spec-
σ8 and the matter density Ωm , which affects the linear variance trum. However, the matter power spectrum is stronger at constrain-
through the growth D(z) according to σL (z, R) ∝ D(z)σ8 . In ing ns by about a factor of two, and adding the PDF improves con-
straints by another factor of 2. As expected, the matter power spec-
trum shape can more easily distinguish between changes in Ωb , h
6 Note that the nonlinear variance was measured directly from the grid of and ns , which all lead to a similar signature in the matter PDF as
smoothed densities instead of from the histogram that gives the PDF. demonstrated in Figure 1.

MNRAS 000, 000–000 (0000)


16 C. Uhlemann et al.
Figure 17 and Figure 15 in Villaescusa-Navarro et al. (2019), it
z=0, 0.5, 1 helps to improve constraints on ΛCDM parameters and combin-
Vtot = 6 (Gpc/h)3 ing it with the matter PDF tightens neutrino mass constraints by
P (k), kmax =0.2h/Mpc 30% to less than 0.01 eV. Combining the matter PDF and matter
power spectrum improves constraints for the matter density Ωm by
PDF, R=10,15 Mpc/h
a factor of 5, and for the clustering amplitude σ8 by a factor of 2
compared to power spectrum only.
The improvement of neutrino mass constraints by considering
the matter PDF compared to the matter power spectrum is much
bigger than the one expected from adding the matter bispectrum.
As demonstrated in Coulton et al. (2019), the tomographic weak
04

lensing convergence bispectrum has a similar degeneracy between


0.

Mν and the clustering amplitude σ8 (or a combination of As and


02
0.

Ωm ) than the power spectrum. This suggest that the additional con-
∆Mν

straining power of the PDF is rooted in its ability to detect dif-


00
0.

ferences in clustering between underdense and overdense regions,


02

which are sensitive to neutrino mass as shown in Figure 3 but get


0.

mixed up in N -point spectra. While focused on different scales,


04

recent simulation results for massive neutrino constraints from the


0.

00

25

50

75

04

02
00

02

04

weak lensing convergence PDF in the presence of shape noise (Liu


83

83

83

83

0.

0.
0.

0.

0.

& Madhavacheril 2019) indeed show a turning of the degeneracy


0.

0.

0.

0.

σ8 ∆Mν
direction between Mν and the clustering amplitude compared to
the power spectrum. On the other hand, the nonlinear redshift-space
Figure 17. Fisher forecast for PDF constraints on {σ8 , Mν } (fixing all halo bispectrum was recently shown to significantly improve con-
other parameters) from a joint analysis at redshifts z = 0, 0.5, 1 for the straints from the redshift space halo power spectrum due to the
matter power spectrum up to kmax = 0.2h/Mpc (blue) and the matter shape-dependent imprint of massive neutrinos measured in simu-
PDF at two radii R = 10, 15 Mpc/h (green). lations (Hahn et al. 2019).

5.6 Applicability to survey data


5.5.4 Joint ΛCDM and neutrino mass constraints
In this study, we have investigated the statistical power of the3D
Having established the complementarity of the matter PDF and matter density PDF as a cosmological probe. In realistic observa-
matter power spectrum at mildly nonlinear scales for ΛCDM pa- tional data we cannot access those 3D matter density in cells di-
rameters,
P we now include the total mass of massive neutrinos rectly. Observables that probe the matter density field fall into two
Mν = mν as additional parameter. broad categories: tracer densities and weak lensing fields.
For the matter power spectrum, the total neutrino mass is One possibility is to avoid tracer bias altogether by extract-
known to be largely degenerate with the amplitude of matter fluctu- ing the weak lensing convergence or shear field, which probes
ations σ8 (Villaescusa-Navarro et al. 2018). In Figure 17, we con- the total projected matter field. In the spirit of our analysis, one
trast constraints from the matter power spectrum up to kmax = could parametrise the cosmology dependence for PDFs of the weak
0.2h/Mpc (blue) to the matter PDF at two radii R = 10, 15 Mpc/h lensing convergence (Valageas 2000; Clerkin et al. 2016; Patton
(green), both using three redshifts and fixing all parameters except et al. 2017; Barthelemy et al. 2019), aperture mass (Bernardeau
for σ8 and Mν . We observe that while the matter power spectrum & Valageas 2000; Reimberg & Bernardeau 2018) or cosmic shear
presents a strong degeneracy (which is not alleviated by extending (Takahashi et al. 2011; Friedrich et al. 2018). Recently, the conver-
the range to kmax = 0.5h/Mpc) the matter PDF can easily disen- gence PDF in tomographic redshift slices has been predicted from
tangle the two parameters, as expected from the imprint of scale- large-deviation statistics and cylindrical collapse (Barthelemy et al.
dependent neutrino clustering shown in Figure 10. Indeed, the im- 2019), using ingredients similar to the ones discussed here. While
pressive constraining power of the matter PDF for the total neutrino weak lensing is insensitive to tracer bias, the presence of baryons
mass is hardly diminished by opening up all ΛCDM parameters. could affect the total matter field on small scales. According to
In Figure 18 we show constraints on the full set of νΛCDM Foreman et al. (2019), baryonic effects do not change the scal-
parameters from an analysis in three redshift slices at z = 0, 0.5, 1 ing relations between the matter bispectrum and the power spec-
with a combined volume of 6 (Gpc/h)3 . We show one and two- trum on mildly nonlinear scales. This suggests that baryonic effects
sigma contours obtained using the matter power spectrum up to on the PDF could potentially be modelled through their impact on
kmax = 0.2h/Mpc (blue), the matter PDF at three radii R = the nonlinear variance, while leaving the reduced skewness S3 un-
10, 15, 20 Mpc/h (green) and their combination (dashed, red). We touched. Shape noise effects that encapsulate the uncertainties on
choose the scales to be in a regime where theoretical predictions, the intrinsic shape of galaxies can be modeled by convolving the
based on perturbation theory for the matter power spectrum, and weak lensing PDF with a Gaussian filter of appropriate width (Liu
large-deviation statistics with spherical collapse for the density & Madhavacheril 2019; Barthelemy et al. 2019).
PDF, can be expected to be accurate. For comparison, we show re- A second option is to extract biased tracer densities from dis-
sults from a Gaussian likelihood approximation of the Planck CMB crete counts of galaxies (Yang & Saslaw 2011; Bel et al. 2016;
data with free neutrino mass using only temperature and low mul- Hurtado-Gil et al. 2017), Lyman-alpha absorption in quasar spectra
tiple polarisation data (yellow). While the matter power spectrum (Munshi et al. 2012; Lidz et al. 2006) or 21cm emission of neu-
has virtually no sensitivity to neutrino mass, as demonstrated in tral hydrogen (Leicht et al. 2019). The impact of tracer bias on

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 17

Planck base mnu plikHM TT lowl lowE


z=0,0.5,1, Vtot = 6 (Gpc/h)3:
P (k), kmax =0.2h/Mpc
PDF, R = 10, 15, 20 Mpc/h
joint P (k)+PDF
08 04 00 04 08 75 90 05 20 0.5 0.6 0.7 0.8 0.9.04 .00 .04 .08 .12 825 830 835 840
0 0 0 0 0 0. 0. 0. 0.
8
⌦b
h100
0. 0. 0. 0. 0. 0. 0. 1. 1.
ns
M⌫ [eV]

28
30
32
34

0. 5
0. 0
0. 5
0
0. 4
0. 0
0. 4
0. 8
12

5
6
7
8
9
75
90
05
20

0. 8
0. 4
0. 0
0. 4
08
82
83
83
84

0
0
0
0

0.
0.
0.
0.
0.

0
0
0
0
0.
0.
0.
0.

0.

0.
0.
1.
1.
0.

h100
0.

⌦m ⌦b ns
8 M⌫ [eV]
∆Ωm ∆σ8 ∆Ωb ∆h100 ∆ns ∆Mν [eV ]
Planck base mnu plikHM TT lowl lowE ±0.0315 ±0.0372 ±0.00317 ±0.0212 ±0.006214 ±0.181
P (k), z = 0, 0.5, 1, kmax = 0.2h/Mpc ±0.0117 ±0.00153 ±0.00575 ±0.0644 ±0.0520 ±0.267
PDF, z = 0, 0.5, 1, R = 10, 15, 20 Mpc/h ±0.00276 ±0.00122 ±0.0324 ±0.267 ±0.108 ±0.0131
joint P (k) + PDF ±0.00232 ±0.000792 ±0.00370 ±0.0380 ±0.0208 ±0.00887

Figure 18. Fisher forecast for constraints on νΛCDM parameters at redshifts z = 0, 0.5, 1 using the matter PDF for three radii R = 10, 15, 20 Mpc/h
(green), the matter power spectrum P (k) up to kmax = 0.2h/Mpc (blue) or both in combination (red dashed). The combination with the power spectrum
improves the constraints on the ΛCDM parameters, while the information on the total neutrino mass Mν comes almost exclusively from the PDF. For
comparison, we also show contours from Planck (yellow) obtained from the chains for free neutrino mass from temperature and low multipole polarisation
data only (base mnu plikHM TT lowl lowE).

the shape of the PDF can be modeled using scatter plots between Pan 2004). As shown in Uhlemann et al. (2018a), suitable weight-
the matter and tracer density in cells extracted from simulations ings by halo mass or galaxy luminosity can substantially reduce the
(Manera & Gaztañaga 2011; Jee et al. 2012), measured moments scatter around the mean bias relation, and redshift space distortions
(Salvador et al. 2018) or abundance-matching inspired techniques can potentially be absorbed in the bias model.
that directly operate on the PDFs (Sigad et al. 2000; Szapudi &
When focusing solely on the tracer density PDF, a partial

MNRAS 000, 000–000 (0000)


18 C. Uhlemann et al.
degeneracy between the linear bias coefficient b1 and the ampli- D±Δ (z)/Dfid (z)-1
tude of matter fluctuations σ8 arises. Breaking this degeneracy
can be achieved by considering the density-dependent clustering 0.02
Δw0=0.01
of spheres (called sphere bias Bernardeau 1996; Codis et al. 2016a;
ΔΩm
Uhlemann et al. 2017, 2018a), which can also be used to quan- 0.01
tify the cosmic error induced by extracting counts-in-cells statistics
from a finite number of tracers in a finite volume (Colombi et al.
0.00
1995; Szapudi & Colombi 1996; Szapudi et al. 1999; Codis et al.
2016a).
A middle ground in between measuring weak lensing and - 0.01
modelling bias are the so-called density-split statistics (Gruen et al. Mν=0.2eV
2016; Friedrich et al. 2018; Gruen et al. 2018) that measure the Mν=0.1eV
- 0.02
galaxy density PDF and use lensing measurements to relate it to
the matter density PDF quantile-by-quantile. Gruen et al. (2018) 0.0 0.5 1.0 1.5 2.0 2.5 3.0
have successfully applied this technique in an analysis of obser- z
vational data taken from the Dark Energy Survey and SDSS, thus
demonstrating that tracer bias can be dealt with in a PDF-based
analyses. With photometric galaxy surveys, we do not have access Figure 19. Fractional changes in the linear growth factor D(z) for cos-
to the undistorted 3D density field, but instead probe line-of-sight mologies with varying Ωm = ±0.01 (red solid/dashed) and w0 = ±0.01
projections of the density field with a certain, irreducible uncer- (black solid/dashed). We also show the impact of massive neutrinos on the
growth of the linear variance for cold dark matter plus baryons at radius
tainty in the radial (redshift) direction. However, the formalism em-
R = 15 Mpc/h for a total neutrino mass of Mν = 0.1eV (blue solid) and
ployed in this work can be extended to account for such projections
0.2eV (green solid).
(Bernardeau 1995; Uhlemann et al. 2018c) and has already been
successfully applied to observational data in Friedrich et al. (2018);
Gruen et al. (2018).
(Friedrich et al. 2018; Gruen et al. 2018), whose combined anal-
ysis of counts- and lensing-in-cells allows to constrain bias and
stochasticity along with cosmological parameters, as discussed in
6 CONCLUSION Section 5.6.

In this work, we determined the information content of the mat-


ter density PDF with regards to all ΛCDM parameters and the
Outlook: primordial non-Gaussianity & dark energy
total neutrino mass. Based on a theoretical model for the matter
PDF from large-deviation statistics, we analysed the impact of cos- In this work we focus on Gaussian initial conditions, while an ac-
mological parameters on the ingredients that determine the shape companying paper (Friedrich et al. 2019) generalises the theoreti-
of the matter PDF. We demonstrated that the ΛCDM parameter cal model to include arbitrary non-Gaussian initital conditions. In
dependence of the matter PDF can be predicted from the scale- particular, the focus is on the imprint of a general primordial bis-
dependence of the linear variance, the growth of structure, spheri- pectrum and illustrate effects of orthogonal and equilateral primor-
cal collapse and the nonlinear variance at the considered radius. For dial non-Gaussianity in the matter PDF, complementing previous
the first time, we modelled the impact of massive neutrinos on the results for local non-Gaussianity (Uhlemann et al. 2018b). We find
total matter PDF, finding that their distinct imprint on the shape is that the amplitude of primordial non-Gaussianity fNL can be con-
due to a massive neutrino background affecting underdensities and strained from the matter PDF at two scales even when marginalis-
their partial clustering along with the cold dark matter plus baryon ing over their variances.
component. In all cases, we find an excellent agreement between While it is beyond the scope of this paper, we plan to include
the theoretically predicted and numerically measured response of changes in the dark energy equation of state in a future analysis.
the matter PDF to changing cosmological parameters. From our theoretical model, we expect that dark energy affects
Finally, we performed a Fisher analysis and demonstrated that the matter PDF trough a redshift-dependent change in the variance
measuring the PDF in multiple redshift slices and at different radii driven by the growth of structure, see equation (A1). In princi-
breaks parameter degeneracies and tightens constraints. In Fig- ple, this allows us to constrain the dark energy equation of state
ure 18, we demonstrated the significant constraining power of the in a multi-redshift analysis (Codis et al. 2016a). In Figure 19 we
matter PDF for the matter density Ωm , the clustering amplitude compare modifications in the linear growth induced by a constant
σ8 and the total neutrino mass Mν , highlighting its complemen- change in the dark energy equation of state (purple) to a change
tarity to the matter power spectrum and cosmic microwave back- in the matter density (black lines) and the imprint total neutrino
ground data from Planck. Combining the total matter density PDFs mass (red, blue, green). Since the characteristic imprint of massive
at three radii and the matter power spectrum up to mildly nonlin- neutrinos in the PDF is mainly driven by a scale-dependent bias in-
ear scales in three redshift slices with a total BOSS-like volume stead of a change in the nonlinear variance induced by the growth,
of 6 (Gpc/h)3 gives a marginalised constraint on the total neutrino we expect it to be distinguishable from dark energy. When focusing
mass of order 0.01 eV. This would allow to place a 5σ constraint on on the time-dependence of the variance at low redshifts, a constant
the minimum sum of the neutrino masses with a rather small vol- change in the dark energy equation of state could be degenerate
ume. Additionally, the inclusion of the PDF improves constraints with a change in matter density. However, since the matter density
on {Ωm , σ8 , ns } by a factor of 5, 2, 2.5 compared to the matter also changes the scale-dependence of the linear variance, one could
power spectrum alone, see Table 18. hope to jointly constrain both parameters without loosing too much
This is an exciting prospect for density-split statistics constraining power.

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 19
ACKNOWLEDGEMENTS Colombi S., Bouchet F. R., Schaeffer R., 1995, ApJS, 96, 401
Cooray A., Sheth R., 2002, Phys. Rep., 372, 1
CU kindly acknowledges funding by the STFC grant RG84196 Coulton W. R., Liu J., Madhavacheril M. S., Böhm V., Spergel D. N., 2019,
‘Revealing the Structure of the Universe’. OF gratefully ac- J. Cosmology Astropart. Phys., 2019, 043
knowledges support by the Kavli Foundation and the Interna- Cuesta A. J., et al., 2016, MNRAS, 457, 1770
tional Newton Trust through a Newton-Kavli-Junior Fellowship Eisenstein D. J., Hu W., 1998, ApJ, 496, 605
and by Churchill College Cambridge through a postdoctoral By- Font-Ribera A., McDonald P., Mostek N., Reid B. A., Seo H.-J., Slosar A.,
Fellowship. We thank Ken Osato and Takahiro Nishimichi for 2014, J. Cosmology Astropart. Phys., 2014, 023
making their codes for computing nonlinear matter power spectra Foreman S., Coulton W., Villaescusa-Navarro F., Barreira A., 2019, arXiv
publicly available. Part of the work of FVN has been supported e-prints, p. arXiv:1910.03597
by the Simons Foundation. SC’s work is partially supported by Friedrich O., et al., 2018, Phys. Rev. D, 98, 023508
the SPHERES grant ANR-18-CE31-0009 of the French Agence Fukuda Y., et al., 1998, Physical Review Letters, 81, 1562
Gatti M., et al., 2019, arXiv e-prints, p. arXiv:1911.05568
Nationale de la Recherche and by Fondation MERAC. The au-
Glazebrook K., Blake C., 2005, The Astrophysical Journal, 631, 1
thors thank Alexandre Barreira, Francis Bernardeau, Daniel Gruen,
Gruen D., et al., 2016, MNRAS, 455, 3367
ChangHoon Hahn, Christophe Pichon and Blake Sherwin for dis- Gruen D., et al., 2018, Phys. Rev. D, 98, 023507
cussions and comments on the draft. Hahn C., Francisco V.-N., Emanuele C., Roman S., 2019, arXiv e-prints, p.
arXiv:1909.11107
Hartlap J., Simon P., Schneider P., 2006, Astronomy & Astrophysics, 464,
References 399404
Hilbert S., Hartlap J., Schneider P., 2011, Astronomy & Astrophysics, 536,
Abazajian K., et al., 2015, Astroparticle Physics, 63, 6680 A85
Abramo L. R., Batista R. C., Liberato L., Rosenfeld R., 2007, Journal of Hurtado-Gil L., Martı́nez V. J., Arnalte-Mur P., Pons-Borderı́a M.-J.,
Cosmology and Astroparticle Physics, 2007, 012012 Pareja-Flores C., Paredes S., 2017, A&A, 601, A40
Ahmed S. N., et al., 2004, Physical Review Letters, 92, 181301 Ichiki K., Takada M., 2012, Phys. Rev. D, 85, 063521
Bailoni A., Spurio Mancini A., Amendola L., 2017, Monthly Notices of the Ivanov M. M., Kaurov A. A., Sibiryakov S., 2019, J. Cosmology Astropart.
Royal Astronomical Society, 470, 688705 Phys., 2019, 009
Banerjee A., Dalal N., 2016, Journal of Cosmology and Astroparticle Ivezić Ž., et al., 2019, ApJ, 873, 111
Physics, 2016, 015015 Jee I., Park C., Kim J., Choi Y.-Y., Kim S. S., 2012, ApJ, 753, 11
Barreira A., Li B., Baugh C. M., Pascoli S., 2013, Journal of Cosmology Kaufman G. M., 1967, Report No. 6710, Center for Operations Research
and Astroparticle Physics, 2013, 056056 and Econometrics, Catholic University of Louvain, Heverlee, Belgium
Barreira A., Krause E., Schmidt F., 2018, J. Cosmology Astropart. Phys., Kayo I., Taruya A., Suto Y., 2001, ApJ, 561, 22
2018, 015 Kopp M., Appleby S. A., Achitouv I., Weller J., 2013, Physical Review D,
Barthelemy A., Codis S., Uhlemann C., Bernardeau F., Gavazzi R., 2019 88
Baumann D., Nicolis A., Senatore L., Zaldarriaga M., 2012, J. Cosmology Krause E., et al., 2017, arXiv e-prints, p. arXiv:1706.09359
Astropart. Phys., 2012, 051
Kreisch C. D., Pisani A., Carbone C., Liu J., Hawken A. J., Massara E.,
Becker-Szendy R., et al., 1992, Phys. Rev. D, 46, 3720
Spergel D. N., Wandelt B. D., 2019, MNRAS, 488, 4413
Bel J., Marinoni C., 2014, A&A, 563, A36
Lam T. Y., Sheth R. K., 2008, MNRAS, 386, 407
Bel J., et al., 2016, A&A, 588, A51
Laureijs R., et al., 2011, preprint, (arXiv:1110.3193)
Bernardeau F., 1992, ApJ, 392, 1
Leicht O., Uhlemann C., Villaescusa-Navarro F., Codis S., Hernquist L.,
Bernardeau F., 1994a, ApJ, 427, 51
Genel S., 2019, MNRAS, 484, 269
Bernardeau F., 1994b, ApJ, 433, 1
Lesgourgues J., Pastor S., 2006, Phys. Rep., 429, 307
Bernardeau F., 1995, A&A, 301, 309
Levi M., et al., 2013, preprint, (arXiv:1308.0847)
Bernardeau F., 1996, A&A, 312, 11
Lewis A., Challinor A., Lasenby A., 2000, ApJ, 538, 473
Bernardeau F., Reimberg P., 2016, Phys. Rev. D, 94, 063520
Lidz A., Heitmann K., Hui L., Habib S., Rauch M., Sargent W. L. W., 2006,
Bernardeau F., Valageas P., 2000, A&A, 364, 1
ApJ, 638, 27
Bernardeau F., Colombi S., Gaztañaga E., Scoccimarro R., 2002,
Phys. Rep., 367, 1 Liu J., Madhavacheril M. S., 2019, Phys. Rev. D, 99, 083508
Bernardeau F., Pichon C., Codis S., 2014, Phys. Rev. D, 90, 103519 LoVerde M., 2014, Phys. Rev. D, 90, 083518
Bernardeau F., Codis S., Pichon C., 2015, MNRAS, 449, L105 Manera M., Gaztañaga E., 2011, MNRAS, 415, 383
Bird S., Viel M., Haehnelt M. G., 2012, MNRAS, 420, 2551 Massara E., Villaescusa-Navarro F., Viel M., Sutter P. M., 2015, J. Cosmol-
Blas D., Lesgourgues J., Tram T., 2011, Journal of Cosmology and As- ogy Astropart. Phys., 11, 018
troparticle Physics, 2011, 034034 Matsubara T., 1995, Progress of Theoretical Physics, 94, 1151
Brouwer M. M., et al., 2018, MNRAS, 481, 5189 Mead A. J., 2017, MNRAS, 464, 1282
Carrasco J. J. M., Hertzberg M. P., Senatore L., 2012, preprint, Mota D. F., van de Bruck C., 2004, Astronomy & Astrophysics, 421, 7181
(arXiv:1206.2926) Munshi D., Coles P., Viel M., 2012, MNRAS, 427, 2359
Carron J., Szapudi I., 2014, MNRAS, 439, L11 Neyrinck M. C., Szapudi I., Rimes C. D., 2006, MNRAS, 370, L66
Cataneo M., Rapetti D., Lombriser L., Li B., 2016, Journal of Cosmology Neyrinck M. C., Szapudi I., Szalay A. S., 2009, ApJ, 698, L90
and Astroparticle Physics, 2016, 024024 Nishimichi T., Bernardeau F., Taruya A., 2016, Physics Letters B, 762, 247
Chan K. C., Moradinezhad Dizgah A., Noreña J., 2018, Phys. Rev. D, 97, Nishimichi T., Bernardeau F., Taruya A., 2017, Phys. Rev. D, 96, 123515
043532 Osato K., Nishimichi T., Bernardeau F., Taruya A., 2019, Phys. Rev. D, 99,
Clerkin L., et al., 2016, Monthly Notices of the Royal Astronomical Soci- 063530
ety, 466, 14441461 Pace F., Waizmann J.-C., Bartelmann M., 2010, Monthly Notices of the
Codis S., Pichon C., Bernardeau F., Uhlemann C., Prunet S., 2016a, MN- Royal Astronomical Society, 406, 1865
RAS, 460, 1549 Pajer E., van der Woude D., 2018, J. Cosmology Astropart. Phys., 2018,
Codis S., Bernardeau F., Pichon C., 2016b, MNRAS, 460, 1598 039
Coles P., Jones B., 1991, MNRAS, 248, 1 Patton K., Blazek J., Honscheid K., Huff E., Melchior P., Ross A. J.,
Colombi S., 1994, ApJ, 435, 536 Suchyta E., 2017, MNRAS, 472, 439

MNRAS 000, 000–000 (0000)


20 C. Uhlemann et al.
Peacock J. A., Smith R. E., 2014, HALOFIT: Nonlinear distribution of cos- 0.006
mological mass and galaxies (ascl:1402.032)
0.004

(R)-1
Peebles P. J. E., 1980, The large-scale structure of the universe. Research
supported by the National Science Foundation. Princeton, N.J., Prince- 0.002

(σ2L )CAMB
ton University Press, 1980. 435 p.

(σ2L )EH
0.000
Δns d
Petri A., Liu J., Haiman Z., May M., Hui L., Kratochvil J. M., 2015, Physi-
ΔΩm
cal Review D, 91 - 0.002
Planck Collaboration et al., 2018, preprint, (arXiv:1807.06209) ΔΩb Δh
Reimberg P., Bernardeau F., 2018, Phys. Rev. D, 97, 023524 - 0.004
Repp A., Szapudi I., 2017, MNRAS, 464, L21
5 8 10 15 20 25 30
0.000

(R)-1
Repp A., Szapudi I., 2018, MNRAS, 473, 3598
Rimes C. D., Hamilton A. J. S., 2005, MNRAS, 360, L82 R [Mpc/h]
Sahlén M., 2019, Phys. Rev. D, 99, 063525
- 0.001

(σ2L+ / σ2L- )CAMB


(σ2L+ / σ2L- )EH
Saito S., Takada M., Taruya A., 2009, Physical Review D, 80
Salvador A. I., et al., 2018, Monthly Notices of the Royal Astronomical
Society, 482, 14351451 - 0.002 ΔΩb Δns
Schäfer B. M., Koyama K., 2008, MNRAS, 385, 411
ΔΩm Δh
Scoccimarro R., Zaldarriaga M., Hui L., 1999, ApJ, 527, 1 - 0.003
Seo H.-J., Sato M., Dodelson S., Jain B., Takada M., 2011, ApJ, 729, L11
Shin J., Kim J., Pichon C., Jeong D., Park C., 2017, ApJ, 843, 73 5 8 10 15 20 25 30
Sigad Y., Branchini E., Dekel A., 2000, ApJ, 540, 62
R [Mpc/h]
Spergel D., et al., 2015, arXiv e-prints, p. arXiv:1503.03757
Szapudi I., Colombi S., 1996, ApJ, 470, 131
Szapudi I., Pan J., 2004, ApJ, 602, 26 2 (R) computed from
Figure A1. Comparison between the linear variance σL
Szapudi I., Colombi S., Bernardeau F., 1999, Monthly Notices of the Royal
the Eisenstein-Hu approximation vs. CAMB for the fiducial cosmology
Astronomical Society, 310, 428
(black) and the derivative cosmologies with varying Ωm (red), Ωb (blue),
Takada M., Hu W., 2013, Phys. Rev. D, 87, 123504
ns (yellow) and h (green) with positive sign (solid lines) and negative sign
Takada M., et al., 2014, PASJ, 66, R1
(dashed lines) as indicated in Table 1 with fixed σ8 . The lower panel shows
Takahashi R., Oguri M., Sato M., Hamana T., 2011, The Astrophysical Jour-
the fractional difference of the ratio of the linear variance between the posi-
nal, 742, 15
tive and negative sign derivative cosmologies. We find that derivatives com-
Taruya A., Bernardeau F., Nishimichi T., Codis S., 2012, Phys. Rev. D, 86,
puted using the Eisenstein-Hu have sub-percent accuracy with respect to
103528
CAMB.
The Dark Energy Survey Collaboration 2005, arXiv e-prints, pp astro–
ph/0510346
Uhlemann C., Codis S., Pichon C., Bernardeau F., Reimberg P., 2016, MN- variance according to equation (8) can be modelled as (Glazebrook
RAS, 460, 1529 & Blake 2005),
Uhlemann C., Codis S., Kim J., Pichon C., Bernardeau F., Pogosyan D.,
5Ωm H(a) a da0 H03
Z
Park C., L’Huillier B., 2017, MNRAS, 466, 2067
D(z) = , (A1)
Uhlemann C., et al., 2018a, MNRAS, 473, 5098 2 H0 0 a03 H 3 (a0 )
Uhlemann C., Pajer E., Pichon C., Nishimichi T., Codis S., Bernardeau F., s  Z z 
2018b, MNRAS, 474, 2853 H(a) Ωm 1 + w(z 0 ) 0
= + Ω Λ exp 3 dz , (A2)
Uhlemann C., Pichon C., Codis S., L’Huillier B., Kim J., Bernardeau F., H0 a3 0 1 + z0
Park C., Prunet S., 2018c, MNRAS, 477, 2772
Valageas P., 2000, A&A, 356, 771 with the dark matter density, Ωm , the dark energy density, ΩΛ , the
Valageas P., 2002, A&A, 382, 412 Hubble constant at zero redshift, H0 , the expansion factor a ≡
Villaescusa-Navarro F., Miralda-Escudé J., Peña-Garay C., Quilis V., 2011, 1/(1 + z) and the dark energy equation of state w(z). In Figure 19
J. Cosmology Astropart. Phys., 2011, 027 in Section 6 we compare the impact of a changed equation of state
Villaescusa-Navarro F., Bird S., Peña-Garay C., Viel M., 2013, J. Cosmol- parameter w = w0 = −1 ± 0.01 on the growth with the change of
ogy Astropart. Phys., 2013, 019 changing Ωm ± 0.01.
Villaescusa-Navarro F., Banerjee A., Dalal N., Castorina E., Scoccimarro
R., Angulo R., Spergel D. N., 2018, ApJ, 861, 53
Villaescusa-Navarro F., et al., 2019, arXiv e-prints, p. arXiv:1909.05273 A2 Approximating the nonlinear variance
Wolk M., Carron J., Szapudi I., 2015, MNRAS, 454, 560
Xavier H. S., Abdalla F. B., Joachimi B., 2016, MNRAS, 459, 3693 In Figure A2 we compare the measured nonlinear power spectrum
Yang A., Saslaw W. C., 2011, ApJ, 729, 123 from the Quijote simulations to standard fitting function (respresso
de Jong J. T. A., Verdoes Kleijn G. A., Kuijken K. H., Valentijn E. A., 2013, and halofit) as well as perturbative methods at 2-loop order (SPT
Experimental Astronomy, 35, 25 and RegPT) at redshifts z = 0 and z = 0.5.
In Figure A3 we show the integrand that enters the computa-
tion of the nonlinear variance for different radii.
APPENDIX A: APPROXIMATING THE VARIANCE
A1 Approximating the cosmology-dependent linear variance
In Figure A1 we show how well the Eisenstein-Hu transfer function
(Eisenstein & Hu 1998) captures the cosmology dependence of the
linear variance.
For a cosmology with dark energy beyond a cosmological con-
stant, the growth function describing the time-dependence of the

MNRAS 000, 000–000 (0000)


Counts-in-cells Cosmology 21

Papp (z,k)/Psim, NL (z,k)-1 P+ (ρ)/P- (ρ)-1


0.15
respresso 2RegPT L 0.04
0.10 ΔΩm Δns

z=0, R=10 Mpc/h


halo t 2SPT
0.05 0.02 ΔΩb Δh
z=0

0.00 0.00
- 0.05 - 0.02
- 0.10
- 0.04
- 0.15
0.10 0.01 0.025 0.05 0.1 0.2 0.3 0.05 - 3 -2 -1 0 1 2 3
respresso 2RegPT L

z=0, R=20 Mpc/h


0.05 halofit 2SPT 0.00
z=0.5

0.00 - 0.05

ΔΩm Δns
- 0.05 - 0.10
ΔΩb Δh
- 0.10 - 0.15
0.01 0.025 0.05 0.1 0.2 0.3 -3 -2 -1 0 1 2 3
k [h/Mpc] (ln ρ-<ln ρ>fid )/σln ρ,fid

Figure B1. The measured ratios of the PDFs for the derivative simulations
Figure A2. Comparison between the measured nonlinear power spectrum at z = 0 with radius R = 10, 20 Mpc/h (top and bottom, points with
averaged over 150000 realisations of the fiducial simulation at redshift z = error bars) compared to the predictiona using the measured nonlinear vari-
0 (upper panel) and z = 0.5 (lower panel) with predictions using respresso ance at the reference scale as input parameter (solid lines) or predicting
(brown), halofit (red), 2-loop perturbation theory in RegPT (green dashed) the nonlinear variance from the measured nonlinear variance of the fiducial
or SPT (blue dashed) and linear theory (black dotted). The thin gray lines model (dashed lines) for changes in Ωm (red), Ωb (blue) ns (yellow) and
indicate 2% error (solid) and 1% error (dashed). h (green).

NL PMν (ρ)/PfidZA (ρ)-1, z=0, R=10 Mpc/h


50 0.2
R [Mpc/h] halofit
40 10 0.1
2SPT
P(k)k2 W 2 (Rk)

15 0.0
30
20 L - 0.1
20 Mν=0.1eV
- 0.2
Mν=0.2eV
10 - 0.3
Mν=0.4eV
- 0.4
0 -3 -2 -1 0 1 2 3
0.01 0.1 0.2 0.3 0.5 1 (ln ρ-<ln ρ>fidZA )/σln ρ, fidZA
k [h/Mpc]

Figure B2. Residuals between the total matter density PDFs with massive
Figure A3. The integrand for the nonlinear variance from equation (12) at neutrinos for a total mass Mν = 0.1, 0.2, 0.4 eV (blue, green, red) at
scales R = 10, 15, 20 Mpc/h at redshift z = 0 using different expressions redshift z = 0 and radius R = 10 Mpc/h.
for the power spectrum as indicated in the legend.

APPENDIX B: RESIDUALS OF MATTER DENSITY PDFS


In Figures B1 and B2, we show the ratios of matter PDFs when
varying ΛCDM parameters and the total neutrino mass.

MNRAS 000, 000–000 (0000)

You might also like