You are on page 1of 29

Version October 17, 2022

Typeset using LATEX style openjournal v. 09/06/15

IT TAKES TWO TO KNOW ONE: COMPUTING ACCURATE ONE-POINT PDF COVARIANCES


FROM EFFECTIVE TWO-POINT PDF MODELS
Cora Uhlemann1,? , Oliver Friedrich2 , Aoife Boyle3 , Alex Gough1 , Alexandre Barthelemy2 , Francis
Bernardeau4,5 , and Sandrine Codis3
1 School of Mathematics, Statistics and Physics, Newcastle University, Herschel Building, NE1 7RU Newcastle-upon-Tyne, U.K.
2 Universitäts-Sternwarte, Fakultät für Physik, Ludwig-Maximilians Universität München, Scheinerstr. 1, 81679 München, Germany
3 AIM, CEA, CNRS, Université Paris-Saclay, Université Paris Diderot, Sorbonne Paris Cité, 91191 Gif-sur-Yvette, France
4 Université Paris-Saclay, CNRS, CEA, Institut de physique théorique, 91191, Gif-sur-Yvette, France and
5 CNRS & Sorbonne Université, UMR 7095, Institut d’Astrophysique de Paris, 75014, Paris, France

Version October 17, 2022


arXiv:2210.07819v1 [astro-ph.CO] 14 Oct 2022

Abstract
One-point probability distribution functions (PDFs) of the cosmic matter density are powerful
cosmological probes that extract non-Gaussian properties of the matter distribution and complement
two-point statistics. Computing the covariance of one-point PDFs is key for building a robust galaxy
survey analysis for upcoming surveys like Euclid and the Rubin Observatory LSST and requires good
models for the two-point PDFs characterising spatial correlations. In this work, we obtain accurate
PDF covariances using effective shifted lognormal two-point PDF models for the mildly non-Gaussian
weak lensing convergence and validate our predictions against large sets of Gaussian and non-Gaussian
maps. We show how the dominant effects in the covariance matrix capturing super-sample covariance
arise from a large-separation expansion of the two-point PDF and discuss differences between the
covariances obtained from small patches and full sky maps. Finally, we describe how our formalism
can be extended to characterise the PDF covariance for 3D-dimensional spectroscopic fields using the
3D matter PDF as an example. We describe how covariances from simulated boxes with fixed overall
density can be supplemented with the missing super-sample covariance effect by relying on theoretical
predictions validated against separate-universe style simulations.

1. INTRODUCTION skewness at one given scale (c.f. Section 5.2 and Figure
The recent rise of non-Gaussian statistics for galaxy 4 of Friedrich et al. 2020). This however also highlights
clustering and weak lensing has highlighted the need for a limitation of those simulations: they potentially fail
accurate models of the likelihood function and, in par- to realistically capture the hierarchy of moments beyond
ticular, the covariance matrices of such statistics. Aside the skewness (for an extension see Baratta et al. 2020) as
from the challenge to model non-Gaussian statistics and well as the full shape of higher order N-point functions.
capture their response to changing cosmological param- With the large data vectors that future multi-probe
eters in the nonlinear regime, getting accurate covari- analyses of upcoming surveys like Euclid (Laureijs et al.
ance matrices can prove difficult and require thousands 2011; Amendola et al. 2018) and the Rubin Observatory
of numerical simulations to ensure convergence (Taylor & LSST (Ivezić et al. 2019) are aiming for, estimated co-
Joachimi 2014; Colavincenzo et al. 2019). For statistics variances - whether from log-normal simulations or other
relying on higher-order correlation functions involving N mock data - can lead to a huge degradation of cosmolog-
points, modelling the covariance generally requires as- ical constraining power (Dodelson & Schneider 2013a;
sumptions for the shape of 2N -point correlations. Even Friedrich & Eifler 2017; Percival et al. 2022). This makes
for Gaussian fields, this task has only been achieved re- it desirable to have an analytic understanding of mea-
cently at all orders N (Hou et al. 2021). To go beyond the surement uncertainties. One-point statistics such as the
assumption of Gaussianity, log-normal models of the cos- PDF of the weak lensing convergence have the advan-
mic density field have become a tool of choice in higher- tage that both their dependence on cosmological param-
order analyses such as studies of the bispectrum (Mar- eters and their covariance, being given in terms of the
tin et al. 2012; Halder et al. 2021), moments of cosmic two-point PDF, can be modelled theoretically (Codis
random fields (Gatti et al. 2019), density split statis- et al. 2016b; Uhlemann et al. 2017). In this work, we
tics (Friedrich et al. 2018; Gruen et al. 2018), the full will focus on the weak lensing convergence smoothed on
probability distribution function (PDF) of cosmic ran- mildly nonlinear scales (of order 100 at source redshifts of
dom fields (Uhlemann et al. 2020; Friedrich et al. 2020; z ' 1 − 2), where the bulk of the PDF and its cosmology
Boyle et al. 2021), and even map-based inference ap- dependence can be predicted at the percent level using
proaches (Sarma Boruah et al. 2022). The reason why large deviation statistics (Bernardeau & Valageas 2000;
(shifted) log-normal simulations (Xavier et al. 2016) have Bernardeau & Reimberg 2016; Barthelemy et al. 2020a;
become so popular is that they obey a similar hierarchy Boyle et al. 2021) and its covariance is well-approximated
between variance and skewness as the physical cosmic by a shifted lognormal distribution (Boyle et al. 2021).
density field. In particular, they can be tuned to match In the opposite regime for small smoothing scales, the
both a desired power spectrum at all scales and a desired high-convergence tail of the weak lensing convergence
PDF and its covariance have been modelled using a halo-
? cora.uhlemann@newcastle.ac.uk
model formalism in Thiele et al. (2020), finding qualita-
2

tive agreement with simulations but significant quanti- ∑ PDF covariance, full sky, O
tative discrepancies attributed to strong sensitivity to Takahashi simulation
small-scale effects. -0.02 0.004
For galaxy clustering, the error on and correlation
among the factorial moments of galaxy counts in cells
Fk = hN (N − 1) . . . (N − k)i have been computed for -0.01 0.002
hierarchical models in Szapudi & Colombi (1996) and
Szapudi et al. (1999). Those studies included cosmic
errors due to a finite volume and discreteness errors 0.0

∑2
0.000
due to a finite number of tracers relevant for cluster-
ing but not the weak lensing convergence, which is in-
stead affected by shape noise due to shear measure- 0.01
°0.002
ments. In Repp & Szapudi (2021a), the variance and
covariance among galaxy counts in cells was obtained for
0.02
the case of a constant separation-independent cell cor- °0.004
relation, which has previously been considered in Codis lognormal theory
et al. (2016b). As we explain here, and derive in Ap- -0.02 -0.01 0.0 0.01 0.02
pendix B, the large-separation expansion in terms of ∑1
bias functions adopted in Codis et al. (2016b); Uhlemann
et al. (2017); Repp & Szapudi (2021a); Repp & Szapudi Fig. 1.— PDF covariance matrix for circular apertures of radius
θs = 7.50 at source redshift zs = 2 comparing the measurement
(2021b) can be generalised to the physically more re- from 108 full sky Takahashi lensing maps (upper triangle) to the
alistic case of a separation-dependent correlation. The shifted lognormal prediction (lower triangle).
large-separation two-point PDF and the resulting super-
sample covariance effect can also be related to the idea of
the position-dependent matter PDF studied in Jamieson
zs = 2, A = 15000 deg2
& Loverde (2020). Recent work of Bernardeau (2022) Lognormal model
determined covariances of density PDFs in hierarchical Takahashi
models, which can be used to understand the dominant FLASK Takahashi
terms in the covariance matrix of cosmological density
fields in the strongly non-Gaussian regime. In this work,
we build on this foundation but focus on the weakly non-
Gaussian weak lensing PDF for which hierarchical mod-
els are less suitable and better results can be achieved
using shifted lognormal models.
86
0.

STRUCTURE
84

In Section 2, we discuss the general covariance mod-


0.

elling for the two-dimensional case of the weak lensing (or


σ8
82

photometric galaxy clustering) PDF based on a model for


0.

the joint two-point PDF. In Section 3, we use a large-


80

separation expansion for this two-point PDF and link it


0.

to dominant terms in an eigendecomposition of the co-


78

variance matrix. In Section 4, we validate our covariance


0.

20

22

24

26

78

80

82

84

86
models using several quantitative tests. In Section 5, we
0.

0.

0.

0.

0.

0.

0.

0.

present the basics for generalising our findings to the case Ωcdm σ8 0.
of the three-dimensional matter (or spectroscopic galaxy
clustering) PDF. We conclude in Section 6 and provide Fig. 2.— Fisher forecasts validating the predicted shifted lognor-
an outlook for how our results can be applied to related mal PDF covariance (purple) against measured covariances from
108 realistic full sky simulated maps (red) and 500 associated mock
one-point observables, include tomography and informa- FLASK maps (green). The parameter errors obtained from the
tion from two-point statistics. raw measured covariances have additionally been corrected by the
Hartlap factor (72). To highlight the variation that can result from
EXECUTIVE SUMMARY small samples, we also show contours for two subsamples of 100 of
the 500 mock FLASK maps (green dotted).
We develop a theoretical model for the covariance of
measurements of the one-point weak lensing convergence
PDF in finite bins. Our model is based on an integra-
tion over the two-point PDF (6) at different separations
and we rely on a joint shifted lognormal model for that inverse of the survey area. We validate our predictions
two-point PDF (28) tuned to match the PDF skewness against large sets of simulated convergence maps both
and shown to capture the density-dependence of the two- from N-body simulations (Takahashi et al. 2017) and us-
point correlation (12) in simulations, as illustrated in ing the FLASK tool for generating log-normal random
Figure 8. We establish expressions for the finite sam- fields (Xavier et al. 2016). To verify the quality of our
pling (11) and super-sample covariance (13) contribu- covariance model we employ a range of tests including
tions to the PDF covariance and their scaling with the visual comparisons, an eigendecomposition of the covari-
3

ance, a band-decomposition of the precision matrix, χ2 et al. 2016) to match the convergence PDF and ob-
tests and Fisher forecasts, with Figures 1 and 2 sum- tain precise covariance estimates in good agreement with
marising two main results. In Figure 1 we show the maps from N-body simulations, of which often only a
measured covariance for the central region of the binned very limited number is available. Following this work, we
κ-PDF for a smoothing radius of θs = 7.50 at source red- will study the covariance of the central region of the one-
shift zs = 2 from Nsim = 108 simulated full sky κ maps point PDF of the weak lensing convergence smoothed
from Takahashi et al. (2017) (upper triangle) along with with a top-hat filter of radius θs = 7.30 at a fixed source
the shifted lognormal prediction (lower triangle, using redshift zs = 2 using 26 linearly spaced bins in the range
equations (6) and (28)). For the predictions, we used the −0.024 < κ < 0.026. In this Section we will use both
cosmological parameters of the simulation along with the Gaussian and shifted lognormal fields generated with
shift parameter s ' 0.085 obtained from the second and FLASK to test our theoretical framework for predict-
third moments measured from the PDF. In Figure 8 in ing the weak lensing PDF covariance based on different
the main text we show that the environment-dependent models for the two-point PDF of the weak lensing con-
clustering that enters the eigendecomposition of the co- vergence in cells of a given separation. While we focus
variance is well captured by the shifted lognormal model. on the two-dimensional case of the weak lensing PDF
Aside from the good visual agreement, we quantify here, the formalism can be adapted to describe the three-
their consistency using two tests. First, we test the com- dimensional clustering PDF as we lay out in Section 5.
patibility of the diagonal elements of the two covariances
shown in Figure 1 with a χ2 test, taking into account 2.1. Covariance prediction from joint 2-point PDF
their uncertainty and correlation among each other1 (co-
variance of the covariance). We use the covariance be- Before calculating the covariance matrix of a PDF mea-
tween the numerically estimated diagonal covariance el- surement, let us first specify how this measurement is
conducted. Let us assume that we have observed the
ements Ĉii and Ĉjj as given in equation (36) following
gravitational lensing convergence field on a part of the
Taylor et al. (2013), where Nsim = 108 is the number of
sky with the area Asurvey . Now let us smooth this con-
simulations used to estimate and the covariance elements
vergence field with a circular aperture of some angular
Cij are evaluated with our theoretical model. Using the
radius θs (potentially cutting away parts of the edges of
above expression to compute a χ2 between the diago- the survey, so that each aperture falls fully inside the
nal of the estimated and predicted covariance we find survey area), and let us pixelise the resulting map. Each
χ2 = 36.8 for a number of 26 bins. This corresponds to pixel of that map contains the value of the average con-
a marginal ∼ 1.50σ deviation between the two matrix vergence within a radius θs around the pixel center.
diagonals. We would now like to estimate the probability density
Finally, we assess the quality of the off-diagonal ele- of a certain value κ for the smoothed convergence field.
ments of our covariance model by performing a Fisher In practice one will need to compute this PDF within a
forecast. Figure 2 compares Fisher contours on the pa- set of finite bins, say around a set of central values κi
rameters σ8 and Ωcdm derived from a number of different and with constant bin width ∆κ, which results in the
covariance matrices: our covariance model, the covari- bins bini = [κi − ∆κ/2, κi + ∆κ/2] . The PDF at each
ance estimated from the N-body simulations of Taka- κi can then be approximated as
hashi et al. (2017) and the covariance estimated from
log-normal simulations. The differences in the width and P (κ ∈ [κi − ∆κ/2, κi + ∆κ/2])
alignment of the constraints obtained from these matri- P(κi ) ≈ , (1)
∆κ
ces is negligible compared to the overall parameter uncer-
tainties. Hence, our covariance model is sufficiently ac- where P(κ) is the PDF (obtained in the limit ∆κ → 0)
curate for application to observational data. We demon- and P (A) is the probability of statement A. With our
strate the flexibility of our covariance model by adapting pixelised map we can estimate the right-hand side of this
it to predict super sample covariance effects for the 3- to obtain the measurement
dimensional matter PDF paving the way to spectroscopic
galaxy clustering. #{pix with |κ − κi | ≤ ∆κ/2}
P̂(κi ) ≡ . (2)
#{pix}∆κ
2. WEAK LENSING PDF COVARIANCE
Following Bernardeau (2022), we introduce the weight
The weak lensing convergence PDF is a well-suited function
complementary summary statistic that can extract non- 
Gaussian information from cosmic shear observations, 1 if κ ∈ [κi − ∆κ/2, κi + ∆κ/2]
thus complementing two-point statistics. In this work, χi (κ) = , (3)
0 else
we focus on mildly nonlinear scales, where weak lens-
ing convergence PDF and its response to fundamental such that the estimator P̂(κi ) is given by
cosmological parameters can be predicted at percent-
level accuracy from large-deviation statistics as shown
P
P χi (κP ) 1 χi (κ(θ))
Z
in Barthelemy et al. (2020a); Boyle et al. (2021). In the P̂(κi ) = ≈ d2 θ , (4)
same work it was shown how one can use shifted log- NP ∆κ Asurvey ∆κ
survey
normal convergence fields generated by FLASK (Xavier
where P labels the NP = Asurvey /Apix pixels of area Apix ,
1 Note that typically only the diagonal elements of estimated θ are angular coordinates on the sky, and κ(θ) is the
covariances are well described with multi-variate Gaussian noise. smoothed, continuous convergence field at that location.
4

7 θ/Θside , is
continuous square patch
discrete, non-overlapping grid
6
continuous full sky
Pd,square (θ̂) = (8)
(
5 2θ̂[π + (θ̂ − 4)
pθ̂] θ̂ ∈ [0, 1[ ,
2 −1

4
2θ̂[π − 2 + 4 θ̂2 − 1 − θ̂ − 4sec (θ̂)] θ̂ ∈ [1, 2] .
Pd (r)

3
This distribution is illustrated as black line in Figure 3,
with a maximum at roughly half the patch size where
2 θ̂max ' 0.479 and a linear growth for small distances,
Pd,square (θ̂) ' 2π θ̂.
1
Before looking at PDF covariance matrix examples in
0
Section 2.2 and introducing models for the two-point
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 PDF in Section 2.3, let us discuss the main universal
r = θ/Θside
effects affecting the PDF covariance: i) the shot-noise
effect related to the finite sampling of cells and ii) the
Fig. 3.— Comparison of the continuous distribution of rela-
tive distances in a square Cartesian survey patch of side length super-sample covariance effect related to a finite survey
Θside = 5deg= 3000 (black) and the discrete version for a coarse area.
grid corresponding to non-overlapping circular cells of radius 7.30
(blue). We also show the distribution of relative distances on the 2.1.1. Finite sampling (shot noise) effects
full sky with maximum Θside = π = 180deg (red).
Previously we discussed the covariance of the PDF in
the continuum limit. In the more realistic discrete case,
we estimate the PDF in a bin centred at κi with width
The second moment of our above estimator is given by ∆i (in our case linearly spaced bins ∆i = ∆), so κ ∈
[κi − ∆i /2, κi + ∆i /2] using a finite sample of NT cells
hP̂(κi )P̂(κj )i that are placed on a regular grid in the survey area. The
histogram-value of the PDF in a bin is obtained from
1 hχi (κ(θ1 ))χj (κ(θ2 ))i
Z Z
≈ 2 d2 θ1 d2 θ2 averaging the continuous PDF P(κ) over that bin
Asurv. (∆κ)2 Z ∆i /2
surv. surv. P(κi + δκi )
1
Z Z Pi = dδκi ' P(κi ) , (9)
≈ 2 2
d θ1 d2 θ2 P(κi , κj ; |θ1 − θ2 |) , (5) −∆i /2 ∆i
Asurv.
surv. surv. and for small bin sizes close to the value of the contin-
where P(κi , κj ; θ) is the joint PDF of two points of the uous PDF at the bin center κ̄i . To compute the co-
smoothed convergence field at distance θ. This equation variance for the discrete case, we can discretise equa-
can be simplified to tion (6) by averaging the continuous joint PDF over
Z the κ-bins and replacing the integration over the dis-
tance distribution by a sum-based average over the set of
hP(κi )P(κj )i = dθ Pd (θ)P(κi , κj ; θ) , (6a) Nd = NT2 = θ∈Θ Nd (θ) distances Θ of all grid cells with
P
multiplicities Nd (θ) that are part of the survey patch
with Pd (θ) indicating the distribution of angular dis-
tances in a given survey area. The covariance of the X Nd (θ)Z ∆i /2 Z ∆j /2 P(κi + δκi , κj + δκj ; θ)
PDF estimator can then be computed from this moment hPi Pj i = dδκi dδκj
Nd −∆i /2 −∆j /2 ∆i ∆ j
via θ∈Θ
X Nd (θ)
cov(P(κi ), P(κj )) = hP̂(κi )P̂(κj )i − P̄(κi )P̄(κj ) , (6b) = Pij (θ) , (10)
Nd
θ∈Θ
where P̄ ≡ hP̂i ≈ P . We see that covariance predictions
rely on two key ingredients: the distance distribution of where for small enough bins Pij (θ) ' P(κi , κj ; θ). In
angular separations and the two-point PDF of the lensing Figure 3 we show a comparison between the continuous
convergence at those separations, which will be discussed distribution of distances in a square patch and the dis-
in more detail in Section 2.3. crete case for a coarser grid of non-overlapping cells. We
Distance distributions. The relevant distance distribu- now split this sum over cell separations showing that the
tions for weak lensing describe a two-dimensional field. contribution from zero separation leads to a finite sam-
For the full sky, the uniform distribution of angular sep- pling or shot-noise term, and the remainder resembles the
arations (measured in radians) reads super-sample covariance contribution that dominates for
1 the case of well-separated cells. Since the one-point PDF
Pd,full sky (θ) =
sin θ , (7) can be obtainedRfrom the two-point PDF by marginali-
2 sation, P(κi ) = dκj P(κi , κj ; θ), for zero separation we
with a maximum at θmax = π/2. This distribution when require that P(κi , κj ; θ = 0) = P(κi )δD (κi − κj ). In the
written in terms of distances relative to the maximal dis- discrete case the joint two-point PDF at zero separation
tance Θside = π is illustrated as red line in Figure 3. is given byP Pij (θ = 0) = P̄i δij /∆i , required by the rela-
For a square survey patch, the distribution of angular tion P̄i = j ∆j Pij (θ = 0). The zero separation mul-
distances relative to the size of the survey patch, θ̂ = tiplicity equals the cell number Nd (θ = 0) = NT , such
5

that the zero-separation term in equation (10) produces cells are at sufficiently wide separations θ & 2θs , the co-
a finite sampling (or shot-noise) covariance contribution2 variance is obtained from integrating the large-separation
limit of the two-point PDF (12) to obtain a super-sample
NT P̄i covariance (SSC) covariance term
covFS (Pi , Pj ) = Pij (θ = 0) = δij . (11)
Nd ∆ i NT
covSSC (P(κi ), P(κj )) = ξ¯ (b1 P)(κi )(b1 P)(κj ) , (13a)
Unsurprisingly, this finite sampling term is inversely pro-
portional to the number of cells. If the cells are placed where we defined the mean correlation
at a fixed separation, an increase of the area corresponds Z θmax
to a suppression of the shot-noise covariance term as ξ¯ = dθ Pd (θ)ξ(θ) , (13b)
2
σFS (P) ∝ NT−1 ∝ Θ−2 −1
side ∝ Asurvey , which is inversely
proportional to the area.
While in principle the finite sampling term can be ren- with Pd (θ) being the probability distribution of angular
dered arbitrarily small by using an increasing number distances and θmax being the maximum distance set by
of cells, this comes at the price of heavy cell overlaps the size of the survey patch, respectively. This mean
at small separations. In this regime, a small-scale ex- correlation agrees with the variance of the mean of κ
1/2 p
2
measured across different patches, so ξ¯ = σκ̄2 .
pansion with small ∆ξ (θ) = σ − ξ(θ)  ξ(θ)  1 For a square patch of side length√Θside , the maxi-
can be used to capture effects of heavy cell overlaps as mal distance in the patch is θmax ' 2Θside where the
discussed in Bernardeau (2022). In the closed-form two-
normalised distance distribution Pd (θ̂) is given in Equa-
point PDF models discussed in Section 2.3 those overlap
tion (8), by conservation of probability we have
effects will be automatically included.
 
2.1.2. Super-sample covariance effects dθ̂ θ 1 θ̂1 −2
Pd (θ) = Pd (θ̂) = Pd θ̂ = ∝ Θside .
Beyond assuming specific (log-)normal models for the dθ Θside Θside
bivariate PDF, one can obtain accurate estimates of the (14)
two-point PDF in the large separation regime illustrated The main contribution in the mean correlation integral
in Figure 4. A large-scale expansion with small corre- comes from small angles θ, where the correlation func-
lation ξ(θ)  ∆ξ (θ)  1 is suitable for well-separated tion is largest and Pd (θ̂) ∝ θ̂. Hence, the SSC covari-
cells where the joint PDF at leading order can be ap- ance term from equation (13) scales as covSSC (Pi , Pj ) ∝
proximated as Codis et al. (2016b) Θ−2 −1
side ∝ Asurvey , which is inversely proportional to the
P(κi , κj ; θ & 2θs ) ' P(κi )P(κj )[1 + ξ(θ)b1 (κi )b1 (κj )] , survey area.
(12) In the discrete case, the covariance can be expressed
as a sum of the previously discussed finite sampling
with some bias function b1 , which describes how the two- term (11) and a super-sample covariance like term
point correlation at a given separation is modulated by
the local weak-lensing convergences. This functional has cov(Pi , Pj ) = hPj Pj i − P̄i P̄j (15a)
been found to be very robust and the previous result = covFS (Pi , Pj ) (15b)
extends what is expected in cosmological density fields
(Bernardeau 1996). When cells are at sufficiently large 1 X h Nd i
+ Pij (θ) − P̄i P̄j .
separations and the initial conditions are Gaussian, this Nd NT (NT − 1)
θ∈Θ\{0}
modulation is independent of separation. We will intro-
| {z }
'1
duce a discretised estimator in equation (16c) and derive
the functional form of those bias functions for specific If cells at separation θ were completely independent, such
forms of the two-point PDF discussed in Section 2.3 by that Pij (θ) = P̄i P̄j , then only the shot-noise term from
an expansion in the correlation ξ(θ). For the related the finite sampling of cells would remain. The second
case of densities in spheres and cylinders, this bias func- term in equation (15) includes significant contributions
tion can be accurately predicted by spherical or cylindri- from cells at wide separation θ & 2θs , where we can use
cal collapse, respectively. The large-deviation statistics the large-separation PDF expansion (12). This leads to a
techniques used to predict the one-point weak lensing super-sample covariance term analogous to equation (13)
PDF (Barthelemy et al. 2020a; Boyle et al. 2021) based
on cylindrical collapse for thin redshift slices could be ¯ 1 P)i (b1 P)j ,
covSSC (Pi , Pj ) = ξ(b (16a)
extended to derive the functional form of this bias, al-
though this is beyond the scope of this work. This large- where (b1 P)i are obtained from a bin-average of the in-
separation expansion predicts a leading order super- volved functions resembling equation (9) and the mean
sample covariance term and further sub-leading contri- correlation can be obtained as
butions that we will describe in Section 3. When all X Nd (θ)
ξ¯ = ξ(θ) . (16b)
2 This shot-noise term was already described in (Codis et al. Nd
2016b). Even if there was no spatial correlation in the sample, in θ∈Θ\{0}
the Poisson limit the covariance of finding Ni cells in convergence
¯ the value
Using the relationship κ0j |κi κ0 = b1 (κi )/ξ,


bin covSN (Ni , Nj ) = N̄i δij , where N̄i is the mean number of cells
expected to be in the convergence bin centred on κ̄i . The number j
of cells, Ni in a given convergence bin of width ∆i centered on κi is of the bias function in the bin κi can be estimated by
related to the histogram-value of the binned PDF from equation (9) comparing the mean of the convergence at separation
such that Ni = NT ∆i Pi and similarly N̄i = NT ∆i P̄i . θ = 2θs , κ0j , from this given convergence κi to the average
6

PDF covariance comparison FLASK vs Gaussian


10 3 NO κ FLASK
0.06 NO κ Gauss
NO κ Gauss pred
10 4
0.05 O κ FLASK
O κ Gauss
κ Gauss pred
10 5 0.04

σ 2 (P (κ))
O κ − κ̄ FLASK
()

unsmoothed grid scale 0.03


O κ − κ̄ Gauss
10 6 smoothed s κ − κ̄ Gauss pred
( ) = 2( s) ( s, ) 0.02
2( s)
10 7
= 2 s = 0.42 2 0.01
= 0.5 2 = 12.470
10 8
0.00
100 101 102
[arcmin] −0.04 −0.02 0.00 0.02 0.04
κ
Fig. 4.— Two-point correlation of circular cells of radius θs '
7.30 (solid black) in comparison to the matter correlation (dashed Fig. 5.— Diagonal of the covariance matrix for the non-Gaussian
black), the variance in cells (dashed blue) and the difference ∆ξ (θ) FLASK PDFs (solid lines) and the Gaussian version (dashed lines)
(red line) that vanishes in the limit of small separations. The blue measured from small patches with non-overlapping cells (red), over-
dotted lines indicate the point at which the cell separation equals lapping cells (blue) and overlapping cells with mean-subtraction in
twice the smoothing scale, while the green dotted lines indicate the every patch (purple). The Gaussian case shows a symmetric co-
point where the correlation is half of the variance, corresponding variance contribution as expected. The corresponding correlation
to slightly smaller scales. matrices are shown in Figure 6. Additionally, we show our the-
oretical predictions for the Gaussian case (dotted lines, partially
obscured by the dashed lines). The dotted vertical lines indicate
the range of κ values considered for the covariance matrices and
Fisher forecasts.

correlation ξ¯ using a sum over the 6 nearest neighbour


(NN) pixels at that separation
P P patches is obtained using the formula
1 P P0 ∈NN(2θs ) χi (κP )κP0 Cij = h(Si − S̄i )(Sj − S̄j )i , S̄i = hSi i , (17)
b̂1 (κi ) = ¯ P . (16c)
ξ 6 P χi (κP )
where h·i indicates an ensemble average over different
This definition is the analogue of the estimator for den- map realisations for the full sky or patches, respectively.
sities in spheres and cylinders (Codis et al. 2016b; Uhle- We rescale all covariances to mimic a total Euclid-like
mann et al. 2017, 2018). survey area Asurvey = 15, 000 deg2 using the rescaling
Since the correlation function decreases with increas- factor Apatch /Asurvey for the patches and Afullsky /Asurvey
ing angular separation and the probability distribution for the full-sky maps. We verified the validity of this
of cell distances peaks at roughly half the patch size, rescaling using covariance measurements from full sky
this super-sample covariance term is the most relevant maps in addition to the theoretical arguments in Sec-
for small patches while it is heavily suppressed for full tion 2.1.2. For visualisation purposes, it is often useful to
sky maps. In some cases, like for simulation-based anal- normalise the covariance matrix, with components Cij ,
ysis or estimation of covariances, it is desirable to con- by its diagonal components, which defines the correlation
struct small patches which closely resemble the full sky matrix R such that
behaviour. The main contribution from the SSC covari- Cij
ance term can be reduced by subtracting the mean κ in Rij = p . (18)
Cii Cjj
every small patch, thus reducing the mean correlation ξ¯
and modifying the shape of the bias function b as we will In Figures 5 and 6 we illustrate two effects that im-
discuss in Section 3.2. pact the overall size and structure of the PDF covari-
ance and the derived correlation matrix, finite sampling
2.2. Illustrative covariance measurements (shot noise) and super-sample covariance. In Figure 5 we
To validate our covariance models, we generate a large show the diagonal of the covariance matrix for FLASK
set of Gaussian as well as shifted lognormal (FLASK) patches (solid lines) and Gaussian patches (dashed lines).
κ maps (with the shift parameter tuned to fit the κ- In Figure 6 we show the corresponding correlation ma-
PDF skewness). We start from full-sky convergence maps trices ordered by decreasing covariance, with the upper
(healpix Nside = 4096 corresponding to 0.850 resolution) triangle showing the FLASK patches and the lower tri-
that we divide into 1200 non-overlapping square patches angle showing the Gaussian patches.
with side-length of 5deg. We build a data vector S ~ We consider the three cases for the covariance in small
from the values of the PDF histogram in the central re- patches of side length 5deg:
gion −0.024 < κ < 0.026 measured from the top-hat • raw κ measured from non-overlapping cells (red
smoothed κ map by considering the regular grid of all lines in Figure 5, left panel in Figure 6),
cells (overlapping) or only cells separated by at least
twice the smoothing radius (non-overlapping). The co- • raw κ measured from overlapping cells (blue lines
variance matrix of the binned κ-PDF across all those in Figure 5, middle panel in Figure 6), and
7

κ PDF correlation, (5deg)2 patches, NO κ PDF correlation, (5deg)2 patches, O κ − κ̄ PDF correlation, (5deg)2 patches, O
1.00
FLASK FLASK FLASK
-0.02
0.75

0.50
-0.01
0.25

0.0
κ2

0.00

−0.25
0.01
−0.50

0.02 −0.75
Gauss Gauss Gauss
−1.00
-0.02 -0.01 0.0 0.01 0.02 -0.02 -0.01 0.0 0.01 0.02 -0.02 -0.01 0.0 0.01 0.02
κ1 κ1 κ1

Fig. 6.— Correlation matrix between the bins of the weak lensing convergence κ PDF measured at a smoothing scale ' 7.30 from (5
deg)2 patches of a full sky FLASK map of κ including the nonzero mean in the patches (left and middle panel) with non-overlapping (NO)
cells (left) and overlapping cells (middle) and κ − κ̄ after subtracting the mean with overlapping cells (right panel). The upper triangles
labelled by FLASK show results from shifted-lognormal realisations, while the lower triangles those of Gaussian realisations. The diagonal
of the covariance matrix is shown in Figure 5.

• mean-subtracted κ̃ = κ − κ̄ from overlapping cells and a correlation structure, for example specified by an
(purple lines in Figure 5, right panel in Figure 6) . expansion for large separations or heavy overlaps.
We see that while the lognormal maps have a visible 2.3.1. Bivariate Gaussian
asymmetry in the covariance diagonal compared to the A Gaussian joint PDF is parametrised by the one-point
Gaussian maps, both lead to very similar correlation ma- PDF variance σ 2 and the two-point correlation between
trices. Comparing the red and blue lines shows that given cells ξ = ξ12 (θ) such that
a fixed total area, increasing the number of cells for the
PDF measurement decreases the finite sampling (or shot PG (κ1 , κ2 ; θ) = PG (κ1 , κ2 ; σ 2 , ξ12 (θ)) (19)
noise term) that determines the covariance matrix diag- h 2
ξ12 (θ)
i
onal, as we explained in Section 2.1.1. The suppression exp − σ2d [κ1 2 + κ2 2 ] + d κ1 κ2
of the diagonal for the case of an increased number of = √ ,
cells comes at the price of making cells overlap which 2π d
creates a strong positive correlation band around the di- (20)
agonal, as can be seen from comparing the left and mid- with the determinant d = σ − ξ12 = σ [1 − 4 2 4
dle panel in Figure 6. Comparing the blue and purple (ξ12 /σ 2 )2 ]. One can perform a large-separation expan-
lines in Figure 5 illustrates that subtracting the mean
in small patches significantly reduces the impact of the sion for ξ12 /σ 2  1 valid in the regime of no overlaps as
super-sample covariance term discussed in Section 2.1.2 shown in Figure 4. At leading order, one can approxi-
and causes the 4-tiled pattern in the middle panel of Fig- mate d ' σ 4 and expand the remaining ξ12 (θ) term in
ure 6 to transform to a 9-tiled pattern in the right panel. the exponential to obtain
Additionally, the removal of this dominant term causes PG (κ1 , κ2 ; θ) κ1 κ2
the overlap contribution in the band along the diago- = 1 + ξ12 (θ) 2 2 + O(ξ 2 ) , (21)
PG (κ1 )PG (κ2 ) σ σ
nal to become more apparent. As we will see later, the
mean-subtracted convergence in small patches provides a which is a meaningful expansion in the regime κ . O(σ)
proxy for the convergence on the full sky, as evident from and reproduces the linear ‘Kaiser’ bias function (Kaiser
Figure 9. While most of our discussion is focused on the 1984; Codis et al. 2016a; Uhlemann et al. 2017)
underlying weak lensing convergence field, we discuss the
inclusion of shape noise arising from cosmic shear mea- b1,G (κ) = κ/σ 2 . (22)
surements in Section 2.3.4. We will discuss an expansion up to higher orders in
ξ(θ)/σ 2 in the context of the large-separation expansion
2.3. Specific models for the joint two-point PDF in Section 3.1.2.
As evident from equation (6), the PDF covariance can The impact of mean-subtraction. For small patches,
be computed directly if the two-point PDF P(κi , κj ; θ) the mean convergence κs will significantly fluctuate from
is known a priori for all separations. In the following patch to patch with a sizeable average correlation ξ¯ = σκ2 s
we will discuss three possible cases relying on closed- causing the super-sample covariance. To better emulate
form expressions for that joint PDF, a bivariate Gaussian measurements of modern surveys on a significant frac-
(normal) PDF, a bivariate shifted lognormal PDF and tion of the full sky, subtracting this mean convergence
a bivariate PDF built from a marginal one-point PDF in each patch in the spirit of the peak-background split
8

(Bardeen et al. 1986; Mo & White 1996; Sheth & Tor- zero-crossings each, see equation (56)].
men 1999) can be desirable. The two-point PDF of the
mean-subtracted lensing convergence κ̃i = κi − κ̄ is ob- 2.3.2. Bivariate shifted lognormal
tained from a trivariate Gaussian PDF for κ1 , κ2 and A shifted lognormal joint PDF for the lensing conver-
the background value in the patch κs = κ̄, which is then gence κ is parametrised by the variance σκ2 and two-point
integrated out correlation ξκ = ξ12,κ (θ) along with a shift parameter s
Z related to the skewness, which we will keep fixed through-
PG (κ̃1 , κ̃2 ) = dκs PG (κ1 = κ̃1 + κs , κ2 = κ̃2 + κs , κs ) . out here. We can define a Gaussian-distributed variable
g(κ) with zero-mean
(23a)
The covariance matrix is of the form κ  κ  σ2
g(κ) = ln + 1 − µG = ln +1 + G , (25)
s s 2
 
σ 2 ξ12 ξ1s
Σκ1 ,κ2 ,κs = ξ12 σ 2 ξ2s  , (23b) where µG = −σG 2
/2 is fixed to ensure hκi
ξ1s ξ2s σs2 2
 = 0 and the
Gaussian variance is σG = ln 1 + σκ2 /s2 . This can be
used to construct the shifted lognormal univariate PDF
where we assume ξ¯ = σs2 = ξ1s = ξ2s , which holds for as
large enough patches compared to the cell size. The in-
−g(κ)2
 
tegration yields a bivariate Gaussian PDF for the mean- 2 Θ(κ + s)
PLN (κ; σG , s) = √ exp , (26)
subtracted convergence P(κ̃1 , κ̃2 ) with reduced variance 2πσG (κ + s) 2σG2
σκ̃2 = σ 2 − ξ¯ and reduced correlation ξκ̃ (θ) = ξ(θ) − ξ, ¯
where the reduction is set by the variance at the patch where Θ indicates the Heaviside step function, such that
scale, ξ¯ = σs2 . The mean-subtracted correlation ξ − ξ¯ the PDF is zero if κ ≤ −s. The shift parameter, s, can
now leads to a vanishing leading order super-sample co- be set to replicate the desired skewness (see the formula
variance term from equation (24) and creates a new ef- for λ in Xavier et al. 2016)
fective leading order term that we will discuss in Sec-
tion 3.2.1. Obviously this above described formalism has σκ σ4
1 + y(µ̃3 )−1 + y(µ̃3 ) − hκi ≈ 3 κ3 , (27)

s=
a long history which goes back to early papers on con- µ̃3 hκ i
strained Gaussian random fields e.g (Bertschinger 1987) s p
2
3 2 + µ̃ + µ̃3 4 + µ̃23
and (Hoffman & Ribak 1991) 3
y(µ̃3 ) = ≈ 1,
Covariance. When obtaining the covariance according 2
to equation (6) from the large-separation expansion of
the Gaussian (21), the leading 1 disappears and the cor- where hκi is the desired mean, σ the target variance and
relation ξ12 (θ) is integrated over the distance distribution µ̃3 = hκ3 i/σκ3 the skewness. The shift parameter is typ-
to yield ξ¯ which sets the amplitude of a super-sample co- ically positive and for our fiducial weak lensing conver-
variance like term we will discuss further in Section 2.1.2 gence case we use s = 0.1145 throughout the main text
as done in Boyle et al. (2021). In the executive summary,
¯ 1 P)G (κi )(b1 P)G (κj ) + O(ξ 2 ) .
covG (P(κi ), P(κj )) = ξ(b we use a different shift parameter because the Takahashi
(24) et al. (2017) simulations have a different cosmology. The
In Figure 7 we compare the correlation matrices as Gaussian limit can be reproduced by sending s/σκ → ∞.
measured (upper triangle) and predicted (lower trian- The bivariate shifted lognormal distribution  also de-
gle) for Gaussian fields of the three cases discussed pends on the correlation ξG = ln 1 + ξκ /s2 and the
before (generated with FLASK, with the same initial determinant d = σG 4
− ξG2
power spectrum and smoothing as before), from non-
overlapping cells with raw κ (left) along with overlapping 2 Θ(κ1 + s)Θ(κ2 + s)
cells of raw κ (middle) and mean-subtracted κ̃ = κ − κ̄ PLN (κ1 , κ2 ; σG √
, ξG , s) = (28)
2π d(κ1 + s)(κ2 + s)
(right panel). In the non-overlapping case on the left-
σ2
 
hand side, we can see how the finite sampling term (11) ξG
× exp − G [g(κ1 )2 + g(κ2 )2 ] + g(κ1 )g(κ2 ) .
causes the diagonal to be most pronounced with a sub- 2d d
dominant 4-tile pattern showing the leading order super-
sample covariance like term from equation (24). When The joint shifted lognormal PDF can be expanded at
2
making cells overlap as shown in the middle panel, the large separations where ξG (ξκ )/σG  1 to obtain up to
finite sampling term becomes irrelevant and the 4-tile leading order in ξκ
pattern [due to the factorisation of that covariance into a PLN (κ1 , κ2 ; θ) ξκ (θ) g(κ1 ) g(κ2 )
product of two independent bias functions with one zero- =1+ 2 2 2 + O(ξκ2 ) .
crossing each] as seen in equation (24)] becomes more PLN (κ1 )PLN (κ2 ) s σG σG
pronounced, while the resulting cell overlap causes a sig- (29)
nificant increase of the correlation in nearby bins as can
be seen in a band along the diagonal. Considering the In particular, it is useful to define a leading order bias
mean-subtracted convergence shown in the right panel function that resembles the Gaussian expression b1,G =
will remove the leading order, super-sample covariance κ/σ 2
term from equation (24) and instead lead to a 9-tile pat- g(κ) sg(κ)
tern [due to the factorisation of that covariance contri- b1,LN (κ) = 2 ' , (30)
bution into a products of two bias functions with two sσG σκ2
9

κ PDF correlation, (5deg)2 , NO κ PDF correlation, (5deg)2 , O κ − κ̄ PDF correlation, (5deg)2 , O


1.00
Gauss patches Gauss patches Gauss patches
-0.02
0.75

0.50
-0.01
0.25

0.0
κ2

0.00

−0.25
0.01
−0.50

0.02 −0.75
Gauss theory Gauss theory Gauss theory
−1.00
-0.02 -0.01 0.0 0.01 0.02 -0.02 -0.01 0.0 0.01 0.02 -0.02 -0.01 0.0 0.01 0.02
κ1 κ1 κ1

Fig. 7.— Correlation matrix between the bins of the Gaussian weak lensing convergence κ PDF at a smoothing scale ' 7.30 measured
from 1200 (5 deg)2 patches for each of 10 full sky Gaussian maps (upper triangle) and the theory prediction (lower triangle) for κ including
the nonzero mean in the patches (left and middle panel) with non-overlapping (NO) cells (left) and overlapping cells (middle) and κ − κ̄
after subtracting the mean with overlapping cells (right panel). The predictions for the diagonal of the covariance matrices are shown in
Figure 5.

To evaluate this, let us assume that κ̃ is a shifted log-


normal random variable (Hilbert et al. 2011), i.e. given
200 Takahashi θ1 =7.320
Takahashi θ2 =10.250
in terms of a zero-mean Gaussian random variable g̃ =
2 2
Gauss pred, σκ =0.0133 ln(κ̃/s̃ + 1) + σ̃G /2 with variance σ̃G = ln 1 + σκ̃2 /s̃2
100 Gauss pred, σκ =0.0117 and a new lognormal shift s̃. The correlationof the Gaus-
shifted LN pred
shifted LN pred
sian variables is given by ξ˜G = ln 1 + ξκ̃ /s̃2 . To find an
0
effective shifted lognormal approximation for κ̃, we must
b(κ)

determine variance, correlation and the shift parameter


s̃. Recall that we know the variance of κ ≡ κs + κ̃ from
−100 any procedure for calculating the matter power spec-
trum. In the same way, we can also know the covariance
−200
hκi κs i = ξis and the variance hκ2s i = σs2 , which we will
approximate as identical ξis = σs2 = ξ. ¯ So we get the fol-
−0.03 −0.02 −0.01 0.00 0.01 0.02 0.03 lowing variance and correlation for the mean-subtracted
κ
κ̃
Fig. 8.— κ-dependent clustering as measured by the cell bias
from one full-sky map in the Takahashi simulation in compari-
σ 2 = hκ̃2 i = hκ2 i − 2hκκs i + hκ2 i ≈ σ 2 − ξ¯ ,
κ̃ s (32a)
κ
son to the theoretical prediction for a shifted lognormal field from ξκ̃ = hκ̃1 κ̃2 i = hκ1 κ2 i − hκ1 κs i − hκ2 κs i + hκ2s i
equation (30).
≈ ξκ − ξ¯ , (32b)
and hence the Gaussian variance σ̃G2 ¯ 2]
= ln[1+(σκ2 − ξ)/s̃
˜ ¯ 2
and correlation ξG = ln[1 + (ξκ − ξ)/s̃ ]. The shift pa-
where for κ  s we have sg(κ) ' κ + σκ2 /(2s) and hence
reproduce the Gaussian result modulo a small shift. For rameter can be related to the skewness as described in
larger κ the shifted lognormal result deviates from the equation B.11 of Hilbert et al. (2011)
Gaussian one to closely resemble measurements from 3 2 2 1 2 3 3 3hκ̃2 i2
simulated κ maps (Takahashi et al. 2017) using the ees- hκ̃3 i = hκ̃ i + 3 hκ̃ i ≈ hκ̃2 i2 ⇒ s̃ ≈ . (33)
timator (16c) as illustrated in Figure 8. s̃ s̃ s̃ hκ̃3 i
The impact of mean-subtraction. When describing Following Friedrich et al. (2018) we use the shift parame-
small patches which will feature different background ter s̃ to match the skewness of κ̃, which is obtained from
‘sample’ convergence values κs , one needs to integrate the third moment of the difference
over a trivariate PDF of the convergences κ1/2 and the
sample value κs , similarly as for the mean-subtraction in hκ̃3 i = h(κ − κs )3 i = hκ3 i − 3hκ2 κs i + 3hκκ2s i . (34)
the Gaussian case (23a)
Z The third moment of κ can be expressed in terms of the
P(κ̃1 , κ̃2 ) = dκs PLN (κ1 = κ̃1 + κs , κ2 = κ̃2 + κs , κs ) . variance σκ and original shift parameter s using hκ3 i ≈
3σκ4 /s. Additionally, we can compute the relevant mixed
(31) moments of a zero-mean normal random variable Y and a
lognormal variable Z = sz (eX −1) given in terms of a nor-
10

mal variable X. The relevant relations are hY (eX−1)i = κ PDF correlation, (5deg)2 patches, O κ PDF correlation, full sky, O
1.00
FLASK FLASK
−2
ξXY and hY 2 (eX − 1)i = ξXY 2
σX2
+ 1 − σX which -0.02
0.75
X 2
yields hY (e − 1) i = 0. This can be used to obtain -0.01
0.50

hκ2 κs i = 0 and hκκ2s i ≈ ξ¯2 σG


2 −2

+ 1 − σG /s. Inserting 0.25

this into equations (33) and (34) leads to a relation be- 0.0

κ2
0.00

tween the shift parameters s and s̃, which we found to −0.25


be practically identical in our case. 0.01
−0.50
Covariance. Again, we can integrate the large- 0.02 −0.75
separation expansion of the shifted lognormal (57a) over theory theory
the distance distribution to obtain the covariance from -0.02 -0.01 0.0 0.01 0.02 -0.02 -0.01 0.0 0.01 0.02
−1.00

equation (6) and eventually get at leading order κ1 κ1

covariance diagonal modelling


¯ 1 P)LN (κi )(b1 P)LN (κj )+O(ξ 2 ) ,
covLN (P(κi ), P(κj )) = ξ(b FLASK patches κ
(35) theory lognormal κ
which is in form analogous to the Gaussian result (24) 0.020 FLASK patches κ − κ̄
despite the modified ingredients. This explains why the theory lognormal patch κ − κ̄
correlation matrix from small patches (with sizable ξ) ¯ in ¯ b̃1 P)2
theory lognormal patch κ − κ̄ + SSC ξ( LN
0.015 FLASK full sky
the FLASK case closely resembles the Gaussian case, as

σ 2 (P(κ))
theory lognormal full sky
seen in Figure 6 while the diagonal is modified due to
the non-Gaussian shape of the final convergence PDF as 0.010
seen in Figure 5.
In Figure 9 we compare the measurements from
FLASK maps to the theoretical predictions for shifted 0.005

lognormal fields. In the left upper panel we show the


correlation matrix obtained from 1200 FLASK patches 0.000
of 5deg each cut from 20 full sky realisations (upper tri- −0.02 −0.01 0.00 0.01 0.02
angle) along with the theoretical prediction (lower tri- κ
angle) finding excellent agreement. We can see that the
Fig. 9.— (Top) Correlation matrix between the bins of the shifted
correlation matrix has a 4-tile structure. This is because lognormal weak lensing convergence κ PDF at a smoothing scale
the mean correlation ξ¯ in the patches is sizeable such ' 7.30 comparing FLASK measurements (upper triangles) to the
that the first order covariance term (35) dominates the theory predictions (lower triangles). (left) Measurements from 1200
visual appearance. In the upper right panel we show the square patches of 5deg size cut from 10 full sky realisations in com-
parison to the lognormal theory prediction in patches. While the
correlation matrix obtained from 500 full sky FLASK first order bias function b1,LN (30) creates the 4-tile pattern, cell
maps (upper triangle) along with the theoretical predic- overlaps create the positive correlation band along the diagonal.
tion (lower triangle) finding excellent agreement. We can (right) Measurements from 500 full sky realisations in compari-
see that the correlation matrix closely resembles the case son to the lognormal theory prediction on the full sky. The 9-
tile pattern is set by the second order bias function b2,LN (58),
of the mean-subtracted convergence measured in small while cell overlaps create the strong correlation along the diagonal.
patches, shown in the lower panel of Figure 6. This is (Bottom) Diagonal of the covariance matrix for the non-Gaussian
because the mean correlation ξ¯ on a full sky is so small FLASK PDFs measured from the full sky (green solid) and as pre-
dicted by the lognormal theory (green dotted). We also show the
that the first order covariance term (35) becomes sub- result for the FLASK patches without/with mean subtraction (pur-
dominant compared to the second order term that we will ple/blue solid) and the lognormal theory computed for the patches
discuss in Section 3.1.4. The latter term also appears as without/with mean-subtraction (purple/blue dotted). The shaded
leading order contribution in the mean-subtracted case bands are indicating estimated errors)
addressed in Section 3.2.3. In the bottom panel we fo-
cus on the diagonal of the covariance matrix, compar-
ing FLASK maps and theoretical predictions for different
cases. This plot shows that while the PDF covariance of ancies for the mean-subtracted κ̃ in small patches. As
the raw κ in small patches looks significantly different, this discrepancy was absent for Gaussian fields shown in
the mean-subtracted κ̃ in small patches resembles the Figure 5, we suspect that it is caused by a more com-
full sky covariance. The theoretical prediction captures plicated patch-to-patch correlation induced by cutting
all three behaviours quite well, Considering relative 1-σ multiple patches from one full sky, which is unaccounted
errors on the diagonal of the covariance matrix (shaded for in the theory describing the mean-subtracted conver-
bands) following from the covariance between the numer- gence as another shifted lognormal random field.
ically estimated diagonal covariance elements Ĉii and Ĉjj
as given in Taylor et al. (2013) 2.3.3. Bivariate Gaussian copula with general marginals
In case neither the bivariate normal nor the shifted
2
  2Cij lognormal distributions can provide a good description
Cov Ĉii , Ĉjj = , (36) of the target PDF as its univariate marginal, one can
Nsim − 1
design a bivariate PDF by specifying the marginal and
the theoretical predictions capture all three behaviours correlation structure independently. A joint PDF can
quite well. We notice that while the lognormal prediction be built from a given marginal one-point PDF P(κ)
for the full sky and the raw κ in patches agree extremely and a correlation structure encoded in a copula density
well with the predictions, there are some minor discrep- c12 which quantifies the correlation between ranked vari-
11

ables obtained
Rκ from the cumulative distribution function diagonal covariance matrix PDF shape noise only
C(κ) = dκ̃ P(κ̃), such that theory
8 measurement
P(κ1 , κ2 ; θ) = P(κ1 )P(κ2 )c12 (C(κ1 ), C(κ2 ); θ) , (37)
where θ indicates angular separation as before. If the
6
κi and κ2 are independent, then the copula density is

σ 2 (P(κSNonly ))
unity c12,indep = 1. A Gaussian copula only depends
on the distance through the pairwise correlation ξ12 (θ), 4
c12,G (C(κ1 ), C(κ2 ); ξ12 (θ)) and is a known function which
can be computed numerically although it has no ana-
lytical expression. Beyond those two cases, systemati- 2
cally constructing the dependence structure is possible
in the large-separation regime, where the copula can be
expanded in powers of the pairwise correlation ξ12 (θ) as 0
we will further discuss in Section 3. −4 −2 0 2 4
κSNonly /σSN
2.3.4. Inclusion of shape noise
Observationally, the weak lensing convergence map is correlation matrix PDF shape noise only
1.00
obtained from cosmic shear measurements, which are measurement
−4
subject to shape noise often constituting the dominant 0.75
source of noise. Since galaxies are intrinsically elliptical, −3
the observed shear contains a contribution from this in- 0.50
trinsic signal, quantified by the variance of the intrinsic −2
ellipticity σ . In simulated maps, shape noise can be in- −1 0.25
cluded by adding a white noise – modeled by a Gaussian κSNonly /σSN
random map κSN – to the ‘raw’ simulated convergence 0 0.00
map κ before smoothing. The standard deviation of this
Gaussian is typically assumed to be 1 −0.25
σ 2
σSN = p ∝ θs−1 , (38) −0.50
n g · Ωθ s 3
−0.75
with the shape-noise parameter σ (here taken to be 4
theory
σ = 0.30), the galaxy density ng (here taken to be 30 −1.00
arcmin−2 ) and the solid angle Ωθs (here taken to corre- −4 −3 −2 −1 0 1 2 3 4
spond to circular apertures of radius θs = 7.30 ). κSNonly /σSN
Smoothing induced shape noise correlations and covari-
ance. At the raw κ map level, the shape noise values Fig. 10.— PDF covariance diagonal (upper panel) and correlation
matrix (lower panel) for the pure shape-noise lensing convergence
in different pixels are uncorrelated and ξ ≡ 0. Top- in circular apertures of radius θs comparing the prediction from
hat smoothing the field at angular size θs will correlate equation (40) to the measurement from shape noise realisations.
shape noise values in overlapping cells (being at separa- The slight apparent asymmetry between positive and negative κ is
tion θ < 2θs ). Their correlation is related to the area of caused by the noise due to the finite number of realisations.
the cell overlap (which is a symmetric lens)
 
2 θ θp 2
A(θ < 2θs ) = 2θs arccos − 4θs − θ2 . (39a)
2θs 2
The resulting correlation of the top-hat smoothed shot for small distances Pd (θ) ≈ 2πθ. This gives the PDF
noise contribution to the weak lensing convergence is covariance for the case of pure shape noise as
2
ξSN (θ < 2θs ) = σSN A(θ)/(πθs2 ) , ξSN (θ ≥ 2θs ) = 0 . Z 2θs
(39b) covSN (P(κi ), P(κj )) = 2
dθ Pd (θ)PG (κi , κj ; σSN , ξSN (θ))
To obtain the covariance of the shot noise contribution 0
alone, we integrate the bivariate Gaussian with variance Z 2θs
2 2 2
σSN and covariance (39) over distances following equa- − PG (κi ; σSN )PG (κj ; σSN ) dθ Pd (θ) .
tion (6) to get 0
(40)
Z 2θs
2
hPSN (κi )PSN (κj )i = dθ Pd (θ)PG (κi , κj ; σSN , ξSN (θ)) In Figure 10 we compare this theoretical result to a mea-
0
Z ∞ surement across 1000 shape noise realisations for square
2 2 maps of length 5deg with a top-hat smoothing of 7.30
+ PG (κi ; σSN )PG (κj ; σSN ) dθ Pd (θ) , finding excellent agreement.
2θs
Shape noise induced convergence PDF convolution.
where Pd (θ) indicates the distribution of angular dis- Shape noise impacts the one-point weak lensing conver-
tances in the survey area, which is approximately linear gence PDF as if it was convolved with a zero-centred
12

cumulant generating function (CGF). We will pay par-


covariance diagonal modelling including shape noise
ticular attention to the leading and next-to-leading order
0.008 FLASK κ noisefree
theory lognormal κ noisefree
contributions to the covariance matrix, which reveals a
0.007 FLASK κ noisy super-sample covariance term that is controlled by the
theory lognormal κ noisy average correlation in the survey patch. We will compute
0.006
the expansion to all orders in ξ12 for a Gaussian field and
the two leading order terms for the general non-Gaussian
σ 2 (P(κ))

0.005
case, illustrated by results for normal and shifted lognor-
0.004
mal fields. This will naturally lead to an expansion of the
0.003 covariance matrix that resembles an eigendecomposition,
which we will discuss in Section 4.1.
0.002

0.001 3.1. Covariance expansion for raw κ


−0.02 −0.01 0.00 0.01 0.02 3.1.1. Leading order determined by κ-dependent clustering
κ
Let us first recall that at leading order in the large-
Fig. 11.— PDF covariance diagonal for the noise-free lensing separation regime, the joint two-point PDF can be ap-
convergence as measured from 500 FLASK full sky maps (green
dashed, shaded band indicating estimated errors) and predicted proximated in terms of the one-point PDF and the
for the shifted lognormal (green dotted) in comparison to the noisy ‘bias’ functions indicating the density-dependent two-
lensing convergence as measured from FLASK (green solid, shaded point clustering
band indicating estimated errors) and predicted from equation (6)
with with the noisy two-point PDF (42) (green dot-dashed). P(κ1 , κ2 , θ) ' P(κ1 )P(κ2 )[1 + ξ(θ)b(κ1 )b(κ2 )] . (43)
Hence, we have for the covariance of the binned proba-
bility
Gaussian (Clerkin et al. 2016)
Pκ+SN (κ̂) = (PG (σSN ) ∗ P)(κ̂) cov(P(κi ), P(κj )) = ξ(¯ P̄b)(κi )¯(Pb)(κj ) + δij P̄(κi ) ,
∆ i NT
(κ̂ − κ)2
 
1 (44)
Z
=√ dκ exp − 2 P(κ) , where NT is the total number of cells and the mean cor-
2πσSN 2σSN
(41) relation ξ¯ is given by equation (13b). Note that the shot-
noise term only acts on the diagonal. For the diagonal,
where the shape noise standard deviation σSN is given in the leading-order prediction for covariance diagonal is
equation (38). For the two-point PDF of the smoothed then
weak lensing we similarly have κ̂i = κi + κi,SN and the ¯ P̄)2 (κ) + P̄(κ) ,
bivariate PDF is given in terms of a 2D-convolution var(P(κ)) = ξ(b (45)
∆NT
Pκ+SN (κ̂1 , κ̂2 ; θ) = (PG (σSN , ξSN (θ)) ∗ P)(κ̂1 , κ̂2 ; θ) , where ∆ is the PDF bin width. For a Gaussian field, the
(42) bias function is given by linear Kaiser bias b(κ) = κ/σκ2 ,
where the shape noise contribution is described by a bi- which seems to be a decent approximation for the central
variate Gaussian PDF from equation (19) with standard region of κ-values, see Figure 8. Note that the cell bias
σSN (38) and correlation ξSN (θ) (39). If the joint PDF b(κ) has at least one zero crossing, for the case of raw κ,
of raw κi was Gaussian, then the noisy PDF of κ̂i aris- around κ ' 0, while for the mean-subtracted case this is
ing from the convolution would be another zero-mean modified as described in Section 3.2.
Gaussian distribution with covariance matrix Σκ+SN = Let us also recall that for non-overlapping cells,
Σκ + ΣSN . typically the ‘finite sampling’ effect dominates (well-
For the case of a shifted lognormal κ PDF, we re- described by Poisson). For overlapping cells, the num-
compute the covariance from the convolved two-point ber of cells is extremely large thus effectively removing
PDF (42) to find that the main impact of adding shape the finite sampling term, but requiring the treatment of
noise is to change the diagonal of the covariance ma- overlaps. The impact of overlapping cells can be cap-
trix, shown in Figure 11, while leaving the correlation tured by either closed-form expressions for the two-point
matrix largely similar. This result was obtained from PDF valid at small distances (like for the Gaussian and
Nsim = 500 FLASK maps of the full sky with a smooth- shifted-lognormal case discussed here) or a heavy overlap
ing of 7.30 applied to the raw κ and the shape-noise added expansion complementary to the large separation one as
κ̂, respectively. Considering relative 1-σ errors on the di- described in Bernardeau (2022).
agonal of the covariance matrix (36) evaluated for Nsim
the predictions are in good agreement with the measure- 3.1.2. Beyond leading order: Gaussian case
ments. For a Gaussian field of zero mean, we have the fol-
lowing simple structure of the joint two-point cumulant
3. LARGE SEPARATION COVARIANCE generating function in terms of the two-point correlation
EXPANSIONS ξ = ξ12 (θ) as follows
In this Section we will perform an expansion of the
joint two-point PDF in powers of the two-point cell cor- ϕG (λ1 , λ2 ) = ϕ0,G (λ1 ) + ϕ0,G (λ2 ) + ξϕ1,G (λ1 )ϕ1,G (λ2 ) ,
relation function ξ12 (θ), partially with the help of the (46)
13

where ϕ0,G = λ2 σ 2 /2, ϕ1,G (λ) = λ and ϕn≥2,G (λ) = 0. averages 2-point cell correlation vs. variance powers
By expanding the last term exp[ξλ1 λ2 ] in a power series ξ n /σ 2n 5deg rescaled
and converting λi to derivatives w.r.t. κi , we obtain ¯ n /(σ 2 − ξ)
¯ n 5deg rescaled
(ξ − ξ)
∞ 10−5
ξ n /σ 2n full sky
PG (κ1 , κ2 , θ) X ξ n (θ)
=1+ bn,G (κ1 )bn,G (κ2 ) (47)
PG (κ1 )PG (κ2 ) n=1
n!
(−1)n ∂ n PG (κ) 1 κ
bn,G (κ) = = He n ,
PG (κ) ∂κn σn σ
(48)
10−6
which can be written in terms of the probabilist’s Her-
mite polynomials Hen
 2  2
x d x
Hen (x) = (−1)n exp exp − , (49)
2 dxn 2 2 4 6 8 10
n
when using x = κ/σ. The first two terms read (Des-
jacques et al. 2018, see also) Fig. 12.— Scaling of the average of powers of the two-point cell
correlation function compared to the variance for the raw κ maps
κ κ2 − σ 2 1 in 5deg patches (red plusses) and the full sky result rescaled by a
b1,G (κ) = 2
, b2,G (κ) = = b21,G − 2 , (50) factor of Apatch /Afullsky (blue plusses). Aside from the removal of
σ σ4 σ the first order term, those hierarchies closely resemble each other
as well as the mean-subtracted κ̃ maps in patches (red crosses).
which can be also obtained from a direct expansion of the
joint PDF (19). This leads to the following expansion for
the covariance following equation (6)

X ξ n ∂ n PG (κ1 ) ∂ n PG (κ2 )
cov(PG (κ1 ), PG (κ2 )) =
n=1
n! ∂κ1 n ∂κ2 n
(51)
∞ hierarchical ordering of ξ n /σ 2n will reflect in the eigen-
X ξn κ 
1 2
κ 
values and the bias functions multiplied by the Gaussian
= 2n
Hen PG (κ1 )Hen PG (κ2 ) ,
n=1
n!σ σ σ PDF evaluated at a given κ-value will yield the eigen-
Z X Nd (θ) vectors (after orthogonalisation and appropriate normal-
ξ := dθPd (θ)ξ n (θ) '
n ξ n (θ) , (52) isation). In this context, the predicted zero-crossings of
Nd the first bias function b1,G (κ) at κ = 0 and of the sec-
θ∈Θ\{0}
ond b2,G (κ) at κ = σ 2 will translate into the approximate
where in the second line we have defined the average of zero crossings of the first two eigenvectors as we will show
powers of the two-point correlation function over the sur- later in Figure 15.
vey patch in analogy to the mean from equation (13b).
The averages of powers of the correlation function ξ n 3.1.3. Beyond leading order: general non-Gaussian case
compared to powers of the variance σ 2n control the large-
separation expansion of the covariance and their hierar- Higher-order terms are obtained from expanding the
chy is illustrated in Figure 12. This plot also shows that joint two-point cumulant generating function in powers
the covariance contributions scale inverse proportional to of the two-point correlation ξ = ξ12 (θ) as follows
the survey area ξ n ∝ A−1 .
Due to the hierarchical ordering of averages of the cor- ϕ(λ1 , λ2 ) = ϕ0 (λ1 ) + ϕ0 (λ2 ) + ξϕ1 (λ1 )ϕ1 (λ2 ) (54)
relation function we expect the two leading order terms ξ2
in the regime κ . O(σ) to be + [ϕ2 (λ1 )ϕ2 (λ2 ) + ϕ2 (λ1 )ϕ21 (λ2 )] + O(ξ 3 )
2 1
cov(PG (κ1 ), PG (κ2 )) ξ¯ κ1 κ2 We are then in position to compute the (connected part)
= 2 (53a)
PG (κ1 )PG (κ2 ) σ σ σ of the joint PDF,
ξ 2 κ21 − σ 2 κ22 − σ 2 Z
dλ1
Z
dλ2
+ (53b) P(κ1 , κ2 ) − P(κ1 )P(κ2 ) =
2σ 4 σ !
2 σ2 2πi 2πi
ξ3 exp [−λ1 κ1 − λ2 κ2 + ϕ(λ1 , −λ1 ) + ϕ(λ2 , −λ2 )] ×
+O ,
σ6  ξ2
ξϕ1 (λ1 )ϕ1 (λ2 ) + ϕ21 (λ1 )ϕ21 (λ2 ) (55a)
which predicts a super-sample covariance leading order 2
term similar to the separate universe picture (81), being ξ2 
proportional to the variance of the mean κ in different + [ϕ21 (λ1 )ϕ2 (λ2 ) + ϕ2 (λ1 )ϕ21 (λ2 )] + O(ξ 3 ) (55b)
2
patches, ξ¯ = σ 2 (κs ), and the derivative of the PDF. This
large-separation expansion of the covariance will be con- We obtain the covariance according to equation (6) af-
nected to an eigendecomposition in Section 4.1. Here, the ter integrating over λi and the distance distribution and
14

symmetrising
cov(P(κ1 ), P(κ2 )) 0.004 σ 2 =10x10-4
¯ 1 (κ1 )b1 (κ2 )
= ξb (56a)
P(κ1 )P(κ2 ) σ 2 =5x10-4
0.003

b2,LN (κ)σ 4 +σ 2
ξ2 σ 2 =2x10-4
+ [(b2 + q1 )(κ1 )(b2 + q1 )(κ2 ) − q1 (κ1 )q1 (κ2 )] (56b)
2 s2 ln(1+κ/s)2
0.002
+ O(ξ 3 ) , κ2
where we defined 0.001

Z
(b1 P)(κ) = ϕ1 (λ) exp[−λκ + ϕ(λ)] (56c)
2πi 0.000

Z
-0.04 -0.02 0.00 0.02 0.04
(b2 P)(κ) = ϕ1 (λ)2 exp[−λκ + ϕ(λ)] (56d) κ
2πi
dλ Fig. 13.— A comparison of a rescaled version of the second order
Z
(q1 P)(κ) = ϕ2 (λ) exp[−λκ + ϕ(λ)] . (56e) bias functions from the shifted lognormal model used here b2,LN
2πi (solid) along with (b2 + q1 )LN (dot-dashed) from equations (58)
(colours for different variances) and the limiting behaviour (59)
For a saddle-point approximation of the integral, we (black solid) in comparison to the Gaussian result (black dashed).
would typically expect b2 ' b21 , although that does not We also show lines of constant σ 2 (dotted) to indicate the zero-
reproduce the Gaussian or shifted lognormal result. The crossings of b2,LN (κ) which appear in the second eigenvectors for
expression (56) consists of dyadic products that resem- the κ PDF covariance in the lower panel of Figure 16.
ble the structure of an eigendecomposition, which we will
look at in Section 4.1, although eigenvectors are addition-
ally pairwise orthogonal.
where we used the result for b1,LN from equation (30).
3.1.4. Beyond leading order: shifted lognormal case We see that in the limit of s → ∞ we recover sg(κ) → κ
The recipe described in the previous paragraph for gen- and the Gaussian result with q1,G = 0, while for mildly
eral non-Gaussian distributions is not valid for a lognor- non-Gaussian fields, we typically have q1  b2 . In a
mal PDF, which is not well-described in terms of a CGF. small variance expansion where σ/s  1, we obtain the
Here we detail an approach that can be used to obtain limiting behaviour
lognormal bias functions by applying a large-separation 2
expansion to the shifted lognormal two-point PDF (28).
σ
s 1
s2 ln 1 + κs − σ 2 1
b2,LN (κ) −→ 4
− 2 (59)
This bias expansion is similar in spirit to the Gaussian σ 4s
case, but requires an additional expansion of the Gaus- Using the all-order results for bn,G (48) together
sian field correlation ξG in terms of the underlying κ  n with
P (−1)n−1 ξκ
correlation ξκ . The joint shifted lognormal PDF can be the expansion of ξG (ξκ ) = n=1 n s2 allows
2
expanded at large separations where ξG (ξκ )/σG  1 to to compute all order expressions for the lognormal bias
obtain up to next-to-leading order in ξκ functions. Note that due to the difference between ξG
PLN (κ1 , κ2 ; ξκ ) g(κ1 ) g(κ2 ) and ξκ , the higher-order terms receive contributions from
= 1 + ξG (ξκ ) 2 2 (57a) lower-order bias functions, as predicted by the CGF-
PLN (κ1 )PLN (κ2 ) σG σG based expansion from before. In Figure 13 we show a
(ξκ ) g(κ1 )2 − σG2
g(κ2 )2 − σG
2
 
2 comparison of the lognormal terms b2 (solid) and b2 + q1
ξG
+ 4 4 (dashed) to the Gaussian result for the shift parameter
2 σG σG
adopted in this work and different variances. We can
3
+ O(ξG ) see that for our case (closely corresponding to the green

ξκ ξκ2 g(κ1 ) g(κ2 )
 line), the q1 contribution is negligible. The crossing be-
=1+ − 2 2 tween the solid b2 line and the dotted line of constant σ 2
s2 2s4 σG σG
indicates the zero-crossing of b2 (κ), which is reflected in
(57b) the second eigenvector for the κ covariance shown in the
2 2 2 2
 
ξκ2 g(κ1 ) − σG g(κ2 ) − σG lower panel of Figure 16. For the case of a survey area
+ 4 4 being a significant portion of the sky, the mean correla-
2s4 σG σG
tion ξ¯ will be extremely small and this second order bias
+ O(ξκ3 ) , function b2 (κ) will determine the covariance structure.
In particular, it is useful to define two additional bias The two zero-crossings of b2 cause the 9-tiled correlation
functions q1,LN and b2,LN following the previous expan- matrix structure displayed earlier, notably in the upper
sion (56), with their combination resembling the Gaus- right panel of Figure 9.
sian expression from equation (50)
3.2. Covariance expansion for mean-subtracted
g(κ)2 − σG
2
[sg(κ)]2 − σκ2 κ̃ = κ − κs
(b2 + q1 )LN (κ) = 4 ' (58a)
s2 σG σκ4 To determine the PDF covariance for the case of a
1 g(κ) g(κ) mean-subtracted κ field, one needs to consider the 3 vari-
q1,LN (κ) = b1,LN (κ) = 2 2 ' 2 , (58b) ables κ1 , κ2 and κs , where κ1/2 are as before and κs is
s s σG σκ the mean convergence in the survey patch. Like the con-
15

vergence field, those variables all have zero expectation to powers of the variances σ 2n and (σ 2 − ξ) ¯ n control
value and we are then interested in the joint PDF of the the large-separation expansion of the covariance (in the
mean-subtracted κ̃i Gaussian case given in equation (51) for raw κ and in
equation (65) for the mean-subtracted κ̃ = κ − κ̄) follow
κ̃1 = κ1 − κs , κ̃2 = κ2 − κs . (60) similar hierarchies as illustrated in Figure 12.
Let us note that this tilde variable is formally equivalent In particular, the leading order term will be
to the slope variable that was introduced in Bernardeau ¯ 2  κ̃2  2
cov(PG (κ̃1 ), PG (κ̃2 )) (ξ − ξ)

et al. (2014, 2015). The CGF for the three variables de- 1 κ̃2
= −1 −1
fined above is written here up to linear order in the two- PG (κ̃1 )PG (κ̃2 ) 2σκ̃4 σ2 σ2
point function at the survey scale, ξs , and up to quadratic | κ̃ {z κ̃ }
4b
σκ̃ 2,G (κ̃1 )b2,G (κ̃2 )
order in the two-point correlation at the given cell sepa- !
ration ξ12 (ξ − ξ)¯3
+O , (66)
1 2 2 σκ̃6
ϕ(λ1 , λ2 , λs ) = σs λs + ϕ0 (λ1 ) + ϕ0 (λ2 ) + ξ12 ϕ1 (λ1 )ϕ1 (λ2 )
2
+ λs [ξ1s ϕ1 (λ1 ) + ξ2s ϕ1 (λ2 )] (61) which predicts a first eigenvalue/-vector similar to the
2
second eigenvalue/-vector of the Gaussian case without
ξ12 2 2 3 ¯ 2 /σ 4 ' ξ 2 /σ 4 as
+ [ϕ (λ1 )ϕ2 (λ2 ) + ϕ2 (λ1 )ϕ1 (λ2 )] + O(ξ12 ) mean-subtraction (53b) because (ξ − ξ) κ̃
2 1 shown in Figure 12. The zero-crossings of the first eigen-
1 2 ¯ 2
vector will be at κ̃ ≈ σ − ξ ≈ σ , closely following the
ϕ(λ1 , λs ) = ξs λ2s + ϕ0 (λ1 ) + λs ξ1s ϕ1 (λ1 ) , (62)
2 result for full sky maps where ξ¯ is negligibly small and
where we assumed that the patches are big enough such eigenvectors are shown as green/olive lines in the middle
that the κs PDF is Gaussian. The one- and two-point panel of Figure 16, as discussed in Section 4.1. Note that
PDFs are obtained from this CGF by performing the as before the joint two-point PDF of the mean-subtracted
change of variable (60) and then marginalising over the κ̃ values could be written in terms of the marginal Gaus-
background κs , leading to sian PDFs and a series of Hermite polynomials as in
equation (48) with prefactors set by averages of powers
dλ1 −λ1 κ̃1 +ϕ(λ1 ,−λ1 ) ¯ 2.
Z
of (ξ − ξ)/σ κ̃
P(κ̃1 ) = e (63)
2πi
ZZ
dλ1 dλ2 −λ1 κ̃1 −λ2 κ̃2 +ϕ(λ1 ,λ2 ,−λ1 −λ2 ) 3.2.2. Leading order: general non-Gaussian case
P(κ̃1 , κ̃2 ) = e . Similarly as for the Gaussian case, we will rewrite the
2πi 2πi
(64) joint CGF of the three variables λ1 , λ2 , λs in terms of
the joint CGF for just two variables λi and λs , where
When all computations are made at linear order in ξ, λs = −(λ1 + λ2 ). We will first discuss the leading order
we can set σs2 = ξ1s = ξ2s = ξ12 = ξ(θ). To obtain result which reads
the second order, one needs to start from ξ¯ = σs2 =
ξ1s = ξ2s 6= ξ12 = ξ(θ), which we will use to illustrate ϕ(λ1 , λ2 , −λ1 − λ2 ) − ϕ(λ1 , −λ1 ) − ϕ(λ2 , −λ2 ) (67)
the similarities and differences between the Gaussian and ¯ 1 λ2 − λ1 ϕ1 (λ2 ) − λ2 ϕ1 (λ1 )] + O(ξ 2 )
=ξϕ1 (λ1 )ϕ1 (λ2 ) + ξ[λ
non-Gaussian cases. ¯ 1 (λ1 )ϕ1 (λ2 ) + ξ[ϕ
¯ 1 (λ1 ) − λ1 ][ϕ1 (λ2 ) − λ2 ] + O(ξ 2 ) ,
=(ξ − ξ)ϕ
3.2.1. Gaussian case
where we used ξ = ξ(θ) for brevity and reorganised the
For a Gaussian field, the two terms λ2s and λs [ϕ1 (λ1 ) + formula such that the for a Gaussian only the first term
ϕ1 (λ2 )] can be summarised and act to reduce the vari- remains, because ϕ1 (λ) = λ. We are then in position to
ance and correlation function by ξ, ¯ such that joint PDF compute the connected part of the joint PDF, given by
P(κ̃1 , κ̃2 ) is a Gaussian with variance σκ̃2 = σ 2 − ξ¯ and
¯ Hence, the covariance can P(κ̃1 , κ̃2 ) − P(κ̃1 )P(κ̃2 )
correlation ξκ̃ (θ) = ξ(θ) − ξ.
dλ1 dλ2 −λ1 κ̃1 −λ2 κ̃2 +ϕ(λ1 ,−λ1 )+ϕ(λ2 ,−λ2 )
ZZ
be obtained in full analogy to the previous computation = e
that led to equation (51) which is modified to read 2πi 2πi
h
X∞ ¯ n ∂ n PG (κ̃1 ) ∂ n PG (κ̃2 )
(ξ − ξ) × (ξ − ξ)ϕ ¯ 1 (λ1 )ϕ1 (λ2 ) + ξ[ϕ
¯ 1 (λ1 ) − λ1 ][ϕ1 (λ2 ) − λ2 ]
cov(PG (κ̃1 ), PG (κ̃2 )) = ,
n! ∂κ̃n1 ∂κ̃n2 i
n=2 + O({ξ 2 , ξ¯2 , (ξ − ξ)¯ 2 }) .
(65)
where due to the definition of the average of powers of When integrated over the distance distribution, the term
the correlation function (52) the n = 1 term disappears. linear in ξ − ξ¯ will vanish and the term involving products
Note that PG (κ̃1 ) does not denote the same PDF as of ϕ1 (λi ) − λi will create a new bias function b̃1,NG (κ̃) for
PG (κ1 ), for the Gaussian case they differ by the vari- the mean-subtracted convergence κ̃ only present for non-
ance being σκ̃2 = σκ2 − ξ¯ rather than σκ2 . Aside from this Gaussian fields. In terms of this, the PDF covariance can
(small) difference and the removal of the n = 1 term, we be written as
now have a different prefactor consisting of averages of ¯ b̃1,NG P)(κ̃1 )(b̃1,NG P)(κ̃2 ) (68a)
cov(P(κ̃1 ), P(κ̃2 )) =ξ(
powers of ξ − ξ¯ rather than ξ. Those averages of pow-
¯ n compared
ers of the correlation function ξ n and (ξ − ξ) + O({ξ 2 , ξ¯2 , (ξ − ξ)
¯ 2 }) ,
16

ingredients for non-Gaussian bias b̃1 for κ̃ = κ − κ̄


Again, we compute the connected part of the joint PDF,
200 given by
∂ log P̄
FLASK b̃1,NG FLASK b1 FLASK
∂κ

150
∂ log PLN
∂κ
b̃1,LN
b1,LN P(κ̃1 , κ̃2 ) − P(κ̃1 )P(κ̃2 )
dλ1 dλ2 −λ1 κ̃1 −λ2 κ̃2 +ϕ(λ1 ,−λ1 )+ϕ(λ2 ,−λ2 )
ZZ
100 = e
2πi 2πi
50 h
× (ξ − ξ)ϕ ¯ 1 (λ1 )ϕ1 (λ2 ) + ξ[ϕ
¯ 1 (λ1 ) − λ1 ][ϕ1 (λ2 ) − λ2 ]
0

ξ¯2
−50 + [ϕ1 (λ1 ) − λ1 ]2 [ϕ1 (λ2 ) − λ2 ]2 (70a)
2
−100 ¯ − ξ)ϕ
+ ξ(ξ ¯ 1 (λ1 )ϕ1 (λ2 )[ϕ1 (λ1 ) − λ1 ][ϕ1 (λ2 ) − λ2 ]
−150
ξ2 2
−0.03 −0.02 −0.01 0.00 0.01 0.02 0.03 + [ϕ (λ1 )ϕ2 (λ2 ) + ϕ2 (λ1 )ϕ21 (λ2 )] (70b)
κ̃ 2 1
¯2
(ξ − ξ) i
Fig. 14.— Additional first-order bias b̃1,NG (κ̃) (green) of the + ϕ21 (λ1 )ϕ21 (λ2 ) + O(ξ 3 ) .
mean-subtracted κ̃ = κ − κ̄ obtained from the sum of the ‘raw’ 2
bias b1 (green) and the response of the PDF ∂ log P̄/∂κ (red) as When integrated over the distance distribution, all terms
given in equation (68b). We also show predictions from the shifted
lognormal case (dashed) relying on the PDF from equation (26) linear in ξ − ξ¯ will vanish. The remaining contributions
and the bias b1,LN from equation (30) are
a) terms in the first two lines, involving powers of
ϕ1 (λ) − λ creating new bias functions, b̃n,NG , only
present for non-Gaussian fields.3
b) terms in the last two lines containing products of
powers of ϕn that yield the bias functions encoun-
tered for the covariance of the raw κ PDF, at sec-
where we introduced the leading order composite bias ond order b2 from equation (56d) and q1 from equa-
tion (56e).

b̃1,NG (κ̃) = b1 (κ̃) + log P(κ̃) . (68b)
∂κ̃ This yields
In Figure 14 we illustrate the first composite non- cov(P(κ̃1 ), P(κ̃2 )) = ξ( ¯ b̃1,NG P)(κ̃1 )(b̃1,NG P)(κ̃2 )
Gaussian bias function (green) along with its ingredi- ξ¯2
ents, the first order bias b1 (blue) and the logarithmic + (b̃2,NG P)(κ̃1 )(b̃2,NG P)(κ̃2 ) (71a)
derivative of the PDF (red) as measured in one FLASK 2
realisation (data points) and predicted for a shifted log- ξ2
normal (dashed lines). We can see that for a mildly non- + [(b2 P)(κ̃1 )(q1 P)(κ̃2 ) + (q1 P)(κ̃1 )(b2 P)(κ̃2 )]
2
Gaussian field, this composite non-Gaussian bias func- ¯2
tion is more than an order of magnitude smaller than (ξ − ξ)
+ (b2 P)(κ̃1 )(b2 P)(κ̃2 ) (71b)
the raw bias b1 , such that there is no guarantee that 2
our result at leading order in ξ¯ constitutes the most + O({ξ 3 , ξ¯ ξ 2 , ξ¯3 , ξ¯ (ξ − ξ)
¯ 2 , (ξ − ξ)
¯ 3 }) ,
relevant correction. This can be established analyti-
cally for a shifted lognormal distribution using equa- where we defined the following NLO composite bias func-
tions (26) and (30), which gives tion only present for non-Gaussian fields

1 x(κ̃)

1
 1 ∂(b1 P) 1 ∂2P
b̃1,LN (κ̃) = − + 2 2 1− . (68c) b̃2,NG (κ̃) = b2 (κ̃) + 2 + . (71c)
κ̃ + s sσG (σκ̃ ) 1 + κ̃/s P(κ̃) ∂κ̃ P(κ̃) ∂κ̃2
For a Gaussian field, the first three lines in equation (71)
The minimum value is b̃1,LN (κ̃ ' 0) ' −1/s and in the do not contribute because ϕ1 (λ) = λ and ϕn≥2 = 0, so
limit of s → ∞ we recover the Gaussian result with van- we recover the leading order of equation (65). Note that
ishing b̃1 . while we ordered terms according to powers of ξ here,
there is no guarantee that the leading order term pro-
3.2.3. Next-to-leading order: general non-Gaussian case portional to ξ¯ will be larger than the next-to-leading or-
der terms proportional to ξ 2 and (ξ − ξ)¯ 2 , as highlighted
We now generalise the previous result up to next-to- by the Gaussian case where the leading order term van-
leading order, where we find ishes and the shifted lognormal case where we found the
ϕ(λ1 , λ2 , −λ1 − λ2 ) − ϕ(λ1 , −λ1 ) − ϕ(λ2 , −λ2 ) (69) term to be subdominant. Typically, power of averages
of the mean correlation are much smaller than averages
¯ 1 (λ1 )ϕ1 (λ2 ) + ξ[ϕ
=(ξ − ξ)ϕ ¯ 1 (λ1 ) − λ1 ][ϕ1 (λ2 ) − λ2 ]
3 At higher orders, this will generalise to a class of terms with
ξ2 products of positive powers of ϕ1 (λ − λ) and powers of ϕ1 (λ)2
+ [ϕ21 (λ1 )ϕ2 (λ2 ) + ϕ2 (λ1 )ϕ21 (λ2 )] + O(ξ 3 ) . and/or ϕn≥2 (λ).
2
17

of powers, so ξ¯2  ξ 2 ≈ (ξ − ξ) ¯ 2 as evident from Fig- which can partially counteract the previous effect, but
ure 12 and for a weakly non-Gaussian field the b̃2,NG term was found to be negligible in our setting. Beyond direct
will only give a small contribution. For the mildly non- effects for the width of the likelihood contours, noise also
Gaussian κ PDF we are considering here, we observe that leads to an uncertainty for their location in parameter
the first order term is effectively removed by the mean- space (Dodelson & Schneider 2013b), which is only neg-
subtraction, as evident from the visual inspection of the ligible if Nsim − Nd  Nd − Np as is the case for our
covariance matrix shown in Figure 6 and the dominant simplified setting with a short data vector.
eigenvectors, which resemble the full sky case shown in
4.1. Eigendecomposition of the covariance matrix
green/olive in the lower panel of Figure 16. Given the
similarity between the Gaussian and FLASK correlation The eigendecomposition of the covariance matrix of di-
matrices, we can deduce that the b2 term given by equa- mension n × n (set by the length n of the data vector)
tion (58) for the shifted lognormal case is in fact the can be written in terms of a complete set of eigenval-
dominant contribution. ues λi (for convenience ordered in descending order, so
Our results from this section can largely be carried over λ1 > . . . > λn ) and associated normalised eigenvectors
to the covariance modelling for the 3-dimensional mat- ei arranged in an orthogonal matrix O
ter PDF on mildly nonlinear scales (roughly 10Mpc/h
at low redshifts), as we will discuss in Section 5 be- cov(P, P) = O · diag(λ1 , . . . , λn ) · OT , O = (e1 , . . . , en )
n
low. Typically, the matter PDF features stronger non- X
Gaussianities, such that the lognormal shift parameter = λk ek ek T . (73)
will be close to unity s3D . 1. In Appendix A we ad- k=1
ditionally discuss tree models as a useful class of non- In terms of components, we can write this as
Gaussian models for which all order covariance expansion
n n
results can be obtained. X X
λk ek ek T

cov(Pi , Pj ) = ij
= λk (ek )i (ek )j ,
4. QUANTITATIVE COVARIANCE MODEL TESTS k=1 k=1

Beyond performing a visual comparison between the (74)


covariances obtained from simulated maps and the theo- where (ek )i denotes the i-th component of the eigen-
retical models, we quantitatively compare their accuracy vector ek . The eigendecomposition can be understood
using in the context of the analytical results from the large-
separation expansion in Section 3. For the Gaussian
1. an eigendecomposition of the covariance validating
case, we obtained an expansion consisting of dyadic prod-
the set of eigenvalues and leading order eigenvec-
ucts with building blocks of the form Hen (x) exp −x2 /2
tors,
with x = κ/σ corresponding to (bn P)(κ). While the
2. a comparison of the dominant bands in the almost Hermite polynomials are orthogonal with respect to the
Gaussian weight function exp −x2 /2 corresponding to

band-diagonal precision matrix as inverse of the co-
variance, the PDF PG (κ), the dyadic product building blocks are
not generally orthogonal as they carry a combined weight
3. a χ2 -test (advocated in Friedrich et al. 2021) to of exp −x2 . However, the orthogonality of neighbour-
establish the Gaussianity of the PDF data vector ing bias contributions bn P and bn+1 P is ensured by the
and accuracy of the model covariance, and alternating symmetry of the Hermite functions.
4. a Fisher forecast for two cosmological parameters Considering the binned PDF means that there is just
{Ωm , σ8 }. a finite set of eigenvectors, which receive contributions
from a whole series of bn P terms. We illustrate that
When comparing our measurements to the predictions in Figure 15 by comparing the first three eigenvectors
in tests involving the inverse covariance or precision ma- (solid/dashed/dotted) of the PDF covariance for a Gaus-
trix, we account for that fact that matrix inversion is a sian κ field obtained from the full theory expression
non-linear operation, so the noise in the estimate of the (cyan) with the normalised bias functions (light grey).
covariance elements will lead to a bias in the elements of The leading order bias contribution b1 P agrees with the
the precision matrix. This bias from covariance estima- first eigenvector almost perfectly, while b2 P resembles
tion noise turns out to be just a factor multiplying the the overall shape of the second eigenvector with slightly
entire inverse covariance matrix, the so called Kaufman- modified zero crossings. This agreement reflects that
Hartlap factor (Kaufman 1967; Hartlap et al. 2006), there is a hierarchy in the contributions from different
powers of the correlation function compared to the vari-
h = (Nsim − Nd − 2)/(Nsim − 1) , (72)
ance ξ n /σ 2n as shown in Figure 12. The slight difference
where Nsim is the number of samples used and Nd is between b2 P and the second eigenvector is due to con-
the length of the data vector. This effect closely re- tributions from higher-order terms as we illustrate by
sembles the impact of the non-Gaussian distribution of showing the eigenvectors of the Hermite expansions (51)
a data vector that is a consequence of a finite number up to a finite nmax (shaded from dark grey to cyan with
of realisations (Sellentin & Heavens 2017) and of size increasing nmax ). The large difference between b3 P and
(Nsim − Nd + Np − 1)/(Nsim − 1), where Np is the num- the third eigenvector comes from its non-orthogonality
ber of parameters. Another effect is due to the nonlinear with b1 P as can be seen from the difference between
relationship between the precision matrix and the result- the dotted lines in light and darker grey. For the case
ing parameter covariance matrix (Percival et al. 2014), of the full sky (or the mean-subtracted κ in patches),
18

comparison covariance matrix eigenvectors & bias expansion


0.4 eigenvalues covariance patches 5deg, overlapping cells
Gauss
Gauss κ
0.3
κ, pred
10−1
κ full sky
0.2
κ full sky, pred

0.1
10−2
0.0

−0.1 EV Gauss theory


EV Hermite nmax = 26 10−3
−0.2 EV Hermite nmax = 13
EV Hermite nmax =6
−0.3 EV Hermite nmax =3
normalised bi P 10−4
−0.4
−0.02 −0.01 0.00 0.01 0.02 0 5 10 15 20 25
κ

Fig. 15.— First (solid), second (dashed) and third (dotted) eigenvectors covariance κ, overlapping cells
eigenvectors of the Gaussian κ PDF covariance predicted for small
Gauss
patches (cyan) and the normalised bias functions (bn P)G (κ) (light 0.3
grey). We also show the eigenvectors obtained from a Hermite ex-
pansion (51) up to order nmax (shaded from dark grey to cyan). 0.2
The vertical black dotted lines indicate where κ = ±σ.
0.1

0.0

−0.1
the smallness (or vanishing) of the mean correlation ξ¯ −0.2
κ patches
κ patches, pred
will practically remove the contribution of b1 P such that κ full sky
the first and second eigenvectors will be resembling b2 P −0.3
κ full sky, pred
and b3 P, respectively. This qualitative statement re- −0.4
mains true for mildly non-Gaussian fields such as our FLASK
shifted lognormal (FLASK) case. We show the eigen- 0.3
values and dominant eigenvectors for the Gaussian and 0.2
FLASK maps in Figure 16. The eigenvalues in the up-
0.1
per panel of Figure 16 show a good agreement between
measurements and theory and suggest a clear ordering 0.0
between contributions from different bias functions aris- −0.1
ing from a large-separation expansion, qualitatively re- κ patches
sembling the behaviour seen in Figure 12 for the aver- −0.2 κ patches, pred
ages of powers of the cell correlation in comparison with −0.3 κ full sky
the powers of the variance. The middle and lower pan- κ full sky, pred
−0.4
els of Figure 16 show the leading three eigenvectors for −0.02 −0.01 0.00 0.01 0.02
Gaussian (middle) and lognormal (bottom) κ fields in κ
small patches (blue/cyan) and the full sky (green/olive).
The first eigenvector (solid) is determined by b1 P and Fig. 16.— (Upper panel) Eigenvalues of the Gaussian κ PDF
covariance for κ from small patches (measurement in blue, predic-
hence shows one zero-crossing close to κ ' 0 as pre- tion in cyan) and the full sky (measurement in green, prediction
dicted by b1,G (22) and b1,LN (30). The second eigen- in olive). (Middle and lower panel) The dominant eigenvectors of
vector (dashed) is driven by b2 P with two zero-crossings the Gaussian (middle) and FLASK (lower) κ PDF covariance, in-
around κ ' σκ as predicted by b2,G (50) and b2,G (58). cluding the 1st eigenvector (solid lines), 2nd eigenvector (dashed
lines) and 3rd eigenvector for raw κ from small patches (dotted
The third eigenvector (dotted) involves b3 P, which is pre- lines). The move from small patches to the full sky (solid and
dicted to have three zero-crossings for the Gaussian case dashed lines) essentially removes the first term in the eigendecom-
from equation (48), but is also sensitive to higher-order position and replaces it with the second term as evident from the
terms as we illustrated in Figure 15. For the covariance agreement of the first green cross with the second blue cross, and
the agreement of the solid green and the dashed blue lines. The
of the κ PDF on the full sky we show the first two eigen- FLASK results are qualitatively similar to the Gaussian case ex-
vectors (measurements in green, predictions in olive). cept for subtle shifts in the location of the zero-crossings and the
We observe that the first eigenvector closely resembles asymmetry between positive and negative κ, both related to the
non-Gaussianity of the underlying field.
the second eigenvector of the κ PDF from small patches
(measurements in blue, predictions in cyan), as predicted
for Gaussian fields in equation (65) and for mildly non-
Gaussian fields in equation (71). Overall, the theoreti-
cal predictions are in good agreement with the measure- ing the skewness of the PDF. Due to the nature of the
ments from Gaussian and FLASK maps. The Gaussian eigendecomposition, the contribution of the finite sam-
and the lognormal cases are qualitatively similar with a pling term on the diagonal of the covariance matrix (11)
few important differences appearing mostly as shifts to- to individual eigenvectors is typically small, such that
wards negative κ (reflecting the PDF peak location) and the leading order results for the non-overlapping case
an asymmetry between positive and negative κ reflect- (not shown) resemble the overlapping case aside from
19

additional noise caused by the small number of sampling precision matrix, lognormal full sky
cells. FLASK 8000
In principle, one can reconstruct the covariance matrix -0.02
from the terms in the eigendecomposition, although in 6000
the presence of an incomplete set of eigenvectors (a num-
ber less than the length of the data vector), this matrix -0.01 4000
will not be invertible and hence unsuitable for statistical
2000
analysis. For the case of non-overlapping cells, there is
a large contribution to the diagonal from shot noise, so 0.0

κ2
0
one can attempt a covariance reconstruction using only
the first few eigenvectors along with a matched diago- −2000
nal mostly arising from shot noise. In practice, we opt 0.01
for the approach to use an invertible covariance baseline −4000
model, such as the ones obtained from the Gaussian or −6000
shifted lognormal model. If this was found to be insuffi- 0.02
ciently accurate, it could be successively improved using theory −8000
the precision matrix expansion (Friedrich & Eifler 2017)
-0.02 -0.01 0.0 0.01 0.02
or iteratively corrected by adjusting the dominant eigen- κ1
vectors to their desired values as predicted by a large-
scale expansion discussed in Section 3.
precision matrix diagonal bands, lognormal fullsky

4.2. Bands of the precision matrix 8000


diag+0, FLASK
diag+0, prediction
The accuracy of cosmological parameter estimation re- 6000
diag+1, FLASK
diag+1, prediction
lies on an accurate modelling of the precision (inverse diag+2, FLASK
covariance) matrix which enter the χ2 tests and Fisher 4000 diag+2, prediction
forecasts (Taylor et al. 2013). We show this precision 2000
matrix in the case of shifted lognormal fields from the
full sky (rescaled to the survey area) in the upper panel 0

of Figure 17 as measured in FLASK maps (upper trian- −2000


gle) and predicted (lower triangle). We notice that the
inverse covariance is close to band-diagonal with alter- −4000

nating signs. In the lower panel we focus on the three −6000


dominant bands (black, blue, red) comparing the predic- −0.02 −0.01 0.00 0.01 0.02
κ
tion (solid) to the measurements (dotted). We find sim-
ilar results for the precision matrix shape obtained from Fig. 17.— (Upper panel) Impression of the approximately band-
FLASK patches, and for both the raw κ and the mean- diagonal precision matrix obtained as inverse of the κ covariance
subtracted κ̃ = κ − κ̄ from overlapping cells. We can matrix for overlapping cells for shifted lognormal fields as predicted
(lower triangle) and measured in 500 FLASK full sky realisations
qualitatively understand the agreement between κ and κ̃ corrected by the Hartlap factor (72) (upper triangle). This matrix
in terms of the previously discussed eigendecomposition. is close to penta-diagonal, featuring 5 central bands, one diagonal
In principle, the precision matrix can be obtained from band (black) and 2 neighbouring bands above/below (blue, red).
the eigenvectors and eigenvalues of the covariance matrix We highlight those bands with with solid/dotted lines for the mea-
sured and predicted case. On the diagonal we show the mean of
as the FLASK measurements and the prediction. The measurement
from a finite number of realisations in the upper panel exhibits ad-
cov−1 (P, P) = O · diag(λ−1 −1 T
1 , . . . , λn ) · O , ditional noise off the diagonal. (Lower panel) Comparison of the
bands in the precision matrix between the the shifted lognormal
which means eigenvectors are retained but their order is theory prediction (dotted lines) and the measurement from FLASK
reversed. From the upper panel of Figure 16 we expect a realisations (solid lines).
flat hierarchy of the highest eigenvalues that determine
the inverse, with the lowest order eigenvectors (associ-
ated with the largest eigenvalues) playing a minor role
for the shape of the precision matrix. top-hat window function of each cell. We thus have
The band-diagonal structure of the precision matrix is
sin(kxi ) sin(kxj )
Z
mainly driven by the strong correlations of neighbouring −1
cov1D (xi , xj ) ∝ dk ∝ [max(xi , xj )] .
bins induced by the overlap of cells arising at small sepa- kxi kxj
rations. This can be inferred from a comparison with the (75a)
non-overlapping case and could be formalised by eval- This special covariance shape is called a hook matrix
uating the small separation contributions described in
1/x1 1/x2 1/x3 . . . 1/xn
 
Bernardeau (2022) for the analogous case of densities.
We can illustrate this effect in a simpler setting by com-  1/x2 1/x2 1/x3 . . . 1/xn 
 1/x 1/x 1/x . . . 1/x 
puting the covariance between 1D self-centered cells for cov1D ∝  3 3 3 n . (75b)
 . .. 
a Poisson (constant) power spectrum with no correlation  .. . 
between non-overlapping cells. Thus for two cells of half- 1/xn ... 1/xn
length xi and xj , their covariance is given by the integral
over the (constant) power spectrum multiplied by the 1D This matrix can be inverted by obtaining the cofac-
20

tor matrix C which determines the inverse as cov−1 1D = χ2 test covariance κ, O


CT / det cov1D . The elements of the cofactor matrix are 0.10
chi2, dof=26
given by Cramer’s rule in terms of determinants of sub- draw original cov
matrices obtained from cov1D by deleting the ith row 0.08 FLASK patches κ − κ̄
and jth column lognormal theory patches κ − κ̄
FLASK patches κ

Ci,j = (−1)i+j det(cov1D |i,j ) . (76) 0.06 lognormal theory patches κ

It can then be shown that indeed the cofactor elements


0.04
vanish everywhere except on the diagonal and the di-
rectly adjacent bands to the main diagonal. This is be-
cause the removal of a row i and column j in the co- 0.02
variance (75) leads to linearly dependent rows/columns
and hence Ci,j = 0 for |i − j| < 2. As the covariance is 0.00
symmetric, so is the cofactor matrix and hence the in- 5 10 15 20 25 30 35 40 45 50
χ2 for dof 26
verse of the covariance will have a band-diagonal struc-
ture. When the values of the radii xi (resembling κi ) are
an increasing sequence, the diagonal will have a positive χ2 test covariance κ, fullsky
0.10
sign and the band adjacent to the diagonal will have a chi2, dof=26
draw original cov
negative sign. FLASK fullsky
0.08
lognormal theory fullsky
4.3. χ2 tests
2
We will use a χ -test (advocated in Friedrich et al. 0.06

2021) to establish the Gaussianity of the data vector and


assess the accuracy of our predicted covariances. Let us 0.04
consider the χ2 associated with the original N samples
of the data vector. For every sample, labelled by i, of the 0.02
data vector, Dorg [i], of length n (encoding the number of
degrees of freedom), one computes the χ2
0.00
5 10 15 20 25 30 35 40 45 50
χ2org [i] = (Dorg [i] − µ)T · C −1 · (Dorg [i] − µ) , (77) χ2 for dof 26

using the mean data vector, µ (also of length n) and data Fig. 18.— χ2 -tests for the validation of our predicted covariances
covariance C (of dimension n × n). The distribution of and the Gaussianity of our data vector showing the χ2 distribution
(black line) and the result for a data vector drawn from a Gaussian
χ2 follows a χ2 distribution with the appropriate number with the measured covariance (black histogram). (Upper panel)
of degrees of freedom n (here the number of PDF bins). Covariance matrix test for the PDF of the convergence κ (blue) or
To test the Gaussianity of the data vector distribution, mean-subtracted convergence κ − κ̄ (purple) measured from over-
one can create another sample of data vectors, DG , by lapping cells computed in small patches comparing the FLASK
measurements (solid) to the predictions from a shifted lognormal
drawing N times from a multivariate Gaussian using the (dotted). (Lower panel) Covariance matrix test for overlapping
mean data vector, µ and covariance C. Then one recom- cells computed for the full sky comparing the measurements from
putes the χ2 for every sample using the covariance to 500 FLASK realisations (solid) to the lognormal prediction (dot-
check whether it follows the same distribution ted).

χ2G [i] = (DG [i] − µ)T · C −1 · (DG [i] − µ) . (78)


If so, one can assert that the data vector is ‘close enough’
to Gaussian distributed around the mean with the mea- heavily overlapping cells from FLASK patches was visu-
sured covariance. Both of those quantities will have the ally different from the Gaussian case, mainly driven by
mean χ2 = χ2org = n, since the covariance matrix is iden- the asymmetry in the diagonal of the covariance matrix
tical and is reproduced in the expectation value of the shown in Figure 5. In Figure 18 we show the χ2 test
fluctuations of the data vector around the mean. results for the overlapping case with FLASK covariances
One can also use a χ2 test to compare the closeness for patches (upper panel) and the full sky (lower panel).
of two covariances, where C denotes the original ‘true’ The agreement of the χ2 obtained from the measured
(or target) covariance and Cmod the model (or approx- covariances (coloured solid) and a draw from a Gaus-
imated) covariance. By recomputing the χ2 using the sian with the same variance (black) shows that the data
model covariance vectors are close to Gaussian distributed. We find good
χ2mod [i] = (Dorg [i] − µ)T · (Cmod )−1 · (Dorg [i] − µ) (79) agreement between the FLASK covariance measurements
(coloured solid) and the shifted lognormal predictions
and checking its distribution compared to the original (coloured dotted). The results for patches (upper panel)
one χ2org mentioned above, one can assess how well the show a slight (6-9%) bias of the theory towards higher χ2 ,
model and true covariance agree. We found that the which could be caused by the patch-to-patch correlation
covariance of non-overlapping cells from the FLASK arising from cutting multiple patches from each FLASK
patches is well captured in both the Gaussian and shifted full sky that is unaccounted for in the theory. Interest-
lognormal theory, as it is dominated by the finite sam- ingly, we find no evidence for a corresponding discrep-
pling term and thus not shown here. The covariance of ancy in the Fisher forecasts for those two cases, demon-
21

strating the complementarity of the two tests adopted in


Friedrich et al. (2021). zs = 2, A = 15000 deg2
FLASK patches NO
4.4. Fisher forecast validation FLASK patches O
FLASK patches O κ − κ̄
To forecast the errors on a set of cosmological param- FLASK full sky
eters, p~, we use the Fisher matrix formalism. The Fisher
matrix given a (set of) summary statistics in the data
vector D ~ is defined as
X ∂Dα ∂Dβ
Fij = (C −1 )αβ , (80)
∂pi ∂pj
α,β

where Dα is the α-th element of the data vector D, ~ C −1

90
0.
denotes the precision matrix introduced in Section 4.2,
which is the matrix-inverse of the data covariance C

σ8
85
0.
given in equation (17) and is was assumed that the covari-
ance matrix is independent of cosmology. The Fisher for-

80
0.
malism rests on the assumption that the summary data
vector is Gaussian distributed which we have explicitly

75
0.
checked in the last subsection for the central region of

20

24

28

32

36
75

80

85

90
the weak lensing PDF considered here. The parameter

0.

0.

0.

0.

0.
0.

0.

0.

0.
covariance matrix C(~ p) is then obtained as inverse of the Ωcdm σ8
Fisher matrix, CijP
= (F −1 )ij . In the Fisher formalism,
Fig. 19.— Fisher forecasts comparing different FLASK-generated
marginalisation over a subset of parameters is achieved covariances using non-overlapping cells (cyan), overlapping cells
by simply selecting the appropriate sub-elements of the (blue), overlapping cells with mean-subtraction (purple) and over-
parameter covariance. lapping cells on full sky maps (green). All covariances were rescaled
To further validate our theoretical covariances, we per- to mimic the same survey area.
form simple Fisher forecasts for two cosmological param-
eters {Ωcdm , σ8 } with the measured and predicted co-
variance matrices, while keeping the data vector and its zs = 2, A = 15000 deg2
dependence on cosmological parameters in (80) as pre- FLASK patches κ − κ̄
dicted from large-deviation statistics (see Boyle et al. Lognormal patches κ − κ̄
2021).4 The Fisher forecasts in Figure 19 demonstrate FLASK full sky
Lognormal full sky
that measuring non-overlapping cells only (cyan) leads FLASK full sky noisy
to much wider parameter errors than including cell over- Lognormal full sky noisy
laps (blue), such that capturing cell overlaps is imper-
ative to avoid artificially inflating the error budgets (as
already pointed out in Szapudi & Colombi 1996). Addi-
tionally we show that when measuring covariances from
small patches obtained from FLASK and subtracting the
88
0.

mean in each small patch (purple) this reproduces the re-


sult obtained from full sky maps rescaled to the survey
86
0.

area (green). In Figure 20 we show that our shifted log-


σ8
84

normal predicted covariances (dotted) for small patches


0.

with mean-subtraction (purple) and the full sky (green)


82

agree well with the FLASK measurements (solid). We


0.

also illustrate the impact of shape noise (orange) and


80

validate our convolution approach from equation (42)


0.
22

24

26

28

30

80

82

84

86

88

against FLASK measurements with added shape noise.


0.

0.

0.

0.

0.

0.

0.

0.

0.

0.

Ωcdm σ8
5. EXTENSION TO 3D CLUSTERING PDF
COVARIANCE Fig. 20.— Fisher forecasts validating the predicted covariances
(dotted) against FLASK measurements (solid) for shifted lognor-
In this Section, we describe the essential steps needed mal fields on small patches (purple), the full sky (green) and in-
to adapt our covariance model to 3-dimensional fields cluding shape noise (orange).
using the 3D matter PDF as an example. While the
line-of-sight projection involved in computing the weak
4 The predicted derivatives rely on large-deviations theory for
the convergence PDF (Barthelemy et al. 2020a). Using CLASS lensing convergence and photometric galaxy density on
(Blas et al. 2011), linear and non-linear (Halofit, Peacock & Smith
2014) matter density variances are calculated to rescale the matter
mildly nonlinear scales tends to Gaussianise the under-
density CGF in each redshift slice. The convergence CGF is ob- lying field, 3D matter and spectroscopic galaxy densities
tained by integration over the redshift slices, and translated to the typically show a greater degree of non-Gaussianity on
final PDF by an inverse Laplace transform. mildly nonlinear scales. This might complicate the two-
22
1.2
2020), and can be improved by including a shift factor
large-deviation theory, σμ =0.61 below unity (dashed black line).6 The good match for
1.0
the fiducial PDF in Figure 21 and the leading order bias
shifted lognormal, σμ =0.61, s=0.9
in Figure 23 suggests that the shifted lognormal model
(ρ), z=0, R=10 Mpc/h

0.8 can be used to obtain accurate analytical PDF covari-


lognormal, σμ =0.61
ance estimates. This can be done by computing the joint
0.6 PDF (82) following a similar lognormal matching proce-
dure as sketched for the mean-subtracted weak lensing
0.4 convergence κ̃. For 3D observables for which the shifted
lognormal model would not yield a good match of the
0.2 PDF and leading order bias, the class of tree models dis-
cussed in Appendix A and Bernardeau (2022) offer an
0.0 alternative with the capability of skewness matching.
0.0 0.5 1.0 1.5 2.0 2.5 3.0
ρ 5.2. Relevance and prediction of super-sample
Fig. 21.— 3D matter density PDF at redshift z = 0 and smooth- covariance
ing radius R = 10 Mpc/h as measured from the Quijote simula-
tions (data points, with error bars indicating the sample variance), Measurements. The covariance measured from sim-
predicted by large-deviation theory (black line) and approximated ulated boxes with fixed mean density (due to a fixed
with a lognormal distribution (black dotted) and a shifted lognor- particle number) ρ̄ cannot capture the super-sample ef-
mal distribution with tuned shift parameter (black dashed). fect created by having different sample densities ρs . To
estimate this super-sample covariance effect for the Qui-
jote simulation suite, separate-universe style ‘DC’ runs
point PDF modelling, but we will show that the previ- emulating a background density contrast δb = ±0.035
ously discussed shifted lognormal model from Section 2.3 have been performed using changed cosmological param-
has flexibility to match the skewness of the one-point eters and simulation snapshot times from the separate
distribution and capture the clustering properties. Fi- universe approach (Sirko 2005). The super-sample co-
nally, we will focus on the theoretical modelling of the variance between two data vector entries Di and Dj can
super-sample covariance, which can be used to comple- be estimated by
ment simulated covariances from periodic boxes missing ∂Di ∂Dj
this effect. covSSC 2
SU (Di , Dj ) = σb , (81)
∂δb ∂δb
5.1. Applicability of the shifted lognormal model where σb2 is the variance of δb (here just δb2 ), and the
The three-dimensional matter PDF is typically ex- other two terms encode the linear response of the data
tracted in terms of the normalised cell density ρ̂ = ρ/ρs vector, which can be determined from the simulations
as ratio of the physical cell density ρ and the mean using finite differences.
sample density ρs , or the associated zero-mean density In the lower triangle of the lower panel of Figure 22 we
contrast δ = ρ̂ − 1. For illustration purposes, we use show the covariance matrix of the matter density PDF
previously obtained measurements for the matter PDF showing the correlation between different density bins
at redshift z = 0 in spheres of radius R = 10 Mpc/h in the central region of the PDF. We see that, as ex-
(Uhlemann et al. 2020) in the Quijote simulation suite pected, neighbouring bins are positively correlated, while
(Villaescusa-Navarro et al. 2019). The one-point PDF of intermediate underdense and overdense bins are anticor-
the normalised density5 ρ̂ was extracted in logarithmi- related with each other. The strong positive correlation
cally spaced bins designed to equally capture under- and in the bands along the diagonal is induced by cell over-
overdensities in the strongly non-Gaussian distribution laps, similarly as discussed previously for the analogous
and means and covariances obtained from 15,000 realisa- weak lensing case. We illustrate the separate universe
tions at the fiducial cosmology. super-sample effect on the PDF covariance in the up-
In Figure 21 we give a visual impression of the mean per triangle of Figure 22, showing that the super-sample
PDF for the normalised matter density ρ̂ as measured covariance term completely dominates the overall covari-
from the simulations (data points) and large-deviation ance for a volume of 1(Gpc/h)3 and will still lead to a
theory prediction (black line, following Uhlemann et al. significant effect for a volume of O(10)(Gpc/h)3 .
2016, 2020). To describe clustering observables with the Predictions. Predicting the covariance of the PDF of
shifted lognormal model, we can simply replace the weak the normalised matter density ρ̂ in a three-dimensional
lensing convergence κ with the density contrast δ in equa- volume is conceptually analogous to the previous two-
tions (26) and (28). Choosing a unit shift parameter dimensional case relevant for weak lensing, so we can
s = 1 corresponds to the usual lognormal approximation write the covariance in terms of the two-point PDF as in
(dotted black line, Coles & Jones 1991). This already equation (6) by replacing the weak lensing convergence
provides a decent phenomenological fit to first princi- with the density κ → ρ and angular by 3D-distances
ple predictions for the matter PDF from large deviation θ → r, where Pd (r) indicates the distribution of distances
statistics (Bernardeau et al. 2014; Uhlemann et al. 2016, r in a given survey volume. The distance distribution in
5 ensuring the compatibility of measurements in simulations with 6 The lognormal model is however insufficient to accurately pre-
different particle numbers (changing the mean density of particles dict the response to cosmological parameters (see Uhlemann et al.
in a sphere). 2020).
23

matter PDF covariance P(Ω), z=0, R=10 Mpc/h super-sample covariance ingredients, z=0, R=10 Mpc/h
3
0.3 SSC ∂ log P
pred b̂1 pred b1 pred
∂ log ρ̂
∂ log P b̂1 sim b1 lognormal
2 sim
0.0004 ∂ log ρ̂
∂ log P b1 sim
0.39 ∂δb sim

1
0.52 0.0002

0
Ω2

0.76 0.0000
−1
1.0
°0.0002
−2
1.32
°0.0004
−3
0.3 0.39 0.52 0.76 1.0 1.32 1.92
1.92 sim x10 ρ̂
0.3 0.39 0.52 0.76 1.0 1.32 1.92
Ω1 Fig. 23.— Comparison of the theoretically predicted sphere bias
(blue solid), the lognormal approximation (blue dashed) and the
measurement in one realisation of the fiducial cosmology of Qui-
Fig. 22.— Comparison of the PDF covariance matrix contribu- jote (blue data points) at radius R = 10 Mpc/h, distance d = 20
tions for z = 0, R = 10 Mpc/h as measured from the 1(Gpc/h)3 Mpc/h and redshift z = 0. Additionally shown is the logarithmic
Quijote simulations and the super-sample covariance contribution derivative of the PDF (red) as predicted (solid) and measured (data
constructed from the DC mode following equation (81). The upper points) and the resulting prediction from the effective bias (83) us-
triangle shows the super-sample covariance contribution while the ing the theoretical ingredients (green solid) or the simulated ones
lower triangle shows the measured covariance from the simulations (green dotted). The predictions agrees very well with the mea-
enhanced by a factor of 10, while the diagonal from the upper panel surements from separate universe simulations (green data points).
is masked. The vertical dotted line indicates the peak location of the PDF.
The ρ-range shown equals the one used in the covariance plots in
Figure 22.

a cubic box is easily obtained from the one for a 3D


unit cube that corresponds to the problem of cube line
picking for which a closed-form expression is available
(Mathai et al. 1999; Weisstein 2022). Similarly to the covariance (81) is well predicted by the effective bias b̂1
weak lensing case, where the mean-subtraction leads to from equation (83) using theoretical ingredients (green
a modification of the relevant bias functions determining solid line) or simulated ones (green dotted line). We also
the covariance, the relative density definition will result show the b̂1 ingredients consisting of the first order sphere
in modified bias functions. This can be predicted by bias function b1 as measured from the correlation be-
considering the matter PDF in different sample densities tween neighbouring spheres in one realisation (blue data
ρs , where the joint PDF of normalised densities ρ̂i is points) in comparison to the large-deviation theory pre-
obtained from the trivariate PDF of physical densities diction (blue solid line, following Uhlemann et al. 2017)
ρi = ρs ρ̂i and the background density ρs by integrating and the lognormal approximation (blue dashed line), and
out the latter the logarithmic derivative of the one-point PDF as mea-
Z sured from finite differences (red data points) and pre-
P(ρ̂1 , ρ̂2 ) = dρs ρ2s P(ρs , ρ1 = ρs ρ̂1 , ρ2 = ρs ρ̂2 ) . (82) dicted from large-deviation theory (red solid line). We
conclude that the super-sample covariance effect can be
The super-sample covariance effect (13) for the PDF of robustly predicted by the theory presented here and that
the normalised matter density ρ̂ is obtained from the the shifted lognormal model holds promise to compute
leading order effective bias function in the resulting co- analytical covariances for the PDF of 3D clustering ob-
variance, given by equation (28) in Bernardeau (2022) servables.
Applications. The total covariance matrix is a sum
∂ log P(ρ̂) of the simulated covariance, Csim and the super-sample
b̂1 (ρ̂) = 1 + + b1 (ρ̂) . (83) covariance term consisting
∂ log ρ̂ p of a dyadic product, CSSC =
T
vSSC vSSC with vSSC = ξ( ¯ b̂1 P)(ρ̂). This sum can be
Note that while the expression looks similar to the one inverted using the Sherman-Morrison formula (Sherman
for the mean-subtracted weak lensing case b̃1,NG (κ̃) from & Morrison 1950) to obtain the precision matrix
equation (68b), this contribution does not vanish even
−1 −1
for a Gaussian field. Predictions for the ingredients, the −1 −1 Csim CSSC Csim
one-point PDF P(ρ) and the sphere bias b1 (ρ), can be ob- (Csim + CSSC ) = Csim − T C −1 v
. (84)
1 + vSSC sim SSC
tained from large-deviation statistics (Codis et al. 2016b;
Uhlemann et al. 2017), shifted lognormal models (Sec- The result can be used to predict the super-sample co-
tion 2.3) and tree models (Appendix A and Bernardeau variance impact on statistical tests like χ2 and Fisher
2022). contours, which generally leads to a degradation of pa-
Comparison. In Figure 23 we demonstrate that the rameter constraints as detailed in (Lacasa & Grain 2019)
derivative of the PDF with respect to the background for the case of Euclid-like photometric galaxy clustering
density (green data points) determining the super-sample observables.
24

6. CONCLUSION & OUTLOOK Based on our PDF estimator, one can further attempt to
Conclusion. In this work, we have shown how to com- model cross-covariances between the one-point PDF and
pute covariances for the one-point PDF of the weak lens- 2-point statistics, which in the mildly non-linear regime
ing convergence on mildly nonlinear scales. Our covari- are dominated by the correlation between the PDF and
ance predictions (6) rely on the two-point PDF, which the variance (see Figure 12 in Uhlemann et al. 2020).
is integrated over all separations according to the distri- Note that our covariance model and its possible exten-
bution of distances in the survey area. Our formalism sions can serve as a starting point for improving covari-
includes predictions for the key effects including finite ance estimates from simulations, even in regimes where
sampling (11) and super-sample covariance (13) which the model itself becomes inaccurate. The Gaussian and
were shown to scale with the inverse survey area. We shifted lognormal covariance models can be used as a
presented effective models for this joint PDF for a Gaus- starting point for a precision matrix expansion (Friedrich
sian field (19), a shifted lognormal field (28) matching & Eifler 2017) in which the leading order eigenvectors
the PDF skewness and density-dependence of the two- could iteratively refine this baseline covariance model.
point clustering (Figure 8), and a field affected by shape Additionally, analytical models can be used to construct
noise (42) from galaxy shape measurements. Going be- priors for a Bayesian covariance estimation with the help
yond (log-)normal models, we laid out a large-separation of cheap surrogates (Chartier & Wandelt 2022).
expansion for the two-point PDF that captures the key
effects of the PDF covariance for general non-Gaussian ACKNOWLEDGEMENTS
fields. The validity of our predictions has been estab- This work has made use of the Infinity Cluster
lished through a set of tests including visual inspection hosted by Institut d’Astrophysique de Paris. We thank
of the covariance, leading order eigenvalues and eigen- Stéphane Rouberol for smoothly running this cluster for
vectors (Figure 16), the precision matrix (Figure 17), χ2 us. We thank the authors of Takahashi et al. (2017)
tests (Figure 18) and a two-parameter Fisher forecast for their publicly available lensing maps, Villaescusa-
(Figure 20). With this suite of tests we have demon- Navarro et al. (2019) for their publicly available simu-
strated that our PDF covariance model not only accu- lation suite products, and Xavier et al. (2016) for their
rately describes large sets of simulated FLASK (Xavier publicly available FLASK code. The figures in this work
et al. 2016) maps, but also measurements of realistic full were created with matplotlib (Hunter et al. 2007) mak-
sky convergence maps obtained from full N -body simu- ing use of the numpy (Harris et al. 2020), scipy (Vir-
lations (Takahashi et al. 2017). tanen et al. 2020), sci-kit image (van der Walt et al.
Finally, we have sketched how the formalism described 2014) and ChainConsumer Python packages.
here for mildly non-Gaussian weak lensing observables is CU is supported by the STFC Astronomy Theory Con-
also applicable to more strongly non-Gaussian 3D clus- solidated Grant ST/W001020/1 from UK Research & In-
tering quantities. Most importantly, we showed how this novation. CU’s work was performed in part at Aspen
can yield predictions for the super-sample covariance ef- Center for Physics, which is supported by National Sci-
fect (Figures 22 and 23) to complement covariance mea- ence Foundation grant PHY-1607611 and partially sup-
surements from finite-box simulations. ported by a grant from the Simons Foundation. AG
Outlook. We expect our formalism can be straightfor- is supported by an EPSRC studentship under Project
wardly adapted to describe the one-point PDF of related 2441314 from UK Research & Innovation. SC thanks
weak-lensing observables such as the CMB weak lensing Fondation Merac and the French Agence Nationale de la
convergence (Barthelemy et al. 2020b) and aperture mass Recherche (grant ANR-18-CE31-0009) for funding. The
(Barthelemy et al. 2021), as well as photometric galaxy authors thank the Euclid work packages for additional
clustering. Our model can be extended to describe the galaxy clustering probes and higher-order weak lensing
cross-correlations between weak lensing PDFs at differ- statistics, as well as the Rubin LSST:DESC higher-order
ent smoothing scales and/or tomographic redshift bins. statistics group for useful discussions.
REFERENCES
Amendola L., et al., 2018, Living Reviews in Relativity, 21 Chartier N., Wandelt B. D., 2022, MNRAS, 515, 1296
Baratta P., Bel J., Plaszczynski S., Ealet A., 2020, A&A, 633, A26 Clerkin L., et al., 2016, Monthly Notices of the Royal
Bardeen J. M., Bond J. R., Kaiser N., Szalay A. S., 1986, ApJ, Astronomical Society, 466, 1444–1461
304, 15 Codis S., Pichon C., Bernardeau F., Uhlemann C., Prunet S.,
Barthelemy A., Codis S., Uhlemann C., Bernardeau F., Gavazzi 2016a, MNRAS, 460, 1549
R., 2020a, MNRAS, 492, 3420 Codis S., Bernardeau F., Pichon C., 2016b, MNRAS, 460, 1598
Barthelemy A., Codis S., Bernardeau F., 2020b, MNRAS, 494, Colavincenzo M., et al., 2019, MNRAS, 482, 4883
3368 Coles P., Jones B., 1991, MNRAS, 248, 1
Barthelemy A., Codis S., Bernardeau F., 2021, MNRAS, 503, Desjacques V., Jeong D., Schmidt F., 2018, Phys. Rep., 733, 1
5204 Dodelson S., Schneider M. D., 2013a, Phys. Rev. D, 88, 063537
Bernardeau F., 1996, A&A, 312, 11 Dodelson S., Schneider M. D., 2013b, Phys. Rev. D, 88, 063537
Bernardeau F., 2022, arXiv e-prints, p. arXiv:2207.02500 Friedrich O., Eifler T., 2017, Monthly Notices of the Royal
Bernardeau F., Reimberg P., 2016, Phys. Rev. D, 94, 063520 Astronomical Society, 473, 4150
Bernardeau F., Valageas P., 2000, A&A, 364, 1 Friedrich O., et al., 2018, Phys. Rev. D, 98, 023508
Bernardeau F., Pichon C., Codis S., 2014, Phys. Rev. D, 90, Friedrich O., Uhlemann C., Villaescusa-Navarro F., Baldauf T.,
103519 Manera M., Nishimichi T., 2020, MNRAS, 498, 464
Bernardeau F., Codis S., Pichon C., 2015, MNRAS, 449, L105 Friedrich O., et al., 2021, MNRAS, 508, 3125
Bertschinger E., 1987, ApJ, 323, L103 Gatti M., et al., 2019, arXiv e-prints, p. arXiv:1911.05568
Blas D., Lesgourgues J., Tram T., 2011, Journal of Cosmology Gruen D., et al., 2018, Phys. Rev. D, 98, 023507
and Astroparticle Physics, 2011, 034–034 Halder A., Friedrich O., Seitz S., Varga T. N., 2021, MNRAS,
Boyle A., Uhlemann C., Friedrich O., Barthelemy A., Codis S., 506, 2780
Bernardeau F., Giocoli C., Baldi M., 2021, Monthly Notices of Harris et al. C. R., 2020, Nature, 585, 357
the Royal Astronomical Society, 505, 2886
25

Hartlap J., Simon P., Schneider P., 2006, Astronomy & Sellentin E., Heavens A. F., 2017, MNRAS, 464, 4658
Astrophysics, 464, 399–404 Sherman J., Morrison W. J., 1950, The Annals of Mathematical
Hilbert S., Hartlap J., Schneider P., 2011, Astronomy & Statistics, 21, 124
Astrophysics, 536, A85 Sheth R. K., Tormen G., 1999, MNRAS, 308, 119
Hoffman Y., Ribak E., 1991, ApJ, 380, L5 Sirko E., 2005, ApJ, 634, 728
Hou J., Cahn R. N., Philcox O. H. E., Slepian Z., 2021, arXiv Szapudi I., Colombi S., 1996, ApJ, 470, 131
e-prints, p. arXiv:2108.01714 Szapudi I., Colombi S., Bernardeau F., 1999, Monthly Notices of
Hunter et al. J. D., 2007, Computing in Science & Engineering, 9, the Royal Astronomical Society, 310, 428
90 Takahashi R., Hamana T., Shirasaki M., Namikawa T.,
Ivezić Ž., et al., 2019, ApJ, 873, 111 Nishimichi T., Osato K., Shiroyama K., 2017, The
Jamieson D., Loverde M., 2020, Physical Review D, 102 Astrophysical Journal, 850, 24
Kaiser N., 1984, ApJ, 284, L9 Taylor A., Joachimi B., 2014, MNRAS, 442, 2728
Kaufman G. M., 1967, Report No. 6710, Center for Operations Taylor A., Joachimi B., Kitching T., 2013, MNRAS, 432, 1928
Research and Econometrics, Catholic University of Louvain, Thiele L., Hill J. C., Smith K. M., 2020, Phys. Rev. D, 102,
Heverlee, Belgium 123545
Lacasa F., Grain J., 2019, Astronomy & Astrophysics, 624, A61 Uhlemann C., Codis S., Pichon C., Bernardeau F., Reimberg P.,
Laureijs R., et al., 2011, preprint, (arXiv:1110.3193) 2016, MNRAS, 460, 1529
Martin S., Schneider P., Simon P., 2012, Astronomy & Uhlemann C., Codis S., Kim J., Pichon C., Bernardeau F.,
Astrophysics, 540, A9 Pogosyan D., Park C., L’Huillier B., 2017, MNRAS, 466, 2067
Mathai A. M., Moschopoulos P., Pederzoli G., 1999, Statistica, Uhlemann C., Pichon C., Codis S., L’Huillier B., Kim J.,
59, 61 Bernardeau F., Park C., Prunet S., 2018, MNRAS, 477, 2772
Mo H. J., White S. D. M., 1996, MNRAS, 282, 347 Uhlemann C., Friedrich O., Villaescusa-Navarro F., Banerjee A.,
Peacock J. A., Smith R. E., 2014, HALOFIT: Nonlinear Codis S., 2020, MNRAS, 495, 4006
distribution of cosmological mass and galaxies (ascl:1402.032) Villaescusa-Navarro F., et al., 2019, arXiv e-prints, p.
Percival W. J., et al., 2014, MNRAS, 439, 2531 arXiv:1909.05273
Percival W. J., Friedrich O., Sellentin E., Heavens A., 2022, Virtanen et al. P., 2020, Nature Methods, 17, 261
MNRAS, 510, 3207 Weisstein E. W., 2022, Cube Line Picking. From MathWorld—A
Repp A., Szapudi I., 2021a, MNRAS, 500, 3631 Wolfram Web Resource,
Repp A., Szapudi I., 2021b, Monthly Notices of the Royal http://mathworld.wolfram.com/CubeLinePicking.html
Astronomical Society, 509, 586 Xavier H. S., Abdalla F. B., Joachimi B., 2016, MNRAS, 459,
Sarma Boruah S., Rozo E., Fiedorowicz P., 2022, arXiv e-prints, 3693
p. arXiv:2204.13216 van der Walt et al. S., 2014, PeerJ, 2, e453

The two-point PDF can be obtained from the two-point


CGF which can be written in terms of the individual
one-point CGFs
APPENDIX ϕ(λ1 ) + ϕ(λ2 ) + ξ12 ϕ(λ1 )ϕ(λ2 )
ϕt (λ1 , λ2 ) = 2 ϕ(λ )ϕ(λ )/4 . (A3a)
A. LARGE-SEPARATION EXPANSION IN THE 1 − ξ12 1 2
MINIMAL TREE MODEL By expanding that expression in the correlation ξ12 , one
In addition to shifted lognormal models used heavily in can obtain the two leading order bias functions and the
the main text, let us discuss one representative from the additional q1 term in the large-separation covariance ex-
class of so-called tree models considered in Bernardeau pansion (56)
(2022) in the context of PDF covariance modelling. The 4ρ

minimal tree model is the simplest example for which 0 F1 1, σ 4 2
b1,t (ρ) = 4ρ − σ 2 , (A4a)

analytical results can be obtained at all orders. Here, we 0 F1 2, σ 4
show how the phenomenology of tree models resembles 4[ρ − 1 − σ 2 b1 (ρ)] 4 ρ−1
 
predictions from large-deviation theory for the matter b2,t (ρ) = = − b 1 (ρ)
PDF (see the upper panel of Figure 24). Additionally, σ4 σ2 σ2
we present a closed-form expression for the n-th order
" #


4 ρ + 1 0 F1 1, σ4
bias parameter bn,t (A14) that can be used to construct = 2 − 4ρ (A4b)
σ σ2

the full large separation series expansion yielding an in- 0 F1 2, σ 4
vertible analytical covariance model. q1,t (ρ) = b1,t (ρ)/2 . (A4c)
The one-point cumulant generating function of the
minimal tree model is given by In the limit where σ 2 → 0, those expressions tend to
the following results, which can be further simplified for
λ σ2 2 σ4 3 small density contrasts δ = ρ − 1 to reproduce the Gaus-
ϕt (λ) = ' λ+ λ + λ +O(σ 6 ) , (A1)
1 − λσ 2 /2 2 4 sian result
|{z} √
κ3 /3! σ→0 2( ρ − 1) δ→0 δ
b1,t (ρ) −→ −→ 2 = b1,G (δ) , (A5a)
σ2 σ
where the expansion shows that the mean is one, σ 2 is the √ 2 2 2 2
variance7 and the reduced skewness is S3 = κ3 /σ 4 = 3/2. σ→0 [2( ρ − 1)] − σ δ→0 δ − σ
b2,t (ρ) −→ 4
−→ 4
= b2,G (δ) .
The one-point PDF of the matter density in the minimal σ σ
tree model can be computed in closed form and is given (A5b)
by a hypergeometric function The two stages of this limiting behaviour are illustrated
4

2
 

 in Figure 24 by a sequence of decreasing σ 2 (from red
Pt (ρ) = 4 exp − 2 (1 + ρ) 0 F1 2, 4 . (A2) to green) showing fast approach to the small variance
σ σ σ limit (black solid) which reproduces the Gaussian result
7 Note that Bernardeau (2022) uses ξ̄ to denote the variance, (black dashed) for small density contrast. The quadratic
which should not be confused with our average correlation ξ̄ from q1 term in the covariance (56) can be combined with the
equation (13b). b1 term (and hence be captured by the first eigenvector),
26

1.0
and for the minimal tree model the Ij (ρ) can be obtained
σ 2 =0.45 as
   
σ 2 =0.1 4j 2 4jρ
0.5 Ij (ρ) = 4 exp − 2 (j + ρ) 0 F1 2, 4 . (A9)
σ σ σ
2( ρ -1)
b1,t (ρ)σ 2

0.0
LDT
We can therefore obtain a closed form equation for bn (ρ)
by differentiation of equation A9. The nth derivative of
ρ-1 Ij is given by
-0.5
n X k   
dn Ij X n k (n−k)
-1.0 = f (j)g (k−`) (j)h(`) (j) ,
dj n k `
k=0 `=0
0.5 1.0 1.5 2.0 (A10)
ρ where we’ve split Ij into three functions, f, g, h, which
1.5 we can differentiate separately as follows
 
σ 2 =0.45 4j 2
f (j) = 4 , g(j) = exp − 2 (j + ρ) , (A11a)
σ 2 =0.1
σ σ
1.0
dn f 4
b2,t (ρ)σ 4 +σ 2

[2( ρ -1)]2 = 4 (jδn,0 + δn,1 ) , (A11b)


dj n σ
(ρ-1)2 dn g −2
n
   
2
0.5 = exp − 2 (j + ρ) (A11c)
dj n σ2 σ
For the hypergeometric function we make use of the fact
that d/dz 0 F1 (n, z) = 0 F1 (n + 1, z) /n, where we defined
0.0
z = 4ρj/σ 4 so that d/dj = σ4ρ4 d/dz and
0.5 1.0 1.5 2.0  
ρ 4ρj
h(j) = 0 F1 2, 4 , (A12a)
Fig. 24.— A comparison of the first two leading order bias σ
functions from the tree model (A4), b1,t (upper panel) and b2,t  n
dn h
 
(lower panel), for different variances (coloured lines) approaching 4ρ 1 4ρj
= F
0 1 2 + n, . (A12b)
the small-variance limit (A5) (green) and the Gaussian expectation dj n σ4 (n + 1)! σ4
(dashed line). In the upper panel we also show the measurements
for b1 from the Quijote simulations (data points) and the large- Putting this into the full sum we obtain (using the Kro-
deviation theory prediction (black line) as in Figure 23, but in necker δs to collapse the k sum, as the only surviving
linear rather than logarithmic scale.
terms are when n = k and when n − k = 1).
n k   
dn Ij 4 − 22 (j+ρ) X X n k
= e σ (jδn−k,0 + δn−k,1 )
dj n σ4 k `
while the q1 term adding to b2,t in (56) does not change k=0 `=0
much in the case of a small variance σ 2  1 and an
k−`  k 4jρ

−2 0 F1 k + 2, σ 4


intermediate density regime. × . (A13)
σ2 σ4 (k + 1)!
For the minimal tree model, one can deduce that the
covariance at any order is given in terms of the set of bias Now, we separate this sum into the case where k = n and
functions bn,t (ρ). A recursion relation can be obtained when k = n−1, factor out common terms only depending
as shown in Bernardeau (2022), which can be converted on n and use the binomial theorem to do these sums as
to an explicit closed-form expression as we show here. simply (1 − 2/σ 2 )n and (1 − 2/σ 2 )n−1 respectively. This
Derivation of closed form bias functions. Define an equation is then the form of bn,t (ρ)Pt (ρ) when j = 1. Di-
auxilliary version of the Laplace transform of the CGF viding through by Pt (ρ) from equation (A2) everywhere
as the integral Ij (ρ): and setting j = 1 gives us the functional form of the n-th
order bias function for the minimal tree model as

Z
Ij (ρ) = exp [−λρ + jϕ(λ)] . (A6)    n−1
2πi 1 4ρ 2
bn,t (ρ) = 1− 2 (A14)
n! σ4 σ
The PDF is then the value I1 (ρ). In the minimal tree " #
4ρ 1 − σ22 0 F1 n + 2, σ4ρ4 4ρ
 
model, the bias functions are predicted via 0 F1 n + 1, σ 4

 + 4ρ .
σ 4 (n + 1)

dλ 0 F1 2, σ 4 0 F1 2, σ 4
Z
bn,t (ρ)Pt (ρ) = ϕ0,t (λ)n exp [−λρ + ϕ0,t (λ)] (A7)
2πi Note that, the hypergeometrics can always be re-
where the one-point CGF ϕ0,t (λ) is given by equa- expanded down in terms of 0 F1 (1, z) and 0 F1 (2, z) using
tion (A1). The nth bias function equation can also be the identity
written as z
0 F1 (n − 1, z) − 0 F1 (n, z) = 0 F1 (n + 1, z) .
dn Ij (ρ)
n(n − 1)
bn,t (ρ)Pt (ρ) = . (A8) (A15)
dj n j=1
27

B. REFINED DERIVATION OF COVARIANCE distribution of the N cells. Let us assume that the value
FOR CELLS IN LARGE-SEPARATION ρ̂I is fixed to a grid point I, such that the distance be-
LIMIT tween the densities ρ̂I and ρ̂J are given by the distance
In Appendix C of Codis et al. (2016b) the result of of grid point J to grid point I. Hence, we will drop
{rIJ } from the argument of the joint PDF in the follow-
the covariance of finding Ni cells with density ρi and Nj
spheres of density ρj is given as (with no summation over ing. Then considering different spatial distributions of
the cells corresponds to selecting a subset S of length N
repeated indices implied)
from the set of indices {1, . . . , NT }
¯ i )b(ρj ) + δij N̄i ,
cov(Ni , Nj ) = N̄i N̄j ξb(ρ (B1)
where b(ρ) is the cell bias and ξ¯ is the mean correlation X Y Z  Y Z 
P b (N ) = dρ̂I dρ̂J P(ρ̂1 , . . . , ρ̂NT ) .
of spheres defined in equation (B10). We can rewrite ˆ
∆ ˆ
¬∆
S I∈S J ∈S
/
the ratio of the number of cells in the bin to the total
number in terms of the measured PDF value in the bin (B7)
and associated density bin size ∆i as Ni ' NT P(ρi )∆i
and similarly N̄i ' NT P̄(ρi )∆i . Hence, we have for the Now we can insert equation (B6) and simplify using
covariance of the histogram probability Pi :

¯ i )b(ρj ) + δij P̄(ρi ) ,


Z Z
cov(P(ρi ), P(ρj )) = P̄(ρi )P̄(ρj )ξb(ρ p= dρ̂ P(ρ̂), 1−p= dρ̂ P(ρ̂) (B8)
∆i NT ˆ ˆ
¬∆
(B2) Z∆ Z
Note that the shot-noise term only acts on the diagonal. pb = dρ̂ P(ρ̂)b(ρ̂), −pb = dρ̂ P(ρ̂)b(ρ) , (B9)
Let us first focus on the diagonal, where the prediction ˆ
∆ ˆ
¬∆
for the noise to signal ratio is then is
var(P(ρi )) 1
R
¯ 2 (ρi ) +
= ξb , (B3) where the normalisation of P enforces R+ dρ̂ P(ρ̂)b(ρ̂) =
2
P̄ (ρi ) P̄(ρi )∆i NT 0 and we will further use b̄ := pb/p. This leads to
The correlationp matrix can be obtained by dividing the
covariance by var(P(ρ1 )var(P(ρ2 ). In the limit of X
infinite volume, where one can place NT → ∞ well- P b (N ) = pN (1 − p)NT −N
separated cells, the correlation matrix would read S
X
Nt →∞
corr(P(ρi ), P(ρj )) −→ sgn (b(ρi )b(ρj )) = ±1 , (B4) + ξ(rIJ )(pb)2 pN −2 (1 − p)NT −N
S3I<J∈S
which predicts a characteristic 4-tiled pattern. X
+2 ξ(rIJ )pN −1 pb(−pb)(1 − p)NT −N −1
B.1. Derivation for autocorrelation of PDF bins S3I<J ∈S
/
!
Let us assume that the joint PDF of the density in X
N 2 NT −N −2
NT cells (for simplicity with regular spacing) at large + ξ(rIJ )p (−pb) (1 − p)
separation8 reads S63I<J ∈S
/
 
P(ρ̂1 , . . . , ρ̂NT ; {rIJ }) N NTNT −N
= p (1 − p) ×
NT
" # N
Y X
= P(ρ̂I ) 1 + b(ρ̂I )b(ρ̂J )ξ(rIJ ) . (B5) X X
I=1 I<J
+ b̄2 ξ(rIJ )
S S3I<J∈S
To derive the result for the biased distribution of finding 2
exactly N cells in a given density bin P b (N ), we further N̄ X
+ ξ(rIJ )
expand the product in (B5) assuming b2 ξ  1 as follows (NT − N̄ )2
S63I<J ∈S
/

P(ρ̂1 , . . . , ρ̂NT ; {rIJ })


!)
2N̄ X
"N #" NT X
# − ξ(rIJ ) ,
Y T X NT − N̄
= P(ρ̂I ) 1 + b(ρ̂I )b(ρ̂J )ξ(rIJ ) . (B6) S3I<J ∈S
/

I=1 I=1 I<J

The probability to find exactly N cells in a density bin where in the second step we have converted p = N̄ /NT .
ˆ = [ρ̂ − ∆ρ/2, ρ̂ + ∆ρ/2] is given by a) integrating over
∆ Note that one can rewrite the sums over  all N -
the probabilities of finding N of the values of ρ1 , . . . , component subsets S (of which there are NNT ) in terms
ρNT within ∆ˆ and NT − N outside ¬∆ ˆ = R+ \ ∆, ˆ and of the mean spatial correlation (which agrees with the
b) summing over all different possibilities for the spatial definition used in Codis et al. (2016b))

8 A generalisation can be obtained by restoring the density-


NT PNT
I<J=1 ξ(rIJ )
dependent cell two-point correlation ξR (ρ̂I , ρ̂J ; rIJ ) instead of the
ξ¯ =
X X
factorised version that is valid at large separations. ξ := ξ(rIJ ) , (B10)
NT (NT − 1)/2
I<J=1
28

as follows In the continuous


R rmax limit, one can write the mean correla-
tion as ξ¯ = rmin dr Pd (r)ξ(r) , where Pd (r) is the prob-
NT − 2
X X  X
ξ(rIJ ) = ξ ability distribution of cell distances r. The maximum
N −2 distance is set by the size of the survey patch/volume,
S S3I<J∈S √
NT N (N − 1) ¯
  for a d-dim box of side length L, we have rmax ' dL.
= ξ, (B11) The minimum distance is set by the cell separation un-
N 2 der consideration, so rmin = 2R for non-overlapping cells
NT − 2 X
 
X X and rmin → 0 for heavily overlapping cells. Note that
2 ξ(rIJ ) = 2 ξ this is not the same as attempting to measure the aver-
N −1
S S3I<J ∈S
/ age correlation in a periodic simulation box (with fixed
 
NT particle number) via the following definition
= N (NT − N )ξ¯ , (B12)
N PNT PN P
N

δ i δ j i=1 δi j=1 δj − δi
NT − 2 X
 
i6=j=1
ξ˜ = = (B16a)
X X
ξ(rIJ ) = ξ NT (NT − 1) NT (NT − 1)
N
S S63I<J ∈S
/ PN 2
NT (NT − N )(NT − N − 1) ¯
hδi=0 i=1 δi
 
= − < 0, (B16b)
= ξ. NT (NT − 1)
N 2
(B13) which tends to be small but negative, because by con-
struction the mean density contrast in a box vanishes.
Where the full array of rIJ with I < J is split into three
subsets of separations a) among cells in the density bin B.2. Derivation of cross-correlation between bins
(length N (N − 1)/2), b) between cells in the density bin
and outside (length N (NT −N )), and c) among cells out- Considering different spatial distributions of the cells
side the density bin (length (NT − N )(NT − N − 1)/2). where a given number are in two density bins corresponds
The combinatorical prefactors come from fixing two in- to first selecting a subset of length N1 , SN1 from the set
dices which corresponds to selecting from a set of NT − 2 of indices I = {1, . . . , NT } and then selecting another
remaining indices, N −2 when both indices are in S, N −1 subset of length N2 , SN2 out of the remaining indices
when one index is in S and the other in I \ S, and N I \ SN1 . For brevity, let us denote SN12 = SN1 ∪ SN2
when both indices are in I \ S. With this, we obtain X Y Z  Y Z 
P b (N1 , N2 ) = dρ̂I dρ̂J
ˆ1 ˆ2
 
NT −N NT SN1 ,SN2 I∈SN1 ∆ J∈SN2 ∆
b N
P (N ) = p (1 − p) ×
N Y Z 
" dρ̂K P(ρ̂1 , . . . , ρ̂NT ) . (B17)
N (N − 1) N̄ N (NT − N ) ˆ
¬∆
1 + b̄2 ξ¯ − K ∈S
/ N12
2 NT − N̄
!# As before, the integrals can be simplified
N̄ 2 (NT − N )(NT − N − 1)
+ , P b (N1 , N2 )
2(NT − N̄ )2 (
NT − N1
 
NT
In the Poisson limit this gives 9 = pN 1 N2
1 p2 (1 − p) ∆NT ,12
×
N1 N2
" !#
b Poiss 2 ¯ N (N − 1) N̄ 2 X X X
P (N ) = P (N ) 1 + b̄ ξ − N̄ N + , + b̄21 ξ(rIJ ) + b̄22 ξ(rIJ )
2 2
SN1 ,SN2 SN1 3I<J∈SN1 SN2 3I<J∈SN2
(B14) X
+ 2b̄1 b̄2 ξ(rIJ )
where P Poiss (N ) = N̄ N exp −N̄ /N !. In Codis et al.

SN1 3I<J∈SN2
(2016a) it was assumed that correlations are the same "
p1 b̄1 + p2 b̄2
at all separations, ξ(rIJ ) = ξ¯ with ξ¯ defined in equa-
X
−2 b̄1 ξ(rIJ )
tion (B10). In this case, one can replace all sums, the 1 − p1 − p2
SN1 3I<J ∈S
/ N12
sum over rIJ by a combinatoric factor for choosing N #
cells out of NT , and the sums over I and J by the num- X
ber of pairs. As we showed here, it turns out that this + b̄2 ξ(rIJ )
result seems holds true even for the more general case SN2 3I<J ∈S
/ N12
of correlations that depend on separation. From this ex-  2 !)
pression we can verify that the biased PDF is normalised p1 b̄1 + p2 b̄2 X
+ ξ(rIJ ) ,
and has mean hN i = N̄ and variance 1 − p1 − p2
SN12 63I<J ∈S
/ N12

var(N ) = hN 2 i − hN i2 = N̄ + N̄ 2 ξ¯ b̄2 (B15) The sums over the different subsets SN1 and SN2 can be
simplified using combinatorical properties.
P We use the
9 Note that without the Poisson limit, the variance of the bino-
average correlation ξ¯ and its numerator ξ as defined in
mial distribution becomes N̄ (1 − p). equation (B10), ¬i to indicate the opposite index ¬i = 2
29

if i = 1 and ¬i = 1 if i = 2, respectively. Introducing the Additionally, one can rewrite


short hand notation NT,12 := NT − N1 − N2 , we obtain
p1 b̄1 + p2 b̄2 N̄1 b̄1 + N̄2 b̄2
= . (B22)
1 − p1 − p2 NT − N̄1 − N̄2
With this one obtains
X X P b (N1 , N2 )
ξ(rIJ )  (
NT − N1
 
SN1 ,SN2 SNi 3I<J∈SNi NT
= pN1 N2
1 p2 (1 − p) ∆NT ,12
× 1+

NT − 2

NT − Ni X
 N1 N2
= ξ
Ni − 2 NT − N¬i
"
N1 (N1 − 1) 2 N2 (N2 − 1) 2
 
NT NT − N1 ¯Ni (Ni − 1)
 + ξ¯ b̄1 + b̄2 + N1 N2 b̄1 b̄2
= ξ , (B18) 2 2
N1 N2 2
N̄1 b̄1 + N̄2 b̄2
− (N1 b̄1 + N2 b̄2 )∆NT,12
X X
2 ξ(rIJ ) NT − N̄1 − N̄2
SN1 ,SN2 SN1 3I<J∈SN2 2 #)
∆NT,12 (∆NT,12 − 1)

N̄1 b̄1 + N̄2 b̄2
NT − 2 NT − N1 − 1 X
  
+ .
=2 ξ 2 NT − N̄1 − N̄2
N1 − 1 N2 − 1
10
 
NT NT − N1 ¯
 In the Poisson limit this becomes
= ξN1 N2 , (B19) "
N1 N2 b Poiss Poiss
X X P (N1 , N2 ) = P (N1 )P (N2 ) × 1+
2 ξ(rIJ )
!
SN1 ,SN2 SNi 3I<J ∈S
/ N12
N1 (N1 − 1) N̄12

NT − 2 NT − Ni − 1 X
  + ξ¯ b̄21 + − N1 N̄1
=2 ξ 2 2
Ni − 1 N¬i !
N2 (N2 − 1) N̄22
NT − N1 ¯ ξ¯ b̄22
  
NT + + − N2 N̄2
= ξNi ∆NT,12 , (B20) 2 2
N1 N2 !#
X X
ξ(rIJ ) + ξ¯ b̄1 b̄2 N1 N2 + N̄1 N̄2 − N1 N̄2 − N̄1 N2 .
SN1 ,SN2 SN12 63I<J ∈S
/ N12

NT − 2 NT − N1 − 2 X
  
= ξ With this we reproduce the result from Codis et al.
N1 N2 (2016a) for the Poisson limit and the associated covari-
 
NT NT − N1 ¯∆NT,12 (∆NT,12 − 1)
 ance
= ξ . (B21) hN1 N2 i = N̄1 N̄2 (1 + ξ¯ b̄1 b̄2 ) . (B23)
N1 N2 2

10 Note that without the Poisson limit, the covariance of the


multinomial distribution becomes −NT p1 p2 = −N̄1 N̄2 /NT .

You might also like