Prepared for submission to JCAP
Unified galaxy power spectrum
measurements from 6dFGS, BOSS,
and eBOSS
Florian Beutler,
a
Patrick McDonald
b
a
Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edin-
burgh EH9 3HJ, UK
b
Lawrence Berkeley National Lab, 1 Cyclotron Rd, Berkeley CA 94720, USA
E-mail: florian.b[email protected], p[email protected]
Abstract. We make use of recent developments in the analysis of galaxy redshift surveys to
present an easy to use matrix-based analysis framework for the galaxy power spectrum multi-
poles, including wide-angle effects and the survey window function. We employ this framework
to derive the deconvolved power spectrum multipoles of 6dFGS DR3, BOSS DR12 and the
eBOSS DR16 quasar sample. As an alternative to the standard analysis, the deconvolved
power spectrum multipoles can be used to perform a data analysis agnostic of survey specific
aspects, like the window function. We show that in the case of the BOSS dataset, the Baryon
Acoustic Oscillation (BAO) analysis using the deconvolved power spectra results in the same
likelihood as the standard analysis. To facilitate the analysis based on both the convolved
and deconvolved power spectrum measurements, we provide the window function matrices,
wide-angle matrices, covariance matrices and the power spectrum multipole measurements for
the datasets mentioned above. Together with this paper we publish a Python-based toolbox
to calculate the different analysis components. The appendix contains a detailed user guide
with examples for how a cosmological analysis of these datasets could be implemented. We
hope that our work makes the analysis of galaxy survey datasets more accessible to the wider
cosmology community.
arXiv:2106.06324v2 [astro-ph.CO] 27 Oct 2021
Contents
1 Introduction 1
2 The survey window function 3
2.1 Including wide-angle effects 4
2.2 Window function convolution as matrix multiplication 7
2.3 Bin averaging the window function 9
2.3.1 The observational binning βˆ†k
o
9
2.3.2 The theoretical binning βˆ†k
th
10
3 Wide-angle effects as matrix multiplication 11
4 Deconvolution as matrix multiplication 12
4.1 Wide-angle compression 15
5 Datasets 16
5.1 6dFGS DR3 16
5.2 BOSS DR12 17
5.3 eBOSS DR16 QSO 18
6 Data analysis 19
6.1 Setup of power spectrum measurements 19
6.2 Deconvolution 19
6.3 Deconvolution and BAO 21
7 Conclusion 23
A User guide 30
B Window function pre-factors 33
C Derivation of the 2D window function 34
C.1 Excluding wide-angle terms and consistency with [1] 34
C.2 Including wide-angle terms and consistency with [2] 35
D Analytic calculation of the odd power spectrum multipoles 36
E Deconvolution results for all datasets 37
1 Introduction
In the last two decades the analysis of galaxy redshift survey datasets has become one of
the most powerful tools to constrain cosmological models [3, 4]. The next generation of
galaxy redshift surveys, such as the Dark Energy Spectroscopic Instrument (DESI [5]) and
the Euclid space mission [6], aim to observe 20 to 40 million galaxy redshifts, increasing the
largest current dataset (BOSS [7], 1 million galaxies) by more than an order of magnitude.
While observations of the Cosmic Microwave Background (CMB) have reached the sample
– 1 –
variance limit for some observables, galaxy surveys are far away from this limit. Extracting
cosmological information from galaxy surveys does pose significant challenges. Most modes
which can be accessed by galaxy surveys are contaminated by non-linear contributions along
with redshift-space distortions and a complicated relation between galaxy density and matter
density. Many different avenues are currently investigated to tackle these challenges (e.g. [8–
11]). Here we will ignore these issues, but focus on the analysis formalism itself.
Extracting cosmological information from the galaxy power spectrum comes with techni-
cal complications, such as wide-angle effects (e.g., [2, 12, 13]) and the survey window function,
which convolves the measured power spectrum. While the survey window is usually known,
this convolution can add significant modeling challenges. The current standard analysis de-
rives the survey window function from the data and convolves any power spectrum model with
the survey window before comparing it with the power spectrum measurement [1, 14, 15]. In-
stead of convolving the model we can also deconvolve the measured power spectrum. Past
attempts of deconvolution were done on a mode by mode basis [16] or assumed the global
plane parallel approximation (often called distant observer approximation [17–19]). Here we
present a deconvolution formalism, which can be applied to a wide-angle multipole represen-
tation of the power spectrum. Recently, [20] advocated deconvolution in the context of going
back to the quadratic estimator formalism originally derived in [21, 22]. For these estimators
one can naturally quote results convolved with a window, or deconvolved, or something in
between [23, 24]. Note that there is a technical difference between that formalism, which
assumes the k bands in which the quadratic averaging is performed are the same as the k
bands used to model the theory power, so that the window matrix is automatically square
and invertible, and the formalism in this and other recent papers, in which the theory bands
can have arbitrarily fine k resolution, independent of the observational averaging band width.
First we express the power spectrum analysis as two matrix multiplications, one matrix
accounting for wide-angle effects and a second matrix accounting for the window function.
We derive these matrices for some of the largest galaxy redshift surveys currently available
(6dFGS DR3, BOSS DR12 and the eBOSS DR16 QSOs sample) and make these matrices
available together with the power spectrum measurements
1
. The appendix of this paper
provides a step-by-step guide for a galaxy survey analysis, including Python-based examples.
The aim is to make galaxy redshift survey datasets more easily accessible for the wider
cosmology community.
As an example application we perform a BAO analysis on the deconvolved power spec-
trum of BOSS DR12. We show that with the products derived in this paper, the likelihood
derived from the convolved and deconvolved power spectrum multipoles is identical.
This paper is organized as follows. We start with a review of the current formulation
of the survey window function, which we turn into a matrix multiplication in section 2.
In section 3 we derive a similar matrix accounting for wide-angle effects. In section 4 we
deriving our deconvolution procedure. In section 5 we introduce the 6dFGS, BOSS and
eBOSS datasets followed by an application of our matrix based deconvolution procedure in
section 6. We conclude in section 7. Appendix A provides a detailed user guide with Python-
based examples. In appendix B we discuss the correlation matrices needed for the window
function calculations. We also show consistency between the equations used in this paper with
the equations in [1] and [2] in appendix C. In appendix D we derive analytic equations for
wide-angle effects. Finally, appendix E contains a summary of the convolved and deconvolved
1
https://fbeutler.github.io/hub/deconv_paper.html
– 2 –
power spectrum measurements for the different datasets.
Throughout the paper we use a (fiducial) flat Ξ›CDM cosmology with Ω
m
= 0.31 when
transferring observables (RA, DEC, z) into cartesian coordinates (x, y, z). When analysing
mock datasets we use the cosmological parameters of the underlying simulations, which are
different for each of the galaxy samples studied here (see section 5). The Python code used
to calculate the wide-angle and window function matrices discussed in this paper is available
at https://github.com/fbeutler/pk_tools.
2 The survey window function
Most studies involving the galaxy power spectrum include the survey window function by
convolving the model for the power spectrum before comparing it to the measurement. Any
asymmetry in the survey window distributes power between the multipoles. This can bias
measurements of anisotropic observables like redshift-space distortions, if not taken into ac-
count correctly. The convolution of the power spectrum multipoles with the window function
multipoles has been laid out in [1] and is given by
P
conv
`
(k) =
Z
dk
0
k
02
X
`
0
W
``
0
(k, k
0
)P
true
`
0
(k
0
) βˆ’
Q
`
(k)
Q
0
(0)
Z
dk
0
k
02
X
`
0
W
0`
0
(0, k
0
)P
true
`
0
(k
0
) , (2.1)
where W
``
0
(k, k
0
) describes the contribution of multipole `
0
to multipole ` due to the survey
window
2
. The second term on the right hand side describes the integral constraint correc-
tion [25] and includes the 1D survey window function multipoles in Fourier space Q
`
(k). While
the treatment of the integral constraint in eq. (2.1) is a good approximation, [26] showed that
there are additional contributions depending on how the reference (random) catalog has been
generated.
While [1] calculated W
``
0
(k, k
0
) through pair counting (see eq. 33 of that reference), re-
cently it was pointed out that this quantity can be obtained using a double Bessel integral [27]
W
``
0
(k, k
0
) =
2
Ο€
(βˆ’i)
`
i
`
0
Z
ds s
2
j
`
(ks)j
`
0
(k
0
s)
X
L
C
``
0
L
Q
L
(s) , (2.2)
where the matrices C
``
0
L
are given in section 3.1 of [27]. The 1D window function multipoles
in configuration-space are defined as
Q
L
(s) =
(2L + 1)
A
Z
dΩ
s
4Ο€
Z
ds
1
Β―n
w
(s
1
)Β―n
w
(s + s
1
)L
L
(
Λ†
s Β·
Λ†
d) , (2.3)
with Β―n
w
(x) = Β―n(x)w(x), where Β―n(x) is the local mean density of galaxies (i.e., expected
number after selection effects), and w(x) represents a weight applied to the density measured
at x (completeness and signal to noise weight). The normalization factor A here is the same
number that appears in the power spectrum estimate, e.g., eq. (3) of [15]. However, we do
not use the standard value of A given in eq. (13) of [15]. Instead we set A to the value
necessary to enforce Q
0
(s β†’ 0) ≑ 1. The same A value is then also used to normalize the
power spectrum estimate, as required for consistency (as recently emphasized by [26]). We
use this definition because we find it gives much closer to the desired unit normalized window,
i.e., window where integration over theory k gives 1. This is desirable because it makes the
2
Note that [1] uses a slightly different nomenclature with |W (k, k
0
)|
2
``
0
≑ W
``
0
(k, k
0
)
– 3 –
convolved power close to the true power, i.e., just smeared slightly by the window instead of
adding an offset in the overall normalization. Note that, at the continuum limit of eq. (2.3),
our normalization convention can be written as
A =
Z
dx Β―n
2
w
(x) . (2.4)
However, the traditional definition as represented by eq. (13) of [15] is also intended to
represent this equation! In the traditional calculation the integral over position and the Β―n(x)
factors are represented by a sum over randoms times Β―n(z), where Β―n(z) is supposed to be a
redshift dependent average of Β―n(x). This is an approximation that in practice only achieves
the goal of Q
0
(s β†’ 0) = 1 to ∼ 10%. We guarantee Q
0
(s β†’ 0) = 1 by computing Q
L
(s)
first and using it directly to fix A. For comparison with past results we give the relative
normalization between the two methods in Table 1. To be clear, if the normalisation between
the power spectrum and window function are consistent, it will not impact any likelihood
analysis, so it is essentially cosmetic.
In the following section we will demonstrate how our window function formalism can be
extended to include wide-angle effects before moving on to develop a matrix based convolution
and deconvolution procedure.
2.1 Including wide-angle effects
Any measurement of the anisotropic power spectrum has to make a choice regarding the
line-of-sight (LOS) direction for a galaxy pair (see, e.g., [2, 12, 13]). Based on this choice the
triangle configuration between the observer and the galaxy pair is reduced to a separation
amplitude s and a multipole dependent weighting given by the Legendre polynomials L
`
.
This geometric simplification introduces wide-angle effects in the power spectrum multipoles
` > 0 as well as the associated window function multipoles. Crucially, wide-angle effects can
break the symmetry between the galaxy pair and hence can introduce odd multipoles, like
the dipole and octopole. These wide-angle effects can be absorbed into the window function
formalism described above. Based on [2] we can extend eq. (2.2) to get
W
(n)
``
0
(k, k
0
) =
2
Ο€
(βˆ’i)
`
i
`
0
Z
ds s
2
j
`
(ks)j
`
0
(k
0
s)
X
L
C
(n)
``
0
L
Q
(n)
L
(s) , (2.5)
where the index n describes the order in the wide-angle expansion and the C
(n)
``
0
L
matrices are
given in appendix B. Eq. (2.5) includes new window function multipoles Q
(n)
L
at each order
of the wide-angle expansion n given by
Q
(n)
`
(s) =
(2` + 1)
A
Z
dΩ
s
4Ο€
Z
ds
1
(s
1
)
βˆ’n
W (s
1
)W (s + s
1
)L
`
(
Λ†
s Β·
Λ†
d) . (2.6)
A derivation of eq. (2.5) is included in Appendix C.2. The constraints imposed by FFTs
usually enforce the end-point LOS definition
Λ†
d =
Λ†
s
1
where the line of sight is chosen along
the distance vector of one of the galaxies (see [2] for details).
Eq. (2.5) can be calculated using one 1D Fourier transforms (FT) for each k bin with
the complexity O(N
k
Γ— N
k
0
log N
k
0
), where N
k
and N
k
0
are the number of bins in k and
k
0
, respectively. Additionally we need to calculate the Q
`
(k) using a 3D FT, which has the
complexity O(N log N), where N is the number of grid cells in which the random galaxies are
binned (see appendix E.1 of [2]). We then obtain Q
`
(s) through Hankel transforms. In this
– 4 –
Figure 1: The window function multipoles of the low redshift bin 0.2 < z < 0.5 of BOSS
DR12, NGC, calculated following eq. (2.5) at zero order in the wide-angle expansion (n = 0).
The corresponding window function multipoles at first order in the wide-angle expansion
(n = 1) are shown in figure 2. The three columns correspond to the contributions of the three
even multipoles to the five non-zero multipoles, including the dipole and octopole. Each
window function is plotted multiple times as a function of k
0
with fixed values of k (vertical
dashed lines). This window function does not account for any bin averaging, but is evaluated
at specific values k with 16 384 values of k
0
(here we focus on 0 < k
0
< 0.12 h Mpc
βˆ’1
, while
in practice we calculate the full range of 0 < k
0
< 0.4 h Mpc
βˆ’1
). For visual purposes, each
of the sub-panels is re-normalized by the factor written in the right hand corner. The final
window function matrix used for the actual analysis requires bin averaging in both k
o
and
k
th
(see eq. 2.16 for more details).
paper we will use eq. (2.5) to calculate the window function using N
k
0
= 16 384 and N
k
= 400
(from 0 < k < 0.4 h Mpc
βˆ’1
in bins of βˆ†k = 0.001 h Mpc
βˆ’1
), which typically takes < 1 minute
on a single state of the art laptop.
Figure 1 shows the window function multipoles W
(n)
``
0
(k, k
0
) for the low redshift bin of
BOSS DR12 in the North Galactic Cap (NGC). This figure only shows the window function
multipoles at zero order in the wide-angle expansion (n = 0 in eq. 2.5). At this order,
only even multipoles `
0
are generated, meaning we have a contribution from the monopole
(left column), quadrupole (middle column) and hexadecapole (right column). Each window
function is plotted several times with fixed k (dashed lines). These plots agree with the plots
shown in Figure 7 of [1], where these multipoles have first been investigated, but they now
include contributions to the odd multipoles.
Figure 2 shows the corresponding window function multipoles at first order in the wide-
angle expansion (n = 1 in eq. 2.5), in which case we only have odd contributions from
the dipole (left column) and octopole (right column). Also note that any window function
multipole W
(n)
``
0
in which the combination `+`
0
is a odd number, results in a complex quantity,
– 5 –
Figure 2: The window function multipoles of the low redshift bin 0.2 < z < 0.5 of BOSS
DR12, NGC, calculated following eq. (2.5) at first order in the wide-angle expansion (n = 1).
The corresponding window function multipoles at zero order in the wide-angle expansion
(n = 0) are shown in figure 1. The two columns correspond to the contributions of the two
odd multipoles to the five non-zero multipoles. Each window function is plotted multiple
times as a function of k
0
with fixed values of k (vertical dashed lines). For visual purposes,
each of the sub-panels is re-normalized by the factor written in the right hand corner.
to account for the fact that the estimated odd power spectrum multipoles are complex [2, 28].
Based on the discussion above we can extend eq. (2.1) in section 2, which describes
the convolution of the power spectrum multipoles with the window function multipoles by
including wide-angle effects:
P
conv
`
(k) =
Z
dk
0
X
`
0
,n
k
02βˆ’n
W
(n)
``
0
(k, k
0
)P
(n),true
`
0
(k
0
)
βˆ’
Q
(0)
`
(k)
Q
(0)
0
(0)
Z
dk
0
X
`
0
,n
k
02βˆ’n
W
(n)
0`
0
(0, k
0
)P
(n),true
`
0
(k
0
) .
(2.7)
We now define a new multipole expansion including wide-angle effects as
P (k, Β΅) =
X
`,n
(kd)
βˆ’n
P
(n)
`
(k)L
`
(
Λ†
k Β·
Λ†
d) , (2.8)
where d is the LOS vector and d = |
Λ†
d|. Based on this definition the power spectrum multi-
poles, P
(n)
`
(k), at different order in the wide-angle expansions n are given by [2, 13]
P
(n)
`
(k) = 4Ο€(βˆ’i)
`
Z
ds s
2
(ks)
n
ΞΎ
(n)
`
(s)j
`
(ks) , (2.9)
where ΞΎ
(n)
`
are the configuration space multipoles at wide-angle order n.
– 6 –
For the remainder of this paper we will use the end-point LOS definition, meaning the
LOS for each galaxy pair is oriented along the distance vector of one of the galaxies (d = s
1
).
In this case the zero order term (n = 0) is equivalent to the commonly used power spectrum
multipoles and at this order only the even multipoles are present. At n = 1 only the odd
multipoles are present and can be obtained as linear combinations of the even multipoles in
configuration space
ΞΎ
(1)
1
= βˆ’
3
5
ΞΎ
(0)
2
(s) , (2.10)
ΞΎ
(1)
3
=
3
5
ΞΎ
(0)
2
(s) βˆ’
10
9
ΞΎ
(0)
4
(s) . (2.11)
The second order terms (n = 2) are given in eq. (2.16) - (2.18) of [2]. The analysis in
this paper will be limited to n ≀ 1, even though including higher order terms (n > 1) is
straightforward. Note that the dipole and octopole power spectra can be calculated directly
in Fourier space without any need for a Hankel transform (see also eq. 3.56 of [29]):
P
(1)
1
(k) = βˆ’i
3
5
h
3P
(0)
2
(k) + kβˆ‚
k
P
(0)
2
(k)
i
, (2.12)
P
(1)
3
(k) = βˆ’i
ξ€”
3
5

2P
(0)
2
(k) βˆ’ kβˆ‚
k
P
(0)
2
(k)

+
10
9

5P
(0)
4
(k) + kβˆ‚
k
P
(0)
4
(k)


. (2.13)
We included a derivation of these equations in appendix D. Using linear theory we obtain
P
(1)
1
(k) = βˆ’if
ξ€’
4
5
b
1
+
12
35
f

[3P
m
(k) + kβˆ‚
k
P
m
(k)] , (2.14)
P
(1)
3
(k) = βˆ’i4f
ξ€”
1
5
ξ€’
b
1
+
3
7
f

(2P
m
(k) βˆ’ kβˆ‚
k
P
m
(k)) +
4
63
f (5P
m
(k) + kβˆ‚
k
P
m
(k))

. (2.15)
2.2 Window function convolution as matrix multiplication
If we assume that we calculate our power spectrum model in bins of βˆ†k
th
and intend to
compare this model with power spectrum measurements in bins of β€˜observed’ bandpowers
βˆ†k
o
, the window function needs to be integrated over k and k
0
as
W
(n)
``
0
(k
o
, k
th
) =
Z
dk k
2
Θ(k
o
, k)
Z
dk
0
k
02βˆ’n
W
(n)
``
0
(k, k
0
)Θ(k
th
, k
0
) , (2.16)
where the step function Θ is defined as
Θ(k
x
, k) =
(
1 if |k
x
βˆ’ k| <
βˆ†k
x
2
0 otherwise.
(2.17)
We will test the bin averaging employed in our analysis in the next subsection. Before we get
to that, let us define our matrix nomenclature, starting with the vectors
P
true
=







ο£­
P
(0),true
0
(k)
P
(1),true
1
(k)
P
(0),true
2
(k)
P
(1),true
3
(k)
P
(0),true
4
(k)
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£·
ο£·
ο£Έ
, P
conv
=







ο£­
P
(0),conv
0
(k)
P
(1),conv
1
(k)
P
(0),conv
2
(k)
P
(1),conv
3
(k)
P
(0),conv
4
(k)
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£·
ο£·
ο£Έ
, Q =







ο£­
Q
(0)
0
(k)
Q
(1)
1
(k)
Q
(0)
2
(k)
Q
(1)
3
(k)
Q
(0)
4
(k)
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£·
ο£·
ο£Έ
, (2.18)
– 7 –
Figure 3: The window function matrix (W
``
0
, see section 2.2) and wide-angle matrix (M
``
0
,
section 3) of BOSS DR12 NGC in the low redshift bin. The wide-angle transformation matrix
maps the theoretical power spectrum multipoles with three even multipoles to a vector with
five multipoles (only accounting for wide-angle effects at order n = 1). The window function
maps the five (theoretical) power spectrum multipoles (in bins of βˆ†k
th
) to the observed/con-
volved power spectrum multipoles (in bins of βˆ†k
o
), which can be compared to the data. For
plotting purposes here we use βˆ†k
th
= βˆ†k
o
= 0.01 h Mpc
βˆ’1
, while the analysis in this paper
uses βˆ†k
o
= 0.01 h Mpc
βˆ’1
and βˆ†k
th
= 0.001 h Mpc
βˆ’1
. The window function matrix is not
symmetric, since the contribution of the monopole to the dipole is different to the contribu-
tion of the dipole to the monopole. The diagonal elements are not 1, since they correspond
to the bin integral of the window function within the bin βˆ†k
th
(see eq. 2.16), which is only 1
for βˆ†k  k
f
as shown in figure 6. The z-axis is logarithmic for better contrast.
where P
true
is a vector of model power spectrum multipoles and the convolution matrix is
given by
W =







ο£­
W
(0)
00
W
(1)
01
W
(0)
02
W
(1)
03
W
(0)
04
W
(0)
10
W
(1)
11
W
(0)
12
W
(1)
13
W
(0)
14
W
(0)
20
W
(1)
21
W
(0)
22
W
(1)
23
W
(0)
24
W
(0)
30
W
(1)
31
W
(0)
32
W
(1)
33
W
(0)
34
W
(0)
40
W
(1)
41
W
(0)
42
W
(1)
43
W
(0)
44
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£·
ο£·
ο£Έ
, (2.19)
where each sub-matrix W
(n)
``
0
is of size N
th
Γ—N
o
. Based on the definitions above, we can write
the convolution of the power spectrum multipoles with the window function multipoles as a
matrix multiplication
P
conv
= WP
true
βˆ’
Q
Q
0
(0)
W
0`
0
(0, k
0
)P
true
`
0
(k
0
) , (2.20)
– 8 –
where
Q
Q
0
(0)
W
0`
0
(0, k
0
) is a matrix representing the integral constraint correction. For the
remainder of this paper we will absorb the integral constraint into W meaning we re-define
W ≑ W βˆ’
Q
Q
0
(0)
W
0`
0
(0, k
0
) . (2.21)
This works exactly because the integral constraint effect is linear in the power spectrum just
like the standard window function. In this paper the window function matrix W is generally
defined in the k-range 0.0 < k < 0.4 h Mpc
βˆ’1
, while the recommended k-range for any fit is
0.01 < k < 0.3 h Mpc
βˆ’1
. The increased k-range accounts for the redistribution of power due
to the window function, which connects modes inside the fitting range with modes beyond
that range. The equations above include wide-angle effects at order n = 0 and n = 1, even
though this formalism could easily be extended to n > 1. Figure 3 shows the matrix W after
performing the integral in eq. (2.16) (and including the integral constraint correction).
2.3 Bin averaging the window function
In eq. (2.16) we wrote down how to take the window function (as derived in eq. 2.5 and
plotted in figure 1 and 2) an05d account for the bin averaging to match the layout of the
data vector. The default choices for our window functions are βˆ†k
th
= 0.001 h Mpc
βˆ’1
and
βˆ†k
o
= 0.01 h Mpc
βˆ’1
. Here we will demonstrate how we implemented the bin averaging of the
window function and we will test certain assumptions made in our formalism.
2.3.1 The observational binning βˆ†k
o
All power spectrum multipole measurements provided with this paper use k-bins of βˆ†k
o
=
0.001 h Mpc
βˆ’1
. However, for practical purposes
3
we use 10 times larger bins of βˆ†k
o
=
0.01 h Mpc
βˆ’1
. The measured power spectra can easily be re-binned into larger bins using
P(k
i
o
) =
10
X
bin j within βˆ†k
i
o
N
modes
j
N
modes
i
P(k
j
o
) , (2.22)
where βˆ†k
j
o
=
1
10
βˆ†k
i
o
= 0.001 h Mpc
βˆ’1
and N
modes
j
is the number of observed k modes in bin
j with
N
modes
i
=
10
X
bin j within βˆ†k
i
o
N
modes
j
. (2.23)
One crucial point to highlight here is that the fundamental mode for all galaxy datasets
included in this paper is smaller than our choice for βˆ†k
o
. This means we need to bin average
our theoretical model in accordance with the average of eq. (2.22). For that reason we measure
the window function in bins of βˆ†k
o
= 0.001 h Mpc
βˆ’1
and average as
W(k
i
o
, k
th
) =
Z
dk k
2
W(k, k
th
)Θ(k
i
o
, k) (2.24)
β‰ˆ
10
X
bin j within βˆ†k
i
o
N
modes
j
N
modes
i
W(k
j
o
, k
th
) . (2.25)
Figure 4 shows the window function contributions to the monopole, W
`0
, with βˆ†k
o
=
0.001 h Mpc
βˆ’1
(yellow line) and βˆ†k
o
= 0.01 h Mpc
βˆ’1
(red line) for the low redshift (z1)
sample of BOSS DR12 NGC.
3
Mainly to reduce the size of the associated covariance matrix.
– 9 –
Figure 4: Comparison of the window function terms W
`0
for different binning schemes at
k
o
= 0.055 h Mpc
βˆ’1
for BOSS DR12 NGC in the low redshift bin (z1). The yellow line shows
the window with bins of βˆ†k
o
= βˆ†k
th
= 0.001 h Mpc
βˆ’1
. The red line shows the window with
larger observational bins of βˆ†k
o
= 0.01 h Mpc
βˆ’1
, by averaging 10 bins of the yellow line. The
green line is averaged over both k
o
and k
th
. The black dashed lines show the edges of the bin
at k
min
= 0.05 h Mpc
βˆ’1
and k
max
= 0.06 h Mpc
βˆ’1
. The relative heights of the three windows
are scaled to make them comparable. The red line corresponds to our default choice.
2.3.2 The theoretical binning βˆ†k
th
The multiplication of the power spectrum model with the window function matrix effectively
calculates a convolution. The question we want to address here is, what would be the required
resolution for this convolution, meaning how sensitive is the convolved power spectrum to
the choice of βˆ†k
th
? In figure 5 we show convolved power spectrum multipoles using our
default binning of βˆ†k
o
= 0.01 h Mpc
βˆ’1
but different βˆ†k
th
. The reference model is based on
the current standard method, where the power spectrum model is (1) Hankel transformed
into configuration-space, (2) multiplied by the configuration-space window and (3) Hankel
transformed back into Fourier space [14, 30]. The result is than bin-averaged to βˆ†k
o
=
0.01 h Mpc
βˆ’1
. The differences between our default choice of βˆ†k
th
= 0.001 h Mpc
βˆ’1
and the
reference model are < 1% of the measurement uncertainties for the monopole on most scales
and even smaller for all other multipoles.
Finally we note that the above binning tests for the power spectrum and window function
are based on a Ξ›CDM model (with the default cosmology given at the end of the introduction).
Our default binning choices might not be optimal for non-standard analysis that searches for
variations in the power spectrum smaller than the chosen bin width. Examples for such cases
are primordial features [31] or primordial non-Gaussianity [32]. However, our binning choice
– 10 –
Figure 5: Comparison of convolved power spectrum multipole models based on different
window function binning choices of βˆ†k
th
using BOSS DR12 NGC in the low redshift bin
(z1). The window function transforms the theoretical binning (βˆ†k
th
) into the binning used
for the data (βˆ†k
o
). The reference model is based on the current standard method [14, 30],
which Hankel transforms the model into configuration-space, multiplies with the real-space
window and Hankel transforms back into Fourier space. The window function uses βˆ†k
o
=
0.01 h Mpc
βˆ’1
in all cases. The window functions provided with this paper uses a default
binning of βˆ†k
th
= 0.001 h Mpc
βˆ’1
, except of the case of deconvolution, where we use βˆ†k
th
=
0.01 h Mpc
βˆ’1
.
should be perfectly suitable for the standard RSD and BAO analysis.
3 Wide-angle effects as matrix multiplication
The odd power spectrum multipoles sourced by wide-angle effects are linearly related to
the even multipoles through eq. (2.14) and (2.15). This implies that we can define a linear
transformation between β€˜ο¬‚at-sky’ statistics (no wide-angle effects) and β€˜curved-sky’ statistics
(including wide-angle effects) as
MP
true,flat-sky
= P
true
, (3.1)
where P
true,flat-sky
= (P
0
, P
2
, P
4
) is a vector of the predicted even multipoles and M is a
5N
k
th
Γ—3N
k
th
matrix, which transforms this vector into a new vector P
true
, which contains 5
– 11 –
multipoles, including the dipole and octopole. We can define
M =





ο£­
I 0 0
0 K
2β†’1
0
0 I 0
0 K
2β†’3
K
4β†’3
0 0 I
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
, (3.2)
where I is the identity matrix of size N
th
Γ— N
th
and
K
2β†’1
lm
= βˆ’i
3
5d
ξ€”
3
k
l
Θ(k
l
, k
m
) + βˆ‚
k
m
Θ(k
l
, k
m
)

,
= βˆ’i
3
5d
ξ€”
3
k
l
Θ(k
l
, k
m
) +
Θ(k
l
, k
m
βˆ’ βˆ†k) βˆ’Ξ˜(k
l
, k
m
+ βˆ†k)
2βˆ†k

, (3.3)
K
2β†’3
lm
= βˆ’i
3
5d
ξ€”
2
k
l
Θ(k
l
, k
m
) βˆ’
Θ(k
l
, k
m
βˆ’ βˆ†k) βˆ’Ξ˜(k
l
, k
m
+ βˆ†k)
2βˆ†k

, (3.4)
K
4β†’3
lm
= βˆ’i
10
9d
ξ€”
5
k
l
Θ(k
l
, k
m
) +
Θ(k
l
, k
m
βˆ’ βˆ†k) βˆ’Ξ˜(k
l
, k
m
+ βˆ†k)
2βˆ†k

, (3.5)
with Θ defined in eq. (2.17). As shown in these equations, we can implement derivatives within
a matrix multiplication by including off-diagonal terms. Here we use two-sided derivatives
except for the first and last bin in the data vector, where we use forward and backwards
derivatives instead. The transformation matrix M does depend on the order of wide-angle
effects n, since at n = 2 there are new even multipoles sourced by wide-angle effects. Following
the rest of this paper, we only include terms at n ≀ 1. The plot on the right in figure 3 shows
the matrix M for the low redshift bin of BOSS DR12 NGC. The only survey specific parameter
in this matrix is the amplitude of the LOS vector d. The code to calculate the matrix M is
publicly available
4
.
4 Deconvolution as matrix multiplication
Usually the aim of a power spectrum analysis is to obtain the likelihood
L ∝ exp
ξ€”
βˆ’
1
2
(P
conv
o
βˆ’ WP
true
)
T
C
βˆ’1
conv
(P
conv
o
βˆ’ WP
true
)

, (4.1)
where C
βˆ’1
conv
is the inverse covariance matrix of the (convolved) power spectrum multipoles,
P
conv
o
represents the measured power spectrum and P
true
is the unconvolved power spectrum
model.
Using eq. (4.1) together with the maximum likelihood condition βˆ‚Ο‡
2
/βˆ‚P
true
= 0 im-
plies that the power spectrum before convolution with the window function is related to the
convolved power spectrum by
(W
T
C
βˆ’1
conv
W)
βˆ’1
W
T
C
βˆ’1
conv
P
conv
o
≑ P
true
o
, (4.2)
where
(W
T
C
βˆ’1
conv
W)
βˆ’1
W
T
C
βˆ’1
conv
= W
βˆ’1
(4.3)
4
https://github.com/fbeutler/pk_tools/blob/master/wide_angle_tools.py
– 12 –
Figure 6: Left: Diagonal of the window function W
(n)
``
(k
o
= k
th
) for the low redshift bin
of BOSS DR12 NGC (solid lines) and SGC (dashed lines). The differently colored lines
correspond to the different multipoles. All lines in the plot on the left use βˆ†k
o
= βˆ†k
th
=
0.01 h Mpc
βˆ’1
. The SGC has smaller values for W
(n)
``
(k
o
= k
th
) because its fundamental mode
is larger. Right: The window function W
(n)
``
(k
o
= k
th
= 0.1 h Mpc
βˆ’1
) as a function of the
size of the bandpower bin width βˆ†k. The quantity plotted on the y-axis of these two figures
scales the bandpower uncertainty for the deconvolved power spectrum multipoles as shown
in eq. (4.5).
Figure 7: Left: Correlation matrix for the 5 power spectrum multipoles of BOSS DR12
NGC in the low redshift bin. These matrices are based on the 2048 BOSS DR12 MultiDark-
Patchy (MD-Patchy) mock catalogs with a binning of βˆ†k
o
= 0.01 h Mpc
βˆ’1
. Right: The
same correlation matrix using the deconvolved power spectrum multipoles. The wide-angle
expansion used here includes all terms with n < 2.
if W is a square matrix and non-singular. Note that the deconvolution process implied in
eq. (4.2) cannot recover the true underlying power spectrum (which can be calculated in
theory), since the window function did erase all information below the fundamental mode.
– 13 –
Figure 8: Slice through the correlation matrices shown in figure 7 at k
o
= 0.105 h Mpc
βˆ’1
for
the convolved (solid lines) and deconvolved (dashed lines) cases, respectively. We focus on
the diagonal components of the monopole (yellow), dipole (pink) and quadrupole (green).
The covariance matrix of the deconvolved power spectrum is given by
C
βˆ’1
deconv
= W
T
C
βˆ’1
conv
W . (4.4)
The correlation matrix, R = C
ij
/
p
C
ii
C
jj
, for BOSS DR12 NGC in the low redshift bin is
plotted in figure 7 before (left) and after (right) deconvolution. Deconvolution often leads to
anti-correlated bins [33], which can be intuitively explained by imagining a power spectrum
model with very fine k binning (βˆ†k ξ€œ fundamental mode k
f
). If one moves the power
spectrum in one bin up, but compensates by moving the power spectrum in the adjacent
bin down, the convolved version of this power spectrum will look very similar to the original
power spectrum, since the window function averages neighbouring bins. Hence such a mode
of variation is not well constrained in a deconvolved estimate, leading to anti-correlation.
Generally however, there is a decrease in the correlation between bandpowers as clearly
visible in figure 8, which implies an increase in the variance of modes within that bin, given
by
Οƒ
P
deconv
`
(k) ∝
1
W
(n)
``
(k, k)
. (4.5)
Figure 6 shows the behaviour of W
(n)
``
(k, k) as a function of k for different βˆ†k in the low
redshift bin of BOSS DR12. This clearly shows that deconvolution can significantly increase
the uncertainties in the power spectrum multipoles if βˆ†k is small. However, this increase in
the uncertainties does not reflect any loss of information, but only shows that for smaller βˆ†k
the window function has a larger impact. The larger uncertainty is compensated by having
less correlation (or anti-correlation) between the bandpowers. The information content of
P
true
o
and P
conv
o
in terms of the likelihood is identical, since we derived P
true
o
in eq. (4.2) using
the likelihood in eq. (4.1) (assuming P
true
o
is estimated in at least as many bands as P
conv
o
).
It should be noted, however, that our deconvolution procedure is based on a Gaussian
likelihood (as given in eq. 4.1), which might not be true on the largest scales of the survey
(see e.g. [34]). This could bias large-scale signals like primordial non-Gaussianity. Alternative
– 14 –
0.00 0.05 0.10 0.15 0.20 0.25 0.30
k
max
odd-poles [hMpc
βˆ’1
]
10
βˆ’1
10
0
10
1
10
2
Ο‡
2
/Ξ½
BOSS DR12 NGC, z1
BOSS DR12 SGC, z1
BOSS DR12 NGC, z3
BOSS DR12 SGC, z3
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
k [hMpc
βˆ’1
]
0.5
0.6
0.7
0.8
0.9
1.0
Οƒ
2
P
`
(k) without odd poles
Οƒ
2
P
`
(k) with odd poles
BOSS NGC z3
` = 0
` = 2
` = 4
Figure 9: Left: The reduced Ο‡
2
for fits of band-parameterized even poles (P
true,flat-sky
) to all
the measured even poles, and odd poles up to the maximum k on the horizontal axis. Right:
Diagonal covariance for the inferred flat sky even poles (see eq. 4.7), including the constraint
from k < 0.1 h Mpc
βˆ’1
odd poles, divided by the same covariance with no odd pole constraint.
to our procedure one could derive the deconvolved bandpowers using an MCMC approach
without the need to assume a Gaussian likelihood or include a Gaussianisation step as e.g.
suggested in [35].
4.1 Wide-angle compression
In eq. (4.2) we can replace W with
˜
W ≑ WM to compress our vector of 5 multipoles to a
vector of 3 even multipoles, i.e.,
P
true,flat-sky
= (
˜
W
T
C
βˆ’1
conv
˜
W)
βˆ’1
˜
W
T
C
βˆ’1
conv
P
conv
o
, (4.6)
while the corresponding covariance matrix is
C
βˆ’1
true,flat-sky
=
˜
W
T
C
βˆ’1
conv
˜
W . (4.7)
This is equivalent to a minimum Ο‡
2
fit of the band-parameterized model for P
true,flat-sky
to all
multipoles. This is an over-constrained fit with the degrees of freedom equal to the number of
bins in the odd multipoles
5
. Because the prediction of the odd multipoles from P
true,flat-sky
is model-independent, the reduced Ο‡
2
(Ο‡
2
/ν) for this fit is a model-independent test of the
various quantities we report (C, M, etc.). Unfortunately, we find that when we try to fit all
multipoles on all scales this way, Ο‡
2
is unacceptably bad, i.e., the measured points are not
consistent with any model, given our M, W, and C. As shown in figure 9 (left), this problem
is driven by the high k bins in the odd multipoles.
Figure 9 shows that Ο‡
2
/Ξ½ is near unity if we drop the odd multipoles with k & 0.1 h Mpc
βˆ’1
,
but quickly grows when including higher k. This is not a straightforward thing to understand,
or even see, by looking at the measured points, covariance matrix, etc. It appears to be driven
by a few special linear combinations of even and odd multipoles which have very low variance
in the mocks used to compute the covariance matrix. These are then not well-predicted by
5
Assuming the theory is band-parameterized using the same set of k bins as the measurement, i.e., the
number of free parameters is equal to the number of bins in the even power spectrum multipoles so that they
cancel in the calculation of the degrees of freedom.
– 15 –
Survey/Sample N
gal
[z
min
z
max
] z
eff
V
eff
k
f
(50%) k
f
(95%) N
m
A/A
old
[Gpc
3
] [hMpc
βˆ’1
] [hMpc
βˆ’1
]
6dFGS DR3 75 117 [0.01, 0.2] 0.096 0.12 0.00514 0.00142 600 1.0168
BOSS DR12 NGC z1 429 182 [0.2, 0.5] 0.38 2.6 0.00222 0.00062 2048 0.9032
BOSS DR12 SGC z1 174 820 [0.2, 0.5] 0.38 1.0 0.00310 0.00086 2048 0.8896
BOSS DR12 NGC z3 435 742 [0.5, 0.75] 0.61 2.8 0.00154 0.00042 2048 0.9104
BOSS DR12 SGC z3 158 262 [0.5, 0.75] 0.61 1.0 0.00226 0.00062 2048 0.9016
eBOSS DR16 QSO NGC 218 209 [0.8, 2.2] 1.52 0.35 0.00094 0.00026 1000 0.9302
eBOSS DR16 QSO SGC 125 499 [0.8, 2.2] 1.52 0.18 0.00114 0.00034 1000 0.8974
Table 1: Properties of the different galaxy samples studied in this paper. The effective
redshift is defined as z
eff
=
P
i
w
i
z
i
/
P
i
w
i
, where w
i
includes any completeness weight as
well as the FKP weight [36]. To calculate the FKP weight as well as the effective volume,
we assume P
0
= 10 000h
βˆ’3
Mpc
3
for BOSS and 6dFGS and P
0
= 6 000h
βˆ’3
Mpc
3
for eBOSS.
The effective volume in this table might differ from other references due to (1) a different
fiducial cosmology and (2) only the raw (unweighted) number density of galaxies is used in
our calculation. k
f
represents the fundamental mode, which is given by the wavenumber at
which the monopole window Q
(0)
0
(k) has 50% of the height of Q
(0)
0
(k β†’ 0) (column 6) or 95%
(column 7). N
m
represents the number of mock realizations available for each sample, and
the last column gives the ratio between our A definition, enforcing Q
0
(s β†’ 0) = 1, and the
normalization used in many previous papers (see section 2). For BOSS we also have catalogs
after applying density field reconstruction. However, only 1000 of the 2048 BOSS mock
catalogs have been processed in this way
6
. More information about the different samples can
be found in the relevant data release papers of 6dFGS [37], BOSS [38] and eBOSS [39].
the theory as multiplied by M and W. E.g., for BOSS DR12 NGC, z3, k < 0.2 h Mpc
βˆ’1
, we
find Ο‡
2
= 1400, of which 1127 is contributed by the single worst eigenvector of the covariance
matrix. We were unable to find any flaw in the calculations that would fix this. Since we
expect the wide-angle effects to be primarily important at low-k, our solution is to simply
drop the odd multipoles at k > 0.1 h Mpc
βˆ’1
from fits. The right panel of figure 9 shows
an example of the impact of the lower k odd multipoles on the inferred deconvolved even
multipole variance. As expected the impact of the odd multipoles is primarily at very low k.
5 Datasets
Here we will introduce the three galaxy redshift survey datasets we analyze in this paper,
namely the 6dFGS DR3 sample, the BOSS DR12 sample and the eBOSS DR16 quasar sample.
5.1 6dFGS DR3
The 6-degree Field Galaxy Survey (6dFGS [37]) is a K-band selected, magnitude limited
(K ≀ 12.9) galaxy survey, based on the 2MASS Extended Source Catalog (2MASS XSC; [40]).
6dFGS covers nearly the entire southern sky and is the lowest redshift sample included in this
paper. The survey made use of the Six-Degree Field (6dF) multi-fibre instrument on the UK
6
We actually have 997 mock catalogs for the NGC in the low and high redshift bins and 1000 mock catalogs
for the SGC in both redshift bins. Some NGC mock catalogs have been excluded due to issues while processing
the catalogs.
– 16 –
Figure 10: Comparison of the monopole window function for 6dFGS DR3, BOSS DR12 and
eBOSS DR16. The dashed lines in the plot on the left indicate the fundamental mode defined
as the 50% height of the monopole. The Figure on the right shows the monopole weighted
by k
2
, which is the quantity which goes in the window function in eq. (2.16) at n = 0. The
double peaked shape of the eBOSS DR16 SGC window is probably caused by the narrow
band geometry and/or the split of the SGC into two separate regions (see figure 2 in [39]).
The curves in the plot on the right hand side are normalized so that the maximum of the
curve is at unity (which allows comparison between the different surveys).
Schmidt telescope at the Siding Spring Observatory. The three data release papers [37, 41, 42]
describe 6dFGS in full detail, including comparisons between 6dFGS, 2dFGRS and SDSS.
Here we use the final K-band selected sample, which contains 75 117 galaxies and has been
used in several galaxy clustering studies [43–48] (see Table 1 for more details).
Mock catalogs: The 6dFGS mock catalogs are based on COLA simulations [49] with 1728
3
particles in boxes with 1.2h
βˆ’1
Gpc on each side. The simulations use 20 time steps down to
z = 0 and a mass resolution of 2.8 Γ—10
10
h
βˆ’1
M
ξ€Œ
. A friends-of-friends (FoF) finder is used to
locate halos with a minimum of 32 dark matter particles per halo. The cosmology used in
these simulations is Ω
m
= 0.3, Ω
b
= 0.0478, h = 0.68, Οƒ
8
= 0.82 and n
s
= 0.96. The derived
halo catalogs are populated with galaxies using an HOD measured on the dataset itself [43]
(for more details see section 3 of [48]).
The 6dFGS sample has by far the smallest volume of all the datasets included in this
analysis (see table 1 and figure 10). However, it does occupy a unique redshift range, which
is not covered by BOSS or eBOSS.
5.2 BOSS DR12
The Baryon Oscillation Spectroscopic Survey (BOSS) was part of SDSS-III [7, 50] and mea-
sured spectroscopic redshifts of 1 198 006 million galaxies [38]. The survey covers 10 252 deg
2
– 17 –
divided in two patches on the sky, the North Galactic Cap (NGC) and the South Galactic Cap
(SGC), over a redshift range of 0.2 - 0.75. Here we split this redshift range into two redshift
bins defined by 0.2 < z < 0.5 and 0.5 < z < 0.75 with the effective redshifts z
eff
= 0.38 and
0.61, respectively. We also include the different incompleteness weights as
w
c
= (w
rf
+ w
fc
βˆ’ 1)w
sys
, (5.1)
which account for redshift failures (w
rf
), fibre collisions (w
fc
) and photometric systematics
related to the observational seeing conditions and correlations with stellar density (w
sys
) [51,
52].
Mock catalogs: The BOSS collaboration provided mock catalogs for the final BOSS DR12
dataset (MD-Patchy mock catalogs [53]). These catalogs have been produced using approxi-
mate gravity solvers and analytical-statistical biasing models calibrated to a reference sample
from the BigMultiDark simulations [54]. The BigMultiDark simulation is based on gadget-
2 [55] with 3840
3
particles in a volume of (2.5h
βˆ’1
Gpc)
3
assuming a Ξ›CDM cosmology with
Ω
m
= 0.307115, Ω
b
= 0.048206, Οƒ
8
= 0.8288, n
s
= 0.9611 and h = 0.6777. The mock
catalogs use halo abundance matching to reproduce the observed BOSS two- and three-point
clustering measurements [56]. This technique is applied as a function of redshift to reproduce
the BOSS DR12 redshift evolution.
We already saw the BOSS DR12 NGC window function W
(n)
``
0
of the low redshift bin
(z1) in figure 1 and 2. We also include the monopole window function, Q
(0)
0
(k), in figure 10.
The BOSS DR12 sample represents the largest galaxy redshift dataset (in terms of effective
volume) [3, 15, 30, 52, 57–59].
5.3 eBOSS DR16 QSO
The extended Baryon Oscillation Spectroscopic Survey ([60], eBOSS) is part of SDSS-IV [61]
and relies on the same optical spectrographs [62] as the SDSS-III BOSS survey. In addition to
observing luminous red galaxies (LRGs) and emission line galaxies (ELGs), eBOSS collected
redshifts for ∼ 500 000 quasars. While the Quasar density is comparatively low, this sample
has the distinction of covering the largest cosmic volume, leading to the smallest fundamental
mode of all samples discussed in this paper (see figure 10). Here we focus on the eBOSS
quasar sample to avoid any overlap with the other samples.
The eBOSS targets [63, 64] are selected from the DR7 [65] and DR8 [66] photometric
catalogs as well as the Wide Field Infrared Survey Explorer (WISE, [67]), as described in [68].
Just like BOSS, the eBOSS quasars are split in two angular regions, the North Galactic Cap
(NGC) and South Galactic Cap (SGC). The effective areas of these regions are 2924 deg
2
and
1884 deg
2
, respectively (see Table 1 for more details). Each eBOSS object has a complete-
ness weight including corrections for fibre collisions, redshift failures [69] and photometric
systematic effects [70]:
w
c
= w
cp
w
npz
w
sys
. (5.2)
We refer to section 5.6 of [71] for details about these weights. The cosmology results for the
DR16 quasar sample were presented in [4, 72]
7
, which reported a BAO distance measurement
in the range 0.8 < z < 2.2.
7
https://www.sdss.org/dr16/
– 18 –
Mock datasets: We make use of a set of 1000 mock catalogs to estimate the covari-
ance matrix of the eBOSS DR16 quasar power spectrum. The mocks are based on the
Extended Zel’dovich (EZ) approximate N-body simulation scheme [73]. These mocks rely
on the Zel’dovich approximation to generate a density field, while including nonlinear and
halo biasing effects through the use of free parameters. These free parameters are tuned to
produce two-point and three-point clustering of a desired data set. The EZ mock catalogs
account for the redshift evolution of the eBOSS quasars by constructing a light-cone out of
7 redshift shells, generated from periodic boxes of side length L = 5000h
βˆ’1
Mpc at different
redshifts. These mock catalogs also mimic the fibre collisions and redshift failures, so that
each object has an associated w
foc
and w
cp
. The cosmology of the EZmocks is Ξ›CDM with
Ω
m
= 0.307115, Ω
b
= 0.048206, h = 0.6777, Οƒ
8
= 0.8255, and n
s
= 0.9611 (same cosmology
as the BOSS DR12 MD-Patchy mocks discussed above).
Figure 10 shows the window function monopole for the different samples where the NGC
sample of eBOSS DR16 indicates the smallest fundamental mode of all samples discussed
in this paper. The fact that the eBOSS quasar sample probes the largest scales currently
accessible with galaxy redshift surveys, makes it a valuable tool to test primordial non-
Gaussianity [32]. However, the high shot noise level does mean that the effective volume of
the eBOSS DR16 QSO sample at most wavenumbers is below BOSS DR12.
6 Data analysis
Here we will apply the deconvolution formalism developed in section 4 to the three datasets
introduced in the previous section.
6.1 Setup of power spectrum measurements
All power spectra discussed in this section are measured using the estimator of [74] and
[75] including the odd multipoles as discussed in appendix E of [2]. For all power spectrum
measurements we use a cubic grid with L = 2000, 3500 h
βˆ’1
Mpc and N = 600, 700 for
6dFGS and BOSS, respectively. For the eBOSS QSO sample we use L = 5400, 6600 h
βˆ’1
Mpc
with N = 1040, 1270 for the SGC and NGC, respectively. This setup ensures that for all
three surveys the Nyquist frequency is > 0.6 h Mpc
βˆ’1
. The galaxies are assigned to the grid
using the triangular shape cloud procedure and we correct for the associated pixel window
function [76]. To further reduce aliasing effects we included the interlacing procedure of [77].
As defined in Section 2, we use a different normalization for the power spectra than other
recent papers, which brings the convolved power closer to the true/deconvolved power. For
comparison with past work, Table 1 gives the ratio, of our definition of the normalisation to
the one that we would compute using eq. (13) of [15].
6.2 Deconvolution
Figure 11 and 12 show a comparison of the convolved and deconvolved power spectrum
multipoles of BOSS DR12 NGC in the low redshift bin. For these plots we combined 10
measured bandpowers of βˆ†k
o
= 0.001 h Mpc
βˆ’1
into larger bins of βˆ†k
o
= 0.01 h Mpc
βˆ’1
. Since
deconvolution requires a square matrix in eq. (4.2), we also use βˆ†k
th
= 0.01 h Mpc
βˆ’1
. In
figure 11 we can see the increase in the variance caused by deconvolution (comparison of
the red and black shaded regions). At the same time deconvolution reduces the correlation
between bandpowers. The odd multipoles in figure 12 also include a best fitting model
– 19 –
Figure 11: Comparison of the convolved and deconvolved even power spectrum multipoles
for the BOSS DR12 low redshift bin (z1) in the NGC. The red line and shaded area shows
the mean of the measured MD-Patchy DR12 power spectrum multipoles together with the
standard deviation. The black line and shaded area shows the equivalent power spectra after
deconvolution. The data points show the actual BOSS DR12 data. The lower panel shows
the difference between the red and black lines in the upper panel, relative to the uncertainties
post-deconvolution, which highlights the impact of the window function.
(green dashed lines) as well as the intrinsic dipole and octopole (blue dashed lines) as given
by eq. (2.14) and eq. (2.15). These models are based on a fit to the even multipoles as
discussed in section 3.6 of [2]. One can clearly see that deconvolving the dipole removes the
window function contributions, which dominate the dipole on most scales and recovers the
intrinsic dipole expected due to wide-angle effects. We included the corresponding results for
6dFGS and eBOSS in appendix E.
Figure 14 in appendix E shows the convolved and deconvolved power spectrum multi-
poles of 6dFGS. From these plots we can see that the monopole power spectrum of the mock
catalogs does not perfectly match the data power spectrum amplitude, while higher order
multipoles agree well [78]. Even though 6dFGS is the lowest redshift sample, the wide-angle
effects seem far less important compared to BOSS DR12, and the dipole is (by eye) consis-
tent with zero. This agrees with [45] where wide-angle effects are discussed in appendix C.
Window function effects also seem to be much smaller than the statistical noise, which might
be caused by the very compact geometry of the 6dFGS survey (see [37] for details about the
angular and redshift distribution of 6dFGS).
Figure 16 in appendix E compares the power spectrum multipoles of the eBOSS DR16
quasar sample with the corresponding mock results before and after deconvolution. While
the dipole and octopole moments are consistent with zero for each bandpower estimate, there
is a clear systematic dipole signal
8
.
– 20 –
Figure 12: Comparison of the convolved and deconvolved odd power spectrum multipoles
for the BOSS DR12 low redshift bin (z1) in the NGC. The red line and shaded area shows
the mean of the measured MD-Patchy DR12 power spectrum multipoles together with the
standard deviation. The black line and shaded area shows the equivalent power spectra after
deconvolution. The data points show the actual BOSS DR12 data. The green dashed line
shows the best fit as discussed in section 3.6 of [2]. The dashed blue line shows the intrinsic
odd multipoles introduced by wide-angle effects based on the same model as the green dashed
line. The lower panel shows the difference between the red and black lines in the upper panel,
relative to the uncertainties post-deconvolution, which highlights the impact of the window
function.
6.3 Deconvolution and BAO
The convolution of the measured power spectrum with the survey window function could
smear out a signal which exists at the same scale as the fundamental mode (or the k
o
band
width, whichever is larger). The BAO signal has a wavelength of ∼ 0.06 h Mpc
βˆ’1
, which is
much larger than the fundamental mode of the surveys discussed in this paper (see table 1).
Nevertheless, deconvolving the power spectrum does increase the BAO signature as
clearly visible in figure 13. This figure shows the BAO signal in the low and high redshift
bins of the BOSS DR12 MD-Patchy mock catalogs together with the best fitting models. The
data points represent the mean and variance of 1000 post-reconstruction MD-Patchy mock
catalogs.
It is important to note that even though the BAO signal appears to be enhanced post-
deconvolution, our deconvolution procedure is based on eq. (4.1) and hence the convolved and
8
Even though the Nyquist frequency is k
Ny
= 0.6 h Mpc
βˆ’1
, the eBOSS power spectrum dipole below
k < 0.3 h Mpc
βˆ’1
does seem to be affected by aliasing, causing the rise in the dipole at high k. This should be
taken into account when analyzing the eBOSS dataset.
– 21 –
Figure 13: Comparison of the convolved and deconvolved BAO signal in the power spectrum
of the low redshift bin (left) and high redshift bin (right) of the BOSS DR12 MD-Patchy mock
catalogs (we are plotting the mean of the 1000 mock catalogs together with the variance).
The red data points correspond to the convolved power spectrum with the best fitting model
shown as the solid red line. The deconvolved measurements are shown by the blue data points
together with the best fitting model (blue solid line). The data points are the weighted mean
of the NGC and SGC. The errorbars correspond to the square root of the diagonal terms of
the covariance matrix.
deconvolved analysis should lead to the same likelihood and the same model parameters, if the
same cuts and theory binning are used. To demonstrate this point we perform an isotropic
BAO analysis based on the mean of 1000 BOSS DR12 MD-Patchy mock power spectrum
monopoles (post-reconstruction). For the pre-deconvolution fit we build the monopole model
following the current standard analysis pipeline described in [30]. We also provide higher
order multipoles for the quadrupole and hexadecapole based on the simple Kaiser model.
We use the window function and wide-angle matrices (W and M) and limit the likelihood
evaluation to the monopole only (see the second example in appendix A). Our fitting range
is 0.01 < k < 0.3 h Mpc
βˆ’1
and we are jointly fitting the NGC and SGC. For the post-
deconvolution fit we only fit the monopole without any window function or wide-angle matrix,
again using the model of [30]. Our fitting procedure is using the Python-based MCMC sampler
zeus [79, 80]
9
.
In the low redshift bin of BOSS DR12 we find α = 0.999 ± 0.013 before deconvolution
and an identical value after deconvolution. Figure 13 (left) compares the best fitting model
and the mean of the MD-Patchy mocks (the plot shows the weighted average of the NGC
and SGC). The equivalent values for the high redshift bin shown in figure 13 (right) are
Ξ± = 1.003 Β± 0.012 before deconvolution and Ξ± = 1.002 Β± 0.012 after deconvolution. Any
observed differences are consistent with noise in the MCMC chain, the slightly different cuts
implied by using the monopole within 0.01 < k < 0.3 h Mpc
βˆ’1
in convolved vs. deconvolved
space, and coarsening of the theory side of the window function used in the deconvolution.
The window matrix used for deconvolution has to be a square matrix with relatively large
k-bins (βˆ†k = 0.01 h Mpc
βˆ’1
in our case). Such large k-bins cannot capture small-scale features
in the theory power spectrum (see figure 4). While these effects should not introduce any
issues within a smooth Ξ›CDM power spectrum (as demonstrated here for the BAO case), it
9
https://zeus-mcmc.readthedocs.io/en/latest/
– 22 –
could become relevant for non-Ξ›CDM models especially if small scale features are present.
7 Conclusion
When analysing galaxy redshift surveys in Fourier space, one needs to account for the survey
window function as well as wide-angle effects. In this paper we leverage recent new develop-
ments dealing with the survey window function and wide-angle effects, to lay out a simple
power spectrum analysis framework based on matrix multiplications. The main results of this
paper are:
(1) We derive a matrix to account for wide-angle effects in the power spectrum multipoles.
We use a new analytic approach rather than the commonly used Hankel transforms.
(2) We expand the window function matrix approach presented in [27] by including wide-
angle effects.
(3) We use this matrix-based analysis framework for the power spectrum multipoles to
demonstrate two possible analysis pipelines, one using the standard path of convolving
the model vector and one based on the deconvolution of the data vector.
(4) We apply the deconvolution procedure to a set of existing galaxy redshift surveys,
namely 6dFGS DR3, BOSS DR12 and eBOSS DR16. Using a BAO analysis we demon-
strate that our deconvolution analysis framework leads to the same likelihood as the
standard analysis.
(5) We provide the power spectrum multipoles as well as the window function matrices,
wide-angle matrices and covariance matrices for 6dFGS DR3, BOSS DR12 and eBOSS
DR16. In the appendix we also provide Python-based examples and a general user guide
for a clustering analysis. These easy to use components hopefully simplify the analysis
of these datasets and make them more accessible for the wider cosmology community.
The deconvolution framework outlined in this paper does not suffer from limitations inher-
ent to other methods presented in the literature, such as the assumption of a global plane
parallel approximation. Nevertheless, the inversion of the window function does require a
square window matrix, which enforces large βˆ†k
th
bins, a limitation which is not present when
convolving the model vector.
Our analysis focuses on the key science targets of galaxy redshift surveys such as RSD
and BAO. Other observables such as primordial non-Gaussianity through the scale-dependent
bias in the power spectrum, naturally requires to focus on the largest scales of the survey.
The products provided with this paper are sub-optimal for such an observable. However, the
formalism presented in this paper can easily be adapted to suit such an observable.
A similar matrix-based analysis approach could also be developed for higher order statis-
tics, like the bispectrum. The extension of our analysis framework to higher order statistics
will be addressed in future work.
Acknowledgments
The authors would like to thank Antonio Cuesta for help with the reconstructed BOSS cat-
alogs and Richard Neveux, Arnaud De-Mattia and Hector Gil-Marin for helpful discussions
– 23 –
regarding the window function normalisation. This project has received funding from the
European Research Council (ERC) under the European Union’s Horizon 2020 research and
innovation programme (grant agreement 853291). FB is a Royal Society University Research
Fellow. PM was supported by the U.S. Department of Energy, Office of Science, Office of
High Energy Physics, under Contract no. DE-AC02-05CH11231.
This work has benefited from a variety of Python packages including numpy [81], scipy [82],
matplotlib [83], zeus [79, 80] and hankl [84].
References
[1] F. Beutler, S. Saito, H.-J. Seo, J. Brinkmann, K. S. Dawson, D. J. Eisenstein et al., The
clustering of galaxies in the sdss-iii baryon oscillation spectroscopic survey: Testing gravity with
redshift-space distortions using the power spectrum multipoles, Monthly Notices of the Royal
Astronomical Society 443 (2014) 1065 [Arxiv:1312.4611v2].
[2] F. Beutler, E. Castorina and P. Zhang, Interpreting measurements of the anisotropic galaxy
power spectrum, Journal of Cosmology and Astroparticle Physics 2019 (2019) 040
[Arxiv:1810.05051v3].
[3] S. Alam, M. Ata, S. Bailey, F. Beutler, D. Bizyaev, J. A. Blazek et al., The clustering of
galaxies in the completed sdss-iii baryon oscillation spectroscopic survey: cosmological analysis
of the dr12 galaxy sample, Monthly Notices of the Royal Astronomical Society 470 (2017) 2617
[Arxiv:1607.03155v1].
[4] [eBOSS Collaboration], S. Alam, M. Aubert, S. Avila, C. Balland, J. E. Bautista et al., The
completed sdss-iv extended baryon oscillation spectroscopic survey: Cosmological implications
from two decades of spectroscopic surveys at the apache point observatory,
Arxiv:2007.08991v1.
[5] A. Aghamousa, J. Aguilar, S. Ahlen, S. Alam, L. E. Allen, C. A. Prieto et al., The desi
experiment part i: Science,targeting, and survey design, Arxiv:1611.00036v2.
[6] R. Laureijs, J. Amiaux, S. Arduini, J. L. Auguères, J. Brinchmann, R. Cole et al., Euclid
definition study report, Arxiv:1110.3193v1.
[7] K. S. Dawson, D. J. Schlegel, C. P. Ahn, S. F. Anderson, Γ‰ric Aubourg, S. Bailey et al., The
baryon oscillation spectroscopic survey of sdss-iii, The Astronomical Journal 145 (2012) 10
[Arxiv:1208.0022v3].
[8] A. Taruya, T. Nishimichi and S. Saito, Baryon acoustic oscillations in 2d: Modeling
redshift-space power spectrum from perturbation theory, Physical Review D 82 (2010) 063522
[Arxiv:1006.0699v1].
[9] L. Senatore and M. Zaldarriaga, Redshift space distortions in the effective field theory of large
scale structures, Arxiv:1409.1225v1.
[10] P. McDonald and A. Roy, Clustering of dark matter tracers: generalizing bias for the coming
era of precision lss, Journal of Cosmology and Astroparticle Physics 2009 (2009) 020
[Arxiv:0902.0991v1].
[11] U. Seljak and P. McDonald, Distribution function approach to redshift space distortions,
Journal of Cosmology and Astroparticle Physics 2011 (2011) 039 [Arxiv:1109.1888v1].
[12] P. H. F. Reimberg, F. Bernardeau and C. Pitrou, Redshift-space distortions with wide angular
separations, Journal of Cosmology and Astroparticle Physics 2016 (2015) 048
[Arxiv:1506.06596v2].
[13] E. Castorina and M. White, Beyond the plane-parallel approximation for redshift surveys,
Monthly Notices of the Royal Astronomical Society (2018) [Arxiv:1709.09730v2].
– 24 –
[14] M. J. Wilson, J. A. Peacock, A. N. Taylor and S. de la Torre, Rapid modelling of the
redshift-space power spectrum multipoles for a masked density field, Monthly Notices of the
Royal Astronomical Society 464 (2015) 3121 [Arxiv:1511.07799v2].
[15] F. Beutler, H.-J. Seo, S. Saito, C.-H. Chuang, A. J. Cuesta, D. J. Eisenstein et al., The
clustering of galaxies in the completed sdss-iii baryon oscillation spectroscopic survey:
Anisotropic galaxy clustering in fourier-space, Monthly Notices of the Royal Astronomical
Society 466 (2017) 2242 [Arxiv:1607.03150v1].
[16] M. Tegmark, A. J. S. Hamilton and Y. Xu, The power spectrum of galaxies in the 2df 100k
redshift survey, Monthly Notices of the Royal Astronomical Society 335 (2001) 887
[Arxiv:astro-ph/0111575v3].
[17] T. Sato, G. Huetsi and K. Yamamoto, Deconvolution of window effect in galaxy power spectrum
analysis, Progress of Theoretical Physics 125 (2010) 187 [Arxiv:1010.0289v2].
[18] T. Sato, G. HΓΌtsi, G. Nakamura and K. Yamamoto, Window effect in the power spectrum
analysis of a galaxy redshift survey, International Journal of Astronomy and Astrophysics 03
(2013) 243 [Arxiv:1308.3551v2].
[19] P. J. Outram, F. Hoyle, T. Shanks, B. J. Boyle, S. M. Croom, N. S. Loaring et al., The 2df qso
redshift survey - vi. measuring lambda and beta from redshift-space distortions in the power
spectrum, Monthly Notices of the Royal Astronomical Society 328 (2001) 174
[Arxiv:astro-ph/0106012v2].
[20] O. H. E. Philcox, Cosmology without window functions: Quadratic estimators for the galaxy
power spectrum, Physical Review D 103 (2021) 103504 [2012.09389].
[21] J. R. Bond, A. H. Jaffe and L. Knox, Estimating the power spectrum of the cosmic microwave
background, Physical Review D 57 (1998) 2117.
[22] U. Seljak, Cosmography and Power Spectrum Estimation: A Unified Approach, The
Astrophysical Journal 503 (1998) 492.
[23] M. Tegmark, A. J. S. Hamilton and Y. Xu, The power spectrum of galaxies in the 2dF 100k
redshift survey, Monthly Notices of the Royal Astronomical Society 335 (2002) 887
[astro-ph/0111575].
[24] M. Tegmark, D. J. Eisenstein, M. A. Strauss, D. H. Weinberg, M. R. Blanton, J. A. Frieman
et al., Cosmological constraints from the SDSS luminous red galaxies, Physical Review D 74
(2006) 123507 [arXiv:astro-ph/0608632].
[25] P. J. A. and N. J. A.Monthly Notices of the Royal Astronomical Society 253 (1991) 307.
[26] A. de Mattia and V. Ruhlmann-Kleider, Integral constraints in spectroscopic surveys,
Arxiv:1904.08851v1.
[27] G. D’Amico, J. Gleyzes, N. Kokron, D. Markovic, L. Senatore, P. Zhang et al., The
cosmological analysis of the sdss/boss data from the effective field theory of large-scale
structure, Arxiv:1909.05271v1.
[28] P. McDonald, Gravitational redshift and other redshift-space distortions of the imaginary part
of the power spectrum, Journal of Cosmology and Astroparticle Physics 2009 (2009) 026
[Arxiv:0907.5220v1].
[29] F. Beutler and E. D. Dio, Modeling relativistic contributions to the halo power spectrum dipole,
Arxiv:2004.08014v1.
[30] F. Beutler, H.-J. Seo, A. J. Ross, P. McDonald, S. Saito, A. S. Bolton et al., The clustering of
galaxies in the completed sdss-iii baryon oscillation spectroscopic survey: Baryon acoustic
oscillations in fourier-space, Monthly Notices of the Royal Astronomical Society 464 (2017)
3409 [Arxiv:1607.03149v1].
– 25 –
[31] F. Beutler, M. Biagetti, D. Green, A. Slosar and B. Wallisch, Primordial features from linear to
nonlinear scales, Physical Review Research 1 (2019) 033209 [Arxiv:1906.08758v1].
[32] E. Castorina, N. Hand, U. Seljak, F. Beutler, C.-H. Chuang, C. Zhao et al., Redshift-weighted
constraints on primordial non-gaussianity from the clustering of the eboss dr14 quasars in
fourier space, Arxiv:1904.08859v1.
[33] A. J. S. Hamilton and M. Tegmark, Decorrelating the power spectrum of galaxies, Monthly
Notices of the Royal Astronomical Society 312 (1999) 285 [Arxiv:astro-ph/9905192v4].
[34] C. Hahn, F. Beutler, M. Sinha, A. Berlind, S. Ho and D. W. Hogg, Likelihood non-gaussianity
in large-scale structure analyses, Monthly Notices of the Royal Astronomical Society 485 (2019)
2956 [Arxiv:1803.06348v1].
[35] M. S. Wang, W. J. Percival, S. Avila, R. Crittenden and D. Bianchi, Cosmological inference
from galaxy-clustering power spectrum: Gaussianization and covariance decomposition, Monthly
Notices of the Royal Astronomical Society 486 (2019) 951 [Arxiv:1811.08155v2].
[36] H. A. Feldman, N. Kaiser and J. A. Peacock, Power spectrum analysis of three-dimensional
redshift surveys, ApJL 426 (1993) 23 [Arxiv:astro-ph/9304022v1].
[37] D. H. Jones, M. A. Read, W. Saunders, M. Colless, T. Jarrett, Q. Parker et al., The 6df galaxy
survey: Final redshift release (dr3) and southern large-scale structures, Monthly Notices of the
Royal Astronomical Society 399 (2009) 683 [Arxiv:0903.5451v1].
[38] B. Reid, S. Ho, N. Padmanabhan, W. J. Percival, J. Tinker, R. Tojeiro et al., Sdss-iii baryon
oscillation spectroscopic survey data release 12: galaxy target selection and large scale structure
catalogues, Monthly Notices of the Royal Astronomical Society 455 (2016) 1553
[Arxiv:1509.06529v2].
[39] M. Ata, F. Baumgarten, J. Bautista, F. Beutler, D. Bizyaev, M. R. Blanton et al., The
clustering of the sdss-iv extended baryon oscillation spectroscopic survey dr14 quasar sample:
First measurement of baryon acoustic oscillations between redshift 0.8 and 2.2, Monthly Notices
of the Royal Astronomical Society 473 (2018) 4773 [Arxiv:1705.06373v2].
[40] T. H. Jarrett, T. Chester, R. Cutri, S. Schneider, M. Skrutskie and J. P. Huchra, 2mass
extended source catalog: Overview and algorithms, The Astronomical Journal 119 (2000) 2498
[Arxiv:astro-ph/0004318v1].
[41] D. H. Jones, W. Saunders, M. Colless, M. A. Read, Q. A. Parker, F. G. Watson et al., The 6df
galaxy survey: Samples, observational techniques and the first data release, Monthly Notices of
the Royal Astronomical Society 355 (2004) 747 [Arxiv:astro-ph/0403501v1].
[42] D. H. Jones, B. A. Peterson, M. Colless and W. Saunders, Near-infrared and optical luminosity
functions from the 6df galaxy survey, Monthly Notices of the Royal Astronomical Society 369
(2006) 25 [Arxiv:astro-ph/0603609v2].
[43] F. Beutler, C. Blake, M. Colless, D. H. Jones, L. Staveley-Smith, L. Campbell et al., The 6df
galaxy survey: Dependence of halo occupation on stellar mass, Monthly Notices of the Royal
Astronomical Society 429 (2013) 3604 [Arxiv:1212.3610v1].
[44] F. Beutler, C. Blake, M. Colless, D. H. Jones, L. Staveley-Smith, G. B. Poole et al., The 6df
galaxy survey: z β‰ˆ 0 measurement of the growth rate and Οƒ
8
, Monthly Notices of the Royal
Astronomical Society 423 (2012) 3430 [Arxiv:1204.4725v1].
[45] F. Beutler, C. Blake, M. Colless, D. H. Jones, L. Staveley-Smith, L. Campbell et al., The 6df
galaxy survey: Baryon acoustic oscillations and the local hubble constant, Monthly Notices of
the Royal Astronomical Society 416 (2011) 3017 [Arxiv:1106.3366v1].
[46] M. I. Scrimgeour, T. M. Davis, C. Blake, L. Staveley-Smith, C. Magoulas, C. M. Springob
et al., The 6df galaxy survey: Bulk flows on 50 βˆ’ 70h
βˆ’1
mpc scales, Monthly Notices of the
Royal Astronomical Society 455 (2016) 386 [Arxiv:1511.06930v1].
– 26 –
[47] A. Johnson, C. Blake, J. Koda, Y.-Z. Ma, M. Colless, M. Crocce et al., The 6df galaxy velocity
survey: Cosmological constraints from the velocity power spectrum, Monthly Notices of the
Royal Astronomical Society 444 (2014) 3926 [Arxiv:1404.3799v2].
[48] P. Carter, F. Beutler, W. J. Percival, C. Blake, J. Koda and A. J. Ross, Low redshift baryon
acoustic oscillation measurement from the reconstructed 6-degree field galaxy survey, Monthly
Notices of the Royal Astronomical Society 481 (2018) 2371 [Arxiv:1803.01746v1].
[49] J. Koda, C. Blake, F. Beutler, E. Kazin and F. Marin, Fast and accurate mock catalogue
generation for low-mass galaxies, Monthly Notices of the Royal Astronomical Society 459
(2016) 2118 [Arxiv:1507.05329v1].
[50] D. J. Eisenstein, D. H. Weinberg, E. Agol, H. Aihara, C. A. Prieto, S. F. Anderson et al.,
Sdss-iii: Massive spectroscopic surveys of the distant universe, the milky way galaxy, and
extra-solar planetary systems, The Astronomical Journal 142 (2011) 72 [Arxiv:1101.1529v2].
[51] L. Anderson, E. Aubourg, S. Bailey, F. Beutler, V. Bhardwaj, M. Blanton et al., The clustering
of galaxies in the sdss-iii baryon oscillation spectroscopic survey: Baryon acoustic oscillations
in the data release 10 and 11 galaxy samples, Monthly Notices of the Royal Astronomical
Society 441 (2014) 24 [Arxiv:1312.4877v2].
[52] A. J. Ross, F. Beutler, C.-H. Chuang, M. Pellejero-Ibanez, H.-J. Seo, M. Vargas-Magana et al.,
The clustering of galaxies in the completed sdss-iii baryon oscillation spectroscopic survey:
Observational systematics and baryon acoustic oscillations in the correlation function, Monthly
Notices of the Royal Astronomical Society 464 (2017) 1168 [Arxiv:1607.03145v2].
[53] F.-S. Kitaura, S. Rodriguez-Torres, C.-H. Chuang, C. Zhao, F. Prada, H. Gil-Marin et al., The
clustering of galaxies in the sdss-iii baryon oscillation spectroscopic survey: mock galaxy
catalogues for the boss final data release, Monthly Notices of the Royal Astronomical Society
456 (2016) 4156 [Arxiv:1509.06400v3].
[54] A. Klypin, G. Yepes, S. Gottlober, F. Prada and S. Hess, Multidark simulations: the story of
dark matter halo concentrations and density profiles, Monthly Notices of the Royal
Astronomical Society 457 (2016) 4340 [Arxiv:1411.4001v2].
[55] V. Springel, The cosmological simulation code gadget-2, Monthly Notices of the Royal
Astronomical Society 364 (2005) 1105 [Arxiv:astro-ph/0505010v1].
[56] S. A. RodrΓ­guez-Torres, C.-H. Chuang, F. Prada, H. Guo, A. Klypin, P. Behroozi et al., The
clustering of galaxies in the sdss-iii baryon oscillation spectroscopic survey: Modeling the
clustering and halo occupation distribution of boss-cmass galaxies in the final data release,
Monthly Notices of the Royal Astronomical Society (2015) stw1014 [Arxiv:1509.06404v3].
[57] J. N. Grieb, A. G. SΓ‘nchez, S. Salazar-Albornoz, R. Scoccimarro, M. Crocce, C. D. Vecchia
et al., The clustering of galaxies in the completed sdss-iii baryon oscillation spectroscopic
survey: Cosmological implications of the fourier space wedges of the final sample, Monthly
Notices of the Royal Astronomical Society 467 (2017) 2085 [Arxiv:1607.03143v2].
[58] A. G. Sanchez, R. Scoccimarro, M. Crocce, J. N. Grieb, S. Salazar-Albornoz, C. DallaVecchia
et al., The clustering of galaxies in the completed sdss-iii baryon oscillation spectroscopic
survey: cosmological implications of the configuration-space clustering wedges, Monthly Notices
of the Royal Astronomical Society 464 (2017) 1640 [Arxiv:1607.03147v1].
[59] S. Satpathy, S. Alam, S. Ho, M. White, N. A. Bahcall, F. Beutler et al., The clustering of
galaxies in the completed sdss-iii baryon oscillation spectroscopic survey: On the measurement
of growth rate using galaxy correlation functions, Monthly Notices of the Royal Astronomical
Society 469 (2017) 1369 [Arxiv:1607.03148v2].
[60] K. S. Dawson, J.-P. Kneib, W. J. Percival, S. Alam, F. D. Albareti, S. F. Anderson et al., The
sdss-iv extended baryon oscillation spectroscopic survey: Overview and early data, The
Astronomical Journal 151 (2016) 44 [Arxiv:1508.04473v2].
– 27 –
[61] M. R. Blanton, M. A. Bershady, B. Abolfathi, F. D. Albareti, C. A. Prieto, A. Almeida et al.,
Sloan digital sky survey iv: Mapping the milky way, nearby galaxies, and the distant universe,
The Astronomical Journal 154 (2017) 28 [Arxiv:1703.00052v2].
[62] S. Smee, J. E. Gunn, A. Uomoto, N. Roe, D. Schlegel, C. M. Rockosi et al., The multi-object,
fiber-fed spectrographs for sdss and the baryon oscillation spectroscopic survey, The
Astronomical Journal 146 (2012) 32 [Arxiv:1208.2233v2].
[63] B. Abolfathi, D. S. Aguado, G. Aguilar, C. A. Prieto, A. Almeida, T. T. Ananna et al., The
fourteenth data release of the sloan digital sky survey: First spectroscopic data from the
extended baryon oscillation spectroscopic survey and from the second phase of the apache point
observatory galactic evolution experiment, The Astrophysical Journal Supplement Series 235
(2018) 42 [Arxiv:1707.09322v3].
[64] I. PΓ’ris, P. Petitjean, E. Aubourg, A. D. Myers, A. Streblyanska, B. W. Lyke et al., The sloan
digital sky survey quasar catalog: Fourteenth data release, Astronomy and Astrophysics 613
(2018) A51 [Arxiv:1712.05029v2].
[65] K. Abazajian and for the Sloan Digital Sky Survey, The seventh data release of the sloan digital
sky survey, The Astrophysical Journal Supplement Series 182 (2008) 543 [Arxiv:0812.0649v2].
[66] S.-I. collaboration, H. Aihara, C. A. Prieto, D. An, S. F. Anderson, Γ‰ric Aubourg et al., The
eighth data release of the sloan digital sky survey: First data from sdss-iii, The Astrophysical
Journal Supplement Series 193 (2011) 29 [Arxiv:1101.1559v2].
[67] E. L. Wright, P. R. M. Eisenhardt, A. Mainzer, M. E. Ressler, R. M. Cutri, T. Jarrett et al.,
The wide-field infrared survey explorer (wise): Mission description and initial on-orbit
performance, The Astronomical Journal 140 (2010) 1868 [Arxiv:1008.0031v2].
[68] A. D. Myers, N. Palanque-Delabrouille, A. Prakash, I. PΓ’ris, C. Yeche, K. S. Dawson et al.,
The sdss-iv extended baryon oscillation spectroscopic survey: Quasar target selection, The
Astrophysical Journal Supplement Series 221 (2015) 27 [Arxiv:1508.04472v2].
[69] I. PΓ’ris, P. Petitjean, N. P. Ross, A. D. Myers, Γ‰ric Aubourg, A. Streblyanska et al., The sloan
digital sky survey quasar catalog: twelfth data release, Astronomy and Astrophysics 597 (2016)
A79 [Arxiv:1608.06483v1].
[70] P. Laurent, S. Eftekharzadeh, J.-M. L. Goff, A. Myers, E. Burtin, M. White et al., Clustering
of quasars in sdss-iv eboss : study of potential systematics and bias determination, Journal of
Cosmology and Astro-Particle Physics 2017 (2017) 017 [Arxiv:1705.04718v1].
[71] A. J. Ross, J. Bautista, R. Tojeiro, S. Alam, S. Bailey, E. Burtin et al., The completed sdss-iv
extended baryon oscillation spectroscopic survey: Large-scale structure catalogs for cosmological
analysis, Arxiv:2007.09000v1.
[72] R. Neveux, E. Burtin, A. de Mattia, A. Smith, A. J. Ross, J. Hou et al., The completed sdss-iv
extended baryon oscillation spectroscopic survey: Bao and rsd measurements from the
anisotropic power spectrum of the quasar sample between redshift 0.8 and 2.2,
Arxiv:2007.08999v1.
[73] C.-H. Chuang, F.-S. Kitaura, F. Prada, C. Zhao and G. Yepes, Ezmocks: extending the
zel’dovich approximation to generate mock galaxy catalogues with accurate clustering statistics,
Monthly Notices of the Royal Astronomical Society 446 (2014) 2621 [Arxiv:1409.1124v2].
[74] D. Bianchi, H. Gil-MarΓ­n, R. Ruggeri and W. J. Percival, Measuring line-of-sight dependent
fourier-space clustering using ffts, Monthly Notices of the Royal Astronomical Society: Letters
453 (2015) L11 [Arxiv:1505.05341v2].
[75] R. Scoccimarro, Fast estimators for redshift-space clustering, Physical Review D 92 (2015)
[Arxiv:1506.02729v2].
– 28 –
[76] Y. P. Jing, Correcting for the alias effect when measuring the power spectrum using fft, The
Astrophysical Journal 620 (2004) 559 [Arxiv:astro-ph/0409240v2].
[77] E. Sefusatti, M. Crocce, R. Scoccimarro and H. Couchman, Accurate estimators of correlation
functions in fourier space, Monthly Notices of the Royal Astronomical Society 460 (2015) 3624
[Arxiv:1512.07295v2].
[78] C. Blake, P. Carter and J. Koda, Power spectrum multipoles on the curved sky: an application
to the 6-degree field galaxy survey, Monthly Notices of the Royal Astronomical Society 479
(2018) 5168 [Arxiv:1801.04969v2].
[79] M. Karamanis and F. Beutler, Ensemble slice sampling, Arxiv:2002.06212v1.
[80] M. Karamanis, F. Beutler and J. A. Peacock, zeus: A python implementation of ensemble slice
sampling for efficient bayesian parameter inference, Arxiv:2105.03468v1.
[81] S. V. D. Walt, S. C. Colbert and G. Varoquaux, The numpy array: a structure for efficient
numerical computation, Computing in Science & Engineering 13 (2011) 22
[Arxiv:1102.1523v1].
[82] P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau et al.,
Scipy 1.0–fundamental algorithms for scientific computing in python, Nature Methods 17
(2020) 261 [Arxiv:1907.10121v1].
[83] J. D. Hunter, Matplotlib: A 2d graphics environment, Computing in Science & Engineering 9
(2007) 90.
[84] M. Karamanis and F. Beutler, hankl: A lightweight python implementation of the fftlog
algorithm for cosmology, Arxiv:2106.06331v1.
[85] J. Hartlap, P. Simon and P. Schneider, Why your model parameter confidences might be too
optimistic – unbiased estimation of the inverse covariance matrix, Astronomy and Astrophysics
464 (2006) 399 [Arxiv:astro-ph/0608064v2].
[86] E. Sellentin and A. F. Heavens, Parameter inference with estimated covariance matrices,
Monthly Notices of the Royal Astronomical Society 456 (2015) L132 [Arxiv:1511.05969v2].
[87] S. Dodelson and M. D. Schneider, The effect of covariance estimator error on cosmological
parameter constraints, Physical Review D 88 (2013) [Arxiv:1304.2593v2].
[88] W. J. Percival, A. J. Ross, A. G. Sanchez, L. Samushia, A. Burden, R. Crittenden et al., The
clustering of galaxies in the sdss-iii baryon oscillation spectroscopic survey: Including
covariance matrix errors, Monthly Notices of the Royal Astronomical Society 439 (2014) 2531
[Arxiv:1312.4841v1].
– 29 –
A User guide
Together with this publication we provide:
1. The window function W, in the form of 5N
o
Γ— 5N
th
= 200 Γ— 2000 and 5N
o
Γ— 5N
th
=
200 Γ— 200) matrices.
2. The wide-angle transformation matrix M (5N
flat-sky
th
Γ— 3N
th
= 2000 Γ— 1200).
3. The covariance matrix C
conv
derived from the mock catalogs (5N
o
Γ—5N
o
= 200 Γ—200).
4. The power spectrum multipole measurements for the mock catalogs and data in βˆ†k
o
=
0.001 h Mpc
βˆ’1
(note that to use them with the products above you need to re-bin to
βˆ†k
o
= 0.01 h Mpc
βˆ’1
, see combine_bins parameter in pk_tools.read_power()). For
BOSS DR12 we also provide the post-reconstruction power spectrum measurements.
All quantities assume 5 multipoles in the k-range 0 < k < 0.4 h Mpc
βˆ’1
in bins of βˆ†k
o
=
0.01 h Mpc
βˆ’1
and βˆ†k
th
= 0.001 h Mpc
βˆ’1
. The only exception is the square window function
(200 Γ—200) meant to be used for deconvolution, which assumes βˆ†k
th
= βˆ†k
o
= 0.01 h Mpc
βˆ’1
.
All products listed above are available at https://fbeutler.github.io/hub/deconv_paper.
html and Python-based modules to read these quantities are available at https://github.
com/fbeutler/pk_tools.
We note that due to the Nyquist frequency in the measured power spectra, any analysis
should be limited to k
max
< 0.3 h Mpc
βˆ’1
. The purpose for the addition of the k-range
0.3 < k < 0.4 h Mpc
βˆ’1
is mainly to allow a sensible window function contribution from outside
the fitting range. We generally advice to exclude the first k-bin (k
min
> 0.01 h Mpc
βˆ’1
), which
could suffer from large scale systematics not studied in detail in this analysis.
Here is an example of a likelihood analysis using Python:
1 import numpy as np
2 import pk_tools # see above
3 # Read data power spectrum (Delta k_o =0.01 h/ Mpc ) as dictionary
4 pk_data_dict = pk_tools . read_power(pkfile , combine_bins =10)
5 # Turn dictionary into vector as [P_0 ,P_1 , P_2 ,P_3 ,P_4]
6 kbins , pk_data_vector = pk_tools . dict_to_vec ( pk_data_dict )
7 # Read covariance matrix
8 C = pk_tools . read_matrix ( covfile )
9 Cinv = np. linalg.inv (C)
10 # Read window matrix ( section 2, eq .~2.14)
11 W = pk_tools . read_matrix ( Wfile )
12 # Read expansion matrix ( section 3, eq .~3.2)
13 M = pk_tools . read_matrix ( Mfile )
14 # Get your favourite power spectrum model as dictionary
15 pk_model_dict = your_favourite_model ()
16 # Turn dictionary into vector as [P_0 ,P_2 , P_4 ]
17 kbins_long , pk_model_vector = pk_tools . dict_to_vec ( pk_model_dict , use_ell =[0 , 2 ,4])
18 # Expand pk model (true ,flatβˆ’sky) βˆ’> ( true) see eq .~3.1
19 expanded_model = np. matmul (M, pk_model_vector )
20 # Convolve with window (true ) βˆ’> (conv ) see eq .~2.18
21 convolved_model = np. matmul(W , expanded_model )
22 # Calculate chi2
23 diff = pk_data_vector βˆ’ convolved_model
24 chi2 = np. dot (diff ,np. dot (Cinv ,diff ))
– 30 –
Note that you can speed up these calculations by multiplying the matrices W and M be-
forehand, which reduces the required matrix multiplications in the likelihood evaluation to
one.
How can I limit the k-range? The power spectrum model should always have a k-range
of 0 < k
th
< 0.4 h Mpc
βˆ’1
in bins of βˆ†k
th
= 0.001 h Mpc
βˆ’1
. If your model cannot predict
the power spectrum up to k
max
= 0.4 h Mpc
βˆ’1
, you should provide some β€œsensible” estimate,
so that the window function contributions from those scales can be included. The likelihood
analysis should be limit to k
max
< 0.3 h Mpc
βˆ’1
, since the power spectra outside this k-range
suffer from aliasing given that k
Ny
β‰ˆ 0.6 h Mpc
βˆ’1
for all measured power spectra. One can
specify the k-range as
1 kmin = 0.01
2 kmax = 0.3
3 # Assuming 40 bins in k_o with Delta k_o =0.01 h/ Mpc
4 krange = np . linspace (0.005 , 0.395 , num =40)
5 fit_selection = np. logical_and ( kmin <krange ,krange < kmax)
6 # Select fittingβˆ’range for data power spectrum and inverse C
7 fit_pk_data_vector = pk_data_vector [ fit_selection ]
8 fit_Cinv = np. linalg . inv (C[np. ix_ ( fit_selection , fit_selection )])
9 # Select fittingβˆ’range for the convolved model power spectrum
10 fit_model = convolved_model [ fit_selection ]
11 # ... and proceed to calculate $\ chi ^2 $
How can I limit my analysis to the monopole? Since the window function couples
the different multipoles you have to provide a model for all multipoles, so that the window
function contributions can be calculated. For the likelihood itself one can limit the analysis
to the monopole by
1 # Select multipoles to be included [P_0 , P_1 , P_2 , P_3 , P_4]
2 pole_selection = [ True , False , False , False , False ]
3 fit_selection = np. repeat(pole_selection , 40)
4 # Adjust data power spectrum and inverse C
5 fit_pk_data_vector = pk_data_vector [ fit_selection ]
6 fit_Cinv = np. linalg . inv (C[np. ix_ ( fit_selection , fit_selection )])
7 # Adjust convolved model power spectrum
8 fit_model = convolved_model [ fit_selection ]
9 # ... and proceed to calculate $\ chi ^2 $
Speeding up the likelihood evaluation: Often it is possible to achieve a speedup in
the likelihood analysis if the convolution of the power spectrum model does not have to be
performed in every model-data comparison. We can re-write the likelihood as [27]
L ∼ exp
ξ€”
βˆ’
1
2
(P
conv
o
βˆ’ WMP
true,flat-sky
)
T
C
βˆ’1
conv
(P
conv
o
βˆ’ WMP
true,flat-sky
)

(A.1)
= exp
ξ€”
βˆ’
1
2
P
conv,T
o
C
βˆ’1
conv
P
conv
o
+ P
true,flat-sky,T
M
T
W
T
C
βˆ’1
conv
P
conv
o
βˆ’
1
2
P
true,flat-sky,T
M
T
W
T
C
βˆ’1
conv
WMP
true,flat-sky

(A.2)
= exp
ξ€”
βˆ’
1
2
P
conv,T
o
C
βˆ’1
conv
P
conv
o
+ P
true,flat-sky
C
βˆ’1
conv,W
P
conv
o
βˆ’
1
2
P
true,flat-sky,T
C
βˆ’1
conv,WW
P
true,flat-sky

(A.3)
– 31 –
with
C
βˆ’1
conv,WW
= M
T
W
T
C
βˆ’1
conv
WM (A.4)
C
βˆ’1
conv,W
= M
T
W
T
C
βˆ’1
conv
. (A.5)
An implementation of these equations could look like this:
1 # Select multipoles to be included [P_0 , P_1 , P_2 , P_3 , P_4]
2 pole_selection = [ True , False , False , False , False ]
3 fit_selection = np. repeat(pole_selection , 40)
4 # Select the fitting range
5 Cinv = np. linalg.inv (C[ np .ix_( fit_selection , fit_selection )])
6 pk_data_vector = pk_data_vector [ fit_selection ]
7 W = W[ np.ix_ ( fit_selection , np .ones (2000) < 1.1)]
8 # Preβˆ’calculate the two covariance matrices and the
9 # dataβˆ’data contribution
10 WCinv = np.matmul(np . transpose (W), Cinv )
11 MWCinv = np .matmul (np. transpose (M), WCinv )
12 MWCinvD = np . matmul (MWCinv , pk_data_vector )
13 MWCinvW = np . matmul (MWCinv , W)
14 MWCinvWM = np. matmul ( MWCinvW , M)
15 data_term = np. dot ( pk_data_vector , np.dot (Cinv , pk_data_vector ))
16
17 ## The following part needs to be run for every parameter evaluation
18 # Get your favourite power spectrum model as disctionary
19 pk_model_dict = your_favourite_model ()
20 # Turn dictionary into vector as [P_0 ,P_2 , P_4 ]
21 kbins_long , pk_model_vector = pk_tools . dict_to_vec ( pk_model_dict , use_ell =[0 , 2 ,4])
22 # Calculate chi2
23 chi2 = data_term
24 chi2 βˆ’= 2.βˆ—np.dot ( pk_model_vector , MWCinvD )
25 chi2 += np. dot (pk_model_vector , np.dot ( MWCinvWM , pk_model_vector ))
This implementation is about a factor of two times faster than the brute-force implementation
shown in the first code example. Of course that is only significant if the window function
convolution is dominating the likelihood evaluation. In many cases most of the time will be
spend in your_favourite_model().
How can I account for uncertainties in the covariance matrix? When deriving the
covariance matrix from a finite set of mock realisations, the resulting likelihood is no longer
Gaussian, but follows a t-distribution. If assuming a Gaussian likelihood the parameter
inference will be biased and this bias depends on the ratio of bins in the data vector and the
number of mock realisations [85]. We can account for this by scaling the likelihood as
ln L ∝ βˆ’Ο‡
2
N
m
βˆ’ N
d
βˆ’ 2
2(N
m
βˆ’ 1)
, (A.6)
where N
m
is the number of mock realisations (given in table 1) and N
d
is the size of the data
vector. Alternatively one can directly account for the non-Gaussian likelihood as proposed in
[86]. For all cases discussed in this paper this approach agrees very well with eq. (A.6).
Assuming your MCMC sampler expects log L as a return value you could implement
this equation as
1 # Hartlap et al. (2007)
2 H = Nmocks βˆ’ len ( pk_data_vector ) βˆ’ 2
3 H /= ( Nmocks βˆ’ 1)
4 return βˆ’chi2βˆ—H /2.
– 32 –
If the final parameter uncertainty is derived from the likelihood itself, you need to account
for a bias caused by the mock based covariance estimate. We can do that with a re-scaling
of the parameter errors [87, 88] by the square root of
m
1
=
1 + B(N
d
βˆ’ N
p
)
1 + A + B(N
p
+ 1)
(A.7)
with
A =
2
(N
m
βˆ’ N
d
βˆ’ 1)(N
m
βˆ’ N
d
βˆ’ 4)
, (A.8)
B =
N
m
βˆ’ N
d
βˆ’ 2
(N
m
βˆ’ N
d
βˆ’ 1)(N
m
βˆ’ N
d
βˆ’ 4)
(A.9)
(we find that the approach of [86] does not take care of this factor).
How can I deconvolve a power spectrum measurement? To perform a deconvolution,
all you have to do is to follow eq. (4.2). The difficulty here is that this equation only holds for
square matrices W. The window functions using βˆ†k
o
= βˆ†k
th
= 0.01 h Mpc
βˆ’1
are available.
1 # Read data power spectrum (Delta k_o =0.01 h/ Mpc ) as disctionary
2 pk_data_dict = pk_tools . read_power(pkfile , combine_bins =10)
3 # Turn dictionary into vector as [P_0 ,P_1 , P_2 ,P_3 ,P_4]
4 kbins , pk_data_vector = pk_tools . dict_to_vec ( pk_data_dict )
5 # Read window matrix (needs to be a square matrix )
6 W = pk_tools . read_matrix ( Wfile )
7 # Invert window function
8 Winv = np. linalg.inv (W)
9 # Deconvolution
10 pk_data_vector_deconvolved = np. matmul(Winv , pk_data_vector )
B Window function pre-factors
Focusing on the first 5 multipoles (0 ≀ L ≀ 4) of the window function, including wide-angle
terms up to second order (n < 3), the weights for the individual contributions in eq. (2.5) are
given by
C
(0)
0`L
= C
(2)
0`L
=





ο£­
1 0 0 0 0
0 0 0 0 0
0 0
1
5
0 0
0 0 0 0 0
0 0 0 0
1
9
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(1)
0`L
=





ο£­
0 0 0 0 0
0
1
3
0 0 0
0 0 0 0 0
0 0 0
1
7
0
0 0 0 0 0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(0)
1`L
= C
(2)
1`L
=





ο£­
0 1 0 0 0
0 0 0 0 0
0
2
5
0
9
35
0
0 0 0 0 0
0 0 0
4
21
0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(1)
1`L
=





ο£­
0 0 0 0 0
1 0
2
5
0 0
0 0 0 0 0
0 0
9
35
0
4
21
0 0 0 0 0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(0)
2`L
= C
(2)
2`L
=





ο£­
0 0 1 0 0
0 0 0 0 0
1 0
2
7
0
2
7
0 0 0 0 0
0 0
2
7
0
100
693
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(1)
2`L
=





ο£­
0 0 0 0 0
0
2
3
0
3
7
0
0 0 0 0 0
0
3
7
0
4
21
0
0 0 0 0 0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
(B.1)
– 33 –
and
C
(0)
3`L
= C
(2)
3`L
=





ο£­
0 0 0 1 0
0 0 0 0 0
0
3
5
0
4
15
0
0 0 0 0 0
0
4
9
0
2
11
0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(1)
3`L
=





ο£­
0 0 0 0 0
0 0
3
5
0
4
9
0 0 0 0 0
1 0
4
15
0
2
11
0 0 0 0 0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(0)
4`L
= C
(2)
4`L
=





ο£­
0 0 0 0 1
0 0 0 0 0
0 0
18
35
0
20
77
0 0 0 0 0
1 0
20
77
0
162
1001
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
C
(1)
4`L
=





ο£­
0 0 0 0 0
0 0 0
4
7
0
0 0 0 0 0
0
4
7
0
18
77
0
0 0 0 0 0
ο£Ά
ο£·
ο£·
ο£·
ο£·
ο£Έ
,
(B.2)
where the wide-angle correction terms at n = 0 and n = 2 have the same shape.
C Derivation of the 2D window function
In this section we derive the equation for the 2D window function W
(n)
``
(k, k
0
). We first derive
this equation for n = 0 showing consistency with [1], where this equation first appeared. We
than include the wide-angle correction terms following [2], which leads to our eq. (2.5).
C.1 Excluding wide-angle terms and consistency with [1]
Here we show the relation between eq. (33) of [1] and eq. (2.2). The convolution of the power
spectrum can be written as (using a LOS of
Λ†
d =
Λ†
s
1
and following eq. B.1 of [1])
P
conv
`
(k) =
2` + 1
2
Z
dΒ΅
Z
dφ
2Ο€
Z
dk
0
(2Ο€)
3
P
true
(k
0
)|W (k βˆ’k
0
)|
2
L
`
(
Λ†
k
0
Β·
Λ†
s
1
) (C.1)
=
2` + 1
2
Z
dΒ΅
Z
dφ
2Ο€
Z
dΒ΅
0
Z
dφ
0
4Ο€
Z
dk
0
k
02
2Ο€
2
P
true
(k
0
, Β΅
0
)Γ—
N
ran
X
i,j,i=j
w
FKP
(x
i
)w
FKP
(x
j
)e
ikΒ·s
e
βˆ’ik
0
Β·s
L
`
(
Λ†
k Β·
Λ†
s
1
) .
(C.2)
Using multipole expansion as well as
e
iksΒ΅
=
X
L
i
L
(2L + 1)j
L
(ks)L
L
(Β΅) , (C.3)
L
`
(
Λ†
s
1
Β·
Λ†
s)Ξ΄
``
0
=
2` + 1
2
Z
dΒ΅
Z
dφ
2Ο€
L
`
(
Λ†
k Β·
Λ†
s)L
`
0
(
Λ†
k Β·
Λ†
s
1
) , (C.4)
we get
10
P
conv
`
(k) = (βˆ’i)
L
i
`
(2` + 1)
Z
dk
0
k
02
2Ο€
2
X
L
P
true
L
(k
0
)j
L
(k
0
s)j
`
(ks)Γ—
N
ran
X
i,j,i=j
w
FKP
(x
i
)w
FKP
(x
j
)L
`
(
Λ†
s
1
Β·
Λ†
s)L
L
(
Λ†
s
1
Β·
Λ†
s) .
(C.5)
10
Note that our Fourier transform convention is P (k) ≑
R
dx e
βˆ’ikΒ·x
ΞΎ(x).
– 34 –
Now using
N
ran
X
i,j,i=j
w
FKP
(x
i
)w
FKP
(x
j
) =
Z
d
3
s
1
Z
d
3
s
2
W (s
1
)W (s
2
)
=
Z
ds s
2
Z
dΩ
s
Z
d
3
s
1
W (s
1
)W (s + s
1
)
(C.6)
and the definition (see eq. 2.21 of [2])
Q
L
(s) = (2L + 1)
Z
dΩ
s
4Ο€
Z
d
3
s
1
W (s
1
)W (s + s
1
)L
L
(
Λ†
s Β·
Λ†
s
1
) (C.7)
as well as
L
`
(
Λ†
s
1
Β·
Λ†
s)L
L
(
Λ†
s
1
Β·
Λ†
s) =
`+L
X
p=|`βˆ’L|
(2p + 1)
ξ€’
` L p
0 0 0

2
L
p
(
Λ†
s
1
Β·
Λ†
s) , (C.8)
results in
P
conv
`
(k) = 4Ο€(βˆ’i)
L
i
`
Z
dk
0
k
02
2Ο€
2
X
L
P
true
L
(k
0
)Γ—
Z
ds s
2
j
L
(k
0
s)j
`
(ks)(2` + 1)
`+L
X
p=|`βˆ’L|
ξ€’
` L p
0 0 0

2
Q
p
(s)
(C.9)
=
Z
dk
0
k
02
X
L
W
`L
(k, k
0
)P
true
L
(k
0
) , (C.10)
where
W
`L
(k, k
0
) = (βˆ’i)
L
i
`
2
Ο€
Z
ds s
2
j
L
(k
0
s)j
`
(ks)A
`L
(s) (C.11)
with
A
`L
(s) = (2` + 1)
`+L
X
p=|`βˆ’L|
ξ€’
` L p
0 0 0

2
Q
p
(s) (C.12)
=
`+L
X
p=|`βˆ’L|
C
`Lp
Q
p
(s) (C.13)
and the factors C
`Lp
= C
(0)
`Lp
are given in appendix B.
C.2 Including wide-angle terms and consistency with [2]
The convolution of the power spectrum multipoles including the wide-angle correction terms
has first been derived in [2] and is given by
11
P
conv
`
(k) = (βˆ’i)
`
(2` + 1)
4Ο€
X
L
`+L
X
p=|`βˆ’L|
ξ€’
L p `
0 0 0

2
Z
ds
X
n
s
n+2
j
`
(ks)ΞΎ
(n)
L
(s)Q
(n)
p
(s) . (C.14)
11
We added a factor of 1/4Ο€ to account for the difference in the window function definition (see eq. 2.21
in [2] and eq. 2.6 in this paper.)
– 35 –
Now using
ΞΎ
(n)
L
= i
L
Z
dk k
2
2Ο€
2
(ks)
βˆ’n
P
(n)
L
(k)j
L
(ks) (C.15)
we get
P
conv
`
(k) = (βˆ’i)
`
2
Ο€
Z
ds
X
L
C
(n)
`L
(s)
X
n
s
n+2
j
`
(ks)Γ—
i
L
Z
k
02
dk
0
(k
0
s)
βˆ’n
P
(n),true
L
(k
0
)j
L
(k
0
s)
(C.16)
=
Z
dk
0
k
02βˆ’n
X
L,n
W
(n)
`L
(k, k
0
)P
(n),true
L
(k
0
) , (C.17)
where
W
(n)
`L
(k, k
0
) = (βˆ’i)
`
i
L
2
Ο€
Z
ds s
2
j
L
(k
0
s)j
`
(ks)A
(n)
`L
(s) (C.18)
and
A
(n)
`L
(s) = (2` + 1)
`+L
X
p=|`βˆ’L|
ξ€’
L p `
0 0 0

2
Q
(n)
p
(s) (C.19)
=
`+L
X
p=|`βˆ’L|
C
(n)
`Lp
Q
(n)
p
(s) . (C.20)
The factors C
(n)
`Lp
are given in appendix B. The equation above is consistent with eq. (C.12)
since
ξ€’
`
1
`
2
`
3
0 0 0

2
=
ξ€’
`
2
`
3
`
1
0 0 0

2
. (C.21)
D Analytic calculation of the odd power spectrum multipoles
Here we derive eq. (2.14) and eq. (2.15) used in section 2.1. The dipole power spectrum is
given by
P
(1)
1
(k) = βˆ’ik
6
5Ο€
Z
dk
0
k
02
P
(0)
2
(k
0
)
Z
s
3
ds j
2
(k
0
s)j
1
(ks) . (D.1)
Following eq. (F.1), (F.6) and (F.10) of [12] we can write
2
Ο€
Z
s
3
ds j
2
(k
0
s)j
1
(ks) = βˆ’k
0
βˆ‚
k
0
ξ€”
1
k
03
Ξ΄
D
(k
0
βˆ’ k)

, (D.2)
where Ξ΄
D
is the Dirac delta function. Using
I
11βˆ’1
(k, k
0
) =
1
k
2
Ξ΄
D
(k
0
βˆ’ k) (D.3)
to replace the integral with a derivative, results in
P
(1)
1
(k) = ik
3
5
Z
dk
0
k
03
P
(0)
2
(k
0
)βˆ‚
k
0
ξ€”
1
k
03
Ξ΄
D
(k
0
βˆ’ k)

(D.4)
= ik
3
5
Z
dk
0
k
03
P
(0)
2
(k
0
)
ξ€”
βˆ’
3
k
04
Ξ΄
D
(k
0
βˆ’ k) +
1
k
03
βˆ‚
k
0
Ξ΄
D
(k
0
βˆ’ k)

. (D.5)
– 36 –
Now we can use
Z
dxf(x)βˆ‚
x
Ξ΄
D
(x βˆ’ a) = βˆ’βˆ‚
a
f(a) , (D.6)
yielding
P
(1)
1
(k) = βˆ’i
3
5
h
3P
(0)
2
(k) + kβˆ‚
k
P
(0)
2
(k)
i
(D.7)
= βˆ’if
ξ€’
4
5
b
1
+
12
35
f

[3P
m
(k) + kβˆ‚
k
P
m
(k)] , (D.8)
where the second line assumes linear theory. This final equation is numerically much easier
to evaluate compared to the equation we started with.
Equivalently we can derive the octopole
P
(1)
3
(k) = βˆ’i
ξ€”
3
5

2P
(0)
2
(k) βˆ’ kβˆ‚
k
P
(0)
2
(k)

+
10
9

5P
(0)
4
(k) + kβˆ‚
k
P
(0)
4
(k)


(D.9)
= βˆ’i4f
ξ€”
1
5
ξ€’
b
1
+
3
7
f

(2P
m
(k) βˆ’ kβˆ‚
k
P
m
(k)) +
4
63
f (5P
m
(k) + kβˆ‚
k
P
m
(k))

.
(D.10)
E Deconvolution results for all datasets
Figure 14, 15 and 16 show the comparisons for the convolved and deconvolved power spectra
for 6dFGS DR3, BOSS DR12 and the eBOSS DR16 QSO samples. Similar plots for the low
redshift bin of BOSS DR12 NGC are included in the main text (see figure 11 and 12). The
remaining plots are included here rather than the main text, to not interrupt the flow of the
paper.
Figure 14: Comparison between the power spectrum multipoles of 6dFGS DR3 measured
in the mock catalogs (gray and red shaded area) and in the data (data points). The results
before deconvolution are shown as the red shaded area and solid red line (mocks) and the
green data points. The deconvolved results are shown as the gray shaded area and solid black
line (mocks) and the blue data points. The residuals in the lower panel show that wide-angle
and window function effects are sub-dominant in 6dFGS.
– 37 –
Figure 15: Comparison between the power spectrum multipoles of BOSS DR12 measured
in the mock catalogs (gray and red shaded area) and in the data (data points). The results
before deconvolution are shown as the red shaded area and solid red line (mocks) and the
green data points. The deconvolved results are shown as the gray shaded area and solid black
line (mocks) and the blue data points. The equivalent BOSS DR12 NGC results for the low
redshift bin are included in the main text (see figure 11 and 12).
– 38 –
Figure 16: Comparison between the power spectrum multipoles of the eBOSS DR16 QSO
sample measured in the mock catalogs (gray and red shaded area) and in the data (data
points). The results before deconvolution are shown as the red shaded area and solid red line
(mocks) and the green data points. The deconvolved results are shown as the gray shaded
area and solid black line (mocks) and the blue data points. The increasing in the dipole power
spectrum at high k seems to be an aliasing effect, even though the Nyquist frequency is twice
as high as the scale range plotted here (k
Ny
> 0.6 h Mpc
βˆ’1
).
– 39 –