DES 2017-0727
FERMILAB-PUB-17-291-E
MNRAS 000, 000–000 (0000) Preprint 7 August 2017 Compiled using MNRAS L
A
T
E
X style file v3.0
Dark Energy Survey Year 1 Results: The Impact of Galaxy
Neighbours on Weak Lensing Cosmology with IM3SHAPE
S. Samuroff
1?
, S. L. Bridle
1
, J. Zuntz
2
, M. A. Troxel
3,4
, D. Gruen
5,6
, R. P. Rollins
1
, G. M. Bernstein
7
,
T. F. Eifler
8
, E. M. Huff
8
, T. Kacprzak
9
, E. Krause
5
, N. MacCrann
3,4
, F. B. Abdalla
10,11
, S. Allam
12
,
J. Annis
12
, K. Bechtol
13
, A. Benoit-L
´
evy
14,10,15
, E. Bertin
14,15
, D. Brooks
10
, E. Buckley-Geer
12
, A.
Carnero Rosell
16,17
, M. Carrasco Kind
18,19
, J. Carretero
20
, M. Crocce
21
, C. B. D’Andrea
7
, L. N. da
Costa
16,17
, C. Davis
5
, S. Desai
22
, P. Doel
10
, A. Fausti Neto
16
, B. Flaugher
12
, P. Fosalba
21
, J. Frieman
12,23
,
J. Garc
´
ıa-Bellido
24
, D. W. Gerdes
25,26
, R. A. Gruendl
18,19
, J. Gschwend
16,17
, G. Gutierrez
12
,
K. Honscheid
3,4
, D. J. James
27,28
, M. Jarvis
7
, T. Jeltema
29
, D. Kirk
10
, K. Kuehn
30
, S. Kuhlmann
31
,
T. S. Li
12
, M. Lima
32,16
, M. A. G. Maia
16,17
, M. March
7
, J. L. Marshall
33
, P. Martini
3,34
, P. Melchior
35
,
F. Menanteau
18,19
, R. Miquel
36,20
, B. Nord
12
, R. L. C. Ogando
16,17
, A. A. Plazas
8
, A. Roodman
5,6
,
E. Sanchez
37
, V. Scarpine
12
, R. Schindler
6
, M. Schubnell
26
, I. Sevilla-Noarbe
37
, E. Sheldon
38
, M. Smith
39
,
M. Soares-Santos
12
, F. Sobreira
40,16
, E. Suchyta
41
, G. Tarle
26
, D. Thomas
42
, D. L. Tucker
12
(DES Collaboration)
Author affiliations are listed at the end of the paper.
7 August 2017
ABSTRACT
We use a suite of simulated images based on Year 1 of the Dark Energy Survey to explore the
impact of galaxy neighbours on shape measurement and shear cosmology. The HOOPOE im-
age simulations include realistic blending, galaxy positions, and spatial variations in depth
and PSF properties. Using the IM3SHAPE maximum-likelihood shape measurement code,
we identify four mechanisms by which neighbours can have a non-negligible influence on
shear estimation. These effects, if ignored, would contribute a net multiplicative bias of
m 0.03 0.09 in the DES Y1 IM3SHAPE catalogue, though the precise impact will be
dependent on both the measurement code and the selection cuts applied. This can be reduced
to percentage level or less by removing objects with close neighbours, at a cost to the effective
number density of galaxies n
eff
of 30%. We use the cosmological inference pipeline of DES
Y1 to explore the cosmological implications of neighbour bias and show that omitting blend-
ing from the calibration simulation for DES Y1 would bias the inferred clustering amplitude
S
8
σ
8
(Ω
m
/0.3)
0.5
by 2σ towards low values. Finally, we use the HOOPOE simulations to
test the effect of neighbour-induced spatial correlations in the multiplicative bias. We find the
impact on the recovered S
8
of ignoring such correlations to be subdominant to statistical error
at the current level of precision.
Key words: cosmological parameters - cosmology: observations - gravitational lensing: weak
- galaxies: statistics
1 INTRODUCTION
A standard and well tested prediction of General Relativity is that
a concentration of mass will distort the spacetime around it, and
thus produce a curious phenomenon called gravitational lensing.
?
NASA Einstein Fellow
The most obvious manifestation is about massive galaxy clusters,
where background galaxies can be elongated into cresent-shaped
arcs. So-called strong lensing of galaxies was first observed in the
late 1980s and has been confirmed many times since. A subtler, but
from a cosmologist’s perspective more powerful, consequence of
gravitational lensing is that background fluctuations in the density
of dark matter will induce coherent distortions to photons’ paths.
This effect is known as cosmic shear, and it was first detected
c
0000 The Authors
arXiv:1708.01534v1 [astro-ph.CO] 4 Aug 2017
2 S. Samuroff, S. L. Bridle, J. Zuntz et al
by four groups at around the same time close to two decades ago
(Bacon et al. 2000; Van Waerbeke et al. 2000; Kaiser et al. 2000;
Wittman et al. 2000). Cosmic shear has the potential to be the sin-
gle most powerful probe in the toolbox of modern cosmology. The
spatial correlations due to lensing are a direct imprint of the large
scale mass distribution of the Universe. Thus it allows one to study
the total mass of the Universe and the growth of structure within
it (Maoli et al. 2001; Jarvis et al. 2006; Massey et al. 2007; Kil-
binger et al. 2013; Heymans et al. 2013; Abbott et al. 2016; Jee
et al. 2016; Hildebrandt et al. 2017; K
¨
ohlinger et al. 2017), or to
map out the spatial distribution of dark matter on the sky (eg Kaiser
1994; Van Waerbeke et al. 2013; Chang et al. 2015). As a probe of
both structure and geometry, cosmic shear is also attractive as a
method for shedding light on the as yet poorly understood compo-
nent of the Universe known as dark energy (Albrecht et al. 2006;
Weinberg et al. 2013). Alternatively, lensing will allow us to place
ever more stringent tests of our theories of gravity (Simpson et al.
2013; Harnois-D
´
eraps et al. 2015; Brouwer et al. 2017). Is also the-
oretically very clean, responding directly to the power spectrum of
dark matter, which is affected by baryonic physics only on small
scales, and avoids recourse to poorly-understood phenomenologi-
cal rules. Indeed galaxy number density enters only at second order
as a weighting of the observed shear due to the fact that one can
only sample the shear field where there are real galaxies (Schmidt
et al. 2009).
Though well modelled theoretically, cosmic shear is techni-
cally highly challenging to measure; as with all these probes it is
not without its own sources of systematic error. It also cannot be
reiterated too many times that the shear component of even the
most distant galaxy’s shape is subdominant to noise by an order of
magnitude. Indeed, the ambitions of the current generation of cos-
mology surveys will require sub-percent level uncertainties (both
systematic and statistical) on what is already a tiny cosmological
ellipticity component g 0.01.
It was realised early on how significant the task of translat-
ing photometric galaxy images into unbiased shear measurements
would be. In response came a series of blind shear measurement
challenges, designed to review, test and compare the best methods
available. The first of these, called STEP1 (Heymans et al. 2006)
grew out of a discussion at the 225th IAU Symposium in 2004.
The exercise was based around a set of simple SKYMAKER sim-
ulations (Bertin & Fouqu
´
e 2010), which were designed to mimic
ground based observations but with analytic galaxies and PSFs and
constant shear. The algorithms at this point represented a first wave
of shear measurement codes and included several moments-based
algorithms (Kaiser et al. 1995; Kuijken 1999; Rhodes et al. 2001),
some early forward modelling methods (Bridle et al. 2002), as well
as a technique called shapelets, which models a light profile as a set
of 2D basis functions (Bernstein & Jarvis 2002; Refregier & Bacon
2003).
The simulations and the codes themselves steadily grew in
complexity. STEP2 was followed by series of GREAT challenges
(Massey et al. 2007; Bridle et al. 2009; Kitching et al. 2010; Man-
delbaum et al. 2014), which focused on different aspects of shape
measurement bias and have been essential in quantifying a num-
ber of significant effects. In recent years the drive to find ever
more accurate ways to measure shear has intensified, with many
novel approaches being suggested. For example Fenech Conti et al.
(2017) use a form of self-calibration, which repeats the shape mea-
surement on a test image based on the best-fitting model for each
galaxy. A related approach, named metacalibration, involves de-
riving corrections to the galaxy shape measurements directly from
the data, using modified copies of the image with additional shear
(Huff & Mandelbaum 2017; Sheldon & Huff 2017). More advanced
moments-based approaches include the BFD method (Bernstein &
Armstrong 2014), which derives a prior on the ensemble elliptic-
ity distribution using deeper fields, and SNAPG (Herbonnet et al.
2017), a similar approach which builds ensemble shear estimates
using shear nulling.
This paper is intended as a companion study to Zuntz et al.
(2017) (Z17), where we present two shear catalogues derived from
DES Y1 dataset. It is also presented alongside a raft of other pa-
pers, which use both catalogues and show them to be consistent in
a number of different scientific contexts (Troxel et al. 2017; Prat
et al. 2017; Chang et al. 2017; DES Collaboration et al. 2017) Con-
taining 22 million and 35 million galaxies respectively, these cat-
alogues are the product of two independent maximum likelihood
codes. The first, called IM3SHAPE, implements simultaneous fits
using multiple models and we calibrate externally using simula-
tions. The second implements a Gaussian model fitting algorithm,
supplemented by shear response corrections using METACALIBRA-
TION. Whereas in Z17 we focus on the catalogues themselves, pre-
senting a raft of calibration tests and a broad overview of the value-
added data products, here we use the same resources to explore
a narrower topic: the impact of image plane neighbours on shear
measurement. Specifically we use the image simulations described
in Z17, from which the Y1 IM3SHAPE calibration is derived, to ex-
plore the mechanisms for neighbour bias, and then propagate the re-
sults to mock shear two-point data to investigate the consequences
for weak lensing cosmology. The results presented in this paper will
be somewhat dependent on the choice of measurement algorithm,
selection cuts and the configuration of the object detection code.
Unlike previous studies on this subject, however, we make use of a
highly realistic simulation and measurement pipeline. Our choices
on each of aspects are realistic, if not unique, for a leading-edge
cosmology analysis.
It is worth remarking, however, that the tests described in this
paper make use of IM3SHAPE only, and should not be assumed to
apply generically to its sister Y1 METACALIBRATION catalogue.
A complementary set of tests using METACALIBRATION are pre-
sented in §4.5 of Z17.
This paper is structured as follows. In Section 2 we briefly re-
view the formalism of lensing, and the observables discussed in this
work. In Section 3 we present a series of numerical calculations us-
ing a toy model to characterise neighbour bias. Section 4 decribes
the simulated DES Y1 datasets, generated using our HOOPOE sim-
ulator. We test the earlier predictions under more typical observing
conditions in Section 5, and extend them into a quantitative set of
results using the more extensive Y1 HOOPOE dataset. Section 6
then presents a numerical analysis designed to test the cosmolog-
ical implications of neighbour bias of the nature and magnitude
found in our simulations. We conclude in Section 7.
2 THE SHEAR MEASUREMENT PROBLEM
The problem of shape measurement is far more intricate than it
might first appear. Any cosmological analysis based on cosmic
shear is reliant on a series of technical choices, which can have
a non-trivial impact on measurement biases, precision and cosmo-
logical sensitivity. Specifically we must choose (a) how to parame-
terize each galaxy’s shape, and which measurement method to use
to estimate it, (b) what selection criteria are needed to obtain data
of sufficiently high quality for cosmology and (c) how biased is the
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 3
measurement and what correction is needed? These choices should
be made on a case-by-case basis, since the optimal solutions are
dependent on a number of survey-specific factors. We discuss each
briefly in turn below.
2.1 Shape Measurement with IM3SHAPE
The shape measurements upon which the following analyses are
based make use of the maximum likelihood model fitting code
IM3SHAPE
1
(Zuntz et al. 2013). It is a well tested and understood
algorithm, which has since been used in a range of lensing studies
(Abbott et al. 2016; Whittaker et al. 2015; Kacprzak et al. 2016;
Clampitt et al. 2017). It was also one of two codes used to produce
shear catalogues in the Science Verification (SV) stage and Year
1 of the Dark Energy Survey. We refer the reader to Jarvis et al.
(2016) (hereafter J16) and Z17 for the most recent modifications to
the code.
We use the definition of the flux signal-to-noise ratio of Z17,
J16 and Mandelbaum et al. (2015):
S/N
N
pix
P
i=1
f
m
i
f
im
i
2
i
!
N
pix
P
i=1
f
m
i
f
m
i
2
i
!
1
2
. (1)
The indices i = (1, 2...N
pix
) run over all pixels in a stack of im-
age cutouts at the location of a galaxy detection. The model pre-
diction and observed flux in pixel i are denoted f
m
i
and f
im
i
re-
spectively and σ
i
is the RMS noise. This signal-to-noise measure
is maximised when the differences between the model and the im-
age pixel fluxes are small. Note that if the best-fitting model f
m
is identical for two different postage stamps, S/N will favour the
image with the greater total flux.
A useful size measure, referred to as R
gp
/R
p
is defined as
the measured Full Width at Half Maximum (FWHM) of the galaxy
after PSF convolution, normalised to the PSF FWHM. Real galaxy
images are are not perfectly symmetric (i.e. size is not independent
of azimuthal angle about a galaxy’s centroid), and single-number
size estimates are obtained by circularising (azimuthally averag-
ing) the galaxy profile and computing the weighted quadrupole mo-
ments of the resulting image. For each galaxy we take the mean
measured size across exposures.
2.2 Shear Measurement Bias
There are many ways bias can enter an ensemble shear estimate
based on a population of galaxies. Although the list is not exhaus-
tive, a handful of mechanisms are particularly prevalent, and have
been extensively discussed in the literature.
Noise Bias: On addition of pixel noise to an image, the best-
fitting parameters of a galaxy model will not scale linearly with
the noise variance. This is as an estimator bias as much as a mea-
surement bias, and results in an asymmetric, skewed likelihood sur-
face (Hirata & Seljak 2003; Refregier et al. 2012; Kacprzak et al.
2012; Miller et al. 2013). Any code which uses the point statis-
tics of the distribution (either mean or maximum likelihood) as a
single-number estimates of the ellipticity results in a bias. This is
1
https://bitbucket.org/joezuntz/im3shape-git
true even in the idealised case where the galaxy we are fitting can
be perfectly decribed by our analytic light profile. The bias is sen-
sitive to the noise levels and also the size and flux of the galaxy,
and thus is specific to the survey and galaxy sample in question.
For likelihood-based estimates one solution would be to impose
a prior on the ellipticity distribution and propagate the full poste-
rior. However, the results can become dependent on the accuracy of
that prior, and such codes require cautious testing using simulations
(Bernstein & Armstrong 2014; Simon & Schneider 2016)
Model Bias: In reality galaxies are not analytic light profiles
with clear symmetries. For the purposes of model-fitting, however,
we are constrained to use models with a finite set of parameters.
A model which does not allow sufficient flexibility to capture the
range of morphological features seen in the images will produce
biased shape measurements (Lewis 2009; Voigt & Bridle 2010;
Kacprzak et al. 2014).
Selection Bias: Even if we were to devise an ideal shape mea-
surement algorithm, capable of perfectly reconstructing the his-
togram of ellipticities in a certain population of galaxies, our at-
tempts to estimate the cosmological shear could still be biased. If
a measurement step prefers rounder objects or those with a partic-
ular orientation, the result would be a net alignment that could be
mistaken as having cosmological origin. In practice selection bias
commonly arises from imperfect correction of PSF asymmetries
(eg Kaiser et al. 2000; Bernstein & Jarvis 2002), and the fact that
many detection algorithms fail less frequently on rounder galaxies
(Hirata & Seljak 2004). It is such effects that make post facto qual-
ity cuts on quantities such as signal-to-noise or size (both of which
correlate with ellipticity) particularly delicate.
Neighbour bias: In practice, galaxies in photometric surveys
like DES are not ideal isolated objects. Rather, they are extracted
from a crowded image plane using imperfect deblending algo-
rithms. The term “neighbour bias” refers to any biases in the re-
covered shear arising from the interaction between galaxies in the
image plane. This can include both the direct impact on the per-
galaxy shapes (e.g. Hoekstra et al. 2017) and changes in the selec-
tion function (e.g. Hartlap et al. 2011). Neighbour bias is the subject
of relatively few previous studies, and is the focus of this paper.
3 TOY MODEL PREDICTIONS
To develop a picture of how image plane neighbours affect shear
estimates with IM3SHAPE, we build a simplified toy model. Using
GALSIM
2
we generate a 48 × 48 pixel postage stamp containing a
single exponential disc profile convolved with a tiny spherically
symmetric PSF (though we confirm that our results are insensi-
tive to the exact size of the PSF). We can then apply a small shear
along one coordinate axis prior to convolution and use IM3SHAPE
to fit the resulting image. In the absence of noise or model bias
the maximum of the likelihood of the measured parameters coin-
cides exactly with the input values. The basic setup then has four
adjustable parameters: the flux and size of the galaxy plus two el-
lipticity components, denoted f
c
, r
c
, g
tr
1
and g
tr
2
. Unless otherwise
stated we fix these to the median values measured from the DES Y1
IM3SHAPE catalogue. We do not model miscentering error between
the true galaxy centroid and the stamp centre.
It is worth noting that neither this basic model nor the more
2
https://github.com/GalSim-developers/GalSim
MNRAS 000, 000–000 (0000)
4 S. Samuroff, S. L. Bridle, J. Zuntz et al
complex simulations that follow attempt to model spatial correla-
tions in shear. Even at different redshifts, a real neighbour-central
pair share some portion of their line of sight. These spatial cor-
relations will amplify the impact of blending, and are worthy of
future investigation. This is, however, likely a second-order effect
of neighbours, and we postpone such study to a future date.
3.1 Single-Galaxy Effects
To explore the interaction in single neighbour-galaxy instances we
introduce a second galaxy into the postage stamp, convolved with
the same nominal PSF. This adds four more model parameters:
neighbour size r
n
, flux f
n
, radial distance from the stamp centre
d
gn
and azimuthal rotation angle relative to the x coordinate axis
θ. At this stage the neighbour has zero ellipticity.
We show this setup at three neighbour positions in Fig. 1. Un-
der zero shear, the system has perfect rotational symmetry, and the
measured ellipticity magnitude ˜g(θ|g
tr
1
= 0) is independent of θ
3
As a first exercise, we generate a circular central galaxy with a cir-
cular Gaussian neighbour, which is gradually shifted outwards from
the stamp centre. Following the usual convention for galaxy-galaxy
lensing, tangential shear is defined such that it is negative when
the major axis of the measured shape is oriented radially towards
the neighbour. The measured two-component ellipticity shown by
the solid and dot-dashed lines in Fig. 2. The decline in the mea-
sured tangential shear to zero at small separations is understand-
able, as there is no reason to expect drawing one circular profile
directly atop another should induce spurious non-zero ellipticity. In
the regime of a few pixels, however, strong blending can increase
the flux of the best-fitting model.
Next, we repeat the calculation, now applying a moderate
cross-component shear to the neighbour (g
2
= 0.1). The result
is shown by the blue lines in Fig. 3. Unsurprisingly the measured
tangential shear is unaffected by a true shear along an orthogonal
axis. In cases where the objects share a large portion of their half-
light radii, we are fitting a strongly blended pair with a single pro-
file, and the neighbour/central distinction becomes difficult to de-
fine. The best-fitting ellipticity recovered from the blended image
is not a pure measurement of either galaxy’s shape; rather it is a
linear combination of the two. We repeat the zero-offset measure-
ment using a range of neighbour fluxes and find that the best-fitting
e
i
follows roughly as a flux-weighted sum over the two galaxies
˜g
i
(f
c
g
tr
i,c
+ f
n
g
tr
i,n
)/(f
n
+ f
c
).
3.2 Ensemble Biases
While useful for understanding what follows, the impact of neigh-
bours on individual galaxy instances is not particularly informative
about the impact on cosmic shear measurements. Even significant
bias in the per-object shapes could average away over many galax-
ies with no residual impact on the recovered shear. More impor-
tant is the collective response to neighbours. To explore this we
build on the toy model concept. To estimate the ensemble effect,
we measure a neighbour-central image at 70 positions on a ring
of neighbour angles. Again, under zero shear g
tr
= 0 the mea-
sured shape is constant in magnitude, and simply oscillates about
0 with peaks of amplitude |˜g(θ|0, d
gn
)|. This sinusoidal variation
is shown by the dotted lines in Fig. 3b at two values of d
gn
(7 and
3
Unless otherwise stated we fix the other model parameters to their fidu-
cial values.
8 pixels). By averaging over a (large) number of neighbours one is
effectively marginalising over θ, which results in an unbiased mea-
surement of the shear h˜g(θ|g
tr
= 0, d
gn
)i
θ
= g
tr
= 0. A non-
zero shear g
tr
6= 0, however breaks the symmetry of the system. A
galaxy sheared along one axis will not respond to a neighbour in the
same way irrespective of θ, which can result in a net bias. To show
this we fix g
tr
= 0.05 and proceed as before. The solid lines in
Fig. 3b show the periodicity in the measured shear at two d
gn
. The
mean value averaged over θ is shifted incrementally away from the
input shear, shown by the horizontal dot-dashed line. Specifically
we should note that the peaks below g
tr
at π/2 and 3π/2 radi-
ans are deeper and narrower than those above it. The cartoon in
Fig. 4 shows how this arises. The purple lines are iso-light con-
tours in a strongly sheared S
´
ersic disc profile (g
1
= 0.3). Clearly
rotating the neighbour from position A to C carries it from the rel-
atively flat low wings of the central galaxy’s light profile closer to
the core. Perturbing an object about C by a small angle results in
a much greater change in the local gradient, 5f
c
(x, y) than doing
the same about A. All other parameters fixed, an incremental shift
along the blue tangent vector will have a larger impact at θ = 0
than at π/2, resulting in asymmetry in the width of the positive and
negative peaks in Fig. 4. The depth of the peak can be explained
qualititatively by similar arguments. At C a neighbour of given flux
is closer to the centre of the light distribution and thus has a greater
flux overlap with the central galaxy than at A. Naturally, then, one
might expect neighbour A to have less impact than C. Returning to
Fig. 3, we can see that the two effects are in competition. Depend-
ing on the exact neighbour configuration, the simultaneous narrow-
ing and deepening the negative peaks can result in a bias in the
neighbour-averaged ellipticity towards large or small values.
The level of this effect will clearly correlate with the magni-
tude of the shear, and so induce a multiplicative bias. To illustrate
this point the above exercise is repeated with a range of different
input shears. The results for our fiducial setup are shown in in Fig.
5. Each point on these axes corresponds to a ring of neighbour po-
sitions for a given input shear. The equivalent measurements with-
out the neighbour are indistinguishable from the x axis. At small
shears, the neighbour induced bias ˜g g
tr
is well aproximated
as a linear in g
tr
. We leave exploration of the possible nonlinear
response at large ellipticities for future investigations. Though the
above numerical exercise demonstrates that it is possible for signif-
icant multiplicative bias to arise as a result of neighbours, it does
not make a clear prediction of the magnitude or even the sign. In-
deed, our toy model is effectively marginalised over θ, but there is
nothing to guarantee that fixing the other neighbour parameters to
the median measured values is representative of the real level of
neighbour bias in a survey like DES. Motivated by this observa-
tion we add a final layer of complexity to the model, as follows.
A single neighbour-central realisation is created as before, defined
by a unique set of model parameters. Now, however, the values
of those parameters p = (d
gn
, f
n
, r
n
, f
c
, r
c
) are drawn randomly
from the DES data. As these quantities will, in reality, be corre-
lated we sample from the 5-dimensional joint distribution rather
than each 1D histogram individually. We then fit the model at 70
neighbour angles and two input shears g
±
= ±0.05 (a total of 140
measurements), and estimate the multiplicative bias as a two-point
finite-difference derivative:
m + 1 =
h˜g(θ|g
+
)i
θ
h˜g(θ|g
)i
θ
g
+
g
. (2)
This process is repeated to create 1.33M unique toy model reali-
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 5
Figure 1. Postage stamp snapshots of the basic two-object toy model described in Section 3. The overlain ellipse shows the maximum likelihood fit to the
image. The panels show three neighbour positions in the range θ = [0, π/2] rad. The best fit ellipticity and half light radius are shown above each image. In
all cases the input values are e = (0, 0), r = 0.5 arcseconds.
0 5 10 15 20 25 30
Neighbour Distance d
gn
/ pixels
0.4
0.3
0.2
0.1
0.0
Tangential or Cross Shear
˜
g
1
(round neighbour)
˜
g
2
(round neighbour)
˜
g
1
(g
tr
2,n
=0.1)
˜
g
2
(g
tr
2,n
=0.1)
Figure 2. Tangential shear measured using the numerical toy model de-
scribed in Section 3.1 as a function of radial neighbour distance. The solid
purple line shows the shape component aligned with the central-neighbour
separation vector and the dot-dashed line is measured along axes rotated
through 45
. Note that the latter is smaller than 10
6
at all points on this
scale. The dashed and dotted black lines show the same ellipticity compo-
nents when the neighbour is sheared in the e
2
direction by g
2
= 0.1.
sations. Binning by neighbour distance we can then make a rough
prediction for the level of neighbour-induced bias and the angular
scales over which it should act. The result is shown in Fig. 6, where
full results using all model realisations are indicated by the dashed
blue line. The majority of cases yield a negative bias, particularly
at low neighbour separation (referring back to Fig. 4, the broad-
ening of the peak around position A dominates over the increased
flux overlap at C). In the real data, of course, we apply a quality
based selection and
¨
uberseg object masking (J16), both of which
are neglected here. We can, however, test the impact of selecting
on fitted quantities that respond to neighbour bias. Imposing a flat
prior on the centroid offset r
0
= (x
2
0
+ y
2
0
)
1
2
(i.e. discarding
randomly generated model realisations where the galaxy centroid
is displaced from the stamp centre by more than a fixed number of
pixels) changes the shape of this curve significantly, as illustrated
by the thick purple line.
We can understand the difference between the results with and
without the centroid cut as a form of selection bias, whereby the cut
preferentially removes toy model realisations in which the neigh-
bour is bright relative to the central galaxy. At any given d
gn
we are
left with a relative overrepresentation of galaxies with f
n
/f
c
1.
Figure 3. Best-fit galaxy ellipticity as a function of neighbour position angle
at fixed neighbour distance d
gn
from the toy model described in the text.
The two panels (left, right) show the same central-neighbour system (g
tr
=
0.05), but with different d
gn
(7 and 8 pixels) and biases m (shown atop
each panel). The solid line in each case is the recovered galaxy shape at each
θ, and the integrated mean along this range is shown by the horizontal dot-
dashed line. The dotted lines show the zero-shear shape (ie. the ellipticity
that would be measured if the input shear were zero), but shifted downwards
such that the mean is at 0.05. Finally, to illustrate the (a)symmetry of the
system we show the solid line flipped about y = g
tr
1
and shifted by π/2
radians as a dashed curve.
Faint neighbours, which in reality tend to be compact high redshift
objects, have little impact when they sit on the outskirts of the cen-
tral profile (A in the cartoon picture in Fig. 4; the regime which
produces negative m). The same faint galaxy has a stronger im-
pact if it is rotated to a position closer to the centre of the central’s
flux profile. Thus one might expect a selection on r
0
to make
the mean m in a particular bin less negative (or even positive) by
preferentially removing brighter galaxies.
4 HOOPOE IMAGE SIMULATIONS
In this section we provide a brief overview of the simulation
pipeline. The process is the same as that described in §4 of Z17,
and we refer the reader to that work for more detail. The end point
of the pipeline is a cloned set of survey images with many of the
observable characteristics of a chosen set of parent images, but for
which we know the input noise properties and galaxy population
pefectly. The simulated images inherit the pixel masking, PSF vari-
MNRAS 000, 000–000 (0000)
6 S. Samuroff, S. L. Bridle, J. Zuntz et al
A
B
C
Figure 4. Cartoon diagram of a neighbour-central system. The purple con-
tours show the lines of constant flux in a S
´
ersic disc profile with extreme
negative ellipticity (g
1
= 0.3). The blue crosses labelled A, B and C are
points on a ring of equal distance from the centre of the profile. The blue
arrows show the local unit vector along a tangent to the ring.
Figure 5. Measured shear minus input shear plotted as a function of input
shear. The purple points show the recovered ˜g
1
from averaging over ring
of 70 neighbour positions. The dark blue lines show the linear relation ˜g
g
tr
= mg
tr
at m = (0.4, 0.45, 0.5). The dotted line shows what
would be measured using the same central profile in the absence of the
neighbour, and is near indistinguishable from the x axis line on all points
within this range of g
tr
.
ation and noise maps measured from the progenitor data. Each sim-
ulated galaxy is then inserted into a subset of overlapping exposures
and into the coadd at the position of a real detection in the DES Y1
data. Object detection is rerun on the new coadd images and galaxy
cutouts and new segmentation masks are extracted and stored in the
MEDS format decribed by J16. The mock survey footprint is shown
in Fig. 7. In the lower panels we show an example of a simulated
coadd (left) and the spatial variation in PSF orientation within the
same image (right).
4.1 Parent Data
We use reduced images from Year One of the Dark Energy Survey
(DES Y1; Diehl et al. 2014) as input to the simulations discussed
in this paper. The Dark Energy Survey is undertaking a five year
0 5 10 15 20 25 30 35 40 45
Neighbour Distance d
gn
/ pixels
1.0
0.8
0.6
0.4
0.2
0.0
Multiplicative Bias m
No r
0
Cut
r
0
< 4 pixels
Figure 6. Multiplicative bias estimated using the Monte Carlo toy model
described in the text. For each neighbour realisation, defined by a particular
distance, flux and size we compute the average of the measured ellipticity
components over 70 rotations on a ring of neighbour angles. To estimate the
bias we perform this averaging twice at two non-zero shears, g
+
and g
,
and compute the finite-difference deriviative using equation 2. The dashed
thin blue line shows the result of using all measurements, while the bold
purple line has a cut based on the offset between the centroid position of the
best-fitting model and the stamp centre.
programme with the ultimate aim of observing 5000 square de-
grees of the southern sky to 24th magnitude in five optical bands,
grizY, covering 0.401.06 microns. The dataset is recorded using a
570 megapixel camera called DECam (Flaugher et al. 2015), which
has a pixel size of 0.26 arcseconds. In full it will consist of 10
interwoven sets of exposures in the g, r, i, z and Y bands.
The Y1 data were collected between August 2013 and Febru-
ary 2014, and cover a substantially larger footprint than the prelim-
inary Science Verification (SV) stage at 1500 square degrees, albeit
to a reduced depth. Details of the reduction and processing are pre-
sented in Z17. Our HOOPOE simulations use a selection of the total
3000 0.75 × 0.75 degree coadded patches known as “tiles”.
4.2 Input Galaxy Selection
For populating the mock survey images a sample of real galaxy pro-
files from the HST COSMOS field, imaged at significantly lower
noise and higher resolution than DES by the Hubble Space Tele-
scope Advanced Camera for Surveys (HST ACS) (Scoville et al.
2007). The COSMOS catalogue extends significantly deeper than
the Y1 detection limit of M
r,lim
= 24.1, extending to roughly 27.9
mag in the SDSS r-band. A main sample for our DES Y1 simula-
tions is defined by imposing a cut at < 24.1 mag.
Since the DES images do not cut off abruptly at 24th magni-
tude, in reality they contain a tail of fainter galaxies that contribute
flux are not identifiable above the pixel noise. To assess the impact
of these objects on shape measurements in Y1, we simulate a pop-
ulation of sub-detection galaxies in addition to the main sample.
In brief we use the full histogram of COSMOS magnitudes to esti-
mate the number of faint galaxies within a given tile. The required
profiles are selected randomly from the faint end of the COSMOS
distribution. Each undetected galaxy is paired with a detection, and
inserted at a random location within the overlapping bounds of the
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 7
0
10
20
30
40
50
60
70
350
50
40
30
2.4 3.2 4.0 4.8 5.6 6.4 7.2 8.0
n
g
/arcmin
2
(a) (b)
Figure 7. Top: The projected footprint of the simulated survey, visualised using the SKYMAPPER package
a
. The colour indicates the local raw number density
in HEALPIX cells of nside = 1024. The axes shown are right ascension and declination in units of degrees. The full simulation comprises 1824 0.73 × 0.73
degree tiles drawn randomly from the DES Y1 area. The solid blue line indicates the bounds of the planned area to be covered by the complete Y5 dataset.
Bottom: A random tile (DES0246-4123) selected from the HOOPOE area. The left panel (a) shows a square subregion of approximately 4 × 4 arcminutes.
The right hand panel (b) shows a PSF whisker plot covering the full 0.73 × 0.73 tile. The length and orientation of each line represents the magnitude and
position angle of the spin-2 PSF ellipticity at that position. Only galaxies which pass IM3SHAPE quality cuts are shown. The white patches show the spatial
masking inherited from the GOLD catalogue, and correspond to the positions of bright stars in the parent data.
a
https://github.com/pmelchior/skymapper
same (subset of) single-exposure images. A more detailed descrip-
tion of this process can be found in Z17.
If these galaxies were present in the data they would enter
the background flux calculation, and thus the subtraction applied
would change due to their presence. Since the simulation pipeline
produces images effectively in a post-background subtraction state
this is not captured. To test this we rerun the SEXTRACTOR back-
ground calculation on a handful of tiles drawn with and without
the faint galaxies. The impact was found to be well approximated
as a uniform shift in the background correction. A flux correction
equal to the pixel-averaged flux of the sub-detection galaxies over
each image plane is, then, applied to postage stamps prior to shape
measurement.
In reality the overdensity of sub-threshold galaxies will be
coupled to the density of detectable objects, which is clearly not
the case in our simulations. To gauge the impact of this we per-
form the following test. Each tile is divided into a 6 × 6 grid, and
the mean multiplicative bias is calculated in each sub-patch. We
bin sub-patches according to the ratio f
faint
N
faint
/N
det
, or the
total number of faint galaxies relative to the number of detectable
ones. The impact is significant, but not leading order; excluding
patches outside the range 0.9 < f
faint
< 1.1 induces a shift of
m 0.005.
An independent noise realisation is generated for each expo-
sure using the weight map from the parent data. We simulate the
noise in each pixel by drawing from a Gaussian of corresponding
width. The coaddition process is not rerun, but rather we compute
an independent noise field by drawing the flux in each pixel from a
MNRAS 000, 000–000 (0000)
8 S. Samuroff, S. L. Bridle, J. Zuntz et al
Figure 8. An example of an object in the main DES Y1 calibration simula-
tion and the neighbour-free resimulation. The upper panels show the coadd
cutout in the original simulated images (left, labelled HOOPOE) and in the
neighbour-subtracted version (right, labelled WAXWING). The lower panels
are the segmentation masks for the same galaxy. A number of neighbours,
both masked (upper left and centre left) and unmasked (lower right) are
visible within the stamp bounds.
zero-centred Gaussian of width determined by the measured vari-
ance in that pixel.
4.3 Neighbour-Free Resimulations
For the purpose of untangling the impact of image plane neighbours
we use the simulated HOOPOE images to create a new spin-off
dataset. In a subset of a little over 500 tiles we store the (convolved)
input profile for each object and the noise-only cutout, taken from
the same position in the image plane prior to objects being drawn.
By adding together these two components we can generate a suite
of spin-off MEDS files, which are equivalent to the results of a
simpler neighbour-free simulation (eg Miller et al. 2013, J16). The
pixel noise realisation, COSMOS selection and input shears, how-
ever, are identical to the progenitor HOOPOE simulations.
We will call this process “resimulating”, and the basic concept
is illustrated in Fig. 8. The 506-tile set of neighbour-free data are
named the WAXWING resimulations. Finally the (now empty) seg-
mentation masks corresponding to the subtracted neighbours are
also removed. In subsequent IM3SHAPE runs on these data we ig-
nore the SEXTRACTOR flags obtained from the main simulations.
5 QUANTIFYING NEIGHBOUR BIAS WITH HOOPOE
Equipped with qualitative predictions from Section 3, we now turn
to the question of neighbour bias in the more complete simulations
described in Section 4. The mock survey was designed to capture
as much of the complexity of shape measurements on real photo-
metric data as possible. We refer to Section 4 of this paper for a
short overview and to §5 of Z17 for a more detailed discussion of
the simulation pipeline and validation tests. The simulated galaxy
catalogue used in the following is identical to the one used to cali-
brate the DES Y1 IM3SHAPE catalogue, including quality cuts and
selection masks.
5.1 Single-Galaxy Effects
The most straightforward way to assess the impact of neighbours
on individual shape measurements in our simulations is to rotate
the measured shapes into a frame defined by the central-neighbour
separation vector. Whereas in the earlier toy model we had only one
neighbour per galaxy, we now have a crowded image plane contain-
ing many objects simultaneously. For simplicity, in the earlier case
we included no masking. For HOOPOE we wish to mimic the pro-
cess of shape measurement on real data as closely as possible. We
generate new segmentation maps by running SEXTRACTOR on the
simulated images, and incorporate them into our shape measure-
ments using the
¨
uberseg algorithm (J16). Each simulated galaxy is
allocated a nearest neighbour using a k-d tree matching algorithm
constructed on the coadd pixel grid using every galaxy simulated at
r-band magnitude M
r
< 24.1. The quantities d
gn
and θ are now
redefined slightly as nearest-neighbour distance and angle. We de-
fine the tangential shear of a galaxy relative to its nearest neighbour
using the standard convention,
e
+
= [e
1
cos(θ) + e
2
sin(θ)] , (3)
and the cross shear
e
×
= [e
2
cos(θ) e
1
sin(θ)] . (4)
Note that negative values of e
+
imply a net tangential alignment
of the measured shapes towards neighbours. By analogy, we de-
fine e
1,n
and e
2,n
, which are the measured ellipticity components,
rotated into a reference frame defined by the major axis of the
neighbour. Non-zero e
i,n
would indicate leakage of the neighbour’s
shape into the measurement, which might conceivably be induced
by inadequate deblending of very close neighbours or by extensive
non-circular masking. We first divide the main simulated catalogue
into bins according to d
gn
, and measure the tangential shear about
nearest neighbours in each bin. The result is shown by the pur-
ple curve in Fig. 9. Note that the statistical uncertainty is within
the width of the line in all bins. The results here show qualitative
agreement with the numerical predictions in Fig. 3. As we found
earlier, the exact shape of this curve is sensitive to the properties
of both the neighbour and the central galaxy. Despite small dif-
ferences, the range of variation is comfortably within the scale of
the postage stamp for the bulk of galaxies in DES Y1. Repeating
the measurement, rotated into the plane of the neighbour shape re-
sults in the dotted and dot-dash lines in this figure. As noted above,
there are not necessarily reliable ellipticity measurements for each
neighbour, so we instead use the sheared input ellipticities. Both
components of e
i,n
are seen to be negligible over all scales.
5.2 Neighbour Ensemble Biases
To explore the more practical question of how neighbours impact
shear estimates we divide the catalogue into bins according to
neighbour distance. Within each d
gn
bin, the galaxies are further
split into twelve bins of input shear, which are fitted to estimate the
multiplicative and additive bias. We show the result as the purple
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 9
0 5 10 15 20 25 30
Nearest Neighbour Distance d
gn
/ pixels
0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.05
Tangential or Cross Shear
Tangential Shear e
+
Cross Shear e
×
Tangential Shear e
+,n
Cross Shear e
×,n
0.004
0.002
0.000
0.002
Figure 9. Tangential shear around image plane neighbours in the full
HOOPOE simulation. The purple solid line shows the mean component of
the measured galaxy shapes radial to the nearest image plane neighbour.
Dashed blue shows the component rotated by 45
, which we have no rea-
son to expect should be non-zero. The dotted and dot-dash lines show the
measured ellipticity components when rotated into a coordinate frame de-
fined by the major axis of the neighbour. The inset shows the same range
in d
gn
(the x-axis tick markers are the same), but with a magnified vertical
axis.
points in Fig. 10, which can be compared with the earlier numerical
model prediction in Fig. 6. The horizontal band on these axes shows
the 1σ mean m measured using all galaxies in the HOOPOE cat-
alogue, and sits at m 0.12. We note a steeper decline than in
the bold line (without the centroid cut), more akin to the case with
the centroid cut (r
0
< 1 arcsec). This is not surprising given
that the quality selection implemented by IM3SHAPE includes ex-
actly this cut. We do not report a local peak at 11 pixels, which
we saw before in Fig. 6. We suggested previously that effect was
the result of positive m in galaxies where the nearest neighbour is
relatively faint and at middling distance. It is likely that many of
these objects manifest themselves as large changes in other quan-
tities to which IM3SHAPEs INFO FLAG (see Z17) is sensitive such
as ellipticity magnitude and fit likelihood, or are flagged by the
SEXTRACTOR deblending cuts.
When divided into broad bins according to the r-band magni-
tude of the nearest neighbour M
r,neigh
(the coloured stripes in Fig.
10) we find the surviving objects show relatively weak dependence
on neighbour brightness, except at the neighbour distances, where
bright neighbours have a slightly stronger (negative) impact than
faint ones.
We measure the additive bias components in the same bins,
but find no systematic variation with d
gn
above noise.
Finally we show the analogous measurement in bins of galaxy
magnitude in Fig. 11. The steep inflation of |m| at the faint end
of this plot has been seen elsewhere (e.g. Zuntz et al. 2017; Fenech
Conti et al. 2017), and is easily understandable as the result of noise
bias. We find that splitting by neighbour magnitude does not reveal
any obvious trend here.
5.3 Untangling the Knot of Neighbour Bias
A central plank of this analysis rests on a comparison of the main
HOOPOE simulations with the neighbour-free WAXWING resimu-
0 5 10 15 20 25 30
Neighbour Distance d
gn
/ pixels
0.8
0.6
0.4
0.2
0.0
Multiplicative Bias m
M
r,neigh
= 21.5
M
r,neigh
= 23.0
M
r,neigh
= 23.6
M
r,neigh
= 24.0
all galaxies
Figure 10. Multiplicative bias as a function of separation from the nearest
image plane neighbour. The purple points show the bias calculated in bins
of neighbour distance using the main HOOPOE simulated shape catalogue.
The coloured bands show the same dataset divided into four equal-number
bins according to the r-band magnitude of the neighbour. As shown in the
legend, the median values in the four bins are 21.5, 23.0, 23.5 and 24.0.
The mean bias and its uncertainty across all distance bins is indicated by
the horizontal band.
20.5 21.0 21.5 22.0 22.5 23.0 23.5 24.0
r-band Magnitude
0.8
0.6
0.4
0.2
0.0
Multiplicative Bias m
M
r,neigh
= 21.5
M
r,neigh
= 23.0
M
r,neigh
= 23.6
M
r,neigh
= 24.0
all galaxies
Figure 11. Multiplicative bias as a function of r-band magnitude. As in Fig.
10 the four coloured bands represent equal number bins of neighbour mag-
nitude. Purple points show the full catalogue, with no magnitude binning.
The mean bias and its uncertainty are shown by the purple horizontal band.
lations described in Section 4.3. The simplest comparison would
be between multiplicative bias values, calculated using all galax-
ies in each catalogue after cuts. These values are shown by the two
upper-most lines (purple) in Fig. 12. The difference is an indicator
of the net impact of neighbours through any mechanism, which we
find to be m 0.05.
To untangle the various contributions to this shift, we con-
struct a matched catalogue. Galaxies in the overlap between
HOOPOE and WAXWING (12M galaxies over 183 square degrees)
are matched by ID; quality cuts are calculated for each set of mea-
surments (see Appendix E from Z17). Geometric masking from the
DES Y1 GOLD catalogue (Drlica-Wagner et al. 2017) and SEX-
MNRAS 000, 000–000 (0000)
10 S. Samuroff, S. L. Bridle, J. Zuntz et al
Figure 12. Graphical illustration of the measured multiplicative bias in the various scenarios considered in this paper. The upper two lines show the mean m in
the main DES Y1 HOOPOE simulations and a spin-off neighbour-free resimulation named WAXWING, as described in Section 4.3. The middle section (green)
shows results using only galaxies which appear in both the HOOPOE and WAXWING simulations. The matching process alone does not imply any quality-
based selection function. The final three lines in red are from a similar matching between a smaller rerun of the simulation with and without sub-detection
limit galaxies. See the text for details about each of these cases.
TRACTOR deblending flags are included for HOOPOE. Since the
latter are irrelevant to WAXWING, we omit them from quality flags
on that dataset. For conciseness we will refer to the two measure-
ments as “matched HOOPOE and “matched WAXWING”, and their
cuts as “HOOPOE cuts” and “WAXWING cuts”. Since the images
are identical in all respects, but for the presence of neighbours, the
statistical noise on the change in measured quantities should be
smaller than the face-value uncertainties.
The appropriate cuts are first applied to each catalogue, then
the results are divided into equal number signal-to-noise bins and
fitted for the multiplicative bias in each. The result is shown by
the points in the upper left-hand part of Fig. 13. The equivalent in
bins of PSF-normalised size is shown on the right. The difference
between the blue and the purple points gives an indication of the
total effect of all neighbour-induced effects on m, indicated by the
solid purple line in the lower panel. The generic shift attributed to
“neighbour bias” is in reality a collection of distinct effects. By
comparing the matched catalogues we identify four main mecha-
nisms: direct contamination, selection bias, S/N bin shifting and
neighbour dilution. Each of these components that we describe is
shown by one of the lines in Fig. 13. For a visual summary of the
various tests designed to isolate them see Fig. 12.
5.3.1 Direct Flux Contamination
The most intuitive form of neighbour bias arises from the fact
that, even after masking, neighbours contribute some flux to the
cutout image of a galaxy. To gauge its impact we take the com-
mon sample of galaxies, which pass cuts in both datasets. The com-
parison is complicated somewhat by binning in measured S/N or
R
gp
/Rp; for this test, we divide both sets of galaxies using the
WAXWING-derived quantities. The resulting m measured using the
HOOPOE galaxies is unrealistic in the sense that we are binning
measurements made in the presence of neighbours by quantities de-
rived from neighbour-subtracted images. This exercise does, how-
ever, isolate the impact of the neighbour flux on the measured el-
lipticity. The result is shown by the purple dotted and purple dot-
dashed lines in the upper and lower panels of Fig. 13. The effect
scales significantly with signal-to-noise and size. Faint small galax-
ies are affected strongly by neighbour light, while larger brighter
ones are relatively immune.
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 11
0.25
0.20
0.15
0.10
0.05
0.00
Multiplicative Bias m
HOOPOE, matched per galaxy
HOOPOE, matched per galaxy (diluted)
HOOPOE, matched per bin
WAXWING, matched per galaxy
HOOPOE, full selection
WAXWING, full selection
1.0 1.2 1.4 1.6 1.8 2.0 2.2
Signal-to-Noise log(S/N)
0.08
0.06
0.04
0.02
0.00
0.02
m
Direct Neighbour Bias
Selection Bias
Neighbour Dilution
Bin Shifting
Total Neighbour Bias
0.15
0.10
0.05
0.00
Multiplicative Bias m
1.2 1.3 1.4 1.5 1.6 1.7
R
gp
/R
p
0.08
0.06
0.04
0.02
0.00
0.02
0.04
0.06
m
Figure 13. Top half of each panel Multiplicative bias as a function of signal-to-noise and size. The purple circles show the measured bias using the main
HOOPOE simulation, and the blue diamonds are the resimulated neighbour-free version. The lines show permutations of the same measurements to highlight
the neighbour-induced effects causing the two to differ. The dot-dashed blue and dashed purple lines show the impact of applying the HOOPOE selection mask
to WAXWING and vice versa. The impact of bin shifting is shown by the purple dotted line, which is calculated from the same matched galaxies, using the
HOOPOE shape measurements for the bias and WAXWING size and S/N for binning. The pink curve is the same as the dashed purple, but with a fraction of
heavily blended galaxies added back with randomised shear (see Section 5.3.4). Bottom half of each panel The change in bias due to the effects described
above. The green (dashed) line shows the impact of selection effects (the difference between the blue diamonds and the dashed line in the top panel). The direct
neighbour bias due to light contamination is shown by the purple dash-dotted line (purple dotted minus blue dash-dot top). The impact of shifting between
bins is shown by the blue dotted (dashed minus purple dotted, top). The pink dot-dot-dashed line illustrates the impact of adding back randomised shears, as
described. Finally the solid line represents the total neighbour bias, which includes all these effects (circles minus diamonds, top).
5.3.2 Neighbour-Induced Selection Bias
To gauge the neighbour-induced selection effect, we take the
WAXWING catalogue but now impose, in addition to its own qual-
ity cuts, the selection function derived from the with-neighbour
HOOPOE dataset. The double masking removes an additional 0.5M
galaxies, which survive cuts in WAXWING but would be cut from
the HOOPOE catalogue. The resulting change in m is shown by the
dot-dash blue lines in the upper panels of Fig. 13 (dashed green in
the lower). The multiplicative bias arising from this cut is less than
one percent in all but the faintest and smallest galaxies, where it
can reach up to m 0.02.
5.3.3 Bin Shifting
The above two tests encapsulate the impact on the measured ellip-
ticities, and the selection flags from neighbour flux. An additional
subtlety arises from the fact that the measured quantities used to
bin galaxies (S/N and R
gp
/R
p
) are themselves affected by the
presence of neighbours. To test this we recalculate m using the
same galaxy selection as in Section 5.3.1 (i.e. passing both sets of
cuts), but now binned by the appropriate measured S/N. For clar-
ity, the bin edges are unchanged, defined to contain equal numbers
of WAXWING galaxies. The result is shown by the dashed lines in
Fig. 13. The difference compared with the case using fixed binning
is purely the result of galaxies moving between bins. This shift-
ing contributes multiplicative bias if one bins galaxies by observed
quantities such as S/N, as we do in order to calibrate IM3SHAPEs
shear estimates. The amplitude of this is illustrated by the blue dot-
ted line in the lower panels. Such neighbour-induced shifting is
noticable if we plot out the S/N of objects in HOOPOE against the
S/N of the same objects in WAXWING. Objects which are strongly
shifted in S/N are more likely to scatter upwards than downwards.
A similar skew can be seen in the R
gp
/R
p
plane; when galaxies are
scattered in size they tend to be thrown further and more often up-
wards than downwards. Small galaxies (which we know already are
more strongly affected by noise bias) are shifted strongly upwards
by the presence of neighbour flux in the HOOPOE images. The re-
sult is a net negative m added to the upper R
gp
/R
p
bins, and a
simultaneous positive shift in the lowest bins from which galaxies
are lost. In the case of galaxy size we see the effects of bin scatter
and direct neighbour bias almost negate each other, although the
degree of cancellation is likely dependent on the specifics of the
measurement code and the dataset.
5.3.4 Neighbour Dilution
A final point can be gleaned from Fig. 13: that applying the
WAXWING cuts to HOOPOE induces a shift in m. Naively one
might expect the HOOPOE selection function, which includes
neighbours, to remove the same galaxies as the WAXWING selec-
tion, plus some extra strongly blended galaxies. It is true that a
sizeable number of galaxies are cut in the presence of neighbours,
but would otherwise not be. There is also, however, a smaller popu-
lation that survive cuts because they have image plane neighbours.
We can see this clearly from the fact that the purple points and
the dashed purple lines Fig. 13 are non-identical. We identify three
separate (but partially overlapping) galaxy selections in this figure:
(a) galaxies passing both sets of cuts, (b) galaxies passing cuts in
the absence of neighbours, but cut by the HOOPOE selection and (c)
galaxies which pass cuts in the presence of neighbours, but cut by
the WAXWING selection. We find that populations (b) and (c) have
much smaller mean neighbour separation than the full population
(the histograms of d
gn
show a sharp peak at under 10 pixels). In
MNRAS 000, 000–000 (0000)
12 S. Samuroff, S. L. Bridle, J. Zuntz et al
contrast, both the full catalogue and population (a) objects a much
broader distribution (
¯
d
gn
24 pixels).
Based on the toy model predictions in Section 3 we set out a
working proposal: that population (c), objects cut out only when
neighbours are removed, are extreme blends dominated by a super-
posed neighbour. We will assume these objects are boosted con-
siderably in size, S/N or both, such that what would otherwise
be a small faint galaxy is now sufficiently bright to pass quality
cuts. In these cases the measured shape of a simulated galaxy might
be expected to be only weakly linked with the input ellipticity. To
approximate this effect we take population (a) HOOPOE galaxies,
subject to both sets of cuts, and add back some of the population
(c) galaxies. Specifically, we include any objects shifted in S/N or
R
gp
/R
p
by more than 20%. The true shears associated with these
galaxies are now randomised to eliminate any correlation with the
measured ellipticity. The result is shown as a pink dot-dot-dashed
line in Fig. 13. We can see that this effect, which we call neigh-
bour dilution, to good approximation accounts for the residual dif-
ference between the population (a) and (c) samples. Particularly
in the upper size bins of the right hand panel the differences are
not eliminated entirely. This is thought to be the result of resid-
ual (albeit weakened) covariance between the measured shapes of
strongly blended objects and the input shears. Clearly the scenario
in which a neighbour totally overrides the original shape of a galaxy
is extreme, and there will be an indeterminate number of moderate
blends which are boosted sufficiently to survive cuts but which re-
tain some correlation with their original unblended shapes. Such
cases are, however, extremely difficult to model accurately with the
resources available for this investigation.
5.4 Isolating the Impact of Sub-detection Galaxies
A handful of previous studies have attempted to quantify the impact
of galaxies below, or close to, a survey’s limiting magnitude. For
example, Hoekstra et al. (2015) and Hoekstra et al. (2017) suggest
they can induce a non-trivial multiplicative bias, which is depen-
dent on the exact detection limit. They recommend using a shear
calibration sample at least by 1.5 magnitudes deeper than the sur-
vey in question (which ours does). Their findings, however, make
exclusive use of the moments-based KSB algorithm (see Kaiser
et al. 1995); such techniques are known to probe a galaxy’s elliptic-
ity at larger radii than other methods, which could in principle make
them more sensitive to nearby faint galaxies. It is thus a worthwhile
exercise to to gauge their impact in our case with IM3SHAPE.
5.4.1 Impact on Multiplicative Bias
We first compare our HOOPOE simulations with the neighbour-free
WAXWING resimulations. Since WAXWING postage stamps con-
sist of only a single profile added to Gaussian pixel noise, they are
unaffected by neighbours of any sort (faint or otherwise). We have
seen that the impact of neighbours is strongly localised, with the
excess m converging within a nearest neighbour distance d
gn
of a
dozen pixels or so. Thus selecting galaxies that are well separated
from their nearest visible neighbour will isolate the impact of the
undetected ones.
A further cut is thus imposed on d
gn
< 20 pixels. Relative to
the case with quality cuts only, the global multiplicative bias now
shifts from m 0.119 to m = 0.064 ± 0.006 (the first and
second lines in green on Fig. 12). This measurement is in mild ten-
sion with the value measured from WAXWING (again under its own
0 10 20 30 40 50 60 70 80
Nearest Subdetection Galaxy d
gf
/ pixels
With Faint, Own Cuts
Only Without Faint
Only With Faint
Figure 14. Histogram of radial distances between galaxies in our measured
shape catalogues (the full HOOPOE simulations) and the nearest object be-
low the DES detection limit. The dotted line includes all objects prior to
quality cuts, while the solid line shows the impact of applying IM3SHAPE’s
INFO FLAG cuts (see J16). The dashed blue line shows the population of
galaxies which survive cuts only in the presence of the faint galaxies.
cuts, with the selection on d
gn
). This difference, which amounts to
a negative shift in m of 0.01 is, we suggest, the net effect of the
sub-detection galaxies. From these numbers alone we cannot tell if
this is a result of selection effects, flux contamination, bin shifting
or some combination thereof.
Interestingly we find that imposing both the HOOPOE and
WAXWING selection functions in addition to the cut on d
gn
brings
m into agreement to well within the level of statistical precison
(compare the final and penultimate lines in green in Fig. 12). That
is, when restricted to a subset of galaxies that pass quality cuts in
both simulations the flux contributed by the faint objects has little
impact. Their main impact is rather that they allow a population
of marginal faint galaxies which would otherwise be flagged and
removed by quality cuts to pass into the final HOOPOE catalogue.
To test this idea further we rerun a subset of 100 random tiles
from the simulated footprint, without the final step of adding sub-
detection galaxies. To minimise the statistical noise in this com-
parison we enforce the same COSMOS profiles, shears and ro-
tations as well as the per-pixel noise realisation as before. SEX-
TRACTOR source detection is applied and the blending flags are
propagated into the postprocessing cuts.
The raw m values calculated from the rerun and the main
HOOPOE simulations, matched to the same tiles, are shown by the
upper two red lines on Fig. 12. The downward shift of 0.01 is
consistent with the previous result based on the main simulation.
This comparison should encapsulate the full effect of the faint ob-
jects (since there are no other differences between these datasets).
For each galaxy we next measure the distance to the nearest
faint object d
gf
, the distribution of which is shown under vari-
ous selections in Fig. 14. Like in the comparison in Section 5.3.4,
there is a population of galaxies that survive cuts only in the sim-
ulation with the sub-detection objects, and these galaxies tend to
be ones with extremely close faint neighbours. Interestingly the
inverse population surviving only when they are removed do not
preferentially have small d
gf
. This is intuitively understandable: a
faint object might boost its neighbour’s apparent size or S/N if it
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 13
were centred within a few pixels. Otherwise it would act as a source
of background noise, which would reduce the quality of the fit.
Finally we find that if we apply both selection functions to
the with-faint galaxies, the measured biases become roughly con-
sistent. These findings, combined with the observations in the pre-
vious section lead us to an interesting conclusion: the major effect
of the faint galaxies in the DES Y1 IM3SHAPE catalogue is to al-
low a population of small faint galaxies to pass quality cuts, where
otherwise they would have been removed. This is analogous to the
neighbour dilution effect described above, but is subdominant to
the influence of visible neighbours.
5.4.2 Impact on Background Flux Subtraction
As a test of the robustness of this result we recompute our
IM3SHAPE fits on the faint-free images, with and without the cor-
rection for the shift in the background flux that would have been
applied if the sub-detection galaxies had been drawn. The mean
per-tile correction is f 0.05, against typical noise fluctua-
tions σ
n
6.5. Matching galaxies and examining the histograms
of S/R and R
gp
/R
p
reveals weak downwards scatter in both
quantities (i.e. the flux subtraction alone makes galaxies appear
smaller and fainter). The magnitude of the shift is, however, tiny,
peaking at 0.1 and 0.005 respectively. This is logical given
the definition in equation 1. If the change is small enough such
that the best-fitting model is stable, then an incremental reduction
in flux will reduce the signal-to-noise of the measurement. Looking
at the best-fit shapes, we find a small shift towards high ellipticities,
which can likewise be understood as a numerical effect; imposing
a flat positive field of zero ellipticity will dilute the measured shear,
producing a bias towards round |e|. The reverse logic applies with
the flux correction, and subtracting a flat value from all pixels will
make galaxies appear slightly more elliptical. In practice we find a
sharp peak at e 0.001.
5.5 Suppressing Neighbour Bias
There is no universal definition for the shape-weighted effective
number density commonly used as proxy for cosmological con-
straining power in a shear catalogue. One which is particularly use-
ful in the context of weak lensing, and which has been adpoted in
DES Y1 is the prescription of Chang et al. (2013), which is de-
signed to account for shape noise and fitting error (see equation
7.5 in Zuntz et al. 2017). A second useful definition is set out by
Heymans et al. (2012) in terms of the (see also Z17). We com-
pute a neighbour distance d
gn
for every object in the real data,
which allows us to cut on this quantity. Removing any galaxy with
a neighbour detected within a radius of d
gn
= 20 pixels reduces
the effective number density of sources using either definition to
about 70% of its initial value, from n
H13
eff
= 5.48 to n
H13
eff
= 3.68
arcmin
2
using Heymans et al. (2012)’s definition. Using the pre-
scription of Chang et al. (2013), the equivalent density drops from
n
C13
eff
= 3.18 prior to cuts and n
C13
eff
= 2.18 arcmin
2
afterwards.
This cut is stringent, as we have shown that beyond 12 pixels
the multiplicative bias becomes insensitive to further selection on
d
gn
. There are, however, a number of limitations in our analysis,
including the fact that d
gn
is defined using the true input positions,
and indeed that we are using only the detected positions in DES to
draw our simulated M
r
< 24.1 galaxies. We thus judge that a level
of conservatism is appropriate here. Relaxing the cut to d
gn
> 14
pixels leaves n
eff
at 84% of its full value.
6 COSMOLOGICAL IMPLICATIONS
As we have shown in the previous sections, if ignored completely
image plane neighbours can induce negative calibration biases in
IM3SHAPE of a few per cent or more. The earlier part of the inves-
tigation focused on when and how neighbour bias can arise, first
in the context of single-galaxies and then on ensemble shear esti-
mates. We now turn to a more pressing question from the general
cosmologist’s perspective: how far should I be concerned about
these effects in practice? We present a set of numerical forecasts us-
ing the MULTINEST nested sampling algorithm (Feroz et al. 2013)
to sample trial cosmologies. Each of the likelihood analyses pre-
sented in this paper has been repeated using a Markov Chain Monte
Carlo sampler (EMCEE). Although we see the same small shift in
contour size noted by DES Collaboration et al. (2017) (see their
Appendix A), which diminishes as the length of the MCMC chains
increases, we find our conclusions are robust to the choice of sam-
pler. Our basic methodology here follows previous numerical fore-
casts (e.g. Joachimi & Bridle 2010; Krause & Eifler 2016; Krause
et al. 2017a). We construct mock DES Y1 cosmic shear measure-
ments using a matter power spectum derived from the Boltzmann
code CAMB
4
with late-time modifications from HALOFIT. The
cosmic shear likelihood surface is sampled at trial cosmologies us-
ing COSMOSIS
5
. The final data used for the likelihood calcula-
tion have the form of real-space ξ
±
correlations. For the photo-
metric redshift distributions we use the measured estimates in four
tomographic bins, obtained from runs of the BPZ code on the Y1
IM3SHAPE catalogue, as described by Hoyle et al. (2017). Since
this analysis was completed before the details of the photometric
redshift calculation for DES Y1 had been finalised, these distribu-
tions differ marginally from (but are qualitatively the same as) the
final version used in Troxel et al. (2017) and DES Collaboration
et al. (2017). In all chains which follow we maginalise over two
nuisance parameters (an amplitude and a power-law in redshift) for
intrinsic alignments, photo-z bias and shear calibration bias. In to-
tal this gives 10 extra free parameters in addition to six for cosmol-
ogy (
m
,
b
, A
s
, n
s
, h,
ν
h
2
), which are also allowed to vary.
Apart from the difference in redshift distributions remarked upon
above, our analysis choices match the DES Y1 cosmic shear anal-
ysis of Troxel et al. (2017). We refer the reader to that paper for
details of the priors and scale cuts, and their derivation. Finally,
the following adopts shear-shear covariance matrices derived from
the analytic halo model calculations of Krause et al. (2017b). We
assume a fiducial ΛCDM cosmology σ
8
= 0.83,
m
= 0.295,
b
= 0.047, n
s
= 0.97, h = 0.688, τ = 0.08, with non-zero
comoving neutrino density
ν
h
2
= 0.00062.
6.1 Mean Multiplicative Bias
We first seek to quantify the bias that would be present in a
cosmic shear analysis in a survey like DES, if we were to use
a simple postage stamp simulation of the sort presented in J16
and Miller et al. (2013). To this end we use the neighbour-free
WAXWING dataset to construct an alternative shear calibration. In
Z17 we compared three methods for shear calibration using the
HOOPOE simulations and found our results to be robust to the
differences. We now use the fiducial (grid-based) scheme to de-
rive an alternative set of bias corrections from WAXWING. These
4
http://camb.info
5
https://bitbucket.org/joezuntz/cosmosis
MNRAS 000, 000–000 (0000)
14 S. Samuroff, S. L. Bridle, J. Zuntz et al
10
1
10
2
Angular Scale / arcseconds
4
2
0
2
4
ab
() / 10
3
hmmi
¯
m
2
h
g
mi
-0.1
0.0
0.1
(4, 1) (3, 1) (2, 1)
10 100
(1, 1)
waxwing Calibration
Spatial Correlations
-0.1
0.0
0.1
(4, 2) (3, 2)
10 100
(2, 2)
-0.1
0.0
0.1
(1, 1)
-0.1
0.0
0.1
(4, 3)
10 100
(3, 3)
-0.1
0.0
0.1
(2, 2) (1, 2)
1 10 100
θ / arcmin
-0.2
-0.1
0.0
0.1
(4, 4)
-0.1
0.0
0.1
(3, 3) (2, 3) (1, 3)
10 100
θ / arcmin
-0.2
-0.1
0.0
0.1
(4, 4)
10 100
(3, 4)
10 100
(2, 4)
10 100
(1, 4)
ξ
obs
+
(θ)
+
(θ) 1
ξ
obs
(θ)
(θ) 1
Figure 15. Top: The observed two point correlation of multiplicative bias, as measured from the main HOOPOE simulation set presented in this paper. Sub-
patches are used to compute m in spatial patches of dimension 0.15 × 0.15 degrees and the correlation function calculated as described in the text. The
dashed vertical line shows the diagonal scale of the sub-patches, below which we do not attempt to directly measure spatial correlations. The shaded blue
bands show the minimum and maximum scales used in the DES Y1 cosmic shear analysis of Troxel et al. (2017). Bottom: Residuals between the mock two
point shear-shear data used in this paper, before and after different forms of bias have been applied. The upper-left and lower-right triangles show the ξ
+
and ξ
correlations respectively, calculated using the redshift distributions of Hoyle et al. (2017). The dotted black lines, which are flat across all scales but
vary between panels, show the result of calibrating our Y1 shear measurements with a simple postage stamp simulation without image plane neighbours. The
dashed lines illustrate the impact of ignoring scale-dependent selection effects, which are not captured by our simulation-based calibration. The shaded blue
regions of each panel show the excluded scales for each particular tomographic bin pairing.
are then applied to the same galaxies in the matched HOOPOE sim-
ulation, and residual biases are measured in four DES-like tomo-
graphic bins. The process is very similar to the diagnostic tests in
§5 of Z17, and so we defer to that work for details of the redshift
bin assignment of simulated galaxies.
Using the neighbour-free simulation we under-correct the
measurement bias by several percent in each bin. The remeasured
residual bias after calibration provides an estimate for the level of
systematic that would be present were we to calibrate DES Y1 us-
ing the simpler WAXWING simulations. In the four tomographic
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 15
Figure 16. Expected cosmology constraints from DES Y1 cosmic shear
only. The purple (solid) contours show the results of calibrating using a sim-
ulation which fully encapsulates all biases in the data, leaving no residual
m in the final catalogue. In blue (dash-dotted) we show the result of cal-
ibrating with an insufficiently realistic simulation, which leaves a residual
bias between 0.03 and 0.08 in each of the redshift bins. For reference
we mark the input cosmology with a black cross.
bins used in DES Y1 we find (∆m
(1)
, m
(2)
, m
(3)
, m
(4)
) =
(0.037, 0.044, 0.064, 0.073), and apply these biases to our
mock data. The resulting shift in the shear two-point correlations is
shown by the black dotted lines in the lower panel of Fig. 15. Since
the calibration scheme does not explicitly include neighbour dis-
tances, but rather orders galaxies into cells of S/N and R
gp
/R
p
,
this test does not include any scale dependent neighbour effects.
The calibration effectively marginalises out d
gn
, and the residual
biases are an average over the survey. For the moment we will as-
sume this mean shift in m is sufficient, and return to the question
of scale dependence in the following section.
Our predicted cosmology constraints with weak lensing alone
in DES Y1 are shown in Fig. 16. In purple we show the results of
the fiducial analysis, in which the shear calibration fully captures
all neighbour effects and leaves no residual multiplicative bias. The
blue (solid) contours then show the impact of residual neighbour
biases per bin at the level described. As we can see, even when
marginalising over m
i
with an (erroneously) zero-centred Gaussian
prior of width σ
m
= 0.035, our cosmology constraints are shifted
enough to place the input cosmology outside the 1σ confidence
bounds. We reiterate here that this calculation highlights the bias
that would arise were we to naively apply a calibration of the sort
used in DES SV based on neighbour-free simulations to the Y1
data. Since we are confident that the HOOPOE code captures the
effects of image plane neighbours correctly (at least to first order)
this is a hypothetical scenario only and not a prediction of actual
bias in DES Y1.
6.2 Scale Dependence
It is not trivial that including an mean neighbour-induced compo-
nent to m over the entire survey will be sufficient to mitigate all
forms of neighbour bias. The local mean m on a patch of sky is sen-
sitive to spatial fluctuations in source density, which could induce
scale dependent bias on arcminute scales. Clearly, when correlat-
Figure 17. The same as Fig. 16, but now showing the impact of residual
scale dependent selection bias. The two sets of confidence contours repre-
sent different assumptions about the small scale extrapolation of the ξ
mm
correlation, as outlined in the Section 6.2. In green (dashed) we show a
mildly optimistic case, using the linear fit shown in Fig. 15. The pink dot-
ted contours show a (strongly pessimistic) power law extrapolation. The
dark blue solid line makes identical assumtions to the pink, but incorporates
small-scale information, to a minimum separation of θ
+
min
= 0.5 arcmin-
utes in ξ
+
(θ) and θ
min
= 4.2 arcminutes in ξ
(θ). As in Fig. 16 the input
cosmology is shown by a black cross.
ing galaxy pairs on small scales one can expect a larger fraction in
which the objects come from a similar image plane environment,
and more often than not that enviroment will be densely populated.
Thus the true multiplicative bias should become more negative on
small scales.
Two subtly different effects emerge from this thought experi-
ment. First, the multiplicative bias of galaxies will be spatially cor-
related i.e. a correlation involving two galaxy populations
m
i
m
j
is not just the product of the means ¯m
i
¯m
j
. Second, in the small
θ bins one is selecting galaxies with close partners with which to
correlate, and thus oversampling the dense parts of the image. To
gauge the level of these effects, we divide each simulated coadd tile
into a grid of 25 square sub-patches with dimension 0.15 × 0.15
degrees. We fit for m using the galaxies in each sub-patch and as-
sign the resulting value to these objects. While this only allows a
noisy measurement of m, it should capture the spatial variations
in number density to the level of a few percent. We next measure
the two-point correlation function of multiplicative bias values as-
signed in this way using TREECORR
6
, excluding galaxy pairs at
angular separation smaller than the scale of the sub-patches. We re-
fer to this bias-bias autocorrelation as ξ
mm
, which we show as a
function of angular scale in the upper panel of Fig. 15. analogously
one could use the sub-patches to construct correlations between m
and galaxy number density ξ
gm
or density with density ξ
gg
. The
statistical noise on these correlations is significantly lower than that
on the individual sub-patches by virtue of the large simulation foot-
print. Note that in Fig. 15 we subtract ¯m
2
, measured from all galax-
ies in the simulation, from the measured ξ
mm
. If there were no θ
dependence the correlation hm
i
m
j
i should simply average to the
square of the global mean in all scale bins. As we can see from
6
http://rmjarvis.github.io/TreeCorr
MNRAS 000, 000–000 (0000)
16 S. Samuroff, S. L. Bridle, J. Zuntz et al
the circular points in this figure, scales larger than the diagonal size
the sub-patches (shown by the vertical dashed line) exhibit non-
negligible excess ξ
mm
. One obvious question is whether this could
be the result of finite binning error, which scatters galaxies in the
same sub-patch into different θ bins. To verify this is not the case
we repeat the measurement as before, but halve the parameter con-
trolling binning error tolerance (“bin slop”) and obtain the same
results.
To extend this measurement down to scales below the sub-
patch size we must make some assumptions about the functional
form of the mm correlation. We fit a power law, ξ
mm
(θ)
ξ
mm
(θ) ¯m
2
:
ξ
mm
(θ) = βθ
α
, (5)
which is shown by the dotted purple line in this figure. This pro-
vides a qualitiatively good fit to the measured points, but as we can
see implies a rather dramatic inflation on small scales.
In the limited range over which we have a nonzero measured
correlation, however, a linear function of θ (truncated at θ = 27
arcminutes) also provides a reasonable by-eye fit. The small-scale
extrapolation in this case is significantly milder. The h
g
i and
hδ
g
δ
g
i measurements are linear with θ to good approximation, and
so we use linear fits to extrapolate them below the patch size.
Assuming the bias per tomographic bin can be written as the
sum of a redshift dependent contribution (i.e. a scale invariant mean
in each bin), and a scale dependent term, one can write the corre-
lation per bin as m
i
m
j
= ¯m
i
¯m
j
+ ∆ξ
mm
(θ). A more complete
derivation of this expression can be found in Appendix A. The first
part can be extracted from the DES Y1 calibration, and we can fit
for ξ
mm
(θ) as described above. A set of modified ξ
ij
±
are thus
computed. These appear in the lower panels of Fig. 15 as dashed
lines. As we can see, the scale cuts of Troxel et al. (2017) (ex-
cluded scales are shaded in blue) are sufficiently stringent to re-
move almost all of the visible scale dependence. Though reassur-
ing for the immediate prospects of DES Y1, this will not trivially
be true for all future (or indeed ongoing) lensing surveys. It is thus
important that the effects we identify here are properly understood
at a level beyond the resources of the current paper. These biased
data are then passed through our likelihood pipeline to gauge the
cosmological impact, which is shown in Fig. 17. In the linear case
(dashed green) there is no discernable bias in the σ
8
m
pair; even
the much harsher power-law extrapolation (pink dotted) induces
only an incremental shift along the degeneracy direction. In both
cases the input cosmology still sits comfortably within the 1σ con-
fidence contour.
Finally we test the impact of relaxing the stringency of our
scale cuts. The minimum scales used for ξ
ij
+
and ξ
ij
are shifted
downwards to 0.5 and 4.2 arcminutes respectively, irrespective of
bin pair, which are the cut-off values used in fiducial cosmic shear
analysis of Hildebrandt et al. (2017). This increases the size of our
datavectors considerably. Incorporating smaller angular scales will
clearly improve the constraining power of the data to an extent. Pri-
marily the effect is to shorten the lensing degeneracy ellipse, cutting
out much of the peripheral curvature, but it also reduces the width
in the S
8
direction. These scales, however, contain biased informa-
tion, which induces tension between the small and large angular
scales. With the strongest (power law) scale dependence consid-
ered, the input cosmology is displaced marginally along the degen-
eracy curve, though it remains comfortably within the 1σ confi-
dence bound.
7 CONCLUSIONS
The Dark Energy Survey is the current state of the art in cosmo-
logical weak lensing. Multi-band imaging down to 24th magnitude
across 1500 square degrees of the southern sky has yielded hith-
erto unparalleled late-time constaints on the basic parameters of
the Universe (see Troxel et al. 2017 and DES Collaboration et al.
2017).
In this paper we have used one of two DES Y1 shear cata-
logues, and large-area simulations based upon them, to quantify
the impact of image plane neighbours on both ensemble shear mea-
surements, and on the inferred cosmological parameters.
In order to properly mitigate the influence of galaxy neigh-
bours, and thus avoid drawing flawed conclusions about cosmology
from the data, it is important to first understand the mechanisms
by which they enter the shape measurement. Using a simple toy
model of the galaxy-neighbour system we have shown that shear
bias can arise even when the distribution of neighbours is isotropic
(i.e. there is no preferred direction). This is the result of a small
difference in the impact of the same neighbour, when it is placed
on or away from the axis of the shear. We have furthermore shown
that the resulting multiplicative shear bias m can be either positive
or negative, depending on the model parameters. With slight mod-
ifications to the toy model, whereby we Monte Carlo sample input
parameters from the joint distribution of the equivalent properties
measured in DES Y1, we have shown that a mild negative m is
dominant when marginalising over a realistic ensemble of neigh-
bours. This was seen to be strongly dependent on the distance of
the neighbour, and to be mitigated but not eliminated by basic cuts
on the centroid position of the best-fitting model.
Using the DES Y1 HOOPOE simulations, which were
also used to derive shear calibration corrections for the Y1
IM3SHAPE catalogue of Z17, we have presented a detailed study of
the ensemble effects of galaxy neighbours. In this analysis we have
identified four mechanisms for neighbour bias, which we call flux
contamination, selection effects, bin shifting and neighbour dilua-
tion. All can be understood in intuitive terms, resulting from close-
by or moderately close neighbours. Our results from the full sim-
ulation are consistent with the toy model calculation. Though we
have shown strong dependence on distance to the nearest neigh-
bour (and thus on number density) we found only weak sensitiv-
ity to neighbour brightness, when averaged across broad bins of
magnitude. In addition to this, cuts on the DES Y1 catalogue suf-
ficient to null the impact of the detectable neighbours would result
in a degradation of over 20% in source number density. We can-
not recommend such measures for a code like IM3SHAPE, in part
because the data contains correlations between shear and number
density. Unless the link is preserved in the calibration simulations,
such selection could conceivably induce additional bias towards
low shear
7
.
Our investigation also assessed the impact of the faintest
galaxies, which are not reliably detected but nonetheless contribute
flux to the survey images. Via two different routes, first using a
spin-off neighbour-free resimulation, and also using a subset of im-
ages simulated again with the sub-detection galaxies missing, our
findings suggested a net contribution to the multiplicative bias bud-
get of m 0.01.
Unlike most earlier works on shear measurement, we have
7
Although the sister catalogue to Y1 IM3SHAPE uses a form of internal
calibration, which should allow one to correct for the additional selection
bias.
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 17
propagated these findings to the most meaningful metric for cos-
mic shear: bias on the inferred cosmological parameters. The study
we have presented here uses the DES Y1 cosmology pipeline, as
well as real non Gaussian shear covariance matrices and photo-
metric redshift distributions to implement MCMC forecasts. In the
first case considered, the data included a (different) multiplica-
tive bias in each redshift bin, designed to approximate the resid-
ual m that would arise were we to calibrate DES Y1 with a simple
neighbour-free simulation. Even marginalising over m with a prior
of N (0, 0.035) this scenario was demonstrated to result in a shift in
the favoured cosmology towards low clustering amplitude of more
than 1σ.
Finally, we have explored a second possible source of mea-
surement bias arising from the link between number density and
neighbour bias. This enters two-point measurements as an addi-
tional correlation between the multiplicative bias in galaxy pairs
at small angular separation. In the final section we have measured
such a correlation from the HOOPOE mock images. With the most
pessimistic small-scale extrapolation, this was found to result in
a shift in the best-fitting cosmology of under 1σ in the negative
S
8
direction, which is not remedied by marginalising over m. A
less dramatic, though still considerable, increase in the correlation
strength on small scales was demonstrated to result in no discern-
able cosmological bias.
Both of these effects are of primary concern for the next gen-
eration of cosmological surveys. By the end of their lifetime KiDS,
DES and HSC are set to offer lensing-based cosmological con-
straints comparable to the CMB. The first, dominant, effect can be
remedied relatively easily by calibrating our shear measurements
with sufficiently complex image simulations. Indeed, the most re-
cent shear constraints of Hildebrandt et al. (2017), K
¨
ohlinger et al.
(2017) and Troxel et al. (2017) have done just that. Unfortunately,
the correct treatment of scale dependent bias is not as clear, though
it should be captured at some level by the per-galaxy responses
upon which METACALIBRATION relies. Though further statements
about the likely small scale dependence of the mm correlation are
beyond the scope of the present study, understanding this intricate
topic will be crucial for future surveys if we are to fully exploit the
constraining power of the data. The massive simulation efforts of
LSST and Euclid, combined with advancement in neighbour miti-
gation using techniques such as multi-object fitting will be invalu-
able in this task. With the enhanced understanding these will pro-
vide and the exquisite data of the next generation surveys, the com-
ing decade will be an exciting time for cosmology.
8 ACKNOWLEDGEMENTS
We thank Nicolas Tessore, Catherine Heymans and Rachel Man-
delbaum for various insights that contributed to this work. We are
also indebted to the many DES “eyeballers” for lending their holi-
day time to help us understand and validate our simulations. The
HOOPOE simulations were generated using the National Energy
Research Scientific Computing Center (NERSC) facility, which is
maintained by the U.S. Department of Energy. The likelihood cal-
culations were performed using NERSC and the Fornax computing
cluster, which was funded by the European Research Council. SLB
acknowledges support from the European Research Council in the
form of a Consolidator Grant with number 681431. Support for DG
was provided by NASA through Einstein Postdoctoral Fellowship
grant number PF5-160138 awarded by the Chandra X-ray Center,
which is operated by the Smithsonian Astrophysical Observatory
for NASA under contract NAS8-03060.
Funding for the DES Projects has been provided by the
U.S. Department of Energy, the U.S. National Science Founda-
tion, the Ministry of Science and Education of Spain, the Sci-
ence and Technology Facilities Council of the United Kingdom, the
Higher Education Funding Council for England, the National Cen-
ter for Supercomputing Applications at the University of Illinois at
Urbana-Champaign, the Kavli Institute of Cosmological Physics at
the University of Chicago, the Center for Cosmology and Astro-
Particle Physics at the Ohio State University, the Mitchell Institute
for Fundamental Physics and Astronomy at Texas A&M Univer-
sity, Financiadora de Estudos e Projetos, Fundac¸
˜
ao Carlos Chagas
Filho de Amparo
`
a Pesquisa do Estado do Rio de Janeiro, Con-
selho Nacional de Desenvolvimento Cient
´
ıfico e Tecnol
´
ogico and
the Minist
´
erio da Ci
ˆ
encia, Tecnologia e Inovac¸
˜
ao, the Deutsche
Forschungsgemeinschaft and the Collaborating Institutions in the
Dark Energy Survey.
The Collaborating Institutions are Argonne National Labora-
tory, the University of California at Santa Cruz, the University of
Cambridge, Centro de Investigaciones Energ
´
eticas, Medioambien-
tales y Tecnol
´
ogicas-Madrid, the University of Chicago, Univer-
sity College London, the DES-Brazil Consortium, the University
of Edinburgh, the Eidgen
¨
ossische Technische Hochschule (ETH)
Z
¨
urich, Fermi National Accelerator Laboratory, the University of
Illinois at Urbana-Champaign, the Institut de Ci
`
encies de l’Espai
(IEEC/CSIC), the Institut de F
´
ısica d’Altes Energies, Lawrence
Berkeley National Laboratory, the Ludwig-Maximilians Univer-
sit
¨
at M
¨
unchen and the associated Excellence Cluster Universe, the
University of Michigan, the National Optical Astronomy Observa-
tory, the University of Nottingham, The Ohio State University, the
University of Pennsylvania, the University of Portsmouth, SLAC
National Accelerator Laboratory, Stanford University, the Univer-
sity of Sussex, Texas A&M University, and the OzDES Member-
ship Consortium.
The DES data management system is supported by the Na-
tional Science Foundation under Grant Numbers AST-1138766
and AST-1536171. The DES participants from Spanish institu-
tions are partially supported by MINECO under grants AYA2015-
71825, ESP2015-88861, FPA2015-68048, SEV-2012-0234, SEV-
2016-0597, and MDM-2015-0509, some of which include ERDF
funds from the European Union. IFAE is partially funded by the
CERCA program of the Generalitat de Catalunya. Research leading
to these results has received funding from the European Research
Council under the European Union’s Seventh Framework Pro-
gram (FP7/2007-2013) including ERC grant agreements 240672,
291329, and 306478. We acknowledge support from the Australian
Research Council Centre of Excellence for All-sky Astrophysics
(CAASTRO), through project number CE110001020.
This manuscript has been authored by Fermi Research Al-
liance, LLC under Contract No. DE-AC02-07CH11359 with the
U.S. Department of Energy, Office of Science, Office of High En-
ergy Physics. The United States Government retains and the pub-
lisher, by accepting the article for publication, acknowledges that
the United States Government retains a non-exclusive, paid-up, ir-
revocable, world-wide license to publish or reproduce the published
form of this manuscript, or allow others to do so, for United States
Government purposes.
Based in part on observations at Cerro Tololo Inter-American
Observatory, National Optical Astronomy Observatory, which is
operated by the Association of Universities for Research in As-
MNRAS 000, 000–000 (0000)
18 S. Samuroff, S. L. Bridle, J. Zuntz et al
tronomy (AURA) under a cooperative agreement with the National
Science Foundation.
REFERENCES
Abbott T. et al., 2016, Phys. Rev. D, 94, 022001
Albrecht A. et al., 2006, arXiv:astro-ph/0609591
Bacon D. J., Refregier A. R., Ellis R. S., 2000, MNRAS, 318, 625
Bernstein G. M., Armstrong R., 2014, MNRAS, 438, 1880
Bernstein G. M., Jarvis M., 2002, AJ, 123, 583
Bertin E., Fouqu
´
e P., 2010, SkyMaker: Astronomical Image Simulations
Made Easy. Astrophysics Source Code Library
Bridle S. et al., 2009, Annals of Applied Statistics, 3, 6
Bridle S. L., Kneib J.-P., Bardeau S., Gull S. F., 2002, in The Shapes of
Galaxies and their Dark Halos, Natarajan P., ed., pp. 38–46
Brouwer M. M. et al., 2017, MNRAS, 466, 2547
Chang C. et al., 2013, MNRAS, 434, 2121
Chang C. et al., 2015, Physical Review Letters, 115, 051301
Chang C., et al., 2017, submitted to Mon. Not. R. Astron. Soc.
Clampitt J. et al., 2017, MNRAS, 465, 4204
DES Collaboration, et al., 2017, to be submitted to Phys. Rev. D
Diehl H. T. et al., 2014, in Proc. SPIE, Vol. 9149, Observatory Operations:
Strategies, Processes, and Systems V, p. 91490V
Drlica-Wagner A., et al., 2017, submitted to Astrophys. J. Suppl. Ser.
Fenech Conti I., Herbonnet R., Hoekstra H., Merten J., Miller L., Viola M.,
2017, MNRAS, 467, 1627
Feroz F., Hobson M. P., Cameron E., Pettitt A. N., 2013, arXiv:1306.2144
Flaugher B. et al., 2015, AJ, 150, 150
Harnois-D
´
eraps J., Munshi D., Valageas P., van Waerbeke L., Brax P., Coles
P., Rizzo L., 2015, MNRAS, 454, 2722
Hartlap J., Hilbert S., Schneider P., Hildebrandt H., 2011, A&A, 528, A51
Herbonnet R., Buddendiek A., Kuijken K., 2017, A&A, 599, A73
Heymans C. et al., 2013, MNRAS, 432, 2433
Heymans C. et al., 2006, MNRAS, 368, 1323
Heymans C. et al., 2012, MNRAS, 427, 146
Hildebrandt H. et al., 2017, MNRAS, 465, 1454
Hirata C., Seljak U., 2003, MNRAS, 343, 459
Hirata C. M., Seljak U., 2004, Phys. Rev. D, 70, 063526
Hoekstra H., Herbonnet R., Muzzin A., Babul A., Mahdavi A., Viola M.,
Cacciato M., 2015, MNRAS, 449, 685
Hoekstra H., Viola M., Herbonnet R., 2017, MNRAS, 468, 3295
Hoyle B., et al., 2017, to be submitted to Mon. Not. R. Astron. Soc.
Huff E., Mandelbaum R., 2017, arXiv:1702.02600
Jarvis M., Jain B., Bernstein G., Dolney D., 2006, ApJ, 644, 71
Jarvis M. et al., 2016, MNRAS, 460, 2245
Jee M. J., Tyson J. A., Hilbert S., Schneider M. D., Schmidt S., Wittman D.,
2016, ApJ, 824, 77
Joachimi B., Bridle S. L., 2010, A&A, 523, A1
Kacprzak T., Bridle S., Rowe B., Voigt L., Zuntz J., Hirsch M., MacCrann
N., 2014, MNRAS, 441, 2528
Kacprzak T. et al., 2016, MNRAS, 463, 3653
Kacprzak T., Zuntz J., Rowe B., Bridle S., Refregier A., Amara A., Voigt
L., Hirsch M., 2012, MNRAS, 427, 2711
Kaiser N., 1994, in Clusters of Galaxies, Durret F., Mazure A., Tran Thanh
Van J., eds., p. 269
Kaiser N., Squires G., Broadhurst T., 1995, ApJ, 449, 460
Kaiser N., Wilson G., Luppino G. A., 2000, ArXiv Astrophysics e-prints
Kilbinger M. et al., 2013, MNRAS, 430, 2200
Kitching T. et al., 2010, arXiv:1009.0779
K
¨
ohlinger F. et al., 2017, arXiv:1706.02892
Krause E., Eifler T., 2016, arXiv:1601.05779
Krause E. et al., 2017a, arXiv:1706.09359
Krause E., et al., 2017b, submitted to Phys. Rev. D
Kuijken K., 1999, A&A, 352, 355
Lewis A., 2009, MNRAS, 398, 471
Mandelbaum R. et al., 2015, MNRAS, 450, 2963
Mandelbaum R. et al., 2014, ApJS, 212, 5
Maoli R., Van Waerbeke L., Mellier Y., Schneider P., Jain B., Bernardeau
F., Erben T., Fort B., 2001, A&A, 368, 766
Massey R. et al., 2007, ApJS, 172, 239
Massey et al., 2007, MNRAS, 376, 13
Miller L. et al., 2013, MNRAS, 429, 2858
Prat J., et al., 2017, to be submitted to Phys. Rev. D
Refregier A., Bacon D., 2003, MNRAS, 338, 48
Refregier A., Kacprzak T., Amara A., Bridle S., Rowe B., 2012, MNRAS,
425, 1951
Rhodes J., Refregier A., Groth E. J., 2001, ApJ, 552, L85
Schmidt F., Rozo E., Dodelson S., Hui L., Sheldon E., 2009, ApJ, 702, 593
Scoville N. et al., 2007, ApJS, 172, 1
Sheldon E. S., Huff E. M., 2017, arXiv:1702.02601
Simon P., Schneider P., 2016, arXiv:1609.07937
Simpson F. et al., 2013, MNRAS, 429, 2249
Troxel M. A., et al., 2017, to be submitted to Phys. Rev. D
Van Waerbeke L. et al., 2013, MNRAS, 433, 3373
Van Waerbeke L. et al., 2000, A&A, 358, 30
Voigt L. M., Bridle S. L., 2010, MNRAS, 404, 458
Weinberg D. H., Mortonson M. J., Eisenstein D. J., Hirata C., Riess A. G.,
Rozo E., 2013, Phys. Rep., 530, 87
Whittaker L., Brown M. L., Battye R. A., 2015, MNRAS, 454, 2154
Wittman D. M., Tyson J. A., Kirkman D., Dell’Antonio I., Bernstein G.,
2000, Nature, 405, 143
Zuntz J., Kacprzak T., Voigt L., Hirsch M., Rowe B., Bridle S., 2013, MN-
RAS, 434, 1604
Zuntz J., et al., 2017, submitted to Mon. Not. R. Astron. Soc.
APPENDIX A: DERIVATION OF A TWO-POINT
MODIFIER FOR SCALE DEPENDENT BIAS
In the following we set out a brief derivation of the analytic modifi-
cations to account for scale-dependent neighbour effects the shear-
shear two-point correlations used in the earlier section. We do not
claim that this is a precise calculation of the sort that could be used
to derive a robust calibration. Rather it is an order of magnitude
estimate to allow us to assess the approximate size of the cosmo-
logical bias these effects could induce in the data.
First, with complete generality it is possible to write the i com-
ponent of the measured shear at angular position θ as
γ
obs
i
(θ) = [1 + m
i
(θ)] γ
i
(θ), (A1)
where γ
i
is the underlying true shear, which is sensitive to cos-
mology only. Extending this to the level of a two-point correlation
between two populations α and β this implies:
ξ
obs,αβ
i
(θ)
D
γ
obs
i
(θ
0
)γ
obs
i
(θ
0
+ θ)
E
θ
=
D
[1 + m
α
i
(θ
0
)][1 + m
β
i
(θ
0
+ θ)]˜γ
α
i
(θ
0
)˜γ
β
i
(θ
0
+ θ)
E
θ
.
(A2)
Note that the observed shear used in a particular bin correlation
is now weighted by the overdensity of galaxies in the image, in
addition to the calibration bias, such that
˜γ
α
i
(θ)
1 + δ
α
g
(θ)
× γ
α
i
(θ). (A3)
Expanding each of the terms one finds:
MNRAS 000, 000–000 (0000)
Cosmic Shear & Galaxy Neighbours 19
ξ
obs,αβ
i
(θ) =
D
γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
m
α
i
(θ
0
)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
m
β
i
(θ
0
+ θ)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
δ
β
g
(θ
0
+ θ)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
m
α
i
(θ
0
)m
β
i
(θ
0
+ θ)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)δ
β
g
(θ
0
+ θ)γ
α
i
(θ
0
)γ
β
i
(θ
0
+ θ)
E
θ
.
(A4)
The terms contributing to the measured two-point shear correla-
tion, then, is sensitive to both spatial correlations between the m
in different galaxies and to the correlations with the source den-
sity. Note that we’ve chosen to neglect a higher-order (six-point)
term. In reality there will also be a connection between galaxy den-
sity and shear, but we will follow the normal convention and as-
sume the contribution is small enough to be neglected. In simple
terms, an excess in the hmmi term above the product of the mean
m values indpendently could arise because galaxy pairs separated
on small scales tend to come from similar image plane environ-
ments. In contrast the density weighted correlations hδ
g
mi would
be zero, but for a simple observation; selecting a random galaxy
with a suitable correlation pair at a distance θ is not the same as un-
conditionally selecting a random galaxy. In the small scale bins we
will over-sample the dense regions, where m tends to be larger (see
Section 6.2). The angular brackets here indicate averaging over all
galaxy pairs separated by θ. If we can assume the bias is indepen-
dent of the underlying cosmology the above expression simplifies
significantly:
ξ
obs,αβ
i
(θ) = (1 + ¯m
α
i
+ ¯m
β
i
+
D
m
α
i
(θ
0
)m
β
i
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)m
β
i
(θ
0
+ θ)
E
θ
+
D
m
α
i
(θ
0
)δ
β
g
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)δ
β
g
(θ
0
+ θ)
E
θ
) × ξ
αβ
i
(θ|p), (A5)
with ξ
αβ
i
being the true correlation function of cosmological shears
hγ
i
γ
i
i, which is contingent on the underlying cosmological param-
eters p. It can be shown that
ξ
+
(θ)
γ
+
(θ
0
)γ
+
(θ
0
+ θ)
θ
±
γ
×
(θ
0
)γ
×
(θ
0
+ θ)
θ
=
γ
1
(θ
0
)γ
1
(θ
0
+ θ)
θ
±
γ
2
(θ
0
)γ
2
(θ
0
+ θ)
θ
= ξ
1
(θ) + ξ
2
(θ), (A6)
and so one can use equation A5 to construct the observed ξ
±
cor-
relation functions
ξ
obs,αβ
±
(θ) =
1 + ¯m
α
+ ¯m
β
+
D
m
α
(θ
0
)m
β
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)m
β
(θ
0
+ θ)
E
θ
+
D
m
α
(θ
0
)δ
β
g
(θ
0
+ θ)
E
θ
+
D
δ
α
g
(θ
0
)δ
β
g
(θ
0
+ θ)
E
θ
ξ
αβ
±
(θ|p). (A7)
The i subscript has been discarded here under the assumption that
m
1
and m
2
are approximately equal for a given set of galaxies.
Next, let’s say imagine that we have a measured datavector.
Our measurements are biased, but we’ll assume it is possible to
devise a correction that recovers the true cosmological signal pre-
cisely. Our observed datavector is then just,
ξ
obs,αβ
±
(θ) = Υ
tr,αβ
ξ
αβ
±
(θ|p), (A8)
which follows trivially from equation A7. Since we do not trivially
know Υ
tr,αβ
ab initio (that’s why we need simulations!) we can
only construct a best-estimate approximation. By applying a cor-
rection factor to the raw measurements we construct a best-estimate
datavector:
ξ
BE,αβ
±
(θ) =
1
Υ
BE,αβ
ξ
obs,αβ
±
(θ) =
Υ
tr,αβ
Υ
BE,αβ
ξ
αβ
±
(θ|p). (A9)
Of course, if our best correction is perfect then the ratio goes to
unity, and we recover the underlying cosmology. Since we apply
corrections to the single-galaxy shears we will assume Υ
BE,αβ
in-
cludes the hδ
g
δ
g
i term, but neglects the correlations involving m.
We then can write:
Υ
BE,αβ
=
1 + ¯m
α
+ ¯m
β
+ ¯m
α
¯m
β
+
D
δ
α
g
(θ
0
)δ
β
g
(θ
0
+ θ)
E
θ
.
(A10)
We can measure the mean bias in each bin that would be obtained
from the calibration directly. As we show in Z17, using the full
DES Y1 HOOPOE catalogues, these biases are 0.08 to 0.20.
Finally, assume that although m clearly varies between
redhshift bins, the strength of the correlation does not That is, the
bias-bias term is the product of the mean ms (which varies between
z bins) plus a scale dependent shift (which doesn’t). One then has:
D
m
α
(θ
0
)m
β
(θ
0
+ θ)
E
θ
= ¯m
α
¯m
β
+ ξ
mm
(θ). (A11)
The additive part can be measured directly from the simulation us-
ing sub-patches, as described earlier. The density-density correla-
tion can be obtained in the same way. This, then, leaves only the
m × δ
g
cross-correlation. This should vanish in the case of zero
correlation, but it also seems reasonable to assume that the magni-
tude should be proportional to the mean bias ¯m
α
in a particular bin.
This allows the scale dependent (non-tomographic) cross correla-
tion measured from HOOPOE to be rescaled appropriately for each
bin pair:
D
δ
α
(θ
0
)m
β
(θ
0
+ θ)
E
θ
=
¯m
β
¯m
ξ
gm
(θ), (A12)
where ¯m is the global multiplicative bias and ξ
gm
(θ) h
g
i,
each measured using all simulated galaxies. Using the above equa-
tions, with our fiducial calibration and three measured correlations,
one can derive a scale dependent modification to shear-shear two
point correlation data using equation A8.
AFFILIATIONS
1
Jodrell Bank Centre for Astrophysics, School of Physics and
Astronomy, University of Manchester, Oxford Road, Manchester,
M13 9PL, UK
2
Institute for Astronomy, University of Edinburgh, Edinburgh
EH9 3HJ, UK
3
Center for Cosmology and Astro-Particle Physics, The Ohio
State University, Columbus, OH 43210, USA
4
Department of Physics, The Ohio State University, Columbus,
MNRAS 000, 000–000 (0000)
20 S. Samuroff, S. L. Bridle, J. Zuntz et al
OH 43210, USA
5
Kavli Institute for Particle Astrophysics & Cosmology, P. O. Box
2450, Stanford University, Stanford, CA 94305, USA
6
SLAC National Accelerator Laboratory, Menlo Park, CA 94025,
USA
7
Department of Physics and Astronomy, University of Pennsylva-
nia, Philadelphia, PA 19104, USA
8
Jet Propulsion Laboratory, California Institute of Technology,
4800 Oak Grove Dr., Pasadena, CA 91109, USA
9
Department of Physics, ETH Z
¨
urich, Wolfgang-Pauli-Strasse 16,
CH-8093 Z
¨
urich, Switzerland
10
Department of Physics & Astronomy, University College
London, Gower Street, London, WC1E 6BT, UK
11
Department of Physics and Electronics, Rhodes University, PO
Box 94, Grahamstown, 6140, South Africa
12
Fermi National Accelerator Laboratory, P. O. Box 500, Batavia,
IL 60510, USA
13
LSST, 933 North Cherry Avenue, Tucson, AZ 85721, USA
14
CNRS, UMR 7095, Institut d’Astrophysique de Paris, F-75014,
Paris, France
15
Sorbonne Universit
´
es, UPMC Univ Paris 06, UMR 7095,
Institut d’Astrophysique de Paris, F-75014, Paris, France
16
Laborat
´
orio Interinstitucional de e-Astronomia - LIneA, Rua
Gal. Jos
´
e Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil
17
Observat
´
orio Nacional, Rua Gal. Jos
´
e Cristino 77, Rio de
Janeiro, RJ - 20921-400, Brazil
18
Department of Astronomy, University of Illinois, 1002 W.
Green Street, Urbana, IL 61801, USA
19
National Center for Supercomputing Applications, 1205 West
Clark St., Urbana, IL 61801, USA
20
Institut de F
´
ısica d’Altes Energies (IFAE), The Barcelona Insti-
tute of Science and Technology, Campus UAB, 08193 Bellaterra
(Barcelona) Spain
21
Institute of Space Sciences, IEEC-CSIC, Campus UAB, Carrer
de Can Magrans, s/n, 08193 Barcelona, Spain
22
Department of Physics, IIT Hyderabad, Kandi, Telangana
502285, India
23
Kavli Institute for Cosmological Physics, University of Chicago,
Chicago, IL 60637, USA
24
Instituto de Fisica Teorica UAM/CSIC, Universidad Autonoma
de Madrid, 28049 Madrid, Spain
25
Department of Astronomy, University of Michigan, Ann Arbor,
MI 48109, USA
26
Department of Physics, University of Michigan, Ann Arbor, MI
48109, USA
27
Astronomy Department, University of Washington, Box
351580, Seattle, WA 98195, USA
28
Cerro Tololo Inter-American Observatory, National Optical
Astronomy Observatory, Casilla 603, La Serena, Chile
29
Santa Cruz Institute for Particle Physics, Santa Cruz, CA 95064,
USA
30
Australian Astronomical Observatory, North Ryde, NSW 2113,
Australia
31
Argonne National Laboratory, 9700 South Cass Avenue,
Lemont, IL 60439, USA
32
Departamento de F
´
ısica Matem
´
atica, Instituto de F
´
ısica, Univer-
sidade de S
˜
ao Paulo, CP 66318, S
˜
ao Paulo, SP, 05314-970, Brazil
Station, TX 77843, USA
34
Department of Astronomy, The Ohio State University, Colum-
bus, OH 43210, USA
35
Department of Astrophysical Sciences, Princeton University,
Peyton Hall, Princeton, NJ 08544, USA
36
Instituci
´
o Catalana de Recerca i Estudis Avanc¸ats, E-08010
Barcelona, Spain
37
Centro de Investigaciones Energ
´
eticas, Medioambientales y
Tecnol
´
ogicas (CIEMAT), Madrid, Spain
38
Brookhaven National Laboratory, Bldg 510, Upton, NY 11973,
USA
39
School of Physics and Astronomy, University of Southampton,
Southampton, SO17 1BJ, UK
40
Instituto de F
´
ısica Gleb Wataghin, Universidade Estadual de
Campinas, 13083-859, Campinas, SP, Brazil
41
Computer Science and Mathematics Division, Oak Ridge
National Laboratory, Oak Ridge, TN 37831
42
Institute of Cosmology & Gravitation, University of Portsmouth,
Portsmouth, PO1 3FX, UK
MNRAS 000, 000–000 (0000)