6
Reducing the shear Martin White * Department of Astronomy, University of California, 601 Campbell Hall, Berkeley, CA 94720, USA Department of Physics, University of California, 601 Campbell Hall, Berkeley, CA 94720, USA Received 21 January 2005; received in revised form 31 January 2005; accepted 31 January 2005 Available online 8 March 2005 Abstract As gravitational lensing measurements become increasingly precise, it becomes necessary to include higher order effects in theoretical calculations. Here we show how the difference between the shear and the reduced shear manifest themselves in a number of commonly used measures of shear power. If we are to reap the science rewards of future, high precision measurements of cosmic shear we will need to include this effect in our theoretical predictions. Ó 2005 Elsevier B.V. All rights reserved. PACS: 98.65.Dx; 98.80.Es; 98.70.Vc Keywords: Cosmology; Lensing; Large-scale structures 1. Introduction Weak gravitational lensing by large-scale struc- ture [13,2] is becoming a central means of con- straining our cosmological model (see e.g. [8,24], for the current status). Continuing effort has yielded increasingly stringent control of systematic errors and ever larger surveys are decreasing the statistical errors rapidly. As the precision with which the measurements are made increases, one demands ever higher fidelity in the data-analysis, modeling and theoretical interpretation. One such area is the computation of the theo- retical predictions for various statistics involving the measurable Ôreduced shearÕ [13,2] g c 1 j for j gj< 1 ð1Þ where c is the ÔplainÕ shear and j is the convergence which is related to the projected mass along the line of sight. Both j and c are defined through the Jacobian of the mapping between the source and image planes. If the lensing is weak then jjj 1, and to lowest order g = c and the power 0927-6505/$ - see front matter Ó 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.astropartphys.2005.01.008 * Address: Department of Astronomy, University of Califor- nia, 601 Campbell Hall, Berkeley, CA 94720, USA. Tel.: +1 510 642 1969; fax: +1 510 642 3411. E-mail address: [email protected] Astroparticle Physics 23 (2005) 349–354 www.elsevier.com/locate/astropart

Reducing the shear

Embed Size (px)

Citation preview

Page 1: Reducing the shear

Astroparticle Physics 23 (2005) 349–354

www.elsevier.com/locate/astropart

Reducing the shear

Martin White *

Department of Astronomy, University of California, 601 Campbell Hall, Berkeley, CA 94720, USA

Department of Physics, University of California, 601 Campbell Hall, Berkeley, CA 94720, USA

Received 21 January 2005; received in revised form 31 January 2005; accepted 31 January 2005

Available online 8 March 2005

Abstract

As gravitational lensing measurements become increasingly precise, it becomes necessary to include higher order

effects in theoretical calculations. Here we show how the difference between the shear and the reduced shear manifest

themselves in a number of commonly used measures of shear power. If we are to reap the science rewards of future, high

precision measurements of cosmic shear we will need to include this effect in our theoretical predictions.

� 2005 Elsevier B.V. All rights reserved.

PACS: 98.65.Dx; 98.80.Es; 98.70.Vc

Keywords: Cosmology; Lensing; Large-scale structures

1. Introduction

Weak gravitational lensing by large-scale struc-

ture [13,2] is becoming a central means of con-

straining our cosmological model (see e.g. [8,24],

for the current status). Continuing effort hasyielded increasingly stringent control of systematic

errors and ever larger surveys are decreasing the

statistical errors rapidly. As the precision with

0927-6505/$ - see front matter � 2005 Elsevier B.V. All rights reserv

doi:10.1016/j.astropartphys.2005.01.008

* Address: Department of Astronomy, University of Califor-

nia, 601 Campbell Hall, Berkeley, CA 94720, USA. Tel.: +1 510

642 1969; fax: +1 510 642 3411.

E-mail address: [email protected]

which the measurements are made increases, one

demands ever higher fidelity in the data-analysis,

modeling and theoretical interpretation.

One such area is the computation of the theo-

retical predictions for various statistics involving

the measurable �reduced shear� [13,2]

g � c1� j

for jgj< 1 ð1Þ

where c is the �plain� shear and j is the convergence

which is related to the projected mass along theline of sight. Both j and c are defined through

the Jacobian of the mapping between the source

and image planes. If the lensing is weak then

jjj � 1, and to lowest order g = c and the power

ed.

Page 2: Reducing the shear

350 M. White / Astroparticle Physics 23 (2005) 349–354

spectra of the shear and convergence are equal:

Cc‘ ¼ Cj

‘ . Thus measurements are usually com-

pared to predictions of Cj‘ derived from models

of the non-linear clustering of matter. This relation

has corrections however, which need to be takeninto account if precise comparisons with data are

to be made in the future. While there has been pre-

vious analytic work on this subject [19,5] and some

numerical work on magnification statistics [20,1,

14], there has been no comprehensive investigation

of this effect in simulations. In this short paper we

present some results, derived from simulations

similar to those described in [27], on the size ofthe corrections for a variety of commonly used sta-

tistics on the angular scales where most current

lensing work been has focused.

2. Simulation

The results are derived from 7 N-body simula-tions of a KCDM cosmology run with the TreePM

code [26]. Each simulation employed 3843 equal

mass dark matter particles in a periodic cubical

box of side 200h�1Mpc run to z = 0 with phase

space data dumped every 50h�1Mpc starting at

z = 4. The cosmology was the same for each simu-

lation (Xmat = 0.28, XBh2 = 0.024, h = 0.7, n = 1

and r8 = 0.9) but different random number seedswere used to generate the initial conditions. The

purpose of these runs was to allow a study of the

statistical distributions of lensing observables.

Each box was used to generate 16 approximately

independent lensing maps, each 3� · 3�, using the

multi-plane ray-tracing code described in [23].

Each map was produced with 20482 pixels and

then downsampled to 10242 pixels.1 Thus a totalof 112 maps or 1008 square degrees were simu-

lated, all with the same cosmology. We report on

the results for sources at z ’ 1 (specifically all at

a comoving distance of 2400h�1Mpc) here. Other

data can be obtained from http://mwhite.berke-

ley.edu/Lensing.

1 Tests with maps downsampled to 5122 pixels indicate that

our results are converged at the several percent level in the

ratios that we quote, with better agreement at larger scales as

expected.

3. Results

3.1. Two-point statistics

The lowest order information on large-scalestructure comes from studying the two-point statis-

tics of the shear field. A number of different two-

point statistics have been used in the literature,

each with their own strengths and weaknesses.

We survey the most commonly used statistics here.

We first show the results on the variance of the

shear, smoothed on a range of angular scales.

Though it has numerous drawbacks in interpreta-tion, this was one of the first statistics computed

from shear maps due to its ease of computation.

The left panel of Fig. 1 shows the (reduced) shear

variance as a function of smoothing scale. For

each point we compute a map of the amplitude

of the reduced shear, smooth the map with a 2D

boxcar of side length R and then compute the var-

iance of the resulting map. The points are highlycorrelated. Since we have full information about

the shear and convergence from the simulation

we are able to compare this with the plain shear re-

sults which are usually computed. The right panel

of Fig. 1 shows the difference of this statistic com-

puted for the reduced shear (g) to that for plain

shear (c). We see that the reduced shear variance

is larger than the usually predicted shear varianceby a non-trivial amount. The bias very gradually

drops as we go to larger scales (not shown here)

but we are unable to follow it to convergence

due to the finite size of the fields simulated.

Naively one might imagine that the corrections

would be smaller than shown in Fig. 1 because j is

small �on large scales�. However j is only small

when averaged over a sizeable region of sky, andthis smoothing does not commute with the divi-

sion in Eq. (1). This fact also means that perturba-

tive calculations must be used with care, because

the convergence or shear amplitude can be quite

large on small scales leading to a break down of

the approximation.

Though the smoothed shear variance has the

worst behavior of the two-point functions we con-sider, we will see qualitatively similar behavior

below. In general this tendency for the small-scale

power to be increased over the theoretical predic-

Page 3: Reducing the shear

Fig. 1. (Left) The smoothed shear variance, Var[jgjR], as a function of smoothing scale R. (Right) The relative difference of the

smoothed shear variance computed for plain shear and reduced shear, as a function of smoothing scale R. The points show the mean

difference, averaged over 112 maps, each 3� · 3�, while the error bars indicate the standard deviation.

Fig. 2. The relative difference of theMap variance computed for

plain shear and reduced shear. The points show the mean

difference, averaged over 112 maps, each 3� · 3�, while the error

bars indicate the standard deviation of the ratio.

M. White / Astroparticle Physics 23 (2005) 349–354 351

tion for plain shear needs to be considered beforeattributing ‘‘excess small-scale power’’ to addi-

tional physical effects e.g. intrinsic galaxy align-

ments. A signature of the effect might be the

difference in the sensitivity of the various statistics

that we compute here.

Another popular measure of the power is the

aperture mass [19]. This is a scalar quantity which

can be derived from an integral over the shear

Mapðh0;RÞ �Z

d2h Qðj~hj;RÞcTð~hþ~h0Þ ð2Þ

where cT is the tangential shear as measured from

h0 and Q is a kernel.

We have chosen the ‘ = 1 form for definiteness,

so

Qðr ¼ x=R;RÞ ¼ 6

pR2r2ð1� r2Þ for r 6 1 ð3Þ

and Q vanishes for r > 1. The results for the aper-

ture mass variance are shown in Fig. 2. Note that

the correction is quite large on small scales, but

drops rapidly on scales above a few arcminutes.

Finally we show results for the two-point corre-

lation function of the shear in Fig. 3. This is thebest behaved two-point statistic that we consider,

since it explicitly eliminates the contribution from

small angular scales.2 We define n = hc+c+i where

2 For this reason the two-point correlation function is less

sensitive to small-scale observational systematics than other

quantities such as Var[Map].

c+ is (minus) the component c1 of the shear in

the rotated frame whose x̂-axis is the separation

vector between the two points being correlated.

The component at 45� is called c·. We find that

the correction to the correlation function is quite

small on the scales shown here. The correction

to zthe other correlation function, hc·c·i, is even

smaller than for n.It is not too surprising that n receives smaller

corrections than Var[Map] when we recall that

Map probes smaller scales than R by a factor of

roughly 3 [19]. A smaller piece of the difference is

Page 4: Reducing the shear

Fig. 3. The relative difference of n(r) computed for plain shear

and reduced shear. The points show the mean difference,

averaged over 112 maps, each 3� · 3�, while the error bars

indicate the standard deviation of the difference. The effect on

hc·c·i is smaller and is omitted.

Fig. 4. The relative difference of the three-point correlation

function, f+++, for equilateral triangle configurations computed

for plain shear and reduced shear. The points show the mean

difference, averaged over 32 maps, each 3� · 3�, while the error

bars indicate the standard deviation.

352 M. White / Astroparticle Physics 23 (2005) 349–354

that Map can be expressed as an integral over nextending all the way to zero lag,3 making Map

slightly more sensitive to the difference between g

and c on small scales. A similar argument holds

for Var[jgjR], with the effect being more pro-

nounced. A reduction in the sensitivity to small

scale power and to shot-noise could be obtained

from differencing these measures between scales,

with an associated loss of power.

These are all of the two-point functions com-

monly derived from shear data—the power spec-trum has only been derived by two groups [16,4]

despite being almost ubiquitous in forecasts of

the potential of future lensing experiments (e.g.

[10,11,3,9,25,15,17,22,12]).

3.2. Three-point statistics

Since the lensing maps are non-Gaussian thereis information beyond4 the two-point statistics.

3 The variance of any convolution of j must have weight at

zero lag.4 This information is not, however, independent. We find a

strong correlation between the amplitude of the two- and three-

point functions on scales of several arcminutes in our maps.

We show in Fig. 4 the dominant three-point func-

tion for equilateral triangles as a function of scale.

The three-point function is hc+c+c+i with c+ de-

fined with respect to a line joining the triangle

center to each vertex [21,28,18]. In the weakly

non-Gaussian limit one can argue that the correc-

tions to the three-point function from consideringreduced shear could be very large. Counting

powers of our small parameter, j or c, the lowest

order contribution vanishes in the Gaussian limit.

Thus the first non-vanishing term is of order j4

[19,5]. The first correction from the reduced shear

also comes in at order j4. However we see that the

difference is not significantly larger than in the case

of the two-point functions on the scales shownhere.

We also show, in Fig. 5, how the configuration

dependence of the three-point function is modified.

The three-point function for triangles with sides

h1 = 2 0 and h2 = 3 0 is plotted as a function of the

cosine of the included angle. We have chosen here

a triangle with no special symmetries in order to

illustrate a ‘‘typical’’ case. Note that the correctionis dependent on angle, showing that the shear and

reduced shear have different shape dependence

which should be taken into account in cosmologi-

cal inferences (e.g.[7,6]).

Page 5: Reducing the shear

Fig. 5. The relative difference of the three-point correlation

function for one triangle size, f+++(h1 = 2 0,h2 = 3 0), computed

for plain shear and reduced shear. The points show the mean

difference, averaged over 32 maps, each 3� · 3�, while the error

bars indicate the standard deviation.

M. White / Astroparticle Physics 23 (2005) 349–354 353

4. Conclusions

Deflection of light rays by gravitational poten-tials along the line of sight introduces a mapping

between the source and image plane. The Jacobian

of this mapping defines the shear and convergence

as a function of position on the sky. In the absence

of size or magnification information neither the

shear nor the convergence is observable, rather

the combination g = c/(1�j) is. Since on small

scales j can be non-negligible this introduces com-plications in predicting the observables of weak

lensing. We have illustrated specifically the effect

of using g rather than c on a number of commonly

used statistics of weak lensing. We found that the

correlation function is least affected, as expected

since it specifically eliminates small-scale informa-

tion. The variance of the shear amplitude,

smoothed on a scale R, is the most dramaticallyaffected, and the amplitude of the effect declines

very slowly with increasing smoothing scale.

Existing lensing experiments derive most of

their cosmological constraints from angular scales

of a few arcminutes to a few tens of arcminutes.

The corrections described here are below the cur-

rent statistical errors, and so should not affect

existing constraints. The next generation of experi-

ments may have sufficient control of systematic er-

rors, and sufficient statistics, that this effect will

need to be included in the analysis. Note that even

for the correlation function the difference between

the commonly predicted plain shear result and thereduced shear value is comparable to the level of

systematic control that we need to achieve to study

dark energy! If we are to reap the science rewards

of future, we need to include the effect of high pre-

cision measurements of cosmic shear in our theo-

retical predictions.

Acknowledgements

M.W. thanks M. Takada for comparison of

results on the shear two-point function and

D. Huterer, B. Jain and P. Zhang for helpful com-

ments. The simulations used here were performed

on the IBM-SP at NERSC. This research was sup-

ported by the NSF and NASA.

References

[1] A.J. Barber, A.N. Taylor, MNRAS 344 (2003) 789.

[2] M. Bartelmann, P. Schneider, Phys. Rep. 340 (2001) 291.

[3] K. Benabed, F. Bernardeau, Phys. Rev. D 64 (2001)

083501.

[4] M.L. Brown et al., MNRAS 341 (2003) 100.

[5] S. Dodelson, P. Zhang, 2005, preprint [astro-ph/0501063].

[6] D. Dolney, B. Jain, M. Takada, MNRAS 352 (2004) 1019.

[7] S. Ho, M. White, Astrophys. J. 607 (2004) 40.

[8] H. Hoekstra, H.K.C. Yee, M.D. Gladders, New Astron.

Rev. 46 (2002) 767.

[9] W. Hu, Phys. Rev. D 66 (2001) 3515.

[10] L. Hui, Astrophys. J. 519 (1999) 9.

[11] D. Huterer, Phys. Rev. D 65 (2001) 063001.

[12] D. Huterer, M. Takada, 2005, preprint [astro-ph/0412142].

[13] Y. Mellier, Ann. Rev. Astron. Astrophys. 37 (1999) 127.

[14] B. Menard, T. Hamana, M. Bartelmann, N. Yoshida,

A&A 403 (2003) 817.

[15] D. Munshi, Y. Wang, Astrophys. J. 583 (2003) 566.

[16] U.-L. Pen et al., MNRAS 346 (2003) 994.

[17] A. Refregier et al., Astron. J. 127 (2004) 3102.

[18] P. Schneider, M. Lombardi, A&A 397 (2003) 809.

[19] P. Schneider, L. van Waerbeke, B. Jain, G. Kruse,

MNRAS 296 (1998) 873.

[20] M. Takada, T. Hamana, MNRAS 346 (2003) 949.

[21] M. Takada, B. Jain, MNRAS 344 (2003) 857.

[22] M. Takada, M. White, Astrophys. J. 601 (2004) L1.

[23] C. Vale, M. White, Astrophys. J. 592 (2003) 699.

Page 6: Reducing the shear

354 M. White / Astroparticle Physics 23 (2005) 349–354

[24] L. van Waerbeke, Y. Mellier, H. Hoekstra, A&A 429

(2005) 75.

[25] N. Weinberg, M. Kamionkowski, 2002, preprint [astro-ph/

0210134].

[26] M. White, Astrophys. J. Supp. 143 (2002) 241.

[27] M. White, C. Vale, Astropart. Phys. 22 (2004) 19.

[28] M. Zaldarriaga, R. Scoccimarro, Astrophys. J. 584 (2003)

559.