0710.3588/ms.tex
1: \documentclass[twocolumn]{emulateapj}
2: %\documentclass[preprint2]{aastex}
3: \usepackage{epsfig,psfig,amsmath,amsfonts,amssymb}
4: %\newcommand{\com}[1] {{\bf ***#1***}}
5: %\newcommand{\com}[1] {{\bf #1}}
6: \newcommand{\com}[1] {#1}
7: \newcommand{\simgt}{\,\hbox{\lower0.6ex\hbox{$\sim$}\llap{\raise0.6ex\hbox{$>$}}}\,}
8: \newcommand{\simlt}{\,\hbox{\lower0.6ex\hbox{$\sim$}\llap{\raise0.6ex\hbox{$<$}}}\,}
9: \begin{document}
10: \shorttitle{Weak Lensing: Ground {\it vs} Space}
11: \title{A Comparison of Weak Lensing Measurements\\From Ground- and Space-Based Facilities}
12: \shortauthors{M.\ M.\ Kasliwal et al.}
13: \author{Mansi M.\ Kasliwal\altaffilmark{1,2}, 
14:         Richard Massey\altaffilmark{1}, 
15:         Richard S.\ Ellis\altaffilmark{1},
16:         Satoshi Miyazaki\altaffilmark{3} \& 
17:         Jason Rhodes\altaffilmark{4,1}}
18: 
19: \altaffiltext{1}{Astronomy Department, California Institute of Technology, 105-24, Pasadena, CA 91125, 
20: USA}
21: \altaffiltext{2}{George Ellory Hale Fellow of Moore Foundation}
22: \altaffiltext{3}{National Astronomical Observatory of Japan, Mitaka, Tokyo 181-8588, Japan}
23: \altaffiltext{4}{Jet Propulsion Laboratory, California Institute of Technology, 105-24, Pasadena, CA 
24: 91125, USA}
25: 
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: %%%%%
28: 
29: \begin{abstract}
30: 
31: We assess the relative merits of weak lensing surveys, using overlapping imaging data 
32: from the ground-based Subaru telescope and the Hubble Space Telescope (HST). 
33: Our tests complement similar studies undertaken with simulated data. 
34: From observations of 230,000 matched objects in the 2 square degree COSMOS field, 
35: we identify the limit at which faint galaxy shapes can be reliably measured 
36: from the ground. Our ground-based shear catalog achieves sub-percent calibration bias compared to 
37: high resolution space-based data, for galaxies brighter than $i^{\prime}\simeq$24.5 and with 
38: half-light radii larger than $1.8\arcsec$. This selection corresponds to a surface density of 
39: 15 galaxies arcmin$^{-2}$ compared to $\sim 71$ arcmin$^{-2}$ from space. On the other hand
40: the survey speed of current ground-based facilities is much faster than that  of HST, although 
41: this gain is mitigated by the increased depth of space-based imaging desirable for tomographic
42: (3D) analyses. As an independent experiment, we also reconstruct the projected mass distribution
43: in the COSMOS field using both data sets, and compare the derived cluster catalogs with those 
44: from $X$-ray observations. The ground-based catalog achieves a reasonable degree of completeness, 
45: with minimal contamination and no detected bias, for massive clusters at redshifts $0.2<z<0.5$. 
46: The space-based data provide improved precision and a greater sensitivity to clusters of lower 
47: mass or at higher redshift.
48: \end{abstract}
49: 
50: \keywords{cosmology: observations -- gravitational lensing -- instrumentation}
51: 
52: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53: %%%%%
54: 
55: \section{Introduction}
56: \label{sec:Introduction}
57: 
58: Dark matter dominates the gravitational component of the cosmic energy density and thus provides
59: the framework for structure formation in the Universe. However, by its nature, the distribution
60: and cosmic growth are challenging to observe. The most promising probe is weak gravitational
61: lensing: analysis of the distorted shapes of ordinary galaxies behind foreground mass
62: concentrations. Several numerical techniques are now available to recover the projected mass
63: distribution from these distortions, and tests on simulated datasets are underway to verify their
64: precision \citep{step1,step2}. There is great optimism in the weak lensing community that such
65: methods will enable both the tomographic mapping of dark matter structures in time and space. This
66: will also provide a robust statistical measure of the nature of dark energy over redshifts 0$<z<$1
67: \citep{mellier99, refregier03}
68: 
69: Observational progress has been particularly dramatic. The first detections of 
70: statistical ``cosmic shear'' were only published in 2000 \citep{bacon00, kaiser00, wittman00,
71: vanwaerbeke00}. In the subsequent \com{seven} years, weak lensing surveys have measured the dark 
72: matter power spectrum \citep{brown03, heymans05, hoekstra06, sembolini06}, traced 
73: the evolution of structure \citep{bacon05, kitching06, massey07a}, enabled the construction 
74: of lensing-selected cluster catalogs \citep{miyazaki02a, wittman06, schirmer07, miyazaki07}, 
75: and non-parametrically reconstructed the total mass distribution both in clusters 
76: \citep{kneib03, clowe06, jee07} and on larger scales \citep{massey07b}. As a result, 
77: weak lensing has been identified as the most promising route to understanding the nature 
78: of dark energy by the ESA-ESO Working Group on Fundamental 
79: Cosmology\footnote{\tt http://www.stecf.org/coordination/esa\_eso/cosmology.php},
80: joint NSF-NASA-DOE Astronomy and Astrophysics Advisory 
81: Committee\footnote{\tt http://www.nsf.gov/mps/ast/aaac.jsp}, 
82: and NSF-DOE High Energy Physics Advisory Panel 
83: Dark Energy Task Force\footnote{\tt http://www.nsf.gov/mps/ast/detf.jsp}. 
84: 
85: The primary signal of any weak lensing analysis is the statistically coherent 
86: distortion of background galaxies along adjacent lines of sight. The main sources of statistical 
87: noise are the finite density of galaxies that can be sufficiently well-detected and resolved for 
88: accurate shape measurement, plus their intrinsic morphologies. The density of resolved 
89: galaxies also governs the angular resolution and fidelity of a reconstructed mass map
90: which, in turn, determines the limiting halo mass that can be detected. On the other hand, 
91: statistical analyses of the dark matter power spectrum are less concerned with individual halos 
92: but require panoramic fields to counter the effects of cosmic (sample) variance. Minimizing 
93: statistical errors in such an analysis, within a finite survey lifetime, requires an optimal balance 
94: between area and depth.
95: 
96: A key debate in the development of future weak lensing experiments concerns the relative merits of
97: ground- versus space-based platforms. 
98: Ambitious surveys now being planned with dedicated, ground-based facilities (eg VST-KIDS, DES, 
99: Pan-STARRS, LSST). These are driven by technological progress including panoramic cameras with
100: small  optical distortions, highly sensitive imaging detectors, and (in the case of Pan-STARRS) 
101: on-chip active correction to reduce the width of the point spread function (PSF). Future surveys 
102: spanning significant fractions of the celestial sphere are envisaged, promising tight constraints 
103: on the  cosmological parameters.
104: 
105: However, measurements with current ground-based facilities are limited by the size and temporal
106: variations of the PSF. There is concern in many quarters that wide-field facilities operating in
107: space (e.g.\ DUNE, SNAP, JDEM) will ultimately be required to achieve the precision required
108: (particularly) to distinguish between various models of dark energy. Space-based facilities will
109: be more costly but will likely offer increased depth, better photometric performance and a stable
110: PSF. The key issue in gauging their merits is not statistical error, but the extent to which
111: potential biases in ground-based data may act as a ``systematic floor'' to prevent complete
112: exploitation.
113: 
114: Some valuable answers can be obtained by comparing simulated ground and space-based images, 
115: \citep{wittman05,lampton06} and the Shear TEsting Programme \citep[STEP:][]{step1,step2}.
116: However, the input parameters used to generate the simulated data may not be realistic or address
117: all the instrumental idiosyncrasies. Of particular concern are the stability and vagaries of the
118: PSF. No simulations have  yet adequately addressed this point -- which may, ultimately, be 
119: the limiting problem for ground-based data. It is often argued that future facilities will be 
120: carefully designed to mitigate any limitations realized with current observational facilities. 
121: While progress can no doubt be expected, both on the ground and in space, we believe many lessons
122:  can be learned from extant data and hardware with proven engineering pedigree.
123: 
124: In this paper, we present the first direct comparison of weak lensing analysis  {\it for the same
125: sky field} using ground and space-based data. Deep, panoramic imaging has been obtained for the
126: 1.64 deg$^2$ COSMOS field \citep{scoville07a} by both the Advanced Camera for  Surveys ({\it ACS})
127: on board the Hubble Space Telescope (HST) \citep{scoville07b} and the {\it Suprime-Cam} imager at
128: the prime focus of the Subaru 8.2m telescope \citep{taniguchi07}. In both cases, the 
129: entire field was covered by mosaicing many
130: independent exposures.  The SuPrimeCam instrument was constructed with weak lensing analysis
131: particularly in mind, and currently provides the best image performance available from any
132: ground-based telescope, in terms of optical distortions over a large field. A comparison of these
133: datasets should therefore provide a realistic and valuable assessment of the relative performance
134: of state-of-the-art imagers on the ground and in space.
135: 
136: The paper is organized as follows. In \S\ref{sec:theory}, we briefly review the relevant theory. In
137: \S\ref{sec:data}, we describe the two data sets, data reduction pipelines and weak lensing
138: analyses. We then present the results. In \S\ref{sec:analyses}, we compare shear measures on a
139: galaxy-by-galaxy  basis to determine the optimum depth at which the ground-based data matches the 
140: performance of the (deeper) space  based data. This permits us to determine the relative survey
141: speeds of Subaru and HST for high precision cosmic shear experiments. In \S\ref{sec:Maps}, we 
142: construct maps of the mass distribution, treating the Subaru and HST maps as independent probes of
143: the same field, and contrast these against X-ray data. This permits us to evaluate the completeness
144: and reliability of a lensing-selected halo catalog, and evaluate the precision of their inferred
145: masses as a function of redshift. In \S\ref{sec:conc}, we summarize our results and discuss their
146: wider implications for future missions.
147: 
148: \section{Review of Weak Lensing Theory}\label{sec:theory}
149: 
150: Gravitational lensing by foreground mass structures distorts an image plane of distant galaxies
151: $I(\mathbf{x})$ via a coordinate transformation
152: \begin{equation}
153: \label{eqn:psi_def_theory}
154: {\mathcal A}_{ij} = \delta_{ij} + \frac{\partial (\delta x_i)}{\partial x_j}
155:   = \left( \begin{array}{cc}
156: 1 - \kappa - \gamma_{1} & \gamma_{2}\\
157: \gamma_{2} & 1 - \kappa + \gamma_{1} \\
158: \end{array} \right),
159: \end{equation}
160: \noindent where $\delta x_i(\mathbf{x})$ is the deflection angle of the light rays. The {\em
161: convergence}
162: \begin{equation}
163: \label{eqn:kappa}
164: \kappa(\mathbf{x})=\frac{4\pi G}{2c^2}\int g(z) \rho(\mathbf{x},z) {\rm d}z ~,
165: \end{equation}
166: \noindent describes overall dilations and contractions. It is proportional to the 
167: total mass density $\rho$ projected along a line of sight, where the {\em lensing sensitivity function}
168: \begin{equation}
169: g(z)=\frac{2D_LD_{LS}}{D_S}
170: \label{eqn:sensitivitydefn}
171: \end{equation}
172: \noindent reflects the efficiency of foreground gravitational lenses at different redshifts --
173: containing a ratio of the angular diameter distance to a lens, the background source, and between the 
174: two. 
175: This can be more simply written as
176: \begin{equation}
177: \kappa ~ \equiv ~ \frac{1}{2} \left( \frac{\partial^2\Psi}{\partial x^2} + \frac{\partial^2\Psi}{\partial y^2}\right) 
178: ~,
179: \label{eqn:convergencedefn}
180: \end{equation}
181: \noindent in terms of a 2D, projected version $\Psi(\mathbf{x})$ of the Newtonian gravitational potential.
182: Two components of {\em shear}
183: %\begin{eqnarray}
184: %\gamma_1 & \equiv & \frac{1}{2} \left(\frac{\partial^2\Psi}{\partial x^2} - \frac{\partial^2\Psi}{\partial y^2}\right)  \\
185: %\gamma_2 & \equiv & \frac{\partial^2\Psi}{\partial x\partial y} ~,
186: %\label{eqn:sheardefn}
187: %\end{eqnarray}
188: \begin{equation}
189: \big\{\gamma_1,~\gamma_2\big\}
190:  \equiv \Bigg\{ 
191:  \frac{1}{2} \left(\frac{\partial^2\Psi}{\partial x^2} - \frac{\partial^2\Psi}{\partial y^2}\right),~
192:         \frac{\partial^2\Psi}{\partial x\partial y}
193:         \Bigg\}~,
194: \label{eqn:sheardefn}
195: \end{equation}
196: \noindent describe stretches and compressions along (at $45^{\circ}$ from) the $x$-axis.
197: 
198: The observed shapes of background galaxies can be described by combination of their
199: Gaussian-weighted quadrupole moments
200: \begin{equation}
201: \label{eqn:ksbd}
202: d\equiv\frac{\iint I({\mathbf x}) ~W({\mathbf x})~r^2~{\mathrm d}^2{\mathbf x}}
203:              {\iint I({\mathbf x}) ~W({\mathbf x})~{\mathrm d}^2{\mathbf x}} ~,
204: \end{equation}
205: \begin{equation}
206: \label{eqn:ksbe}
207: \big\{\varepsilon_1,~\varepsilon_2\big\}
208:   \equiv\frac{\iint I({\mathbf x}) ~W({\mathbf x})~r^2\big(\cos{(2\theta)},~\sin{(2\theta)}\big)~{\mathrm d}^2
209: {\mathbf x}}
210:              {\iint I({\mathbf x}) ~W({\mathbf x})~r^2~{\mathrm d}^2{\mathbf x}} ~,
211: \end{equation}
212: \noindent where 
213: \begin{equation}
214: \label{eqn:ksbweight}
215: W({\mathbf x})=e^{-r^2/2r_g^2} ~.
216: \end{equation}
217: Although $\kappa$ is generally the desired quantity, and could be obtained in principle from
218: measurements of galaxy sizes \eqref{eqn:ksbd} or fluxes, this has proved difficult in practice,
219: because expectations for these quantities prior to lensing are unknown. On the other hand, while
220: galaxies have a natural dispersion of intrinsic ellipticities \eqref{eqn:ksbe}, they are (almost)
221: uncorrelated with each other in the absence of lensing, i.e.\ $\langle\varepsilon_i\rangle=0$. Any
222: correlation between the {\em observed} ellipticities of galaxies seen along adjacent lines of sight
223: arises because their light has traversed similar intervening large-scale structure
224: $\rho(\mathbf{x},z)$. In practice, corrections to measured ellipticities also need to be made for
225: the smearing of galaxies by the PSF, and for the differing susceptibilities of some galaxy
226: morphologies to an input shear. For more details of this procedure, see e.g.\ \citep{ksb}.
227: 
228: The observed shear can finally be transformed into convergence through their close relation in
229: Fourier space
230: \begin{equation}
231: \label{eqn:gammatokappa}
232: \tilde{\kappa}=
233: \frac{(\ell_1^2-\ell_2^2)\tilde{\gamma_1}+(2\ell_1 \ell_2)\tilde{\gamma_2}}
234:      {(l_1^2+l_2^2)}
235: \end{equation}
236: \noindent \citep{kaiser93}.  This is typically some amount, to reduce noise. 
237: Furthermore, like any scalar quantity extracted from a vector field, a
238: convergence signal can also be split into two independent components, 
239: $\kappa = \kappa^{E} + i~\kappa^{B}$ \citep{king01}.
240: The grad-like ``$E$-mode'' is the signal produced by weak lensing.
241: The curl-like ``$B$-mode'' is not produced by physical processes \citep[except at very low levels,
242: as described by][]{schneider02}, and therefore ought to be consistent with zero in the absence of
243: systematics. Usefully, it contains the same noise properties as the $E$-mode signal -- so it acts
244: as an independent realization of noise in the field, and any significant deviations from zero
245: alert to the presence of residual systematics (such as imperfect correction for the PSF).
246: 
247: \section{Observations and Data Reduction}
248: \label{sec:data}
249: 
250: \subsection{The COSMOS Data Sets}
251: 
252: Our data all cover the COSMOS survey field, a 1.64 deg$^2$ contiguous square, centered at
253: 10:00:28.6, +02:12:21.0 (J2000) \citep{scoville07a}. The ground-based imaging was obtained in eleven 
254: mosaiced pointings
255: of the {\it Suprime-Cam} camera at the prime focus of the Subaru telescope on Mauna Kea
256: \citep{miyazaki02b}. These were taken on the 18th and 21st of February 2004, nights selected for
257: their excellent observing conditions: the mean seeing was $0.54\arcsec\pm0.03\arcsec$. The field
258: constitutes part of a larger weak lensing survey discussed, along with full details of the primary data
259: reduction pipeline, \com{in \citep[][Green et al.\ {\it in prep.}]{miyazaki07}. In fact, the relevant field 
260: in that survey covered a slightly larger area than the COSMOS field. 
261: The Subaru imaging was truncated when matching galaxy catalogs and was truncated {\it after} making 
262: convergence maps, to avoid edge effects associated with the
263: Fourier transform operations in equation~\eqref{eqn:gammatokappa}.}
264: 
265: Our comparison is made possible by the unique availability of deep, panoramic space-based imaging of
266: the COSMOS field \citep{scoville07b}. During HST cycles 13 and 14, 577 slightly overlapping
267: pointings were obtained from the {\it Advanced Camera for Surveys} ({\it ACS}) on board the Hubble
268: Space Telescope. Four dithered exposures at each pointing were stacked using the {\sc drizzle}
269: algorithm \citep{drizzle} to improve the native pixel scale of $0.05\arcsec$ and recover a final
270: pixel scale of $0.03\arcsec$. Full details of the primary data reduction pipeline for the HST images
271: are given in \citep{koekemoer07}.
272: 
273: It is important to emphasize that both the {\it ACS} and {\it Suprime-Cam} data exhibit
274: idiosyncrasies that present significant challenges for weak lensing analysis. For example, the
275: atmospheric seeing varied during the two nights over which the Subaru data were obtained; and the
276: distortions of the telescope's primary mirror under a gravity load were only passively corrected via
277: a look-up table as it followed the field. In a future ground-based experiment, such as LSST or
278: Pan-STARRS, seeing  variations could be normalized over a survey  by stacking a very large number of
279: short, independent  exposures taken over a long time period. Dome seeing could likewise be improved
280: with future  technologies. And while the telescope superstructure is particularly rigid at Subaru,
281: active correction  of the mirror support could undoubtedly improve future designs. Equivalently, the
282: sky background  seen from HST is affected by Earthshine that depends on the telescope pointing
283: \citep{leauthaud07}. The HST PSF also varies over time due to thermal fluctuations during each
284: low-Earth orbit  \citep{jee07, rhodes07}. Finally, the charge transfer efficiency of the {\it ACS}
285: CCD detectors had been  significantly degraded by high energy particles by the time the COSMOS data
286: were obtained, and worsened during the observing window \citep{rhodes07}. None of these problems
287: are inherent to all space-based observations: future missions might minimize or eliminate all three
288: effects by adopting a regular observing pattern, orbiting the Lagrange point L2, and using
289: radiation-hardened CCDs. However, the weak lensing analysis of existing space-based data is indeed
290: compromised by the extent to which such hardware variations can be modeled. In this sense, our
291: comparison is actually more informative than one based on simulated data that reproduces only
292: idealized and mean instrumental characteristics. 
293: 
294: The relevant characteristics of the two data sets are summarized in Table~\ref{tab:observations},
295: including limiting depths for a point source at $5\sigma$, in a $3\arcsec$ aperture from the ground 
296: and a $0.15\arcsec$ aperture from space \citep{capak07}. 
297: In addition to these images, the COSMOS field has been observed across all wavelengths from radio to
298: $X$-rays. Of particular relevance here are ($i$) deep $X$-ray observations by XMM
299: \citep{hasinger07}, which can be used to locate massive structures via thermal emission from hot
300: gas; and ($ii$) multicolor optical and near-IR imaging campaigns from the Subaru, Canada France
301: Hawaii, Cerro Tololo and Kitt Peak telescopes, which provide 15 additional bands and photometric
302: redshifts \citep{capak07, mobasher07}. The photometric redshift estimation code uses Bayesian priors
303: based on an adopted luminosity function, and includes reddening based on both Galactic and Calzetti
304: extinction laws. The results were calibrated using 868 galaxies in the field brighter than
305: $i^\prime=24$ and with spectroscopic redshifts. For galaxies closer than $z=1.2$, the rms scatter in
306: $(z_{phot} - z_{spec})\, /\, (1 + z_{spec})$ is 0.031.
307: 
308: \begin{deluxetable}{lcc}
309: \tabletypesize{\footnotesize}
310: %\tablenum{1}
311: \footnotesize
312: %\tabletypesize{\scriptsize}
313: \setlength{\tabcolsep}{0.1in}
314: \tablecaption{Survey Characteristics}
315: \tablewidth{0pc}
316: \tablehead{& \colhead{Ground} & \colhead {Space}}
317: \startdata
318: Instrument                  & Subaru/{\it Suprime-Cam} & HST/{\it ACS}   \\
319: Primary aperture            & 8.2m                     & 2.4m            \\
320: Exposure time               & 5 $\times$ 360s          & 4 $\times$ 507s \\ 
321: Total survey time           & 5 hours                  & 325 hours       \\ 
322: Filter                      & $i^\prime$               & F814W           \\
323: Limiting AB magnitude       & 26.2                     & 26.6            \\
324: %~~(5$\sigma$, point source) & \citep{capak07}          & 24.9 in $3\arcsec$ ap. \\
325: Field of View               & 2.14 deg$^2$             & 1.67 deg$^2$    \\
326: Pixel Scale                 & 0.202$\arcsec$           & 0.03$\arcsec$   \\
327: Point Spread Function       & 0.68$\arcsec$            & 0.12$\arcsec$   \\
328: \enddata
329: \label{tab:observations}
330: \end{deluxetable}
331: 
332: \subsection{Object Detection}
333: 
334: Objects were detected in the Subaru images using {\sc hfindpeaks} from the {\sc imcat}
335: package\footnote{Nick Kasiser's {\tt imcat} software package is available from  {\tt
336: http://www.ifa.hawaii.edu/$\sim$kaiser/imcat/}}. This finds the centroid and scale size $r_g$ that
337: maximizes the peak S/N of the image after smoothing with a Gaussian. The code also returns the
338: half-light radius, $r_h$, of each galaxy. Galaxies were initially detected to magnitudes fainter
339: than those for which it is possible to accurately measure shapes. To reduce noise in the final
340: analysis, weights were given to each galaxy, and galaxies with a detection S/N$<14$ were removed
341: from the catalog altogether. The resulting surface density is $n_\mathrm{gal}=42$ galaxies arcmin$^{-2}$.
342: 
343: %\com{Satoshi to add details as required.}
344: 
345: Objects were detected in the {\it ACS} images using SExtractor \citep{BertinArnouts} in a dual
346: Hot/Cold configuration \citep{leauthaud07}, designed to identify both large and small objects while
347: avoiding fragmentation of the former, or merging  of the latter. The SExtractor centroids were then
348: improved, and the best-fitting scale size was selected, via an iterative process during shape
349: measurement. Galaxies smaller than $d=0.11\arcsec$ or fainter than $S/N\approx20$ were removed 
350: from
351: the catalog, and a weighting scheme was applied to faint galaxies as a function of their detection
352: S/N \citep{leauthaud07}. Note that an absolute calibration of the S/N was difficult to determine in
353: practice, because flux in adjacent pixels becomes correlated during DRIZZLE. The S/N cut corresponds
354: approximately to a limiting magnitude $F814W(AB)<26.5$ for a point source. The resulting surface
355: density is $n_\mathrm{gal}=71$ galaxies arcmin$^{-2}$, with a median redshift $z_\mathrm{med}=1.2$.
356: 
357: 
358: \subsection{Shear Measurement}
359: \label{sec:Shear}
360: 
361: Because the image characteristics of the two data sets are quite different, we adopted separate
362: methods to measure galaxy shapes, remove PSF effects, and ultimately obtain the weak lensing shear
363: signal. Each of these methods has been optimized for the respective data sets, so our comparison
364: will necessarily incorporate the limitations of each pipeline. We believe this is in the spirit of a
365: fair comparison of ground versus space. To minimize any differences arising entirely from the
366: algorithms themselves however, we have intentionally adopted related methods from the same
367: generation of software development and codes that have been well-tested.  Although newer shear
368: measurement methods \citep{k2k, im2shape, dahle02, shapelets2, bj02, hirata03, kuijken06,
369: shapelets4, reiko} may offer improved performance, none has yet been sufficiently tested  across
370: both observing regimes.
371: 
372: The Subaru images were analyzed with the \citep{hamana03} implementation of the widely used
373: \citep[][hereafter KSB]{ksb} shear measurement method. This particular implementation is a
374: derivative of the ``LV'' pipeline tested in the Shear TEsting Program
375: \citep{step1,step2}.
376: 
377: The HST images were analyzed with the \citep[][hereafter RRG]{rrg} shear measurement method. This is
378: a perturbation of the KSB method for space-based data. It calculates the same quadrupole moments,
379: but corrects them individually for the effects of convolution with the PSF, and only in the final
380: stage takes the ratio~\eqref{eqn:ksbe}. This is necessary because the small and cuspy
381: diffraction-limited PSFs otherwise introduce divisions by very small (and noisy) numbers. RRG been
382: applied to HST {\it WFPC2} \citep{rrg}, {\it STIS} \citep{rhodes04}, and {\it ACS} data
383: \citep{massey07a}. The {\it ACS} pipeline was thoroughly tested on simulated images during the
384: creation of the COSMOS catalog \citep{leauthaud07}, and also for a continuation of STEP using
385: simulated space-based images. 
386: 
387: %In the following analyses it is important to distinguish between comparisons
388: %where HST and Subaru data are compared independently  (such as in 
389: %the convergence maps), and those where matched catalogs are 
390: %necessary (for comparing shear on a galaxy-by-galaxy basis). In the
391: %case of matched catalogs, we have experimented with various procedures,
392: %typically demanding spatial co-location within 1$\arcsec$. 
393: 
394: 
395: 
396: 
397: 
398: 
399: 
400: 
401: \section{Statistical Applications}
402: \label{sec:analyses}
403: 
404: \subsection{Shear-shear comparisons}\label{sec:shearbyshear}
405: 
406: We shall now compare the global properties of our ground- and space-based shear catalogs, to  
407: determine the depth (and galaxy surface density) at which reliable shear measurement is possible
408: from the ground. This will be relevant for many statistical applications, including measurements of
409: the angular shear-shear correlation function that are typically used to constrain cosmological
410: parameters. In such analyses, where statistical noise is reduced by averaging over many lines of
411: sight, the key issue is the reliability and level of residual systematics in the shear measurement.
412: 
413: We asses the performance of the ground-based shear measurements against those from {\it the same
414: galaxies} in space-based data, making the necessary but reasonable assumption that the shapes are
415: much more reliable when measured from the much higher resolution images with a smaller PSF. Such a
416: comparison is clearly only possible for the subset of galaxies contained in both catalogs. The two
417: quantities of interest will be linearity in the comparison (the slope of the shear-shear
418: comparison is equivalent to the STEP ``calibration bias'' parameter $m$) and the scatter (which
419: represents the combined shear measurement noise from both HST and Subaru, plus any systematic
420: effects).
421: 
422: We match galaxies whose positions agree to within 1$\arcsec$, and produce a common catalog 
423: containing $n_{gal}=32$ galaxies arcmin$^{-2}$. Many objects in the Subaru galaxy catalog without
424: matched counterparts in HST galaxy catalog have half-light radii on the limits of seeing and are likely to have been
425: revealed as stars by the higher resolution data; in any case, the omitted galaxies had below-average weights in
426: the Subaru catalog. The remaining unmatched objects are a combination of noisy/skewed galaxies with
427: offset centroids, or galaxies that lie in regions of the HST images masked because of scattered
428: light from nearby bright stars. For the following tests tests, we shall ignore the weights on remaining
429: galaxies, and treat all objects equally.
430: 
431: Figure~\ref{fig1} shows a comparison of the shear signal for the matched galaxies. Since errors  are
432: present in both axes, to calculate the best-fit linear relationship we adopt a least squares  
433: method that minimizes the perpendicular distance to the best fit line (instead of one that assumes 
434: one variable is error-free). Ideal shear measurement from both instruments would yield  a best-fit
435: slope of unity. \com{There will inevitably be a small amount of scatter, because the weight function~\eqref{eqn:ksbweight}
436: is not necessarily the same size $r_g$ in the ground- and space-based analyses.
437: In practice, we find a best-fit slope of 0.87, indicating that the shears have been
438: underestimated from the ground. The measurement noise is also
439: problematic, with $\sigma_\gamma=0.16$ per component (perpendicular to the best-fit line; 
440: note that this does not include intrinsic source ellipticity variance because the galaxies are matched) and
441: a skewed, non-Gaussian distribution of outlying shear estimates} that would render a cosmic shear analysis less stable. 
442: 
443: \begin{figure}
444: \centerline{\epsfig{file=f1.eps,width=3.2in}}
445: \caption[]{Comparison of shear measured from galaxies seen in both Subaru and Hubble Space Telescope images.
446: The greyscale shows the number of galaxies with different shear measurements. The outer contour
447: includes 90\% of the galaxies, and successive inner ones include 10\% fewer. 
448: The solid line is the least squares linear relation.
449: Its slope of 0.87 indicates that shears have been underestimated in the ground-based analysis, 
450: or that the catalog is still partially contaminated by stellar sources.
451: This value is insensitive within 0.01 to the reintroduction of galaxy weights.
452: Furthermore, the non-Gaussian
453: wings of the scatter extend well beyond the rms error of 0.16, shown as dashed lines.}
454: \label{fig1}
455: \end{figure}
456: 
457: The overall performance in figure~\ref{fig1} is a superposition of good shears from bright and (in
458: particular) large galaxies, plus smaller objects that cause most of the bias and scatter. Indeed,
459: systematic errors could be completely eliminated by using only the very largest galaxies. However,
460: the statistical noise in a cosmic shear analysis of shear-shear correlation functions scales as
461: $\sigma_\gamma/\sqrt{n_\mathrm{gal}}$. An optimal strategy for any particular ground-based survey
462: will involve catalog cuts requiring a trade-off between systematic and statistical errors. However,
463: the optimal cuts will vary as a function of survey area and depth. To produce a result of general
464: interest, we therefore show in Figures~\ref{fig:calibration} and \ref{fig:noise}, the resulting
465: calibration bias, scatter and galaxy density for a range of possible cuts in galaxy size and
466: magnitude. 
467: 
468: \begin{figure}
469: \centerline{\psfig{file=f2.eps,width=3.2in,angle=90}}
470: \caption[]{Relative calibration between shear measurements from galaxies in Subaru and Hubble Space Telescope data, 
471: for galaxies of different sizes and magnitudes. 
472: The contours show deviations from a slope of unity in figures~\ref{fig1} and
473: \ref{finalcuts}, which would have been ideal. For faint galaxies, these deviations tended to be an
474: underestimation in the Subaru pipeline relative to HST. There is some evidence that shears are
475: overestimated in large, bright
476: galaxies, although the small number of these objects means that the extrapolation is less certain.
477: The calibration biases are calculated locally, for
478: galaxies only in a given cell of \{size,magnitude\} space.
479: On the other hand, the grey-scale shows the cumulative number density of galaxies $n_{gal}$  that 
480: would remain in a ground-based catalog, were cuts to be applied at the local values 
481: (i.e.\ including all larger and brighter galaxies).}
482: \label{fig:calibration}
483: \end{figure}
484: 
485: \begin{figure}
486: \centerline{\psfig{file=f3.eps,width=3.2in,angle=90}}
487: \caption[]{Combined noise from shear measurements of galaxies matched in catalogs from Subaru
488: and Hubble Space Telescope data. The contours show $\sigma_\gamma$ as a function of galaxy size and
489: magnitude. As in figure~\ref{fig:calibration}, these are calculated only for galaxies with that
490: particular size and magnitude. The contours close at the top merely because there are very few
491: large, faint galaxies, so the rms scatter increases. 
492: The grey scale again shows the total number density of available
493: galaxies.}
494: \label{fig:noise}
495: \end{figure}
496: 
497: A simple result emerges from Figure~\ref{fig:calibration}. It is noticeable from the horizontal and
498: vertical contours that size and magnitude cuts seem to neatly parametrize independent sources of
499: error. Using existing shape measurement methodology, shear can be measured from galaxies brighter
500: than $i^\prime=24.5$ and larger than $r_h=1.8$, with measurement noise $\sigma_{\gamma}\simeq
501: 0.03$ and a calibration bias less than 3\% \com{(and only 1\% with galaxy weighting)},
502: which is acceptable for competitive constraints from future surveys \citep{refregier04}. This leaves
503: a surface density of $n_\mathrm{gal}=15$ galaxies arcmin$^2$ from the ground, with a median redshift
504: of $z_{\rm med}=$0.8. The comparison with space-based data for these cuts is shown in
505: figure~\ref{finalcuts}. 
506: 
507: \begin{figure}
508: \centerline{\epsfig{file=f4.eps,width=3.2in}}
509: \caption[]{As for figure \ref{fig1}, but for the subset of galaxies brighter than $i^\prime=24.5$
510: and larger than $r_h=1.8\arcsec$.
511: The total least squares slope is 0.97, implying an almost unbiased recovery of the shear signal
512: from Subaru, and the data is better and more symmetrically enclosed within the rms scatter of 0.11.}
513: \label{finalcuts}
514: \end{figure}
515: 
516: Note that we have not been able to test the reliability of space-based shear measurements using this
517: method, nor even considered the population of small galaxies resolved only from space. Without data
518: even better than {\it ACS} imaging to compare to, we resort to simulations. The full RRG pipeline
519: was calibrated against simulated images by \citep{leauthaud07}. However, of the 71 galaxies
520: arcmin$^{-2}$ successfully in real {\it ACS} images, only the brightest 40 could be used by 
521: \citep{massey07a} to minimize the $B$-mode signal and overcome problems of CCD charge transfer
522: inefficiency (CTI). This limitation clearly needs to be overcome: perhaps via a CTI correction
523: algorithm like that developed for {\it STIS} by \citep{cte}, and radiation-hardened detectors in
524: future telescopes.
525: 
526: \subsection{Survey speed}\label{sec:surveyspeed}
527: 
528: Our ground versus space shear comparisons have important implications when it comes to
529: considering the optimal  approach for measuring shear for cosmological applications.
530: %In this section we consider the question of balancing {\it depth} and {\it area}.
531: %Although this 
532: %combination (the {\it \'etendue}) can be expected to increase with both future ground and 
533: %space-based instruments, we will base our discussion on the current imaging depth
534: %and fields of view of HST and Subaru.
535: Although the {\it \'etendue} of instruments can be expected to increase both on the ground and in space,
536: we will base our discussion on the imaging depth and fields of view of our HST and Subaru surveys.
537: 
538: An important criterion is what can be accomplished in a given amount of observing time; the HST and
539: Subaru requirements for our comparison are summarized in Table~\ref{tab:observations}).  HST 
540: overheads
541: approximately tripled the on-source exposure time, and, at Subaru,  high quality imaging was
542: secured during what might be considered a fortuitous observing window. Noting that \citep{bacon01}
543: found images with seeing worse than $\sim 0.8\arcsec$ of little use for weak lensing analysis, 
544: coupled with observational visibility, it
545: seems reasonable to incorporate a factor of at least four inefficiency for a generic survey: even for a
546: superb ground-based facility such as Subaru, on an excellent site such as Mauna Kea\footnote{{\tt
547: http://www.cfht.hawaii.edu/Instruments/Imaging/Megacam/ observingstats.html}. Future surveys such as
548: Pan-STARRS and LSST, which plan to co-add  many short exposures with independent PSFs, may 
549: achieve
550: near-uniform image quality  by rejectng a certain fraction of exposures. But the relevant figure of
551: merit is still the fraction of time spent with seeing better than  $0.8\arcsec$.}.  Based on its
552: superior field of view, Subaru is then $\sim$ 24 times  faster than HST in  useful mapping  speed. As,
553: to  first order, the signal to noise in statistical analyses increases as $\sqrt{n_\mathrm{gal}}$,  for
554: a fixed survey lifetime, this  corresponds to a $\sim5$-fold  improvement in signal to noise. 
555: 
556: This simplistic analysis is of course mitigated by the higher resolution available from space. We next
557: insert the gain in surface density, viz  71 galaxies arcmin$^{-2}$ resolved in our space-based imaging 
558: c.f.\ 15 arcmin$^{-2}$ from the ground. We will assume that the additional, small galaxies have a similar
559: distribution of intrinsic 
560: ellipticities as the larger ones \citep[c.f.][]{snap2, leauthaud07} and that the measurement noise on an 
561: average survey galaxy is constant (since the size distribution of resolved galaxies compared to the PSF 
562: size is roughly independent of the PSF size). Incorporating this increased background surface density,
563: the ground-based gain per unit time drops to only a factor of 2.3. 
564: 
565: Equally important to the increased density of galaxies are their higher redshifts. Distant galaxies are 
566: more
567: sensitive to
568: low-redshift lenses, and sensitive to more total lenses. The shear signal grows proportionally to the
569: median source redshift $z_{\rm med}^{0.6-0.8}$ \citep{jainseljak97}. With the redshift distributions
570: for galaxies shown in figure~\ref{fig:zsensitivity}, the total gain in signal to noise for a 2D weak
571: lensing survey conducted from Subaru over one conducted with HST is only about 1.7.
572: 
573: 
574: Perhaps the most important advantage of space is the increased redshift range of resolved galaxies. 
575: This better enables their stratification into redshift bins for tomographic (3D) analyses. Deep infrared 
576: imaging, needed for accurate photometric redshifts, is also likely to remain the province of space-based
577: observatories.  Tomographic techniques can tighten the constraints on cosmological parameters $
578: \Omega_{\mathrm
579: M}$ 
580: and $\sigma_8$ by a factor of at least three \citep{massey07a} and potentially as much as 
581: five \citep{heavens06}. 
582: Further advantages of these techniques includes the elimination of unwanted signal from 
583: adjacent galaxies' intrinsically-correlated shapes \citep{king03, heymans03}. While wide-field ground-based
584: instruments may therefore yield significant improvements for  Dark Energy Task Force ``Stage 3'' 
585: surveys, 
586: advanced analysis techniques for ``Stage 4'' surveys will realistically be possible only with space-based
587: facilities.
588: These will bring new scientific opportunities, cross-checks for systematics, and greater efficiency.
589: 
590: 
591: \begin{figure}
592: \centerline{\psfig{file=f5.eps,width=2.in,angle=90}}
593: \caption[]{The dotted lines show the redshift distribution of source galaxies from 
594: Hubble Space Telescope (black) and Subaru (red) imaging surveys, for the catalog cuts used in figure~\ref{finalcuts}.
595: The jagged lines show the measured photo-$z$s, and the smooth curves assume a simple parametric 
596: form for the background galaxy redshift distribution from \citep{smailzdist}, 
597: with $\alpha=2$, $\beta=1.5$, $z_{\rm med}=$0.8 or 1.2, and an overall
598: normalization to reproduce the observed number density of galaxies.
599: The solid lines show the corresponding lensing sensitivity functions calculated from the analytic curves.
600: Thees lie always in front of the
601: source galaxies but are notably higher for a space-based survey, particularly at redshifts greater
602: than 0.5. Extending this redshift coverage is crucial for 3D tomographic analysis techniques
603: to measure the growth of structure.}
604: \label{fig:zsensitivity}
605: \end{figure}
606: 
607: 
608: 
609: 
610: 
611: 
612: 
613: 
614: 
615: 
616: 
617: 
618: 
619: 
620: \section{Mass Maps and Halo Detection}
621: \label{sec:Maps}
622: 
623: We now investigate the reconstruction of maps of the mass distribution (figures~\ref{emode}  and
624: \ref{bmode}), and the detection of individual mass peaks. The mass and redshift distribution
625: $N(M,z)$ of several thousand lensing-selected clusters could be used to constrain cosmological
626: models \citep[][Green et al.\ {\it in prep.}]{hamana03, wang04}. Additionally, the physical properties of the 
627: dark matter
628: particles can be investigated by comparing the detailed distribution of dark matter with that of
629: baryons \citep{clowe06, dmring}. The key issues will be the angular resolution of reconstructed mass
630: maps, as well as the mass and redshift range in which halos can be successfully detected. We treat
631: this as an independent experiment from the previous section, beginning the comparison of ground- and
632: space-based data afresh. In particular, we do {\it not} cut the Subaru data  to the shallower depth
633: discussed in \S\ref{sec:shearbyshear}, to eliminate the last few systematic biases. 
634: %We therefore relax the galaxy size and magnitude cuts. 
635: The intent is  not to align our two comparisons but rather to optimize each analysis as an 
636: independent experiment -- as would be the case if either were being undertaken as a self-contained 
637: survey.
638: 
639: Unfortunately, even with the unprecedented investment of HST time for the 
640: COSMOS survey, we can expect the number of lensing-detected structures in this finite field 
641: to be modest. At the Subaru depth, a surface
642: density of $\sim$5 halos deg$^{-2}$  \citep{miyazaki07} implies only 
643: around eight halos are likely to be found in the COSMOS field. 
644: Thus we recognize in this comparison that the statistical significance 
645: of our results will be quite limited.
646: 
647: 
648: \begin{figure*}
649: %\centerline{\epsfig{file=f4.eps,width=5.0in, height=7.0in,angle=270}}
650: \centerline{\epsfig{file=f6a.eps,width=3.5in,angle=0}\epsfig{file=f6b.eps,width=3.5in,angle=0}}
651: \caption[]{Convergence $E$-mode maps from the Subaru (left) and
652: Hubble Space Telescope (right), \com{after smoothing by a $1\arcmin$ Gaussian kernel.
653: The data presented in the left panel are identical to that in figure 13 of \citep{miyazaki07}, except that the
654: field has been slightly truncated to match the right panel.
655: Convergence is proportional to the total projected mass along a line of sight, 
656: modulated by the
657: lensing sensitivity function \eqref{eqn:sensitivitydefn} plotted in figure~\ref{fig:zsensitivity}. 
658: Contours are drawn at detection significances of $3\sigma$, $4\sigma$ and $5\sigma$, with dashed lines for underdensities.
659: Clusters A, B, C and D are detected in both maps.
660: Other peaks E-L are only detected in one of the two.
661: White enclosing circles denote clusters deemed ``secure'' by the rigorous standards of \citep{miyazaki07}, and cyan circles
662: denote ``unsecure'' clusters. The size of the circles shows the size of the smoothing kernel that maximises detection significance,
663: enlarged by a factor of 2 for clarity. 
664: }}
665: \label{emode}
666: \end{figure*}
667: 
668: \begin{figure*}
669: \centerline{\epsfig{file=f7a.eps,width=3.5in,angle=0}\epsfig{file=f7b.eps,width=3.5in,angle=0}}
670: \caption[]{Convergence $B$-mode maps from Subaru (left) and Hubble Space 
671: Telescope (right) data. This is not
672: produced by physical gravitational lensing, so deviations from zero include a combination
673: of spurious effects from e.g.\ imperfect PSF correction, plus a realization of statistical noise.
674: \com{The smoothing scales, color ramps and contours are identical to those in figure~\ref{emode}.}}
675: \label{bmode}
676: \end{figure*}
677: 
678: 
679: 
680: 
681: 
682: 
683: 
684: 
685: 
686: 
687: \subsection{Residual Systematics}
688: 
689: First, we consider the $B$-mode signal. As discussed in \S\ref{sec:theory}, the $B$-modes act as an
690: independent realization of noise in the mass map, and locally highlight any problems with the
691: correction for PSF or other effects peculiar to the (two very different) instruments.
692: Unsurprisingly, a visual inspection of Figure~\ref{bmode} shows that the $B$-mode signal is
693: significantly lower in our space-based data, with fewer $B$-mode peaks. The overall noise level is
694: reduced, and holes arising from masked foreground stars are also smaller and less frequent. In the
695: ground-based maps, these create additional edges that lead to spurious effects during the Fourier
696: transforms required by equation~\eqref{eqn:gammatokappa}. 
697: The extended gaps are caused by difficulties modeling the PSF near the edge of the field of view,
698: \com{and could be eliminated in future surveys by more conservative tiling strategies}.
699: %Figure~\ref{bpeaks} compares the
700: %distribution of local maxima in the $B$-mode signal and their peak detection significance.
701: 
702: %
703: % NOTE: the next paragraph really needs to stay in because it is referred to later
704: %
705: \com{The southwest corner of the field has been
706: troublesome throughout our analysis. This pointing was observed in slightly worse seeing, so the
707: density of galaxies is reduced and the noise in the mass reconstruction is higher.}
708: %There could also be a potential problem with the PSF model as there are spurious.
709: 
710: %\begin{figure}
711: %\centerline{\psfig{file=f8.eps,width=2.3in, angle=90}}
712: %\caption[]{Histogram of $B$-mode peaks in convergence maps \com{smoothed on $1\arcmin$ scales} from 
713: %Subaru data (dotted red) and Hubble Space Telescope data (solid black).
714: %The ground-based map contains more peaks of higher significance, suggesting it is more affected by systematics.}
715: %\label{bpeaks}
716: %\end{figure}
717: 
718: \subsection{Halo Detection}
719: \label{sec:halos}
720: 
721: The higher surface density of background galaxies from space also improves the reconstruction of the
722: $E$-mode ``mass map''convergence field. The noise is lower and the angular resolution higher
723: (although to aid comparison, both panels in figure~\ref{emode} are smoothed to the same scale).
724: Several of the key features are qualitatively similar but we are struck by the significant differences
725: in the prominence of other mass peaks. 
726: \com{To evaluate the robustness of detections, we shall now employ an automated peak-finding
727: algorithm.}
728: 
729: \com{ Following \citep{miyazaki07}, we smooth the convergence maps by a Gaussian kernel of rms 
730: width 1$\arcmin$ and find local maxima with detection significance $\nu>4$ (assuming Gaussian errors 
731: on the shear measured within $0.7\arcmin$ cells on the sky equal to the dispersion of those galaxy shears). 
732: Five peaks (marked A, B, C, D and K in figure~\ref{emode}) are then identified in the ground-based data. 
733: However, two of these are near boundaries of the field mask. Imposing the rigorous restrictions discussed
734: by \citep{miyazaki07}, we find that only peaks A, B and C remain \citep[c.f.\ table~3 in ][]{miyazaki07}.  All 
735: three are also detected in a space-based lensing analysis, the 3D distribution of galaxies and as extended 
736: sources in  $X$-ray data \citep{hasinger07,finoguenov07}.} Assuming the mass-luminosity relation adopted 
737: by \citep{finoguenov07}, the detection threshold of this very deep $X$-ray data is well below that expected
738: for lensing up to redshift $\sim1$, so this acts as an ideal external arbiter (of course, $X$-ray 
739: mass-observable relations are somewhat uncertain). The properties of the three clusters are summarized 
740: in table~\ref{tab:mass} and demonstrate excellent agreement between the ground- and space-based
741: data using the formalism of \citep{miyazaki07}.
742: 
743: Cluster A (SJ J0959.6+0231) is the most massive structure inside the COSMOS field, easily detected
744: at many wavelengths. It appears to be in the process of a major merger, and has been studied
745: individually by \citep{guzzo07}, who also obtained a spectroscopic redshift of $z=0.73$.
746: Cluster B (SL J1001.4+0159) is associated with an $X$-ray peak
747: and overdensity of galaxies at $z_\mathrm{phot}=0.35$. There is a second set of galaxies at
748: $z_\mathrm{phot}=0.85$ within $2\arcmin$, which undoubtedly complicates the interpretation a
749: little, but our results are consistent with this high redshift projection being a minor
750: perturbation. Cluster C (NSC J100047+013912) is yet more local \citep[$z=0.22$,][]{miyazaki07} and 
751: appears large on the sky. Only part of this cluster is inside the region of HST imaging, so the space-based 
752: signal is significantly weakened, and the mass is potentially underestimated by HST.
753: %\com{This highlights the need to for any future facility to be able to observe a large, contigous area.}
754: 
755: \com{ To broaden our search, and test the limits of detectability, we additionally investigate the 
756: multi-scale procedure of \citep{hamana03}. For this, we smooth the convergence maps with Gaussian 
757: filters of rms width between 0.5$\arcmin$ and 4$\arcmin$, identifying local maxima inside the mask on 
758: each scale. For each peak with a detection signal to noise $\nu>4$ on any scale, we record $\nu$
759: and the smoothing scale that maximizes $\nu$. We also drop the restrictions on distance from the 
760: mask boundaries. This will increase the number of detected peaks, but at the expense of potentially 
761: introducing some spurious features. We then search for counterparts in the other data set, within 
762: 3$\arcmin$ of detected peaks.}
763: 
764: With the above criteria, we identify four mass peaks common to both convergence maps (A, B, C and D).
765: Cluster D is within 3$\arcmin$ of an $X$-ray peak, and an overdensity of galaxies at photometric 
766: redshifts $z_\mathrm{phot}=0.93$.  This redshift is rather high for a lensing analysis, and it was flagged 
767: as ``unsecure'' by \citep{miyazaki07} because it is near a boundary in the image mask.
768: However, the tentative Subaru detection is strengthened by the confirmation from HST, and appears 
769: to be robust. 
770: 
771: %With the above criteria, we identify four halos (A, B, C and D) common to both data sets. All four
772: %are discussed in more detail as part of the larger Subaru survey from which this data is drawn
773: %\citep{miyazaki07,green07}. Cluster A, the most massive structure in the COSMOS field, and one
774: %seemingly undergoing a major merger, is also discussed in \citep{guzzo07}. An overdensity of
775: %galaxies is present at all four locations. Spectroscopic redshifts are available for clusters A
776: %($z=0.73$) and D ($z=0.22$) \citep{miyazaki07}, and photometric redshift estimates for clusters B
777: %($z\approx0.93$) and C ($z=0.35$) \citep{finoguenov07}. Cluster C is the most problematic: there is
778: %a second set of galaxies with photometric redshifts at $z=0.85$ within $2\arcmin$. This undoubtedly
779: %complicates the interpretation a little, but our results are consistent with this high redshift
780: %projection being a minor perturbation.
781: 
782: %To make sense of the remaining diverse detections, we also use the XMM data for this field.
783: %\citep{finoguenov07} compiled a comprehensive list of mass concentrations in the COSMOS field by
784: %cross-correlating the $X$-ray data \citep{hasinger07} with the galaxy photometric redshift catalog
785: %\citep{mobasher07}. We search for counterparts of our lensing-selected halos in the XMM catalog by
786: %spatial colocation within 3$\arcmin$. Assuming their adopted mass-luminosity relation, the detection
787: %threshold of this very deep $X$-ray data is below that expected for lensing up to redshift $\sim1$,
788: %so this acts as an ideal arbiter (of course, $X$-ray mass-observable relations are notoriously
789: %uncertain).
790: 
791: Peaks E, F, G and H (also marked on Figure~\ref{emode}) are seen only in the space-based map. The first three
792: correspond to extended $X$-ray emission from clusters with masses $M_{500}$ between $2-4\times10^{13}M_\sun$
793: \citep{finoguenov07}. Peak H is more massive ($M_{500}=1.8\times10^{14}M_\sun$), but is at very high
794: redshift. \com{All four of these peaks are real detections; however no counterparts within 3$\arcmin$
795: are seen in the ground-based map, even down to $\nu>$3. Most likely, this is because of their lower mass and
796: higher redshift \citep{hamana03}.} The detection of peak G
797: was prevented in the ground-based data by a bright foreground star.
798: 
799: %Conversely, peaks J, K, L, M, N, O and P are detected in the ground-based map but have no
800: %corresponding space-based detection. Halos J and N are right on the edge of the HST field. There is a
801: %real cluster just outside the HST imaging from J, and indeed a weak signal is detected from its
802: %wings, although at $\nu<4$. Peak N is likely to be spurious: such noise artifacts are more common
803: %near the edge. Peak L was also detected in the \citep{miyazaki07} analysis, but again flagged as
804: %``unsecure''. It does align with a slight detection in the space-based map, but below $\nu=3$, and
805: %there is no $X$-ray counterpart. This may be coincidence, and a spurious signal, or perhaps a very
806: %distant object. Peaks K, M, O, P are in a southwest 1/9 portion of the field that has been
807: %troublesome throughout our analysis. This pointing was observed in slightly worse seeing, so the
808: %density of galaxies is reduced. Peaks K, O and P in particular are also only detected at very large
809: %smoothing scales, nearly the size of individual {\it Suprime-Cam} CCDs. There could be a potential
810: %problem with the PSF model. While all of the peaks seen in the space-based data (A-I) have
811: %counterpart detections in the $X$-ray catalog, none of the peaks seen only in the ground-based map
812: %(J-P) do. This suggests that peaks K-P are spurious features.
813: 
814: \com{
815: Conversely, peaks I, J, K and L are detected only in the ground-based map.
816: Peaks I and J are real, but lie just outside the HST imaging.
817: There is an extended $X$-ray source at peak I, with unknown redshift, and 
818: a projection of two $M_{500}\approx2\times10^{13}M_\sun$ clusters at redshifts $z=0.40$ and $z=0.75$ at peak J \citep{finoguenov07}.
819: In both cases, there is a weak, $\nu<3$ signal in the HST data, from the wings of the cluster.
820: Peak K was detected in the \citep{miyazaki07} analysis, but again flagged as
821: ``unsecure'' because it is near the edge of a pointing.
822: It does align with a slight, $\nu<3$ detection in the space-based map, 
823: but there is no $X$-ray counterpart. This may be a spurious peak with chance coincidence, or perhaps a very
824: distant object.
825: Peak L appears to be spurious: such noise artifacts are more common near the edge of the field. 
826: }
827: 
828: In summary, to the extent that we can draw conclusions from such a small sample, there
829: is very good agreement between the primary halo catalog drawn from the ground-based
830: data and that independently found from the space-based data. Additional halos of
831: lower mass and higher redshift are seen in the space-based catalog, and those located uniquely in the
832: ground-based data can be understood in the context of either being outside the space-based
833: region or close to its periphery.
834: 
835: \begin{deluxetable*}{cccccccccccc}
836: \tabletypesize{\footnotesize}
837: %\tablenum{3}
838: \footnotesize
839: %\tabletypesize{\scriptsize}
840: \setlength{\tabcolsep}{0.07in}
841: \tablecaption{Cluster Masses}
842: \tablewidth{0pc}
843: \tablehead{
844: % first line of header
845: \colhead{Cluster}   &\colhead{RA}  &\colhead{Dec} &\colhead{Redshift}  &  \colhead{XMM mass} & \colhead{HST mass}  & \colhead{Subaru mass} \\ %& \colhead{HST Peak Mass}  &  \colhead{Subaru Peak Mass}  \\ %&\colhead{XMM size}  & \colhead{HST size}  &  \colhead{Subaru size} \\
846: % second line of header
847:    & &   & & \colhead{10$^{14} M_{\sun}$} & \colhead{10$^{14} M_{\sun}$} & \colhead{10$^{14} M_{\sun}$}}% &  \colhead{10$^{14} M_{\sun}$} & \colhead{10$^{14} M_{\sun}$ % & $r_{500}$ [$\arcmin$] & $r_{200}$ [$\arcmin$] & $r_{200}$ [$\arcmin$] 
848: \startdata
849:  A  & 149.917 & 2.515   &  0.73 & $1.90\pm0.05$ & 23$^{+13}_{-8}$  & $13^{+33}_{-9}$   \\%& 1.37 &  6.6  &  16.6    1.3 $\pm$ 0.5   &   2.7 $\pm$ 0.5   \\
850: %H  & 150.092 & 2.207   &  0.93 & $0.33\pm0.04$ & 15$^{+76}_{-13}$ & $29$^{+54}_{-19}$ \\%& 0.65 & 4.1  &  3.2   \\% 3.2 $\pm$ 0.5   &   4.2 $\pm$ 0.5   \\
851:  B  & 150.359 & 1.999   &  0.35 & $0.10\pm0.01$ & 9$^{+7}_{-4}$    & $17^{+13}_{-7}$   \\%& 0.88 &  3.4 &  3.9   \\% 1.4 $\pm$ 0.5   &   1.1 $\pm$ 0.5   \\
852:  C  & 150.184 & 1.657   &  0.22 & $1.01\pm0.02$ & 17$^{+17}_{-9}$  & $55^{+55}_{-27}$ %\\%& 2.75 & 3.8  &  7.5   \\% 1.3 $\pm$ 0.5   &   0.8 $\pm$ 0.5   \\
853: %Add XMM Luminosity here 1.56e44, 2.85e43, 3.8e42, 4.1e43
854: %Size 0.0228, 0.0109, 0.0147, 0.0458 degrees i.e. 1.368, 0.654, 0.882, 2.748
855: \enddata
856: \label{tab:mass}
857: \end{deluxetable*}
858: 
859: 
860: %\begin{deluxetable}{ccccl}
861: %\tabletypesize{\footnotesize}
862: %%\tablenum{2}
863: %\footnotesize
864: %\setlength{\tabcolsep}{0.07in}
865: %\tablecaption{Cluster Detections\\
866: %{\bf ***Detection S/N could replace ticks.\\
867: %Could also add size, RA, Dec, z,...?***}}
868: %\tablewidth{0pc}
869: %\tablehead{
870: %\colhead{Cluster}   &\colhead{Space}  &\colhead{Ground} &\colhead{$X$-ray}  &  \colhead{Comments}}
871: %\startdata
872: %A & \checkmark &  \checkmark  & \checkmark &  \\
873: %B & \checkmark &  \checkmark  & \checkmark &  \\
874: %C & \checkmark &  \checkmark  & \checkmark & At edge of HST field \\
875: %D & \checkmark & (\checkmark) & \checkmark &  \\
876: %E & \checkmark &       ~      & \checkmark &  \\
877: %F & \checkmark &       ~      & \checkmark &  \\
878: %G & \checkmark &       ~      & \checkmark &  \\
879: %H & \checkmark &       ~      & \checkmark &  \\
880: %I &      ~     & (\checkmark) & \checkmark & Outside HST field \\
881: %J &      ~     & (\checkmark) & \checkmark & Outside HST field \\
882: %K &      ~     & (\checkmark) &      ~     & Uncertain         \\
883: %L &      ~     & (\checkmark) &      ~     & Spurious          
884: %\enddata
885: %\label{tab:mass}
886: %\end{deluxetable}
887: 
888: 
889: \subsection{Halo Mass Estimation}
890: 
891: We now attempt to measure the total mass of each of the three halos (A-C) securely detected from both
892: the ground and space.
893: We assume that the clusters have an NFW density profile
894: \begin{equation}
895: \rho(r) = \delta_c ~\rho_c/(r/r_s)(1+r/r_s)^2 ~
896: \end{equation} 
897: \noindent \citep{nfw}, where $\delta_c$ is a function of the cluster's concentration $c$
898: and scale size $r_s \equiv r_{200}/c$, \com{and $r_{200}$ is the radius within which the mean density
899: is 200 times the critical density.} 
900: %Since $r_{200}$ is the radius where the mean density
901: %is 200 times the critical density, it can be equivalently expressed in mass units of $M_{200} = 4/3~\pi~r_
902: %{200}^3~\rho_c$. 
903: The shear profile of an NFW cluster is derived by \citep{king01}. 
904: We perform a maximum likelihood fit to the log(mass) and concentration parameters, using the shear 
905: measurements from all galaxies within 10$\arcmin$ of the \com{peak convergence signal}, averaged in radial bins of 0.5$\arcmin$.
906: %Figure~\ref{masses} shows the 68\%, 95\% and 99\% joint confidence limits.
907: \com{
908: It has been variously noted (J.\ Berg\'e, S.\ Paulin-Henrikkson, private comm.) that fitting noisy data 
909: of individual clusters with an NFW profile does permit large (and therefore massive) models with 
910: unnaturally low concentration values.
911: %We are more concerned aboout the errors than the absolute values but, 
912: To counter this effect, we impose a concentration prior, using the lognormal distribution found 
913: for all haloes in the {\it Millenium Simulation} as a function of mass
914: by \citep[][equations~(5), (6) and figure~6]{neto}.
915: The resulting likelihood surfaces are shown in figure~\ref{masses}, 
916: with the effect of the prior being to close the bottom of the contours.
917: Table~\ref{tab:mass} lists the best-fit masses and 68\% confidence limits after marginalizing over
918: concentration between $1<c<10$.
919: }
920: 
921: %%***behind the cluster - did you cut foreground gals? Needs a line saying yes or no, either way.*** 
922: %in a radius of 10$\arcmin$ of the cluster position in our catalogs, averaging 
923: %shear in bins of 0.5$\arcmin$. Next, we compute the residual sum of squares
924: %(RSS), $s^2$, of the theoretical prediction and data for a range of values of mass and
925: %concentration. Our best fit value for the mass and concentration is the value for
926: %which $s^2$ is a minimum.  We then compute the histogram of likelihood, 
927: %$exp(-s^2/2)$, and integrate until the area is 68\%, 95\% and 99\% of the 
928: %total area to determine the contour levels that best represent our RSS grid. 
929: %Our contour plots of the RSS grid for the ground and space-based data are
930: %compared in Figure~\ref{masses}. To estimate the best-fit mass and uncertainty,
931: %we marginalize the 2-d likelihood grid over the concentration parameter to get a
932: %1-d likelihood as a function of mass. Next, we compute the peak in the likelihood
933: %as the best-fit mass and the range for which the area is 68\% of the total area 
934: %to estimate the uncertainty. We summarize the mass estimates
935: %of clusters A, B, C and D in table~\ref{tab:mass}. 
936: 
937: \begin{figure}
938: \centerline{\epsfig{file=f8.eps, width=80mm}}
939: \caption[]{Best-fit mass and concentration index of three clusters in the COSMOS field, assuming
940: NFW radial mass profiles. The contours show the 68\%, 95\% and 99\% confidence regions obtained
941: from Subaru data (dotted red) and Hubble Space Telescope data (solid black).}
942: \label{masses}
943: \end{figure}
944: 
945: 
946: 
947: Although our common sample is small, there is an encouraging agreement between the detailed
948: properties of the clusters recovered from the ground and from space. For the higher redshift cluster A,
949: our space-based data does put significantly tighter constraints on the mass and
950: concentration than our ground-based data. However, for the lower redshift clusters B
951: and C, the results are satisfyingly similar. We note again that cluster C is partially outside the
952: HST imaging. Since shears are only measured around one half of the cluster, 
953: the \com{statistical errors} in the space-based analysis are \com{larger and its mass could be underestimated}.
954: Certainly, for massive clusters with redshifts 
955: $0.2\simlt z\simlt 0.5$, it appears that our ground-based depth and resolution is adequate. The main benefit of 
956: space-based imaging is in the measurement of lower mass halos and higher redshift clusters, plus the 
957: increased resolution to further investigate the distribution of their masses.
958: 
959: 
960: 
961: 
962: 
963: 
964: 
965: 
966: 
967: 
968: \section{Discussion}\label{sec:conc}
969: 
970: We have performed parallel weak lensing analyses of Subaru and Hubble Space Telescope imaging in 
971: the
972: COSMOS field. Our comparisons of the observed shear and convergence signals have revealed a 
973: number of
974: issues, and suggest that such a study with real data usefully complements the independent
975: approach based on blind analyses of simulated data \citep{step1, step2}. 
976: 
977: For statistical ``cosmic shear'' analyses, shear measurement with an existing ground based
978: telescope, using existing measurement techniques, can be achieved with less than 1\% bias relative
979: to higher resolution space-based data, for a galaxy surface density of 15 arcmin$^{-2}$. 
980: One limitation of our approach is that we cannot check the performance of our space-based analysis
981: on the additional, small galaxies.
982: At first sight the factor of $\sim$3 shortfall in surface density seems inconsequential given the
983: lower cost and improved areal mapping speed of existing ground-based cameras
984: such as SuPrime-Cam. However, accompanying the brighter Subaru limit is a
985: reduction in survey depth and hence the redshift distribution of background sources.
986: More distant sources contain a larger signal, and a narrower range of redshifts also hinders
987: tomographical tests \citep{bacon05,massey07b}, which tighten cosmological parameter constraints
988: significantly. 
989: 
990: A key issue is whether this limiting depth is a fundamental one for all future
991: ground-based cameras. PanSTARRS, VST, and even LSST each have significantly
992: smaller primary mirrors than Subaru, so achieving even the $S/N$ discussed here
993: would require formidable exposure times. Most importantly, the deep infrared imaging
994: that is required for photometric redshifts to enable tomographic analyses is likely to be
995: difficult over large survey fields from the ground, because of increased sky background.
996: Recent weak lensing analyses are limited at roughly the same level by uncertainty in
997: galaxy shape measurement and photometric redshift estimation. 
998: %For example, the
999: %relatively shallow $K$-band coverage in even the small COSMOS field limits the tomographic
1000: %analysis \citep{massey07a} to only three redshift bins, and even then contributes one of the
1001: %dominant sources of potential systematic uncertainty. Full implementations of cross-correlation
1002: %cosmography will presumably require deep near-IR imaging from space.
1003: 
1004: Statistical measurements from the ground are also hindered by variable atmospheric seeing. Past
1005: experience has taught the authors that data collected in seeing worse than $0.8\arcsec$ is
1006: of little use for weak lensing analysis. The apparently rapid speed of data collection for our Subaru
1007: data belies the time spent waiting for better seeing, even with the excellent atmospheric
1008: conditions above Mauna Kea and the well-controlled dome seeing of Subaru. For this small-scale
1009: survey, we obtained exceptional quality imaging during a fortuitous observing window. The relevant
1010: quantity for larger-scale surveys in the future will be the time-averaged seeing quality, and the
1011: fraction of time spent with seeing better than $0.8\arcsec$. This is particularly true for surveys
1012: like Pan-STARRS and LSST, that plan to adopt a strategy of co-adding many shorter exposures. 
1013: Their advantage is that the stacked images will achieve a near-uniform image quality, by virtue of the
1014: independent PSFs in each short exposure. This can then be tuned to the required image quality by
1015: rejecting a certain fraction of exposures.
1016: 
1017: Variable seeing conditions is also of concern for the reconstruction of mass maps (c.f.\ Green et al.\ {\it in 
1018: prep.}). Difficulties in
1019: the analysis of one pointing in the SW corner of the Subaru map result in a patchy recovery of
1020: large-scale structure; with \com{more noise and a lower range of probed redshift}
1021: in certain regions. 
1022: 
1023: %Figures~\ref{emode} and \ref{bmode} also give a clear visual indication 
1024: %of the benefits of a space-based platform. The systematics and statistical noise in the $B$-mode 
1025: %reconstruction are
1026: %significantly reduced, and the $E$-modes allow an improved resolution in selecting halos, \com{plus mass
1027: %sensitivity \citep[especially on larger smoothing scales, see][]{massey07b}
1028: %that includes filamentary large-scale structure}. This will be particularly important
1029: %for reconstructions of individual regions of interest like the Bullet cluster, where the detailed
1030: %differences between the location of mass and baryons in a small patch of sky (and in their wider
1031: %setting of nearby large-scale structure) may yield the best constraints on the properties of dark
1032: %matter.
1033: 
1034: Most importantly, \com{however, the four most massive clusters out of eight detected from space are also detected from the ground --
1035: with one intriguing additional signal and two more confirmed clusters just ouside the field of view observed from space.
1036: Reassuringly, the three clusters conservatively deemed ``secure'' by the independent analysis of \citep{miyazaki07} 
1037: have now been confirmed via space-based weak lensing and $X$-ray observations.
1038: The physical properties of the four massive halos in common (A, B and C) are remarkably consistent whether derived from 
1039: from ground- or space-based weak lensing. The measured masses and radial profiles of these clusters are consistent and, for
1040: the lower redshift clusters, the error bars are comparable.}
1041: %There is no 
1042: %significant systematic overestimation or underestimation of the implied mass or size, and
1043: %for lower redshift clusters, the ground- and space-based errors are comparable.
1044: A Dark Energy Task Force ``Stage 3'' survey from the ground appears eminently feasible. 
1045: %For 
1046: %the two higher redshift clusters, the space-based images provide a slightly better constraint 
1047: %on the mass and concentration parameter. 
1048: 
1049: %If we consider completeness of the sample,
1050: %4 out of 9 clusters detected by HST and XMM are recovered by Subaru -- 
1051: %a completeness of 44\%. Conversely, 56\% of the Subaru halos are undetected by HST, 
1052: %but these are likely spurious. Most of these detections lie in the SW corner of  the image 
1053: %which appears to be affected by a single pointing of lower quality. The XMM exposure map also 
1054: %indicates a lower sensitivity in this area.
1055: %Without this SW corner, the degree of contamination in the Subaru catalog with respect to 
1056: %XMM and ACS is only 1 out of 5, i.e.\ 20\%.  We emphasize here that our sample to estimate
1057: %the completeness and contamination is small and further studies of this nature are 
1058: %recommended to improve the statistics.
1059:  
1060: A wide-field space-based platform would open up many new applications. Very important for statistical
1061: applications is the increased redshift range of resolved background galaxies. These not only contain
1062: a larger shear signal, but also more readily split into redshift bins for tomographic analysis.
1063: Three-dimensional analysis techniques will tighten constraints on cosmological parameters by factors
1064: of $3-5$, and directly measure quantities that depend upon the properties of dark energy, like the
1065: growth of structure over cosmic time and the redshift-distance relation. They will also eliminate
1066: sources of error due to the intrinsic correlations of galaxy shapes. Sufficiently good photometric
1067: redshifts require deep, wide-field near-IR imaging, and these are also realistically possible over
1068: large surveys only from space. Recent weak lensing analyses with relatively shallow near-IR coverage
1069: like \citep{massey07a} are limited to roughly the same degree by uncertainty in galaxy shape
1070: measurement and photometric redshift estimation. Full implementations of cross-correlation
1071: cosmography will almost certainly require deep near-IR imaging from space. Such advanced techniques
1072: will become particularly important as ground-based surveys expand to encompass the entire observable
1073: sky.
1074: 
1075: The increased surface density of galaxies resolved from space also improves maps of the mass
1076: distribution. \com{As shown in figures~\ref{emode} and \ref{bmode}, the statistical noise and systematic 
1077: contamination in the $B$-mode are significantly reduced.
1078: Eight clusters are detected in the COSMOS $E$-mode signal without any contamination from spurious peaks.
1079: With the increased mass and spatial resolution of mass reconstructions from space, it becomes possible to detect halos
1080: the size of galaxy groups, as well as clusters over a wide range of redshifts -- thus tracing
1081: their formation,} which is governed by the properties of dark matter and the nature of gravity. 
1082: Space-based data also crosses the threshold to mapping even filamentary large-scale structure in three dimensions
1083: \citep{massey07b}. 
1084: %Performed over large
1085: %expanses of the sky, this} will give us an unprecedented view of the invisible universe, and also
1086: %parametrizing the local environment of other astrophysical phenomena.
1087: Obtaining the detailed, 3D
1088: distribution of mass will be particularly important near regions of interest like the Bullet cluster
1089: \citep{clowe06}, where the small differences between the location of mass and baryons in a small
1090: patch of sky may yield the best possible
1091: information about the properties of dark matter. In this and other astrophysical phenomena, knowledge of the
1092: local mass environment and nearby large-scale structure is critical.
1093: 
1094: Overall, we conclude that ground-based weak lensing surveys can perform several tasks remarkably
1095: well, with sufficiently small amount of systematic bias to easily justify the next generation of
1096: dedicated  ground-based surveys. Two dimensional statistical analyses will be able to produce 
1097: order-of-magnitude improvements in weak lensing constraints, using proven hardware technology and
1098: software pipelines. On the other hand, a wide-field space-based imager would provide important
1099: control over some systematic effects, and open up many new applications that are, at least
1100: currently, unachievable from the ground. For several of the most exciting techniques that will
1101: directly probe the nature of dark matter and dark energy, eventual space-based imaging is likely to
1102: be essential.
1103: 
1104: \section{Acknowledgments}
1105: 
1106: This work is supported by the US Department of Energy under contract DE-FG02-04ER41316. It is based upon observations with the NASA/ESA
1107: Hubble Space Telescope, obtained at the  Space Telescope Science Institute, which is operated by AURA Inc, under NASA contract  NAS 5-26555.
1108: It is also based on data collected at the Subaru Telescope, operated by the  National Astronomical Observatory of Japan. It is our pleasure
1109: to thank Alexie Leauthaud and Takashi Hamana for help with the catalogs and Alexandre Refregier, James Taylor and Alexis Finoguenov for
1110: useful discussions. We also gratefully acknowledge the contributions of the entire COSMOS collaboration, consisting of more than 70
1111: scientists worldwide. More information on the COSMOS survey is available from {\tt \url{http://www.astro.caltech.edu/$\sim$cosmos}}.
1112: 
1113: 
1114: \begin{thebibliography}{}
1115: 
1116: 
1117: \bibitem[{Bacon, Refregier \& Ellis }{2000}]{bacon00} Bacon, D., Refregier, A.\ \& Ellis, R., 2000, \mnras, 
1118: 318, 625.
1119: 
1120: \bibitem[{Bacon et al.\ }{2001}]{bacon01} Bacon, D., Refregier, A., Clowe, D.\ \& Ellis, R., 2001, \mnras, 
1121: 325, 1065.
1122: 
1123: \bibitem[{Bacon et al.\ }{2005}]{bacon05} Bacon, D.\ et al., 2005, \mnras, 363, 723.
1124: 
1125: \bibitem[Bardeau et al.\ 2007]{bardeau07} Bardeau, S., Soucail, G., Kneib, J.-P., Czoske, O., Ebeling, H., 
1126: Hudelot, P., Smail, I.\ \& Smith, G., 2007, \aap, 470, 449.
1127: 
1128: \bibitem[Benjamin et al.\ 2007]{benjamin07} Benjamin, J.\ et al., 2007, \mnras, in press (astro-ph/
1129: 0703570).
1130: 
1131: \bibitem[{Bernstein \& Jarvis }{2002}]{bj02} Bernstein, G.\ \& Jarvis, M., 2002, \aj, 123, 583.
1132: 
1133: \bibitem[Bertin \& Arnouts 1996]{BertinArnouts} Bertin, E.\ \&  Arnouts, S., 1996, A\&AS, 117, 393.
1134: 
1135: \bibitem[{Brown }{et al.\ }{2003}]{brown03} Brown, M., Taylor, A., Bacon, D., Gray, M., Dye, S., 
1136: Meisenheimer, K.\ \& Wolf, C., 2003, \mnras, 3431, 100.
1137: 
1138: \bibitem[Bridle et al.\ 2001]{im2shape} Bridle, S.\ et al., 2001, in Scientific N.~W., Proceedings of the
1139: Yale Cosmology Workshop.
1140: 
1141: \bibitem[Bristow et al.\ 2004]{cte} Bristow, P., 2004, ASP Conf.\ Proc., 314, 780.
1142: 
1143: \bibitem[Capak et al.\ 2007]{capak07} Capak, P.\ et al., 2007, ApJS, 172, 99.
1144: 
1145: \bibitem[Clowe et al.\ 2005]{clowe05} Clowe, D.\ et al., 2005, \aap, 451, 395.
1146: 
1147: \bibitem[Clowe et al.\ 2006]{clowe06} Clowe, D.\ et al., 2006, \apj, 648, L109.
1148: 
1149: \bibitem[Dahle et al.\ 2002]{dahle02} Dahle, H., Kaiser, N., Irgens, R., Lilje, P.\ \& Maddox, S., 2002, \apj 
1150: S, 139, 313.
1151: 
1152: \bibitem[Finoguenov et al.\ 2007]{finoguenov07} Finoguenov, A.\ et al., 2007, ApJS, 172, 182.
1153: 
1154: \bibitem[Fruchter \& Hook 2002]{drizzle} Fruchter, A.\ \& Hook, R., 2002, PASP, 114, 144.
1155: 
1156: \bibitem[Gavazzi et al.\ 2007a]{gavazzi07a} Gavazzi, R. et al., 2007, \aap, 462, 459
1157: 
1158: \bibitem[Gavazzi et al.\ 2007b]{gavazzi07b} Gavazzi, R. et al., 2007, \apj, 667, 176.
1159: 
1160: %\bibitem[Green et al.\ 2007]{green07} Green, A.\ et al., 2007, in preparation.
1161: 
1162: \bibitem[Guzzo et al.\ 2007]{guzzo07} Guzzo, L.\ et al., 2007, ApJS, 172, 219.
1163: 
1164: \bibitem[Hamana et al.\ 2003]{hamana03} Hamana, T.\ et al., 2003, \apj, 597, 98.
1165: 
1166: \bibitem[Hasinger et al.\ 2007]{hasinger07} Hasinger, G.\ et al., 2007, ApJS, 172, 29.
1167: 
1168: \bibitem[{Heavens, Kitching \& Taylor }{2006}]{heavens06} Heavens, A., Kitching, T.\ \& Taylor, A., 2006, 
1169: \mnras, 373, 105.
1170: 
1171: \bibitem[{Heymans \& Heavens }{2003}]{heymans03} Heymans, C.\ \& Heavens, A., 2003, \mnras, 339, 
1172: 711.
1173: 
1174: \bibitem[Heymans et al.\ 2005a]{step1} Heymans, C.\ et al., 2005a, \mnras, 368 1323.
1175: 
1176: \bibitem[Heymans et al.\ 2005b]{heymans05} Heymans, C.\ et al., 2005b, \mnras, 361, 160.
1177: 
1178: \bibitem[Heymans et al.\ 2006]{heymans06} Heymans, C.\ et al., 2006, \mnras, 371, L60.
1179: 
1180: \bibitem[{Hirata \& Seljak }{2003}]{hirata03} Hirata, C., Seljak, U., 2003, \mnras, 343, 459.
1181: 
1182: \bibitem[Hoekstra et al.\ 2006]{hoekstra06} Hoekstra H., Mellier Y., Van Waerbeke L.,
1183: Semboloni E., Fu L., Hudson M., Parker L., Tereno I.\ \& Benabed K., 2006, \apj, 647, 116.
1184: 
1185: \bibitem[{Jain \& Seljak }{1997}]{jainseljak97} Jain, B.\ \& Seljak, U., 1997, \apj, 484, 560.
1186: 
1187: \bibitem[Jee et al.\ 2007a]{jee07} Jee, J.\ et al., 2007a, \apj, 661, 728.
1188: 
1189: \bibitem[Jee et al.\ 2007b]{dmring} Jee, J.\ et al., 2007b, \apj, in press (arXiv:0705.2171).
1190: 
1191: \bibitem[Kaiser et al.\ 2000]{kaiser00} Kaiser N., Wilson G.,\ \& Luppino G.\ 2000, \apj, submitted (astro-
1192: ph/0003338).
1193: 
1194: \bibitem[{Kaiser \& Squires }{1993}]{kaiser93} Kaiser, N.\ and Squires, G.\ 1993, \apj, 404, 441.
1195: 
1196: \bibitem[{Kaiser, Squires \& Broadhurst }{1995}]{ksb} Kaiser, N., Squires, G.\ \& Broadhurst, T, 1995, \apj, 
1197: 449, 460.
1198:  
1199: \bibitem[Kaiser 2000]{k2k} Kaiser, N., 2000, \apj, 537, 555.
1200: 
1201: \bibitem[{King \& Schneider }{2001}]{king01} King, L.\ \& Schneider, P., 2001, \aap, 369, 1.
1202: 
1203: \bibitem[{King \& Schneider }{2003}]{king03} King, L.\ \& Schneider, P., 2001, \aap, 398, 23.
1204: 
1205: \bibitem[{Kitching }{et al.\ }{2007}]{kitching06} Kitching T., Heavens A., Taylor A., Brown M., 
1206: Meisenheimer K., Wolf C., Gray M.\ \& Bacon D., 2007, \mnras, 374, 1377.
1207: 
1208: \bibitem[Kneib et al.\ 2003]{kneib03} Kneib, J-P.\ et al., 2003, \apj, 598, 804.
1209: 
1210: \bibitem[Koekemoer et al.\ 2007]{koekemoer07} Koekemoer, A.\ et al., 2007, ApJS, 172, 196.
1211: 
1212: \bibitem[Kuijken 2006]{kuijken06} Kuijken, K., \aap, in press (astro-ph/0601011).
1213: 
1214: \bibitem[Lampton et al.\ 2006]{lampton06} Lampton, M., Sholl, M., Jelinsky, P.\ \& Stabenau H., 2006, 
1215: AAS, 209, 980
1216: 
1217: \bibitem[Leauthaud et al.\ 2007]{leauthaud07} Leauthaud, A.\ et al., 2007, \apj, 172, 219.
1218: 
1219: \bibitem[Mandelbaun et al.\ 2006a]{mandelbaum06a} Mandelbaum, R., Seljak, U., Kauffmann, G., 
1220: Hirata, C.\ \& Brinkmann, J, 2006, \mnras, 368, 725.
1221: 
1222: \bibitem[Mandelbaun et al.\ 2006b]{mandelbaum06b} Mandelbaum, R., Seljak, U., Cool, R., Blanton, M., 
1223: Hirata, C.\ \& Brinkmann, J, 2006, \mnras, 372, 758.
1224: 
1225: \bibitem[Massey et al.\ 2004]{snap2} Massey, R.\ et al., 2004, \aj, 127, 3089.
1226: 
1227: \bibitem[Massey et al.\ 2007a]{massey07a} Massey, R.\ et al., 2007a, ApJS, 172, 239.
1228: 
1229: \bibitem[Massey et al.\ 2007b]{massey07b} Massey, R.\ et al., 2007b, Nature, 445, 286.
1230: 
1231: \bibitem[Massey et al.\ 2007c]{step2} Massey, R.\ et al., 2007c, \mnras, 376, 13.
1232: 
1233: \bibitem[Massey et al.\ 2007d]{shapelets4} Massey, R., Rowe B., Refregier, A., Bacon, D.\ \& Berg\'e, J., 
1234: 2007d, \mnras, 380, 229.
1235: 
1236: \bibitem[Maturi et al.\ 2007]{maturi07} Maturi, M.\ et al., 2007, A\&A, 462, 473.
1237: 
1238: \bibitem[Mellier 1999]{mellier99} Mellier, Y., 1999, Ann.\ Rev.\ Astr.\ Astrophys., 37, 127.
1239: 
1240: \bibitem[Miyazaki et al.\ 2002a]{miyazaki02a} Miyazaki, S.\ et al., 2002a, \apjl, 580, 97.
1241: 
1242: \bibitem[Miyazaki et al.\ 2002b]{miyazaki02b} Miyazaki, S.\ et al., 2002b, PASJ, 54, 833.
1243: 
1244: \bibitem[Miyazaki et al.\ 2007]{miyazaki07} Miyazaki, S.\ et al., 2007, \apj, in press (arXiv:0707.2249).
1245: 
1246: \bibitem[Mobasher et al.\ 2007]{mobasher07} Mobasher, B.\ et al., 2006, ApJS, 172, 117.
1247: 
1248: \bibitem[{Nakajima \& Bernstein }{2007}]{reiko} Nakajima, R.\ \& Bernstein, G., 2007, \aj, 133, 1763.
1249:  
1250: \bibitem[{Navarro, Frenk \& White }{1996}]{nfw} Navarro, J., Frenk, C.\ \& White, S.\ 2006, \apj 462, p.563.
1251:  
1252: \bibitem[{Neto et al.\ }{2007}]{neto} Neto, A.\ et al., 2007, \mnras, submitted (arXiv:0706.2919).
1253: 
1254: \bibitem[Parker et al.\ 2007]{parker07} Parker, L., Hoekstra, H., Hudson, M., Van Waerbeke, L.\ \& Mellier, 
1255: Y., 2007, \apj, in press (arXiv0707.1698).
1256: 
1257: \bibitem[Refregier 2003]{refregier03} Refregier, A., 2003, Ann.\ Rev.\ Astr.\ Astrophys., 41, 645.
1258: 
1259: \bibitem[{Refregier \& Bacon }{2003}]{shapelets2} Refregier, A.\ \& Bacon ,D., 2003, \mnras, 338, 48.
1260: 
1261: \bibitem[Refregier et al.\ 2004]{refregier04} Refregier, A.\ et al., 2004, \aj, 127, 3102
1262: 
1263: \bibitem[{Rhodes et al.\ }{2001}]{rrg} Rhodes, J., Refregier, A.\ \& Groth, E.\ 2001, \apjl, 552, L85.
1264: 
1265: \bibitem[Rhodes et al.\ 2004]{rhodes04} Rhodes, J.\ et al., 2004, \apj, 605, 29.
1266: 
1267: \bibitem[Rhodes et al.\ 2007]{rhodes07} Rhodes, J.\ et al., 2007a, ApJS, 172, 203.
1268: 
1269: \bibitem[Schirmer et al.\ 2007]{schirmer07} Schirmer, M.\ et al., 2007, A\&A, 462, 872.
1270: 
1271: \bibitem[{Schneider et al. }{2002}]{schneider02} Schneider, P., Van Waerbeke, L.\ \& Mellier, Y., 2002, 
1272: \aap, 389, 729.
1273: 
1274: \bibitem[Schneider 2005]{schneider05} Schneider, P.\ 2005, Gravitational Lensing: Strong, Weak \& 
1275: Micro, Springer-Verlag, Berlin, 273.
1276: 
1277: \bibitem[Schrabback et al.\ 2006]{schrabback06} Schrabback, T.\ et al., 2007, \aap, 468, 823.
1278: 
1279: \bibitem[Scoville et al.\ 2007a]{scoville07a} Scoville, N.\ et al., 2007, ApJS, 172, 1.
1280: 
1281: \bibitem[Scoville et al.\ 2007b]{scoville07b} Scoville, N.\ et al., 2007, ApJS, 172, 38.
1282: 
1283: \bibitem[Semboloni et al.\ 2006]{sembolini06} Semboloni E., Mellier Y., van Waerbeke L.,
1284: Hoekstra H., Tereno I., Benabed K., Gwyn S., Fu L., Hudson M., Maoli R.\ \& Parker L., 2006, \aap, 452, 51.
1285: 
1286: \bibitem[{Smail, Ellis \& Fitchett }{1994}]{smailzdist} Smail I., Ellis R.\ \& Fitchett M., 1994, \mnras, 270, 245.
1287: 
1288: \bibitem[Taniguchi et al.\ 2007a]{taniguchi07} Taniguchi, Y., et al.\ 2007, ApJS, 172, 9.
1289: 
1290: \bibitem[Taylor et al.\ 2004]{taylor04} Taylor, A.\ et al., 2004, \mnras, 353, 1176.
1291: 
1292: %\bibitem[Starck et al.\ 2006]{starck06} Starck, J.\ et al.\ 2006, A\&A, 451, 1139.
1293: 
1294: \bibitem[van Waerbeke et al.\ 2000]{vanwaerbeke00} Van Waerbeke, L.\ et al., 2000, A\&A, 358, 30.
1295: 
1296: \bibitem[Wang et al 2004]{wang04} Wang, S.\ et al., 2004, Phys.\ Review D., 70, 123008.
1297: 
1298: \bibitem[{Wittman }{et al.\ }{2000}]{wittman00} Wittman, D.\ et al., 2000, Nature, 405, 143.
1299: 
1300: \bibitem[Wittman 2005]{wittman05} Wittman, D.\ et al.\ 2005, \apjl, 632, 5.
1301: 
1302: \bibitem[{Wittman }{et al.\ }{2006}]{wittman06} Wittman, D., Dell'Antonio, I., Hughes, J., Margoniner, V., 
1303: Tyson, J., Cohen, J.\ \& Norman, D., 2006, \apj, 643, 128
1304: 
1305: \end{thebibliography}
1306: 
1307: 
1308: 
1309: %\include{tables}
1310: %/scr/mansi/lensing/Emodes.eps -> fig4.eps
1311: %/scr/mansi/lensing/Bmodes.eps -> fig5.eps
1312: %/scr/mansi/lensing/NFWcontours_v3.ps
1313: 
1314: \end{document}
1315: 
1316: