1: %%
2: %% Beginning of file 'sample.tex'
3: %%
4: %% Modified 2004 January 9
5: %%
6: %% This is a sample manuscript marked up using the
7: %% AASTeX v5.x LaTeX 2e macros.
8:
9: %% The first piece of markup in an AASTeX v5.x document
10: %% is the \documentclass command. LaTeX will ignore
11: %% any data that comes before this command.
12:
13: %% The command below calls the preprint style
14: %% which will produce a one-column, single-spaced document.
15: %% Examples of commands for other substyles follow. Use
16: %% whichever is most appropriate for your purposes.
17: %%
18:
19: \documentclass{emulateapj}
20: %\usepackage{apjfonts,graphics}
21:
22: %\documentclass[12pt,preprint]{aastex}
23: \usepackage{graphics,graphicx,rotating}
24:
25: %% manuscript produces a one-column, double-spaced document:
26: %% \documentclass[manuscript]{aastex}
27:
28: %% preprint2 produces a double-column, single-spaced document:
29: %% \documentclass[preprint2]{aastex}
30:
31: %% Sometimes a paper's abstract is too long to fit on the
32: %% title page in preprint2 mode. When that is the case,
33: %% use the longabstract style option.
34:
35: %% \documentclass[preprint2,longabstract]{aastex}
36:
37:
38: %% If you want to create your own macros, you can do so
39: %% using \newcommand. Your macros should appear before
40: %% the \begin{document} command.
41: %%
42: %% If you are submitting to a journal that translates manuscripts
43: %% into SGML, you need to follow certain guidelines when preparing
44: %% your macros. See the AASTeX v5.x Author Guide
45: %% for information.
46:
47: \newcommand{\simgt}{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}
48: \newcommand{\simlt}{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}
49: \newcommand{\gIII}{I\hspace{-.3mm}I\hspace{-.3mm}I}
50:
51: \def\bbeta{\mbox{\boldmath $\beta$}}
52: \def\btheta{\mbox{\boldmath $\theta$}}
53: \def\bnabla{\mbox{\boldmath $\nabla$}}
54: \def\bk{\mbox{\boldmath $k$}}
55: \def\bvtheta{\mbox{\boldmath $\vartheta$}}
56: \def\bdelta{\mbox{\boldmath $\Delta$}}
57: \def\bphi{\mbox{\boldmath $\phi$}}
58:
59: %% You can insert a short comment on the title page using the command below.
60:
61: %\slugcomment{Not to appear in Nonlearned J., 45.}
62:
63: %% If you wish, you may supply running head information, although
64: %% this information may be modified by the editorial offices.
65: %% The left head contains a list of authors,
66: %% usually a maximum of three (otherwise use et al.). The right
67: %% head is a modified title of up to roughly 44 characters.
68: %% Running heads will not print in the manuscript style.
69:
70: \shorttitle{Combining Lens Distortion and Depletion}
71: \shortauthors{Umetsu \& Broadhurst}
72:
73: %% This is the end of the preamble. Indicate the beginning of the
74: %% paper itself with \begin{document}.
75:
76: \begin{document}
77:
78: %% LaTeX will automatically break titles if they run longer than
79: %% one line. However, you may use \\ to force a line break if
80: %% you desire.
81:
82:
83: \title{
84: Combining Lens Distortion and Depletion to Map the
85: Mass Distribution of A1689\altaffilmark{1}
86: }
87:
88: %% Use \author, \affil, and the \and command to format
89: %% author and affiliation information.
90: %% Note that \email has replaced the old \authoremail command
91: %% from AASTeX v4.0. You can use \email to mark an email address
92: %% anywhere in the paper, not just in the front matter.
93: %% As in the title, use \\ to force line breaks.
94:
95:
96: \author{Keiichi Umetsu\altaffilmark{2,3}}
97: \email{keiichi@asiaa.sinica.edu.tw}
98: %\and
99: \author{Tom Broadhurst\altaffilmark{4}}
100: \email{tjb@wise.tau.ac.il}
101:
102: %% Notice that each of these authors has alternate affiliations, which
103: %% are identified by the \altaffilmark after each name. Specify alternate
104: %% affiliation information with \altaffiltext, with one command per each
105: %% affiliation.
106:
107: \altaffiltext{1}{Based in part on data collected at the Subaru Telescope,
108: which is operated by the National Astronomical Society of Japan}
109: \altaffiltext{2}
110: {Institute of Astronomy and Astrophysics, Academia Sinica, P.~O. Box
111: 23-141, Taipei 106, Taiwan}
112: \altaffiltext{3}
113: {Leung center for Cosmology and Particle Astrophysics,
114: National Taiwan University, Taipei 106, Taiwan}
115: \altaffiltext{4}
116: {
117: School of Physics and Astronomy, Tel Aviv University, Israel
118: }
119:
120:
121:
122:
123: %% Mark off your abstract in the ``abstract'' environment. In the manuscript
124: %% style, abstract will output a Received/Accepted line after the
125: %% title and affiliation information. No date will appear since the author
126: %% does not have this information. The dates will be filled in by the
127: %% editorial office after submission.
128:
129: \begin{abstract}
130: We derive a projected 2D mass map of the well studied galaxy cluster
131: A1689 based on an entropy-regularized maximum-likelihood combination
132: of the lens magnification and distortion of red background galaxies
133: registered in deep Subaru images. The method is not restricted to the
134: weak regime but applies to the whole area outside the tangential
135: critical curve, where non-linearity between the surface mass-density
136: and the observables extends to a radius of a few arcminutes. The known
137: strong lensing information is also readily incorporated in this
138: approach, represented as a central pixel with a mean surface density
139: close to the critical value. We also utilize the distortion
140: measurements to locally downweight the intrinsic clustering noise,
141: which otherwise perturbs the depletion signal. The resulting 2D map
142: shows that the surface density of A1689 is smoothly varying and
143: symmetric, similar to the distribution of cluster members, with no
144: apparent substructure at $r \simgt 130 {\rm kpc}/h$ ($\sim 1\arcmin$).
145: The projected mass profile continuously steepens with radius and is well
146: fitted by the Navarro-Frenk-White model, but with a surprising large
147: concentration $c_{\rm vir}=13.4^{+5.3}_{-3.3}$, lying far from the
148: predicted value of $c_{\rm vir}\sim 5$, corresponding to the measured
149: virial mass, $M_{\rm vir}=(2.1\pm 0.2)\times 10^{15}M_{\odot}$,
150: posing a challenge to the standard assumptions defining the $\Lambda$CDM
151: model. We examine the consistency of our results with estimates
152: derived with the standard weak lensing estimators and by
153: comparison with the inner mass profile obtained from strong lensing.
154: All the reconstructions tested here imply a virial mass in the range,
155: $M_{\rm vir}=(1.5-2.1)\times 10^{15}M_{\odot}$, and the combined ACS
156: and Subaru-2D mass reconstruction yields a tight constraint on the
157: concentration parameter, $c_{\rm vir}=12.7\pm 1\pm2.8$
158: ($c_{200}\sim 10$), improving upon the statistical accuracy of our
159: earlier 1D analysis.
160: Importantly, our best fitting profile properly reproduces
161: the observed Einstein radius of $45''$ ($z_s=1$), in contrast to
162: other weak lensing work, reporting lower concentration profiles,
163: which underestimate the observed Einstein radius.
164: \end{abstract}
165:
166: %% Keywords should appear after the \end{abstract} command. The uncommented
167: %% example has been keyed in ApJ style. See the instructions to authors
168: %% for the journal to which you are submitting your paper to determine
169: %% what keyword punctuation is appropriate.
170:
171: %% Authors who wish to have the most important objects in their paper
172: %% linked in the electronic edition to a data center may do so in the
173: %% subject header. Objects should be in the appropriate "individual"
174: %% headers (e.g. quasars: individual, stars: individual, etc.) with the
175: %% additional provision that the total number of headers, including each
176: %% individual object, not exceed six. The \objectname{} macro, and its
177: %% alias \object{}, is used to mark each object. The macro takes the object
178: %% name as its primary argument. This name will appear in the paper
179: %% and serve as the link's anchor in the electronic edition if the name
180: %% is recognized by the data centers. The macro also takes an optional
181: %% argument in parentheses in cases where the data center identification
182: %% differs from what is to be printed in the paper.
183:
184: %\keywords{globular clusters: general ---
185: %globular clusters: individual(\objectname{NGC 6397},
186: %\object{NGC 6624}, \objectname[M 15]{NGC 7078},
187: %\object[Cl 1938-341]{Terzan 8})}
188:
189:
190: %% Modified by KU (2007/07/21)
191: \keywords{cosmology: observations --- dark matter --- galaxies: clusters:
192: individual (A1689) --- gravitational lensing}
193:
194:
195: %% From the front matter, we move on to the body of the paper.
196: %% In the first two sections, notice the use of the natbib \citep
197: %% and \citet commands to identify citations. The citations are
198: %% tied to the reference list via symbolic KEYs. The KEY corresponds
199: %% to the KEY in the \bibitem in the reference list below. We have
200: %% chosen the first three characters of the first author's name plus
201: %% the last two numeral of the year of publication as our KEY for
202: %% each reference.
203:
204:
205: \section{Introduction}
206:
207:
208: Weak gravitational lensing of background galaxies provides a unique,
209: direct way to study the mass distribution of galaxy clusters
210: (Bartelmann \& Schneider 2001) via the systematic shape distortion of
211: background galaxies (Tyson, Wenk, \& Valdes 1990; Kaiser \& Squires
212: 1993; Schneider \& Seitz 1996; Umetsu, Tada, \& Futamase 1999) and
213: also to a lesser extent by the magnification of the background
214: (Broadhurst, Taylor, \& Peacock 1995; Taylor et al. 1998). We have
215: examined both these effects in our earlier work on A1689 (Broadhurst
216: et al. 2005a, hereafter B05a), where we found good consistency between
217: the magnitude of the radial depletion of background red galaxies
218: caused by lens magnification and the weak lensing distortion profile
219: of the same background galaxy population, which we then combined to
220: derive an improved mass profile.
221:
222: A limitation of the magnification technique is the intrinsic
223: clustering of the background, which for red galaxies is mainly in the
224: form of localized groups of modest angular size in the background
225: field. In principle, with redshift information sharp overdensities can
226: be isolated in redshift and down weighted, or with sufficiently large
227: number of redshifts, the shift of the magnified luminosity function
228: can be utilized independently of density fluctuations
229: (Broadhurst et al. 1995; Zhang \& Pen 2005).
230: The combination of all lensing related effects is of course
231: desirable, leading to the derivation of the highest precision feasible
232: when constructing mass maps and density profiles. Furthermore, because
233: lensing effects depend on distance, the cosmological redshift
234: distance relation may be constrained via the geometric scaling of the lensing
235: signal with redshift (e.g., Taylor et al. 2007; Medezinski et
236: al. 2007).
237:
238: Advances in the quality of imaging encourage a closer examination
239: of the empirical effects of lensing and the development of more
240: comprehensive techniques to extract reliable high resolution
241: information. From space, deep multi-color images of massive clusters
242: can be used to identify many sets of multiple images per cluster
243: (Broadhurst et al. 2005b, hereafter B05b; Gavazzi et al. 2003; Kneib et
244: al. 2004; Sand et al. 2004; Smith et al. 2005;
245: Brada${\check {\rm c}}$ et al. 2006).
246: From the ground the stable prime-focus wide-field cameras
247: of Subaru and CFHT are producing data of sufficient quality to examine
248: weak lensing distortions over a wide range of radius.
249: %(e.g.,van Waerbeke et al. 2000; Miyazaki et al. 2002; Sato et al. 2003;
250: %Hamana et al. 2003; Gavazzi et al. 2004; Hoekstra et al. 2006; Mahdavi
251: %et al. 2007; Medezinski et al. 2007; Okabe \& Umetsu 2008).
252: %%%
253: More recently, wide-field near IR cameras, such as MOIRCS on Subaru
254: WIRCAM on CFHT, OMEGA2000 on
255: Calar-Alto and WFCAM on UKIRT, may help improve the accuracy of
256: photometric redshifts for many faint galaxies. However, it is still
257: the case that no set of deep high quality wide field images exists for
258: any massive cluster with full optical-IR coverage, despite all the
259: progress of field surveys.
260:
261:
262:
263: A further motivation for pursuing accurate lensing maps is the
264: increased precision of model predictions for statistical properties
265: of cluster-sized mass halos in the standard Lambda cold dark matter
266: ($\Lambda$CDM) model.
267: Many of the free parameters of this model now rest on a firm empirical
268: foundation with relatively tight constraints on the index and
269: normalization of the power spectrum of density perturbation and the
270: background cosmological model (e.g., Spergel et al. 2003; Tegmark et
271: al. 2004; Spergel et al. 2007).
272: %%%
273: In this context $N$-body simulations have become impressively
274: comprehensive, in particular the recent Millennium simulation
275: (Springel et al. 2005) which simulates a huge volume of $500 {\rm
276: Mpc}/h$, and has been used to predict the mass function and evolution
277: of nearly 100,000 group and cluster sized CDM halos.
278: %%%%
279: This model is tightly defined and hence amenable to comparisons with
280: the real Universe, particularly for the case of clusters where baryons,
281: which are usually omitted from large scale simulation, are not expected
282: to have a significant impact on the shape of gravitational potential of
283: a cluster since the high temperature of the cluster gas prevents
284: efficient cooling, and hence the majority of baryons simply trace
285: the gravitational potential of the dominant dark matter.
286: %This model is tightly proscribed and is amenable to comparisons with
287: %the real Universe, particularly for the case of clusters where baryons
288: %which are absent from this simulation are not expected to have a
289: %significant impact on the gravitational potential of a cluster due to
290: %the high temperature of cluster gas preventing efficient cooling, so
291: %that the majority of baryons simply trace the gravitational potential
292: %of the dominant dark matter.
293: %%%%%
294: Accurate $N$-body simulations based on the $\Lambda$CDM scenario predict
295: a relatively shallow, low-concentration mass profile for massive cluster
296: halos, where the logarithmic gradient flattens continuously toward the
297: center of mass (Navarro, Frenk, \& White 1997, hereafter NFW)
298: with a flatter central slope than a purely isothermal
299: body interior to the inner characteristic radius,
300: $r_s \simlt 100-200 {\rm kpc}/h$.
301: %%%
302: A useful index of the degree of concentration, $c_{\rm vir}$,
303: compares the virial radius, $r_{\rm vir}$, to
304: $r_s$ of the NFW profile, $c_{\rm vir} \equiv r_{\rm vir}/r_s$.
305: %Massive clusters are of particular interest in the context of this
306: %model, because they are predicted to
307: %have a relatively shallow mass profile (or low concentration)
308: %fitted described by the form proposed by Navarro, Frenk, \& White
309: %(1996).
310: This prediction for the CDM halo $c_{\rm vir}$--$M_{\rm vir}$ relation
311: has been established thoroughly with high resolution
312: simulations (e.g., Navarro et al. 1997; Bullock et al. 2001; Wechsler et
313: al. 2002; Neto et al. 2007)
314: with some intrinsic variation related to the individual assembly
315: history of a cluster (e.g., Jing \& Suto 2000; Tasitsiomi et al. 2004;
316: Hennawi et al. 2007).
317: %%%
318: In particular, the detailed $N$-body millennium simulation (Springel et
319: al. 2005) predicts a simple relationship between the halo mass and
320: concentration parameter for halo masses in the range of galaxy groups
321: to massive clusters, as quantified by Neto et al. (2007),
322: who found that the expected median value for cluster sized halos
323: of $M_{\rm vir}\sim 10^{15}M_\odot$ to be $c_{\rm vir}\sim 5$ at $z=0$
324: ($c_{200}\sim 4$),
325: with a spread of the order of $\Delta\log_{10}c_{\rm vir}=0.1$.
326: (see Johnston et al. 2007 for a good summary of the state of art in halo
327: concentrations based on Wechsler et al. 2006 and Neto et al. 2007).
328:
329:
330: In this paper we explore further methods designed to achieve the
331: maximum possible lensing precision by combining all lensing
332: information for A1689. This cluster is among the most massive clusters
333: with the largest known Einstein radius ($\sim 50\arcsec$), and is one of
334: the best studied clusters for lensing work (Tyson et al. 1990;
335: Tyson \& Fisher 1995; Taylor et al. 1998; King, Clowe, \& Schneider
336: 2002; Bardeaul et al. 2005; B05a; B05b; Oguri et al. 2005; Halkola,
337: Seitz, \& Pannella 2006; Bardeau et al. 2007; Limousin et al. 2007;
338: Medezinski et al. 2007; Umetsu, Takada, \& Broadhurst 2007;
339: Okura, Umetsu, \& Futamase 2008), located at a moderately low redshift
340: of $z=0.183$.
341: %%%
342: In B05a we developed a
343: %{\bf maximum likelihood method for}
344: ``model-independent'' method\footnote{We remind the reader that model
345: dependence is unavoidable to some extent in scientific analysis.
346: In this work we define the term ``model independent'' to refer to
347: those methods without prior assumptions about the functional form
348: of the lensing profiles and distributions.}
349: for reconstructing the cluster mass profile using azimuthally-averaged
350: weak-lensing shape distortion and magnification bias measurements, in
351: the wide-field, Subaru images. This together with many multiple images
352: identified in deep {\it Hubble Space Telescope} (HST) Advanced Camera
353: for Surveys (ACS) imaging defined a detailed lensing based
354: cluster mass profile out to the cluster virial radius ($r\simlt 2
355: h^{-1}$ Mpc). The combined strong and weak lensing mass profile is
356: well fitted by an NFW profile (Navarro et al. 1997) with
357: high concentration of $c_{\rm vir}\sim 13.7$, which is significantly
358: larger than theoretically expected ($c_{\rm vir}\sim 5$) for the
359: standard $\Lambda$CDM model (Bullock et al. 2001; Neto et al. 2007),
360: although the degree of concentration is still controversial (B05a;
361: Medezinski et al. 2007; Limousin et al. 2007).
362: %%%
363: Such a high concentration is also seen in other massive clusters
364: from careful lensing work, such as MS 2137-23 ($c_{200}\simeq 12$,
365: Gavazzi et al. 2003) and CL0024+1654 ($c_{200}\simeq 22$, Kneib et
366: al. 2003). These results could raise serious questions regarding the
367: basic assumptions behind the $\Lambda$CDM model. If clusters collapse
368: earlier than predicted then it is expected that denser and hence more
369: concentrated halos will develop in the context of CDM (Wechsler et
370: al. 2002).
371: %%%
372: On the other hand, it has been argued that part of this discrepancy from
373: lensing observations could be reconciled by observational effects such
374: as triaxiality of CDM halos (Oguri et al. 2005;
375: Hennawi et al. 2007; Sereno 2007; Corless \& King 2007), and the
376: projection of structure along the line of sight (e.g., King \& Corless
377: 2007),
378: both of which boost the projected surface mass density and hence the
379: lensing signal. Such observational biases in the lensing-based
380: concentration parameter have been explored in details by
381: Hennawi et al. (2007) on the basis of $N$-body simulations,
382: indicating a positive bias of $\sim 30\%$ in the halo concentration
383: derived from 2D lensing measurements. Although A1689 is a very round
384: shaped cluster with evidence of only modest substructure (Teague,
385: Carter, \& Grey 1990; Girardi et al. 1997; Andersson \& Madejski 2004;
386: Czoske 2004; B05b), such a chance alignment of structure could be a
387: potential source of high concentrations.
388: %%%
389: Furthermore, for a reliable measurement of the cluster mass profile,
390: systematic errors inherent in the lensing measurements, such
391: as the uncertainty in the background redshift distribution
392: and the dilution effect on the lensing signal due to contamination by
393: cluster members (B05a; Medezinski et al. 2007), need to be taken into
394: account.
395:
396:
397:
398: The paper is organized as follows.
399: We briefly summarize in \S 2 the basis of cluster weak lensing.
400: %In \S 3 we describe the observational data, background sample selection,
401: %and joint weak lensing analysis of shape distortion and magnification
402: %bias data.
403: In \S 3 we describe the observational data and
404: the background sample selection for the weak
405: lensing analysis;
406: %%%
407: we then
408: %%% outline
409: summarize
410: our joint weak lensing analysis of shape distortion and magnification
411: bias data.
412: %A combind weak lensing shape analysis of A1689 is presented in \S 3.
413: %In \S 4 we describe our magnification bias measurements in A1689.
414: %%%
415: In \S 4 we present a method for reconstructing the two-dimensional mass
416: distribution of A1689 from combined weak lensing shape distortion and
417: magnification bias measurements.
418: %%
419: In \S 5 we derive mass profiles of A1689 from weak lensing data using
420: three different methods, and compare resulting mass profiles; we also
421: combine our weak lensing mass profiles with strong lensing constraints
422: from previous studies to test the CDM paradigm;
423: then, we assess carefully various
424: sources of potential systematic error in the halo concentration
425: parameter derived from the lensing observations.
426: Finally, summary and
427: discussions are given in \S 6.
428:
429: %%%
430:
431: Throughout this paper, we use the AB magnitude system, and adopt a
432: concordance $\Lambda$CDM cosmology with ($\Omega_{\rm m0}=0.3$,
433: $\Omega_{\Lambda 0}=0.7$, $h=0.7$).
434: %%%
435: In this cosmology one arcminute corresponds to the physical scale
436: $129$kpc$/h$ for this cluster. The reference center of our analysis
437: is fixed at the center of the cD galaxy: ${\rm RA} = 13:11:29.52, {\rm
438: Dec} = -01:20:27.59$ (J2000.0).
439:
440:
441:
442:
443: \section{Cluster Weak Lensing}
444: \label{sec:wl}
445:
446: Weak gravitational lensing is responsible for the weak shape-distortion
447: and magnification of the images of background sources due to the
448: gravitational field of intervening foreground clusters of galaxies
449: and large scale structures in the universe.
450: The deformation of the image can be
451: described by the $2\times 2$
452: Jacobian matrix $\cal{A}_{\alpha\beta}$
453: ($\alpha,\beta=1,2$)
454: of the lens mapping.
455: The Jacobian ${\cal A}_{\alpha\beta}$ is real and symmetric, so that
456: it can be decomposed as
457: \begin{eqnarray}
458: \label{eq:jacob}
459: {\cal A}_{\alpha\beta} &=& (1-\kappa)\delta_{\alpha\beta}
460: -\Gamma_{\alpha\beta},\\
461: \Gamma_{\alpha\beta}&=&
462: \left(
463: \begin{array}{cc}
464: +{\gamma}_1 & {\gamma}_2 \\
465: {\gamma}_2 & -{\gamma}_1
466: \end{array}
467: \right),
468: \end{eqnarray}
469: where
470: $\delta_{\alpha\beta}$ is Kronecker's delta,
471: $\Gamma_{\alpha\beta}$ is the trace-free, symmetric shear matrix
472: %defined as $\Gamma_{\alpha\beta}=\gamma_1\sigma_3+\gamma_2\sigma1$
473: with $\gamma_{\alpha}$ being the components of
474: spin-2
475: complex gravitational
476: shear $\gamma:=\gamma_1+i\gamma_2$,
477: describing the anisotropic shape distortion,
478: %and $\sigma_{\alpha}$ being the $2\times 2$ Pauli matrices,
479: and $\kappa$ is the
480: %$\kappa(\btheta)=\int\! d\Sigma\,\Sigma_{\rm crit}^{-1}$ is the
481: lensing convergence responsible for the
482: trace-part of the Jacobian matrix, describing the isotropic area
483: distortion.
484: In the weak lensing limit where $\kappa,|\gamma|\ll 1$,
485: $\Gamma_{\alpha\beta}$ induces a quadrupole anisotropy of the
486: background image, which can be observed from ellipticities
487: of background galaxy images.
488: %%%%
489: The flux magnification due to gravitational lensing
490: is given by the inverse Jacobian
491: determinant,
492: \begin{equation}
493: \label{eq:mu}
494: \mu =
495: \frac{1}{{\rm det}{\cal A}}
496: =
497: \frac{1}{(1-\kappa)^2-|\gamma|^2},
498: \end{equation}
499: where we assume subcritical lensing, i.e.,
500: ${\rm det}{\cal A}(\btheta)>0$.
501:
502:
503: The lensing convergence is expressed as a line-of-sight projection
504: of the matter density contrast out to the source plane (${\rm S}$)
505: weighted by certain combination $g$ of co-moving angular diameter distances
506: (e.g., Jain et al. 2000),
507: \begin{equation}
508: \label{eq:kappa}
509: \kappa =
510: \frac{3H_0^2\Omega_m}{2c^2}
511: \int_0^{\chi_{\rm S}}\!d\chi\, g(\chi,\chi_{\rm
512: S})\frac{\delta}{a}
513: \equiv \int\!d\Sigma_m\,\Sigma_{\rm crit}^{-1}
514: \end{equation}
515: %$\kappa(\btheta)=\int\! d\Sigma\,\Sigma_{\rm crit}^{-1}$,
516: where $a$ is the cosmic scale factor, $\chi$ is the co-moving distance;
517: $\Sigma_m$ is the surface mass density of matter, $\Sigma_m
518: =\int_0^{\chi_{\rm S}}\!d\chi \, a(\rho_m-\bar{\rho})$,
519: with respect to the cosmic mean density $\bar{\rho}$, and
520: %where
521: $\Sigma_{\rm crit}$
522: is the critical surface mass density for gravitational lensing,
523: \begin{equation}
524: \label{eq:sigmacrit}
525: \Sigma_{\rm crit} = \frac{c^2}{4\pi G}\frac{D_{s}}{D_d D_{ds}}
526: \end{equation}
527: with $D_s$, $D_d$, and $D_{ds}$ being the angular diameter distances
528: from the observer to the source, from the observer to the deflecting
529: lens, and from the lens to the source, respectively.
530: For a fixed background cosmology and a lens redshift $z_d$,
531: $\Sigma_{\rm crit}$ is a function of background source redshift
532: $z_s$.
533: %In the following we closely follow the standard notation of
534: %Bartelmann \& Schneider (2001).
535: For a given mass distribution $\Sigma(\btheta)$, the lensing signal is
536: proportional to the angular diameter distance ratio,
537: $D_{ds}/D_s$.
538: %\begin{equation}
539: %\label{eq:dratio}
540: %D(z_s) \equiv D_{ds}/D_s.
541: %\end{equation}
542:
543: In the present weak lensing study we aim to reconstruct
544: the dimensionless surface mass density $\kappa$
545: from weak lensing distortion and magnification data.
546: To do this, we utilize the relation between the gradients of
547: $\kappa$ and $\gamma$ (Kaiser 1995; Crittenden et al. 2002),
548: \begin{equation}
549: \label{eq:local}
550: \triangle \kappa (\btheta)
551: = \partial^{\alpha}\partial^{\beta}\Gamma_{\alpha\beta}(\btheta)
552: = 2\hat{\cal D}^*\gamma(\btheta)
553: \end{equation}
554: where
555: $\hat{\cal D}$ is the complex differential operator
556: $\hat{\cal D}=(\partial_1^2-\partial_2^2)/2+i\partial_1\partial_2$.
557: The Green's function for the two-dimensional Poisson equation is
558: $\triangle^{-1}(\btheta,\btheta')=\ln|\btheta-\btheta'|/(2\pi)$,
559: so that equation (\ref{eq:local}) can be solved to yield the following
560: non-local relation between $\kappa$ and $\gamma$ (Kaiser \& Squires 1993):
561: \begin{equation}
562: \label{eq:gamma2kappa}
563: \kappa(\btheta) =
564: \frac{1}{\pi}\int\!d^2\theta'\,D^*(\btheta-\btheta')\gamma(\btheta')
565: \end{equation}
566: where $D(\btheta)$ is the complex kernel defined as
567: \begin{equation}
568: \label{eq:kerneld}
569: D(\btheta) = \frac{\theta_2^2-\theta_1^2-2i\theta_1\theta_2}{|\theta|^4}.
570: \end{equation}
571: Similarly, the spin-2 shear field can be expressed in terms of the
572: lensing convergence as
573: \begin{equation}
574: \label{eq:kappa2gamma}
575: \gamma(\btheta) =
576: \frac{1}{\pi}\int\!d^2\theta'\,D(\btheta-\btheta')\kappa(\btheta').
577: \end{equation}
578: Note that adding a constant mass sheet to $\kappa$ in equation
579: (\ref{eq:kappa2gamma}) does not change the
580: shear field
581: $\gamma(\btheta)$ which is observable in the weak lensing limit,
582: leading to the so-called {\it mass-sheet degeneracy}
583: based solely on shape-distortion measurements (e.g., Bartelmann \&
584: Schneider 2001; Umetsu et al. 1999).
585: In general, the observable quantity is not the
586: gravitational shear $\gamma$ but the {\it reduced} shear,
587: \begin{equation}
588: \label{eq:redshear}
589: g=\frac{\gamma}{1-\kappa}
590: \end{equation}
591: in the subcritical regime where ${\rm det}{\cal A}>0$
592: (or $1/g^*$ in the negative parity region with ${\rm det}{\cal A}<0$).
593: We see that the reduced shear $g$ is invariant under the following
594: global transformation:
595: \begin{equation}
596: \label{eq:invtrans}
597: \kappa(\btheta) \to \lambda \kappa(\btheta) + 1-\lambda, \ \ \
598: \gamma(\btheta) \to \lambda \gamma(\btheta)
599: \end{equation}
600: with an arbitrary scalar constant $\lambda\ne 0$
601: (Schneider \& Seitz 1995). This transformation is equivalent to scaling
602: the Jacobian matrix ${\cal A}(\btheta)$ with $\lambda$,
603: $\cal {A}(\btheta) \to \lambda {\cal
604: A}(\btheta)$. This mass-sheet degeneracy can be unambiguously broken
605: by measuring the magnification effects, because the magnification $\mu$
606: transforms under the invariance transformation (\ref{eq:invtrans}) as
607: \begin{equation}
608: \mu(\btheta) \to \lambda^2 \mu(\btheta).
609: \end{equation}
610:
611:
612:
613:
614:
615:
616: \section{Data Analysis}
617: \label{sec:analysis}
618:
619: In this section we present a full technical description of our weak
620: lensing distortion and magnification analyses on A1689 based on the
621: Subaru images, which were analysed in our earlier work of B05a and
622: Medezinski et al. (2007).
623: This work is also based on the same weak lensing
624: shape and magnification measurements as used in
625: B05a. We note that Medezinski et al. (2007) used a slightly different
626: analysis pipeline so as to include the shape measurements of bright
627: cluster members and noisier objects (as well as better resolved red
628: background galaxies), optimizing for the weak lensing dilution analysis
629: including the measurements of cluster light and cluster luminosity
630: function where the completeness is crucial; Medezinski et al. (2007)
631: included a blue background population in the weak lensing shape
632: analysis, in addition to the red background population which B05a and
633: this work are based on; The magnification information, on the other
634: hand, was not taken into account in Medezinski et al. (2007).
635:
636: In B05a we simply assumed that the mean redshift of the red background
637: galaxies is $z_s=1$.
638: % based on deep photo-$z$ estimation for deep field
639: %data ($\bar z_s\simeq 1\pm 0.1$, Benitez et al. 2002).
640: In the present work we improve the accuracy of determination of the
641: cluster mass and concentration parameters, by taking into account the
642: redshift distribution of
643: red background galaxies
644: examined by Medezinski et al. (2007)
645: based on the multicolor photometry of Capak et al. (2004) in
646: the HDF-N (see \S \ref{subsec:red}).
647:
648:
649:
650: \subsection{Subaru Data and Photometry}
651: \label{subsec:data}
652:
653: For our weak lensing analysis
654: we used Subaru/Suprime-Cam imaging data of A1689 in $V$
655: %(1,920s)
656: and SDSS $i'$
657: %(2,640s)
658: retrieved from the Subaru archive,
659: SMOKA (see B05a and Medezinski et al. 2007 for more details).
660: The FWHM in the co-added mosaic image is
661: $0\arcsec\!\!.82$ in $V$ and
662: $0\arcsec\!\!.88$ in $i'$ with $0\arcsec\!\!.202$
663: pix$^{-1}$, covering a field of $\approx 30'\times 25'$.
664: %%%%
665:
666: Photometry is based on a combined $V+i'$ image with SExtractor
667: (Bertin \& Arnouts 1996), where the $i'$ image is used as
668: the source detection image.
669: We adopt the following key configuration parameters
670: in SExtractor:
671: ${\it DETECT\_MINAREA} = 5$,
672: ${\it DETECT\_THRESH}={\it ANALYSIS\_THRESH}=3$.
673: The limiting magnitudes are obtained as
674: $V=26.5$ and $i'=26.0$ for a $3\sigma$ detection within a $2$\arcsec
675: aperture.
676: A careful background selection is critical for a weak lensing analysis
677: so that unlensed cluster members and
678: foreground galaxies to not dilute the true lensing signal of the background
679: (see B05a and Medezinski et al. 2007).
680: %%%
681: We identify an E/S0 sequence of cluster galaxies in the
682: %%% @@ edited by KU (2008/02/18)
683: %% color-magnitude (CM) diagram, which can be defined by the linear
684: color-magnitude (CM) diagram, which can be defined by the linear
685: CM relation:
686: $(V-i')_{E/S}=-0.02094i'+1.255$ (B05a;
687: see Figure 1 of Medezinski et al. 2007 for the CM diagram).
688: %% We note that Medezinski et al. (2007)
689: %% used slightly different values of the CM parameters for A1689.
690: For the number counts to measure lensing magnification,
691: we define a sample of galaxies that are redder than the
692: cluster sequence by
693: $(V-i')>1.0$ and $20<i'<25.5$,
694: which yields a total of $N_\mu= 8,907$ galaxies,
695: or the mean surface number density of $\bar{n}_{\mu}=12.0$ arcmin$^{-2}$.
696: %%%
697: For the magnification bias analysis, a conservative
698: magnitude limit of $i'<25.5$
699: is adopted to avoid incompleteness.
700:
701:
702:
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704:
705:
706: \subsection{Weak Lensing Distortion Analysis}
707: \label{subsec:shear}
708:
709: We use the IMCAT package developed by N. Kaiser
710: \footnote{http://www.ifa.hawaii/kaiser/IMCAT} to perform object
711: detection, photometry and shape measurements, following the formalism
712: outlined in Kaiser, Squires, \& Broadhurst (1995; hereafter KSB).
713: We have modified the method somewhat following the procedures
714: described in Erben et al. (2001). We used the same analysis pipeline
715: as in
716: B05a,
717: Umetsu et al. (2007),
718: and Okabe \& Umetsu (2008).
719:
720:
721: \subsubsection{Object Detection}
722: \label{subsubsec:detection}
723:
724:
725: Objects are first detected as local peaks in the image by
726: using the IMCAT hierarchical peak-finding
727: algorithm {\it hfindpeaks}
728: which for each object yields object parameters such as a peak position,
729: an estimate of the object size ($r_g$), the significance of
730: the peak detection ($\nu$).
731: %%%
732: The local sky level and its gradient are
733: measured around each object from the mode of pixel values
734: on a circular annulus defined by
735: inner and outer radii of
736: $16\times r_g$ and $32\times r_g$ (see Clowe et al. 2000).
737: In order to avoid contamination in the background estimation
738: by bright neighboring stars and/or
739: foreground galaxies,
740: all pixels within $3\times r_g$ of another object are excluded from the
741: mode calculation. Total fluxes and half-light radii ($r_h$) are then
742: measured on sky-subtracted images using a circular aperture of
743: radius $3\sqrt{2}\times r_g$ from the object center. Any pixels within
744: $2.5\times r_g$ of another object are excluded from the aperture.
745: The aperture $i'$-magnitude is then calculated from the measured total
746: flux and a zero-point magnitude.
747: Any objects with positional differences between the
748: peak location and the weighted-centroid
749: greater than $d=0.4$ pixels are excluded from
750: the catalog.
751:
752:
753: Finally, bad objects such as spikes, saturated stars, and noisy
754: detections
755: %and groups
756: %of objects detested as a single object
757: must be removed from the weak lensing object
758: catalog.
759: %%%
760: We removed from our object catalog
761: (1) extremely large objects with $r_g>10$ pixels,
762: (2) objects with low detection significance, $\nu<7$,
763: (3) objects with large raw ellipticities, $|e|>0.5$,
764: (4) noisy detections with unphysical negative fluxes, and
765: (5) objects containing more than $10$ bad pixels, ${\it nbad}>10$.
766: %%%%
767: This selection procedure yields an object catalog with
768: $N=62,384$ ($82.6$ arcmin$^-2$).
769:
770:
771:
772:
773:
774: \subsubsection{Weak Lensing Distortion Measurements}
775: \label{subsubsec:shape}
776:
777:
778: To obtain an estimate of the reduced shear,
779: $g_{\alpha}=\gamma_{\alpha}/(1-\kappa)$ ($\alpha=1,2$),
780: we measure the image ellipticity
781: $e_{\alpha} = \left\{Q_{11}-Q_{22}, Q_{12} \right\}/(Q_{11}+Q_{22})$
782: from the weighted quadrupole moments of the surface brightness of
783: individual galaxies defined in the above catalog,
784: \begin{equation}
785: Q_{\alpha\beta} = \int\!d^2\theta\,
786: W({\theta})\theta_{\alpha}\theta_{\beta}
787: I({\btheta})
788: \ \ \ (\alpha,\beta=1,2)
789: \end{equation}
790: where $I(\btheta)$ is the surface brightness distribution of an object,
791: $W(\theta)$ is a Gaussian window function matched to the size of the
792: object.
793:
794:
795: Firstly the PSF anisotropy needs to be corrected using the star images
796: as references:
797: \begin{equation}
798: e'_{\alpha} = e_{\alpha} - P_{sm}^{\alpha \beta} q^*_{\beta}
799: \label{eq:qstar}
800: \end{equation}
801: where $P_{sm}$ is the {\it smear polarizability} tensor
802: being close to diagonal, and
803: $q^*_{\alpha} = (P_{sm}^*)^{-1}_{\alpha \beta}e_*^{\beta}$
804: is the stellar anisotropy kernel.
805: We select bright, unsaturated foreground stars
806: of $20\simlt i' \simlt 22.5$
807: identified in a branch
808: of the half-light radius ($r_h$) vs. magnitude ($i'$) diagram
809: %($20<i'<22.5$, $\left<r_h\right>_{\rm median}=2.38$ pixels)
810: to measure $q^*_{\alpha}$.
811: %%%
812: In order to obtain a smooth map of $q^*_{\alpha}$ which is used in
813: equation (\ref{eq:qstar}), we divided the co-added mosaic image of
814: $9{\rm K}\times 7.4{\rm K}$ pixels into $5\times 4$ blocks,
815: each with $1.8{\rm K}\times 1.85{\rm K}$
816: pixels. The block length is based on the typical coherent
817: scale of PSF anisotropy patterns. In this way the PSF anisotropy in
818: individual blocks can be well described by fairly low-order
819: polynomials.
820: %
821: We then fitted the $q^*$ in each block independently with
822: second-order bi-polynomials, $q_*^{\alpha}(\btheta)$, in
823: conjunction with iterative $\sigma$-clipping rejection on each
824: component of the residual:
825: $\delta e^* = e^*_{\alpha}-(P_{sm}^*)^{\alpha\beta}q^*_{\beta}(\btheta)$.
826: %%%%
827: The final stellar sample consists of 540 stars
828: (i.e., $N_* \sim 30$ stars per block),
829: or the mean surface number
830: density of $\bar{n}_*=0.72$ arcmin$^{-2}$.
831: %%%%%
832: It is worth
833: noting that the mean stellar ellipticity before correction is
834: $(\bar{e_1}^*, \bar{e_2}^*) \simeq (-0.013, -0.018)$
835: over the data field, while the residual
836: $e^*_{\alpha}$ after correction
837: is reduced to
838: $ {\bar{e}^{*{\rm res}}_1} = (0.47\pm 1.32)\times 10^{-4}$,
839: $ {\bar{e}^{*{\rm res}}_2} = (0.54\pm 0.94)\times 10^{-4}$;
840: The mean offset from the null expectation is reduced down to
841: $|\bar{e}^{* \rm res}| = (0.71\pm 1.12) \times 10^{-4}$.
842: %%%%%%
843: On the other hand, the rms value of stellar ellipticities,
844: $\sigma_{e*}\equiv \sqrt{\left<|e^*|^2\right>}$, is reduced from $2.64\%$ to
845: $0.38\%$ when applying the anisotropic PSF correction.
846: %%%%%
847: We show in Figure \ref{fig:anisopsf1}
848: the quadrupole PSF anisotropy field
849: as measured from stellar ellipticities before and after the anisotropic
850: PSF correction.
851: Figure \ref{fig:anisopsf2} shows
852: the distribution of stellar ellipticity
853: components before and after the PSF anisotropy correction.
854: %%%%
855: From the rest of the object
856: catalog, we select objects with
857: $\bar{r}_{h*} \lesssim r_h \lesssim 10$ pixels
858: as an $i'$-selected weak lensing galaxy sample,
859: where $\bar{r}_{h*}\approx 2.4$ pixels
860: is the median value of stellar half-light radii,
861: corresponding to half the median width of circularized PSF
862: over the data field.
863: An apparent magnitude cutoff
864: of $20 \simlt i' \simlt 26$
865: is also made to remove from
866: the weak lensing galaxy sample bright foreground/cluster galaxies
867: and very faint galaxies with noisy shape measurements.
868: %%%%
869:
870: %%%
871: Second, we need to correct image ellipticities for
872: the isotropic smearing effect
873: caused by atmospheric seeing and the window function used for the shape
874: measurements. The pre-seeing reduced shear $g_\alpha$ can be
875: estimated from
876: \begin{equation}
877: \label{eq:raw_g}
878: g_{\alpha} =(P_g^{-1})_{\alpha\beta} e'_{\beta}
879: \end{equation}
880: with the {\it pre-seeing shear polarizability} tensor
881: $P^g_{\alpha\beta}$.
882: We follow the procedure described in Erben et al. (2000)
883: to measure
884: $P^g$.
885: %%%
886: We adopt the scalar correction scheme, namely
887: \begin{equation}
888: (P_g)_{\alpha\beta}
889: =\frac{1}{2}{\rm tr}[P_g]\delta_{\alpha\beta}\equiv
890: P^{\rm s}_{g}\delta_{\alpha\beta}
891: \end{equation}
892: (Erben et al. 2000;
893: Hoekstra et al. 1998;
894: Hudson et al. 1998;
895: Okabe \& Umetsu 2008).
896:
897:
898: In order to suppress artificial effects due to the noisy
899: $P_g^{\rm s}$ estimated for individual galaxies, we apply filtering
900: to raw $P_g^{\rm s}$ measurements. Firstly, we discard those noisy
901: objects which have negative raw $P_g^{\rm s}$ values.
902: %%%%%%%%%%%%%%%%%%%
903: Secondly,
904: we compute for each object
905: a median value
906: of $P_g^{\rm s}$ among $N$-neighbors in the size $r_g$ and magnitude
907: $i'$ plane to define object parameter space: for each object,
908: $N$-neighbors are identified in the size ($r_g$) and magnitude ($i'$)
909: plane;
910: the median value of $P_g^{\rm s}$ is used as the smoothed
911: $P_g^{\rm s}$ for the object, $\langle P_g^{\rm s}\rangle$,
912: and the variance $\sigma^2_{g}$
913: of $g=g_1+ig_2$ is calculated using equation (\ref{eq:raw_g}).
914: %%%
915: The dispersion $\sigma_g$ is used as an rms error of the shear estimate
916: for individual galaxies.
917: We take $N=30$.
918: After filtering noisy $P_g^{\rm s}$ measurements,
919: the minimum value of $\langle P_g^{\rm s}\rangle$ is $\approx 0.035$.
920: %%%%%
921: Figure \ref{fig:Pg} shows the averaged $\langle P_g^{\rm s}\rangle$
922: as a function of object size $r_g$ for the $i'$-selected
923: weak lensing galaxy sample.
924: The mean of
925: $\langle P_g^{\rm s}\rangle$ over all galaxies in the sample is obtained
926: as $0.307$, mostly weighted by galaxies with $r_g = 2-3$ pixels.
927: %%%
928: The mean variance $\overline{\sigma_g^2}$
929: over the galaxy sample is obtained as
930: $\simeq 0.152$, or $\sqrt{\overline{\sigma_g^2}}\approx 0.39$.
931: %%%
932: Finally, we use the following estimator for
933: the reduced shear:
934: $g_{\alpha} = e'_{\alpha}/\left< P_g^{\rm s}\right>$.
935: %%%
936: The final $i'$-selected galaxy sample
937: contains $30,369$
938: galaxies or $\bar{n}_g\simeq 40.3$ arcmin$^{-2}$.
939:
940:
941:
942:
943:
944:
945: \subsection{Red Background Selection}
946: \label{subsec:red}
947:
948: As demonstrated by B05a
949: and Medezinski et al. (2007), it is crucial
950: in the weak lensing analysis
951: to make a secure selection of background galaxies in order
952: to minimize contamination by
953: cluster/foreground galaxies and hence to make an accurate determination
954: of the cluster mass,
955: otherwise dilution of the distortion signal results from
956: the inclusion of unlensed cluster galaxies, particularly at small radius
957: where the cluster is relatively dense.
958: %%%
959: This dilution effect is
960: simply to reduce the strength of
961: the lensing signal when averaged over a local ensemble of
962: galaxies,
963: in proportion to the fraction of unlensed cluster/foreground galaxies
964: %cluster/foreground galaxies
965: whose orientations are randomly distributed, thus diluting the lensing
966: signal relative to the reference background level derived from
967: the background population (Medezinski et al. 2007).
968: %%%%%
969: With a pure red background sample (B) as a reference,
970: one can quantify the {\it degree of dilution} for a galaxy sample (G)
971: containing $N_{\rm CL}$ cluster galaxies and $N_{\rm BG}$
972: background galaxies
973: in terms of the strengths of the averaged tangential shear signal
974: $\langle g_+(\theta)\rangle$
975: as (Medezinski et al. 2007)
976: \begin{equation}
977: 1+\delta_d(\theta) \equiv \frac{N_{\rm BG}+N_{\rm CL}}{N_{\rm BG}}
978: =\frac{\langle g_+^{(B)}(\theta)\rangle}
979: {\langle g_+^{(G)}(\theta)\rangle}
980: \frac{\langle D_{ds}/D_s\rangle^{(G)}_{z_s>z_d}}
981: {\langle D_{ds}/D_s\rangle^{(B)}_{z_s>z_d}},
982: \end{equation}
983: where $\langle D_{ds}/D_s\rangle_{z_s>z_d}$'s are
984: averaged distance ratios
985: for respective background populations;
986: if the two samples contain the same
987: background population, then
988: $\delta_d =
989: \langle g_+^{(B)}\rangle/\langle g_+^{(G)}\rangle-1$.
990: %%%
991: The degree of dilution thus varies depending on the radius from
992: the cluster center, increasing towards the cluster center.
993: %%%%%%%%%%%%%%%%%%%%%%%%%
994: Medezinski et al. (2007) found
995: for their {\it green} galaxy sample
996: ($[V-i']_{E/S0-0.3}^{+0.1}$) containing
997: the cluster sequence galaxies in A1689
998: that the fraction of cluster membership,
999: $N_{\rm CL}/(N_{\rm BG}+N_{\rm CL})$,
1000: tends $\sim 100\%$ within $\theta \simlt 2'$.
1001:
1002:
1003:
1004: %%%%%
1005: For our weak lensing distortion analysis
1006: we define a sample of {\it red background galaxies}
1007: whose colors are redder due to large $k$-corrections than
1008: %% @@ edited by KU (2008/02/18)
1009: %% the color-magnitude (CM) sequence of cluster member galaxies.
1010: the CM relation, or red sequence, of cluster member galaxies.
1011: These red background galaxies are largely composed of early to mid-type
1012: galaxies at moderate redshifts (Medezinski et al. 2007).
1013: %%%%
1014: Cluster member
1015: galaxies are not expected to extend to these colors in any significant
1016: numbers because the intrinsically reddest class of cluster galaxies,
1017: %% @@ edited by KU (2008/02/18)
1018: %% i.e. E/S0 galaxies, are defined by the CM sequence and lie blueward
1019: %% of
1020: i.e. E/S0 galaxies, are defined by the red sequence and lie blueward of
1021: chosen sample limit, so that even large photometric errors will not
1022: carry them into our red sample.
1023: %%%% @@ KU (2008/03/29)
1024: This can be demonstrated readily, as shown in Figure \ref{fig:dilution},
1025: where we plot the mean tangential shear strength $\langle g_+ \rangle$,
1026: averaged over a wide radial range of $1\arcmin < \theta < 18\arcmin$,
1027: %%the central $10\arcmin$ region ($1\arcmin < \theta < 10\arcmin$) as
1028: as a function of color limit
1029: by changing the lower color limit progressively blueward.
1030: Here we do not apply area weighting to enhance the effect of dilution in
1031: the central region.
1032: %, finding a sharp drop in the lensing signal
1033: %at our limit, $(V-i')<(V-i')_{E/S0}+\Delta(V-i')$
1034: %when the cluster red sequence starts to
1035: %contribute significantly, thereby reducing the mean lensing signal.
1036: %%%
1037: We take the lower (bluer) color limit of $+0.22$ mag
1038: where the cluster contribution is negligible,
1039: %and the upper color limit of $6$ to include the
1040: %majority of the background red population,
1041: defining the red background sample by
1042: $\Delta(V-i')\equiv (V-i')-(V-i')_{E/S0} > 0.22$,
1043: as adopted by B05a.
1044: Figure \ref{fig:dilution} shows a sharp drop in the
1045: lensing signal at $\Delta(V-i') \simlt 0.1$,
1046: when the cluster red sequence starts to contribute significantly,
1047: thereby reducing the mean lensing signal.
1048: At $\Delta(V-i')\simgt 0.1$,
1049: the mean lensing signal of the red background
1050: stays fairly constant, $\langle g_+ \rangle \simeq 0.143$,
1051: ensuring that our weak lensing measurements
1052: are not sensitive to this particular choice of the color limit
1053: (see \S \ref{subsec:sys}).
1054: %%%%
1055: Similarly, on the left of Figure \ref{fig:dilution}, we show the mean
1056: distortion strength for our blue galaxy sample with the magnitude
1057: limit in the interval $23 < i' < 25.5$, so as to take only faint blue
1058: galaxies. The lower color limit is set to $\Delta(V-i')>-1.5$.
1059: For galaxies with colors bluer than the
1060: cluster sequence, cluster galaxies are present along
1061: with background
1062: galaxies, since the cluster population extends to bluer colors
1063: of the later type members.
1064: Consequently, the mean lensing signal of the blue sample
1065: is systematically lower than that of the red sample, unless we take the
1066: upper (redder) color limit around $\Delta(V-i')\sim 0.7$ where, however,
1067: the distortion measurement is quite noisy (Figure \ref{fig:dilution}).
1068: We therefore exclude blue galaxies from our weak lensing analysis.
1069: Note that
1070: the background populations do not need to be complete
1071: in any sense but should simply be well defined and contain only
1072: background.
1073: %%%
1074: This color-magnitude selection criteria yielded a total of
1075: $N_g= 5728$ galaxies, or the mean surface number density of
1076: $\bar{n}_g=7.6$ galaxies arcmin$^{-2}$.
1077: For the red background sample, we found
1078: %the mean variance $\bar{\sigma}_g^2$
1079: %over the red galaxy sample is obtained as
1080: $\overline{\sigma_g^2}\simeq 0.133$, or
1081: $\sqrt{\overline{\sigma_g^2}}\approx 0.36$, which is slightly
1082: smaller than the rms dispersion for the $i'$-selected galaxy sample.
1083:
1084:
1085: We need to estimate the depths of our color-magnitude selected
1086: red samples when measuring the cluster mass profile,
1087: because the lensing signal depends on the source redshifts in proportion
1088: to $D_{ds}/D_s$.
1089: %%%%%
1090: Medezinski et al. (2007) utilized the multicolor
1091: photometry of Capak et al. (2004)
1092: based on Subaru $UBVRIZ$ imaging covering $0.2$ deg$^2$, and
1093: estimated
1094: photometric redshifts for color-magnitude
1095: selected background galaxy samples.
1096: %%%%
1097: Using the Capak et al.'s photometric redshift distributions,
1098: Medezinski et al. (2007) estimated a mean redshift for the red
1099: background with $i'>18$ mag to be
1100: $\bar{z}_s \approx 0.87$ at fainter magnitude limits of $i'=26$--$27$,
1101: and also calculated weighted mean lensing depths $\langle
1102: D_{ds}/D_s\rangle$ for respective background populations as a function
1103: of apparent $i'$ limiting magnitude, and found that the averaged
1104: distance ratio $\langle D_{ds}/D_s\rangle$ grows only slowly with
1105: increasing apparent magnitude limit:
1106: $\langle D_{ds}/D_s \rangle = 0.693 \pm 0.02$ for
1107: the red sample with $i'_{\rm cut}=25.5$--$26.5$
1108: (see Figure 9 of Medezinski et al. 2007),
1109: and the corresponding redshift equivalent to this mean distance is
1110: $z_{s,D}=0.68 \pm 0.05$, where the distance-equivalent redshift $z_{s,D}$ is
1111: defined by the following equation:
1112: \begin{equation}
1113: \label{eq:zD}
1114: \bigg\langle \frac{D_{ds}}{D_s} \bigg\rangle_{z_s} =
1115: \frac{D_{ds}}{D_s}\bigg|_{z_s=z_{s,D}}.
1116: \end{equation}
1117: %%%
1118: Therefore, we can safely assume that our two different red samples for
1119: the distortion and magnification measurements (with $i'_{\rm cut}=26.0$
1120: and $25.5$, respectively) have nearly the same depths.
1121: %%%
1122: We note that B05a assumed that all of the background galaxies are
1123: located at a single redshift of $z_s=1$, corresponding to
1124: $D_{ds}/D_s=0.772$, which underestimates $\Sigma_{\rm crit}$ by
1125: $\sim 11\%$
1126: %%%($0.772/0.693=1.11$),
1127: and hence underestimates the cluster mass accordingly.
1128: %%%%
1129: We will come back to this issue in \S \ref{subsec:sys}.
1130:
1131:
1132:
1133: \section{Weak Lensing Map-making}
1134: \label{sec:map}
1135:
1136: This section describes a maximum entropy method (MEM) for reconstructing
1137: the two-dimensional cluster mass distribution from combined shape
1138: distortion and magnification bias observations.
1139:
1140: As described above (see \S \ref{subsec:red}), only the red selected
1141: background is used for the measurement of the lens distortion and
1142: magnification, in order to minimize the effect of dilution.
1143:
1144:
1145: \subsection{Shape Distortion Data}
1146: \label{subsec:distortion}
1147:
1148: For map-making,
1149: we pixelize the distortion data into a regular grid
1150: of $N_{\rm data} = 21\times 17 = 357$ independent pixels
1151: covering a field of $\approx 30'\times 24'$.
1152: %%%
1153: The pixel size is set to
1154: $\Delta_{\rm pix}=1\farcm 4$,
1155: and the mean
1156: galaxy counts per pixel is $\bar{n}_g\Delta\Omega_{\rm pix} \sim 15$
1157: with $\Delta\Omega_{\rm pix}=\Delta_{\rm pix}^2$ being
1158: the solid angle per pixel.
1159:
1160: We compute an estimate $\tilde g_{\alpha}(\btheta_i)$ for the reduced
1161: shear $g_{\alpha}(\btheta_i)$ on a regular grid of cells
1162: ($i=1,2,...,N_{\rm pix}$) as
1163: \begin{equation}
1164: \label{eq:bin_shear}
1165: \tilde g_{\alpha}(\btheta_i)
1166: =
1167: \sum_{k\in {\rm cell} i} u_k g_{\alpha,k}
1168: \Big/
1169: \sum_{k\in {\rm cell} i} u_k
1170: \equiv \tilde{g}_{\alpha,i},
1171: \end{equation}
1172: where
1173: $g_{\alpha,k}$ is a noisy estimate of the $\alpha$th component of the
1174: reduced shear for the $k$th galaxy, and
1175: $u_k=1/\sigma_{g,k}^2$ is its inverse-variance weight
1176: (see \S \ref{subsec:shear}).
1177: %%%
1178: The error covariance matrix
1179: for the pixelized reduced shear
1180: $\tilde{g}=\tilde{g}_1+i\tilde{g}_2$
1181: (\ref{eq:bin_shear}) is then
1182: diagonal, and
1183: given by
1184: \begin{equation}
1185: \langle
1186: \Delta\tilde{g}_{\alpha,i}\Delta\tilde{g}_{\beta,j}
1187: \rangle
1188: =
1189: \frac{1}{2}
1190: \sigma^2_{\tilde{g},i}
1191: \delta_{\alpha\beta}
1192: \delta_{ij},
1193: \end{equation}
1194: %where $\delta_{ij}$ is the Kronecker's delta, and
1195: where $\sigma^2_{\tilde{g}, j}$ is the error variance for the $j$th
1196: pixel defined as
1197: \begin{equation}
1198: \label{eq:bin_shearvar}
1199: \sigma^2_{\tilde{g}}(\btheta_j) =
1200: %\frac{\sum_{k\in {\rm cell} j} u_{k}^2 \sigma^2_{g,k}}
1201: \frac{1}
1202: %{ \left(
1203: {\sum_{k\in {\rm cell} j} u_{k}}.
1204: % \right)^2}.
1205: \end{equation}
1206: Here we have used $\langle \Delta g_{\alpha,k}\Delta g_{\beta,l}
1207: \rangle = (1/2)\sigma^2_{g,k} \delta_{\alpha\beta}\delta_{kl}$
1208: for individual shear estimates $g_{\alpha,k}$.
1209: %%%%
1210: The per-pixel rms dispersion for $\tilde{g}(\btheta)$
1211: is then reduced down to $\sqrt{\overline{\sigma^2_g}/N}
1212: \sim 0.36/\sqrt{15}\sim 0.092$.
1213:
1214:
1215: %%%%
1216: In Figure \ref{fig:shear} we show the reduced-shear field
1217: obtained from the red galaxy sample, where for visualization purposes
1218: the $\tilde{g}_{\alpha}(\btheta)$ is resampled on to a finer grid and
1219: smoothed with a Gaussian with ${\rm FWHM}=2'$.
1220: A coherent tangential shear pattern is clearly seen
1221: in Figure \ref{fig:shear} around the cluster center.
1222:
1223:
1224: \subsection{Magnification Bias Data}
1225: \label{subsec:mag}
1226:
1227: %The magnification bias arises because gravitational lensing changes the
1228: %apparent solid angle of the background but conserves the surface
1229: %brightness, which influences the observed surface density of
1230: %background sources,
1231: Lensing magnification, $\mu(\btheta)$, influences the observed
1232: surface density of background sources, expanding the area of
1233: sky, and enhancing the observed flux of background sources.
1234: In the subcritical regime
1235: (Broadhurst, Taylor \& Peacock 1995; Umetsu et al. 1999),
1236: the magnification $\mu(\btheta)$ is given by equation (\ref{eq:mu}).
1237: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1238: The count-in-cell statistic is measured from
1239: the flux-limited red galaxy sample
1240: (see \S \ref{subsec:data}) on the same grid as the distortion data:
1241: \begin{equation}
1242: \label{eq:cic}
1243: \tilde{N}(<m_{\rm cut}; \btheta_i) = \sum_{k\in {\rm cell} i}1
1244: \equiv \tilde{N}_i
1245: \end{equation}
1246: with $m_{\rm cut}$ being the magnitude cutoff corresponding to the
1247: flux-limit ($i'_{\rm cut}=25.5$).
1248: %%%
1249: The normalization and slope of the unlensed number counts
1250: $N_0(<m_{\rm cut})$ for our red galaxy sample are reliably estimated as
1251: $n_{\mu,0}=12.6\pm 0.23$ arcmin$^{-2}$ and
1252: $s\equiv d\log{N_0(m)}/dm=0.22\pm 0.03$
1253: from the outer region $\ge 10'$ (B05a).
1254: %%%
1255: The mean galaxy counts per pixel is thus
1256: $N_0 = n_{\mu,0}\Delta\Omega_{\rm pix}\sim 25$.
1257: %%%
1258: The magnification bias at the $i$th cell is then estimated as
1259: $1+\tilde\delta_{\mu,i} =
1260: \tilde{n}_{\mu}/n_{\mu,0}
1261: =
1262: \tilde{N}_i/N_0$,
1263: with $\tilde n_{\mu,i}=\tilde N_{\mu,i} / \Delta\Omega_{\rm pix}$,
1264: where the dilution effect $\delta_d$ is negligible for our
1265: red-background
1266: sample; otherwise $\tilde N_i/N_0 = 1+\tilde{\delta}_\mu+\tilde{\delta}_d$.
1267: %%%
1268: The slope is less than the lensing invariant slope, $s=0.4$,
1269: and hence a net deficit of background galaxies is expected:
1270: \begin{equation}
1271: \label{eq:magbias}
1272: \delta_{\mu}(\btheta) \equiv \langle \tilde\delta_{\mu}(\btheta)\rangle
1273: = \mu_i^{2.5s-1}(\btheta)-1
1274: \end{equation}
1275: (Broadhurst et al. 1995; B05a).
1276: %%%%
1277: In the limit of weak lensing
1278: where $\kappa,|\gamma|\ll 1$,
1279: $\delta_{\mu}(\btheta) \simeq (5s-2) \kappa(\btheta)
1280: \simeq -0.9\kappa(\btheta)$ with $s=0.22$.
1281: %%%
1282: %%%
1283:
1284:
1285: The masking effect due to bright cluster galaxies and bright
1286: foreground objects is properly taken into
1287: account and corrected for (B05a).
1288: %%%%
1289: We conservatively account for the masking of observed sky
1290: by excluding a generous area $\pi a b$ around each masking object, where
1291: $a$ and $b$ are defined as $\nu_{\rm mask}\equiv 3$ times
1292: the major ({\tt A\_IMAGE}) and minor axes ({\tt B\_IMAGE}) computed
1293: from SExtractor, corresponding roughly
1294: to the isophotal detection limit in our configuration
1295: (see \S \ref{subsec:data}).
1296: %%%
1297: We calculate the correction factor for this masking effect
1298: as a function of radius from the cluster center,
1299: and renormalize the number density of each cell accordingly.
1300: %%%%
1301: The masking area is negligible at large radii, and increases up to
1302: $\sim 20\%$ of the sky close to the cluster center, $\theta\simlt 3'$.
1303: %%%
1304: B05a showed that
1305: the magnification bias measurements with and without the masking
1306: correction are roughly consistent with each other, and with the NFW
1307: prediction from the ACS strong lensing observations (B05b)
1308: and the Subaru distortion profile (see Figure 2 of B05a),
1309: even though all the measurements have different
1310: systematics.
1311: Note that if we use the masking factor $\nu_{\rm mask}$ of $2$ or $4$,
1312: instead of 3,
1313: the results shown below remain almost unchanged (see \S \ref{subsec:sys}
1314: for more details).
1315: %%%%
1316: %We found that
1317: %the results are little
1318: %changed when the masking factor of 2 or 4 are adopted instead of 3.
1319: %The magnificaiton-bias map is displayed in Figure \ref{fig:magbias}.
1320: %%%%
1321:
1322:
1323: Figure \ref{fig:magbias} shows the resulting magnification-bias
1324: distribution
1325: derived from the red galaxy sample based on the SExtractor
1326: photometry (\S \ref{subsec:data}).
1327: %%%
1328: A clear depletion of the red galaxy
1329: counts is visible in the central, high-density region of the cluster.
1330: %%% :: Keiichi's original ::
1331: %A clear depletion of the red galaxy
1332: %counts is seen in the central, high-density region of the cluster.
1333: %%%
1334: On the other hand,
1335: it has been argued that estimates of the lensing magnification
1336: based on number counts suffer from noise
1337: arising from the intrinsic clustering of the source galaxies
1338: (e.g., Zhang \& Pen 2006).
1339: Indeed,
1340: some
1341: variance is apparent in the spatial distribution of red galaxies.
1342: In particular, a local enhancement
1343: %$n_{\mu}/n_0\sim 1.5$
1344: of red background galaxies
1345: can be seen in Figure \ref{fig:magbias} around $\btheta \sim (0',5')$,
1346: having an overdensity of $\tilde n_{\mu}/n_0 \sim 1.5$.
1347: This may explain the discrepant point at $\theta\sim 5'$ in the
1348: magnification bias profile of B05a.
1349: %Estimates of the lensing magnification
1350: %based on number counts suffer from noise
1351: %arising from the intrinsic clustering of the source galaxies,
1352: %To overcome this intrinsic clustering problem
1353: %we wil
1354: %intrinsic clustering signal is ..
1355: %Such an intrinsic clustering signal of red galaxies ...
1356: %must be unweighted
1357: We will come back to this issue later in \S \ref{subsec:declustering}.
1358: %%%%
1359: %Indeed, in the one dimensional analysis of
1360: %Broadhurst et al. (2005b), we found a discrepant point at $\theta\sim
1361: %5'$ in the magnification-bias profile (Figure 2 of Broadhurst et
1362: %al. 2005b).
1363:
1364:
1365:
1366: \subsection{Maximum Entropy Mass Reconstruction Method}
1367: \label{subsec:massrec}
1368:
1369: The relation between distortion and convergence is non-local, and
1370: the convergence $\kappa$ derived from distortion data alone suffers from
1371: a mass sheet degeneracy (see \S \ref{sec:wl}).
1372: However, by combining the distortion and magnification measurements the
1373: convergence can be obtained unambiguously with the correct mass
1374: normalization.
1375: %%%%
1376:
1377: Here we combine pixelized distortion and depletion data of red
1378: background galaxies in a maximum likelihood sense, and reconstruct the
1379: two-dimensional distribution of the lensing convergence field,
1380: $\kappa(\btheta)$.
1381: %%%
1382: Several authors have proposed maximum likelihood and maximum entropy
1383: methods for reconstructing the projected mass distribution from
1384: joint weak lensing observations of the shape distortion and
1385: magnification bias effects (Bartelmann 1995; Seitz, Schneider, \&
1386: Bartelmann 1998; Bridle et al. 1998; Umetsu et al. 2007).
1387: %%%
1388: In the present study, we utilize a maximum entropy method
1389: extended to account for positive/negative distributions of the
1390: underlying field starting with the method proposed in the context of
1391: interferometric observations of the Cosmic Microwave Background (CMB)
1392: radiation by Maisinger, Hobson, \& Lasenby (1997) and Hobson \& Lasenby
1393: (1998).
1394:
1395:
1396: Bayes' theorem states that given a hypothesis $H$ and some data $D$ the
1397: posterior probability ${\rm Pr}(H|D)$ is the product of of the
1398: likelihood ${\rm Pr}(D|H)$ and the prior probability ${\rm Pr}(H)$,
1399: with a proper normalization by the evidence ${\rm Pr}(D)$ (Maisinger et
1400: al. 1997):
1401: \begin{equation}
1402: \label{eq:bayes}
1403: {\rm Pr}(H|D) = \frac{{\rm Pr}(H)\, {\rm Pr}(D|H)}{{\rm Pr}(D)}.
1404: \end{equation}
1405: In the context of map-making, the hypothesis $H$ is taken as the
1406: pixelized image ${\bf p}=\{p_i\}$ ($i=1,2,...,N_{\rm pix}$).
1407: Since the evidence in Bayes' theorem is a normalization constant for a
1408: given dataset $D$, we maximize ${\rm Pr}({\bf p}|D)\propto {\rm
1409: Pr}({\bf p})\,
1410: {\rm Pr}(D|{\bf p})$ with respect to ${\bf p}$.
1411: We assume that the errors on the data follow a Gaussian distribution, so
1412: that the likelihood is given as ${\rm Pr}(D|{\bf p})\propto
1413: \exp\left[-\chi^2({\bf p})/2\right]$, with $\chi^2$ being the standard
1414: misfit statistic. Then, for a given form of the entropy function $S({\bf
1415: p}, {\bf m})$, the entropic prior is written as
1416: \begin{equation}
1417: \label{eq:entropy}
1418: {\rm Pr}({\bf p}) \propto
1419: \exp\left[\alpha S({\bf p},{\bf m})\right],
1420: \end{equation}
1421: where ${\bf m} = \{m_i\}$ ($i=1,2,...,N_{\rm pix}$)
1422: is a set of model parameters for the pixelized image, ${\bf p}$.
1423: We follow the prescription given in Maisinger et al. (1997) and Hobson \&
1424: Lasenby (1998), and define the cross-entropy function for ${\bf p}$ as
1425: %the cross-entropy function for a pixelized {\it image} ${\bf p}=\{p_i\}$
1426: %($i=1,2,...,N_{\rm pix}$) is written in discretized form as (Maisinger
1427: %et al. 1997; Hobson \& Lasenby 1998)
1428: \begin{equation}
1429: S({\bf p}, {\bf m}) =
1430: \sum_{i=1}^{N_{\rm pix}}\left[
1431: \psi_i -2m_i-p_i\ln\left(
1432: \frac{\psi_i+p_i}{2m_i}
1433: \right)
1434: \right],
1435: \end{equation}
1436: where $\psi_i\equiv \sqrt{p_i^2+4m_i^2}$.
1437: %and ${\bf m}=\{m_i\}$ ($i=1,2,...,N_{\rm pix}$)
1438: %is a set of model parameters for ${\bf p}$.
1439: %%%
1440: We take $p_i = \kappa_i\equiv \kappa(\btheta_i)$
1441: as the image to be reconstructed,
1442: and express a set of discretized $\kappa$-values as
1443: ${\bf p}=\{\kappa_i\}$ $(i=1,2,...,N_{\rm pix})$.
1444: %%%%
1445: In general, an entropy regularization
1446: helps to reduce the sensitivity of the least $\chi^2$
1447: solutions to small-scale noise in data, by imposing
1448: smoothness constraints on the solutions.
1449: %%%%
1450: This MEM prior ensures that
1451: $\kappa(\btheta) \to 0$ in the noise-dominated regime,
1452: or low-density regions in the outskirts of the cluster.
1453: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1454: Note that unlike conventional MEM functions, this MEM prior is free from
1455: the ``positive bias'' in the reconstructed image ($= {\rm signal} +
1456: {\rm residual \ noise}$), and this bias is more significant in the low
1457: signal-to-noise ratio, or low density regions ($r\to r_{\rm vir}$) where
1458: we are interested in measuring $\kappa$.
1459:
1460:
1461:
1462:
1463: %%%%
1464: We take into account the non-linear, but subcritical, regime of the
1465: lensing properties, $\kappa$ and $\gamma_{\alpha}$,
1466: for a MEM mass reconstruction (see Bridle et al. 1998).
1467: We use equations (\ref{eq:gamma2kappa}) and (\ref{eq:kappa2gamma})
1468: to relate the gravitational shear and the convergence fields,
1469: $\gamma$ and $\kappa$.
1470: %%%%
1471: %%%%
1472: The log posterior probability function,
1473: $F({\bf p}) = -\ln{{\rm Pr}({\bf p}|D)}$, is then expressed as a linear
1474: sum of the shear/magnification data log-likelihoods (Schneider, King, \&
1475: Erben 2000) and the entropy term (Maisinger et al. 1997):
1476: \begin{eqnarray}
1477: \label{eq:loglikelihood}
1478: F({\bf p})& = &
1479: \frac{1}{2} \chi^2({\bf p})
1480: %\left[
1481: %\chi^2_g({\bf p})
1482: %+
1483: %\chi^2_\mu({\bf p})
1484: %\right]
1485: - \alpha S({\bf p}, {\bf m}),\\
1486: \chi^2({\bf p}) &=& \chi^2_g({\bf p}) + \chi^2_\mu({\bf p})\\
1487: %%
1488: \chi^2_g &\equiv&
1489: \sum_{i=1}^{N_{\rm pix}}
1490: \sum_{\alpha=1}^{2}
1491: \frac{ [\tilde{g}_{\alpha,i}-g_{\alpha,i}({\bf p}) ]^2 } {
1492: \sigma_{\tilde{g}_\alpha,i}^2 },\\
1493: %%
1494: \chi^2_\mu &\equiv&
1495: \sum_{i=1}^{N_{\rm pix}}
1496: \frac{ [\tilde{N}_{i}-N_{i}({\bf p})]^2} {\tilde N_i},
1497: \end{eqnarray}
1498: where $\alpha (>0)$ is the dimensionless regularization constant, and
1499: $\sigma_{\tilde{g}_{\alpha,i}}$ is the per-component rms error
1500: for the pixelized distortion measurement ($\alpha=1,2$),
1501: $\sigma_{\tilde{g}_{1,i}} =\sigma_{\tilde{g}_{2,i}} =
1502: \sigma_{\tilde{g},i}/\sqrt{2}$;
1503: %%%%%
1504: $g_{\alpha,i}({\bf p})$ (equation [\ref{eq:redshear}])
1505: and $N_i({\bf p}) = \mu_i({\bf p})^{2.5s-1} N_0$
1506: %n_{\mu,0}\Delta\Omega_{{\rm pix},i}$
1507: are the theoretical expectations for
1508: $\tilde{g}_{\alpha,i}$ and $\tilde{N}_i$, respectively,
1509: where $\Delta\Omega_{{\rm pix},i}$ is the effective observed area
1510: of the $i$th pixel excluding the masking area by the cluster members.
1511: %%%%%%%
1512: %In defining the magnification bias likelihood function $l_\mu$,
1513: In defining $\chi^2_\mu$,
1514: we have used the Gaussian approximation for the
1515: Poisson distribution
1516: of count-in-cell statistics
1517: ($N_0 \sim 25$),
1518: as done in the one-dimensional analysis by B05a.
1519: %%%
1520: We also note that the dispersion for $\tilde{g}$
1521: is modified as $\sigma[\tilde g(\btheta)] \approx \sigma_{\tilde
1522: g}(1-|g({\bf p})|^2)$ in the subcritical, non-linear regime (Schneider
1523: et al. 2000); however, we neglect this non-linear correction for the
1524: dispersion $\sigma_{\tilde g}$ to simplify various calculations, as in
1525: B05a.
1526: %%%%
1527: We found these are indeed good approximations for our combined
1528: distortion and depletion datasets, and the results presented here
1529: are little changed when the full likelihood function is used.
1530: %%
1531: %%
1532:
1533:
1534: The maximum likelihood solution, $\hat{\bf p}$,
1535: is obtained by minimizing
1536: the function $F({\bf p})$ with respect to ${\bf p}$
1537: for given $\alpha$ and ${\bf m}$.
1538: %%%%%
1539: To do this,
1540: we compute numerically the
1541: derivatives $\partial F({\bf{p}})/\partial p_i$ using a
1542: conjugate-gradient algorithm (Press et al. 1992).
1543: %%%
1544: %We take $m_i={\rm const}\equiv m$,
1545: We determine by iteration
1546: the Bayesian value of $\alpha=\hat{\alpha}(\hat{\bf p}, {\bf m})$
1547: by the following equation
1548: (Bridle et al. 1998):
1549: %for a given value of $m$.
1550: \begin{equation}
1551: \label{eq:alpha}
1552: -2\hat\alpha S(\hat{\bf p},{\bf m}) = N_{\rm pix} - \hat\alpha {\rm Tr}({\bf
1553: M}^{-1}) \, \equiv N_{\rm good},
1554: \end{equation}
1555: where ${\bf M}$ is a $N_{\rm pix}\times N_{\rm pix}$ matrix defined by
1556: \begin{equation}
1557: {\bf M} = {\bf G}^{-1/2} {\bf H} {\bf G}^{-1/2},
1558: \end{equation}
1559: with
1560: \begin{eqnarray}
1561: H_{ij}(\hat{\bf p})&=&\frac{\partial^2 F({\bf p})}
1562: {\partial p_i \partial p_j}\bigg|_{{\bf p}=\hat{\bf p}},\\
1563: G_{ij}(\hat{\bf p}) &=& \frac{\partial^2 S({\bf p})}
1564: {\partial p_i \partial p_j}\bigg|_{{\bf p}=\hat{\bf p}},
1565: \end{eqnarray}
1566: evaluated at ${\bf p}=\hat{\bf p}$;
1567: $N_{\rm good}$ is a measure of the
1568: effective number of parameters (Suyu et al. 2006).
1569: With the optimal $\alpha$, we thus expect that
1570: the final value of the misfit statistic $\chi^2$ is close to the {\it
1571: classical number of degrees of freedom} (hereafter NDF, see Suyu et
1572: al. 2006), ${\rm NDF} \equiv N_{\rm data} - N_{\rm good}$ (classic MEM).
1573: We take $m_i={\rm const}\equiv m$,
1574: where
1575: $m$ can be regarded as a characteristic amplitude
1576: of the image (Maisinger et al. 1997).
1577: %%%%
1578: %%%%
1579: We find that
1580: the maximum-likelihood solution $\hat{\bf p}$
1581: for the Bayesian $\hat\alpha$
1582: is insensitive to the choice of $m$ (see \S \ref{subsec:sys}).
1583: In the following we set $m$ to be $0.5$.
1584: %%%
1585: In order to be able to
1586: quantify the errors on the mass reconstruction we
1587: evaluate the Hessian matrix ${\bf H}$ of the function
1588: $F({\bf p})$ at ${\bf p}=\hat{\bf p}$,
1589: %\begin{equation}
1590: %\label{eq:hess}
1591: %H_{ij}(\hat{\bf p})=\frac{\partial^2 F({\bf p})}
1592: %{\partial p_i \partial p_j}\bigg|_{{\bf p}=\hat{\bf p}},
1593: %\end{equation}
1594: from which the covariance matrix
1595: of the reconstructed $\kappa$ map is given by
1596: %parameters ${\bf p}$ is given by
1597: \begin{equation}
1598: C_{ij}
1599: \equiv \langle
1600: \delta\kappa_i
1601: \delta\kappa_j
1602: %(\kappa-\hat\kappa)_i
1603: %(\kappa-\hat\kappa)_j
1604: \rangle=
1605: (H)^{-1}_{ij}(\hat{\bf p}).
1606: \end{equation}
1607: %%%%
1608:
1609: \subsection{Strong Lensing Constraints}
1610: \label{subsec:acs}
1611:
1612: The map-making method described in \S \ref{subsec:massrec}
1613: is only applicable to subcritical regions
1614: lying outside of the critical curves (see Bartelmann \& Schneider 2001).
1615: Therefore, our Subaru distortion and depletion datasets cannot be used
1616: to constrain the cluster mass distribution in the strong lensing region
1617: within the tangential critical curve (see B05b).
1618: %%%%
1619: Instead, we can
1620: pause quadratic constraints on such pixels
1621: to the total log-likelihood function
1622: based on strong lensing models as:
1623: \begin{equation}
1624: F({\bf p})
1625: \to F({\bf p}) + \frac{\beta}{2}\sum_{i=1}^{N_c}
1626: \left(\kappa-\kappa_{\rm c}\right)_i^2,
1627: \end{equation}
1628: where
1629: $\kappa_{{\rm c}.i}$ is the constraint on $\kappa_i$ by strong lensing,
1630: $N_{\rm c}$ is the number of constraints, and $\beta$ is a large number
1631: without causing this external term to become singular to working precision.
1632:
1633:
1634: In the present pixelization scheme, there is one such pixel ($N_{\rm
1635: c}=1$) containing the strong lensing region for which the Einstein
1636: radius of $\theta_{\rm E}\simeq 43''$ corresponds to a mean redshift of
1637: $\bar z_s=0.87$ for our red background population (see \S
1638: \ref{subsec:red}).
1639: To constrain the central $\kappa$ pixel, we utilized a mass model of
1640: A1689, which is well constrained by ACS strong lensing observations
1641: restricted to the central region $\simlt 2'$ (B05b).
1642: The central ACS-derived mass profile of B05b is best described by an NFW
1643: model with a virial mass $M_{\rm vir}=2.6\times 10^{15}M_{\odot}/h$ and a
1644: concentration $c_{\rm vir}=8.2$ having a scale radius of $r_s \equiv
1645: r_{\rm vir}/c_{\rm vir}=310 {\rm kpc}/h$ (B05b).
1646: With this NFW model, we find $\kappa_{\rm c} \approx 0.781 \pm 0.1$ at
1647: the central pixel of $\btheta\sim (0',0')$ for a reference source
1648: redshift of $z_s=1$, corresponding to $\kappa_{\rm c}\approx 0.700$ for
1649: our red background sample (\S \ref{subsec:red}). Hence, assuming $z_s=1$
1650: for the source redshift will overpredict the central density by a
1651: factor of $0.772/0.693\simeq 1.113$, which will affect slightly the
1652: overall amplitude of the reconstructed $\kappa$ map (see \S
1653: \ref{subsec:sys}).
1654:
1655:
1656:
1657:
1658:
1659: \subsection{Downweighting Intrinsic Clustering Contributions and Noisy
1660: Measurements}
1661: \label{subsec:declustering}
1662:
1663: In contrast to the shearing effect, the magnification bias
1664: is a local, direct measure of the lensing convergence,
1665: free from the mass-sheet degeneracy.
1666: However, a practical difficulty of the magnification bias measurement
1667: is the intrinsic clustering contribution which locally can be larger
1668: than the lensing induced signal in a given pixel.
1669: %% :: Keiichi's original ::
1670: %However, the practical difficulty of
1671: %the magnification bias measurement is the intrinsic clustering
1672: %contribution which is usually larger than the lensing induced signal.
1673: In order to obtain a clean measure of the lensing signal,
1674: such intrinsic clustering needs to be eliminated.
1675: %%%%%%%%%%%%%%%%%
1676: %% Zhang \& Pen (2005)
1677: Broadhurst et al. (1995) proposed an active declustering method
1678: based on the facts that
1679: the magnification bias depends strongly on the shape of the luminosity
1680: function, whereas intrinsic clustering depends weakly on the
1681: intrinsic luminosity function,
1682: and that they have different redshift dependence
1683: (see also Zhang \& Pen 2005).
1684: %%%%
1685: This method however requires
1686: the addition of redshift information for
1687: %%% :: Keiichi's original ::
1688: %information of photometric redshifts as well as fluxes of
1689: individual background galaxies.
1690: Furthermore, the shape information of
1691: luminosity functions for respective background populations must be
1692: provided in order to convert a density depletion or enhancement
1693: into the lens magnification, $\mu$.
1694: %Therefore,
1695: %which enables us to
1696:
1697: Alternatively, one may employ prior information from, for example,
1698: the surface luminosity density of cluster member galaxies
1699: to predict the projected mass distribution,
1700: which is then used to
1701: downweight intrinsic clustering contributions.
1702: In the present study, we have adopted
1703: an objective rejection scheme
1704: based on the gravitational shear predictions, summarized in
1705: the following steps:
1706: (1) Using shape distortion data alone
1707: we derive an entropy-regularized
1708: maximum likelihood solution $\bf{p}$ for the mass distribution.
1709: (2) Then, the shear-based $\kappa$ map is used to predict
1710: the magnification bias on the same grid of pixels.
1711: (3) Finally, we make a pixel-to-pixel comparison between
1712: the observed and predicted galaxy counts, and
1713: reject those magnification measurements $\tilde N_i$
1714: which are in conflict with
1715: the shear-based predictions $N_i(\bf{p})$ as:
1716: \begin{equation}
1717: |\tilde N_i - N_i({\bf p})| > \nu_{\rm clust} \sqrt{N_0}
1718: \end{equation}
1719: where $N_0\sim 25$ is the unlensed, mean number counts of red galaxies,
1720: and $\nu_{\rm clust}$ is a rejection threshold in units of $\sigma$. We
1721: set a rejection threshold at $4\sigma$, i.e., $\nu_{\rm clust}=4$.
1722: %%%
1723: Changing this threshold will affect the details of the
1724: reconstruction especially in the outer regions ($\theta \simgt
1725: 7\arcmin$).
1726: However, we confirmed that changing the threshold $\nu_{\rm clust}$ by
1727: $\pm 1$ leaves our results unchanged within our statistical
1728: uncertainties (see
1729: \S \ref{subsec:sys}), and that the resulting mass map is fairly
1730: consistent with the surface luminosity and density distributions of
1731: cluster member galaxies, as described in detail below.
1732:
1733:
1734:
1735:
1736: In addition to the above,
1737: in order to exclude unreliable measurements,
1738: we have assigned zero weight to those pixels which
1739: satisfy either of the following rejection criteria:
1740: \begin{enumerate}
1741: \item Measurement pixel with no usable galaxy,
1742: \item Measurement pixel in the strong lensing region (\S \ref{subsec:acs}),
1743: %\item Magnification measurement rejected according to the shear-based
1744: % prediction (\S \ref{subsec:declustering}),
1745: \item Magnification measurement near boundary regions
1746: with low
1747: completeness at fainter magnitudes.
1748: \end{enumerate}
1749:
1750: The last rejection criterion is
1751: required since the slope of the unlensed number counts,
1752: $s = d\log{N_0(m)}/dm$, could depend on the selection completeness;
1753: such an apparent variation of the slope $s$ could give rise
1754: to additional errors in the mass reconstruction.
1755:
1756:
1757: \subsection{Map-making of A1689}
1758: \label{subsec:a1689map}
1759:
1760:
1761: We have applied our MEM method to the
1762: joint measurements of the shape distortion and magnification bias
1763: effects of red background galaxies
1764: %combined shape distortion and magnification bias measurements
1765: in order to reconstruct the two-dimensional mass distribution of A1689.
1766: We utilized the FFTW implementation
1767: %(Frigo \& Johnson 1998)
1768: of
1769: Fast Fourier Transforms (FFTs) to convert between
1770: the lensing convergence and the gravitational shear fields
1771: using equations (\ref{eq:gamma2kappa}) and (\ref{eq:kappa2gamma}).
1772: The FFT however implies a periodicity in
1773: both horizontal and vertical directions,
1774: producing aliasing effects at the borders of the computational domain.
1775: In order to avoid such aliasing effects,
1776: we used large arrays of $31 \times 27$ pixels
1777: with $N_{\rm pix}=31\times 27=837$,
1778: covering a field of $43\farcm 4\times 37\farcm 8$.
1779: %%%
1780: On the other hand,
1781: the distortion and depletion measurements
1782: are limited to the central $30'\times 24'$ region
1783: ($21\times 17=357$ pixels);
1784: the wide-field Subaru data thus allow us to probe the
1785: projected mass distribution on scales ranging from
1786: $\theta\sim 1'$ up to $\theta\sim 20'$.
1787: %%%
1788: To minimize spurious effects from the periodic boundary
1789: condition,
1790: pixelized maps are further zero padded
1791: by a factor of $2$ in each dimension
1792: (Seljak 1998; Sato et al. 2003).
1793: %%% Seljak 1998, ApJ, 506, 64
1794: %%%%
1795:
1796:
1797: For the spin-2 distortion measurements, we have in total $N_{{\rm
1798: data},g}=2 \times 355 = 710$ usable (real) observations.
1799: %%%%%%
1800: For the magnification measurements,
1801: we have $N_{{\rm data},\mu}=302$ usable observations.
1802: %we discarded depletion measurements outside
1803: %the central $24'\times 24'$ region
1804: %we only used depletion data in the
1805: %central $24'\times 24'$ region,
1806: %yielding a net number of $N_{{\rm data},\mu}=270$.
1807: %%%%%
1808: %The first reason for this masking is
1809: %because, unlike the distortion signal,
1810: %the depletion signal measures the local surface mass density $\kappa$,
1811: %and hence falls rapidly with increasing radius.
1812: %Furthermore,
1813: %Secondly,
1814: %local overdensities of
1815: %red galaxies are apparent in the outer region
1816: %(Figure \ref{fig:magbias})
1817: %where the distortion signal is relatively weak; thus, this intrinsic
1818: %clustering of the red background could dominate over the lensing singal
1819: %in such low density regions, and hence strongly affect the mass
1820: %reconstruction.
1821: %%%%
1822: The total number of constraints is thus $N_{\rm data}=N_{{\rm
1823: data},g}+N_{{\rm data},\mu}=1012$, yielding $N_{\rm data}-N_{\rm
1824: pix}=175$ degrees of freedom (dof).
1825: %%
1826: %%
1827: The Bayesian value of the regularization parameter $\hat{\alpha}$
1828: that satisfies equation (\ref{eq:alpha}) was obtained as
1829: $\hat{\alpha}=96.6$, when $-2\hat\alpha S \approx N_{\rm pix} -
1830: \hat\alpha {\rm Tr}({\bf M}^{-1}) \approx 213.1 (=N_{\rm good})$.
1831: The minimum function value of $F$ was found to be
1832: $F_{\rm min}=F(\hat{\bf p})=
1833: (1/2)\chi^2(\hat{\bf p})-\hat{\alpha}S(\hat{\bf p})
1834: = 651.2$, and $\chi^2(\hat{\bf p})= 1089.3$,
1835: corresponding to ${\rm NDF} = N_{\rm data} - N_{\rm good} = 798.9$,
1836: i.e., $\chi^2(\hat{\bf p})/{\rm NDF} = 1.36$.
1837: Hence, the resulting $\chi^2$ is somewhat large, and is rather close to
1838: historic MEM, i.e., $\chi^2(\hat {\bf p}) \approx N_{\rm data}$.
1839: However, we note that our MEM mass reconstructions without the magnification
1840: data yield $\chi^2(\hat {\bf p})/{\rm NDF} \approx 1$ (see Table
1841: \ref{tab:memmethod}) as expected for classic MEM. Therefore, slightly
1842: large values of this misfit statistic using the magnification data could be
1843: attributed to the fact that the intrinsic clustering noise in the
1844: magnification measurements is not included in the likelihood
1845: calculation, underestimating the errors for the magnification measurements.
1846:
1847:
1848:
1849:
1850: In the left panel of
1851: Figure \ref{fig:rawkappa} we show the resulting
1852: $\kappa$ map
1853: on a grid of $31\times 27$ pixels reconstructed from
1854: the combined distortion and depletion data with the ACS constraint
1855: on the mean value of $\kappa$ for the central pixel.
1856: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1857: The reconstructed spin-2 shear field ($\gamma[\hat{\bf p}]$)
1858: is overlayed up on the $\kappa$ map.
1859: The right panel in Figure \ref{fig:rawkappa}
1860: shows the two-dimensional distribution of the
1861: rms reconstruction error for $\kappa_i$, estimated from the diagonal
1862: part of the pixel-pixel covariance matrix of errors:
1863: $\sigma(\kappa_i) = C^{1/2}_{ii}({\bf \hat p})$.
1864: %%%
1865: Figure \ref{fig:image} presents the contours of the reconstructed
1866: $\kappa$ map
1867: superposed on the $V+i'$ pseudo-color image covering a field of
1868: $30'\times 25'$ around the cluster.
1869: Here,
1870: for visualization purposes,
1871: the $\kappa$ map is resampled on to a finer grid
1872: and convolved with a Gaussian of ${\rm FWHM}=1\farcm 4$,
1873: corresponding to a physical resolution of ${\rm FWHM}\sim 180 {\rm
1874: kpc}/h$ at the cluster redshift.
1875: %%%
1876: The projected mass distribution of the cluster is
1877: smoothly varying and symmetric, with no significant substructure at
1878: $r\simgt 100 {\rm kpc}/h$.
1879: %showing that mass and
1880: %light in A1689 are similarly distributed
1881: %with a fairly round shape,
1882: %and concentrated around
1883: %the cD galaxy.
1884: %We note that brightest cluster galxies tend to deviate from the
1885: %linear CM relation, so that we visually checked if such brightest
1886: %cluster
1887: %galaxies were properly included in our cluster member sample, and
1888: %included them if they were missing.
1889: %A field correction is applied to the measured luminosity and number
1890: %densities of cluster galaxies to account for contamination by field
1891: %galaxies. We estimate the background luminosity/number
1892: %density level from an annular
1893: %region outside of the cluster region ($r\simgt r_{\rm vir}$), and then
1894: %subtract this background contribution from the observed
1895: %luminosity/number density of the cluster.
1896: %%
1897: Figure \ref{fig:lmap} compares the reconstructed lensing fields
1898: and the member galaxy distributions in A1689, Gaussian smoothed
1899: to a resolution of ${\rm FWHM}=2'$.
1900: %%%%
1901: The top panels show the reconstructed $\kappa$
1902: and the magnification-bias $\delta_\mu=\mu^{2.5s-1}-1$
1903: fields in the left and right panels, respectively.
1904: %%%
1905: The bottom left and right panels display the field-corrected
1906: $i'$-band luminosity and number density maps, respectively,
1907: of bright red-sequence galaxies with $i'<23$ mag.
1908: %where the background density levels are estimated from an
1909: %annular region
1910: %$15'\simlt \theta \simlt 20'$
1911: %outside of the cluster region;
1912: %Overlayed up on the images are the mass contours shwon in the top-left
1913: %panel.
1914: For each panel the color scale is linear, and
1915: ranges from $0\%$--$100\%$ of the peak value.
1916: %A field correction is applied to the measured luminosity and number
1917: %densities of cluster galaxies to account for contamination by field
1918: %galaixes.
1919: It is clear from Figure \ref{fig:lmap}
1920: that mass and light in A1689 are similarly distributed
1921: with a fairly round shape, and well centered on the main cD galaxy.
1922:
1923:
1924:
1925: \subsection{Mass Maps from Different Datasets and Boundary Conditions}
1926: \label{subsec:3mem}
1927:
1928: Any mass reconstruction technique based on the shear information
1929: involves a non local process, meaning that one has to assume
1930: certain boundary conditions to convert the gravitational shear field,
1931: $\gamma(\btheta)$,into the lensing convergence field, $\kappa(\btheta)$
1932: (Bartelmann \& Schneider 2001; Umetsu et al. 1999).
1933: %%%
1934: If the data field is sufficiently large so as to ensure that
1935: projected mass fluctuations over the field average out,
1936: or that the mean convergence over the field is zero, then
1937: one may simply use equation (\ref{eq:gamma2kappa}) to invert
1938: the observed shear field into the convergence field (Kaiser \& Squires
1939: 1993).
1940: Or, if magnification data are available, then
1941: a combination of complementary shearing and magnification
1942: information can be used to break the degeneracy between
1943: the observables and the underlying gravitational potential field
1944: (Bartelmann 1995; Seitz et al. 1998).
1945: Besides, strong lensing observations,
1946: if available, will place additional, tight constraints on the
1947: central mass distribution of the cluster where weak lensing
1948: alone cannot constrain (e.g., mass reconstruction of A370 in
1949: Umetsu et al. 1999).
1950:
1951:
1952: Here we consider three sets of combinations of datasets and boundary
1953: conditions
1954: as summarized in
1955: Table \ref{tab:memmethod}:
1956: (i) {\it 2D MEM+} method using
1957: shear and magnification data with the central ACS constraint,
1958: (ii) {\it 2D MEM} method using shear and magnification data without the ACS
1959: constraint,
1960: (iii) {\it 2D MEM-S} method using shear data with the central ACS constraint.
1961: For each MEM reconstruction, we derive an
1962: entropy-regularized maximum likelihood solution for $\kappa$
1963: with the Bayesian value of $\alpha$.
1964: The resulting Bayesian value of $\alpha$ and the minimized
1965: functional values of
1966: $F$ and $\chi^2$ are also listed in
1967: Table \ref{tab:memmethod}.
1968: %%%%
1969:
1970:
1971: In order to quantify the significance of the reconstruction,
1972: we define an estimator for the detection signal-to-noise
1973: ratio (S/N) by the following equation:
1974: \begin{equation}
1975: \label{eq:sn}
1976: \left({\rm S/N}\right)^2 \equiv
1977: \sum_{\kappa_i>0}\sum_{\kappa_j>0}
1978: \kappa_i \kappa_j C^{-1}_{ij},
1979: \end{equation}
1980: where the indices $i$ and $j$ run over all pixels
1981: except those with negative values of $\kappa$
1982: and those in the strong lensing region
1983: (see Table \ref{tab:memmethod}).
1984: The reconstructions based on
1985: both the distortion and depletion data yield a similar
1986: S/N of $\sim 18$--$19$, whereas the reconstruction
1987: from the distortion data alone gives a slightly lower S/N of
1988: $\sim 15$ (see Table \ref{tab:memmethod}).
1989: For comparison we quote the detection significance
1990: from the 1D Subaru analysis
1991: by B05a based on the same red background catalogs for
1992: the weak lensing distortion and depletion analyses:
1993: %%%%%%%%%%%%%%%%
1994: B05a measured the lens distortion and depletion profiles
1995: over a radial range of $1'\simlt \theta \simlt 18'$.
1996: %%%%%%%%%%%%%%%%
1997: We find ${\rm S/N}\simeq 14.2$ and ${\rm S/N}\simeq 9.2$ for the
1998: measurements of the
1999: distortion and depletion profiles,
2000: respectively.
2001: %%%%%%%%%%%%%%%%
2002: Since covariance between the distortion and magnification measurements
2003: can be neglected,
2004: the total S/N is simply obtained as
2005: $\sqrt{14.2^2+9.2^2} \sim 17$.
2006: %%%%%
2007: These numbers are quite comparable to those as measured from the
2008: reconstructed $\kappa$ field and its covariance matrix,
2009: indicating that the lensing information contained in the
2010: red catalogs is properly propagated into the $\kappa$-basis\footnote{
2011: This is not trivial since
2012: the noise level in a MEM-reconstructed
2013: map is affected by the smoothness constraint specified by
2014: $\alpha$. Besides,
2015: MEM is non-linear, so that the resulting reconstruction errors depend
2016: on the signal as well as noise properties.}.
2017: We find it important to use the Bayesian value for $\alpha$ in
2018: order to have an optimal smoothness for the mass reconstruction,
2019: avoiding oversmoothing.
2020: A more quantitative comparison between different reconstructions
2021: will be given in \S \ref{sec:massprofile}.
2022:
2023:
2024:
2025:
2026:
2027: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2028: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2029:
2030: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2031:
2032: \section{Mass Profile of A1689}\label{sec:massprofile}
2033:
2034: In this section, we aim to quantify and characterize
2035: the projected mass distribution ($\kappa$) of A1689
2036: reconstructed from Subaru weak lensing observations,
2037: in order to derive quantitative constraints on the
2038: three-dimensional mass distribution.
2039: Specifically,
2040: we will adopt the NFW density profile
2041: $\rho(r)\propto r^{-1}(1+r/r_s)^{-2}$
2042: (Navarro et al. 1997) to describe the cluster mass distribution,
2043: characterized by the virial mass $M_{\rm vir}$ and the concentration
2044: parameter $c_{\rm vir}=r_{\rm vir}/r_s$,
2045: defined as the ratio of the
2046: virial radius $r_{\rm vir}$ to the scale radius $r_s$.
2047: The best-fitting NFW parameters $(M_{\rm vir},c_{\rm vir})$
2048: can be then compared with $\Lambda$CDM predictions
2049: based on $N$-body simulations (Bullock et al. 2001; Neto et al. 2007).
2050: %%%
2051: As a test for the consistency and reliability, we will compare
2052: the best-fitting NFW parameters derived from different combinations of
2053: datasets, boundary conditions, and weak lensing techniques.
2054: Here we introduce three different methods to
2055: derive a convergence profile $\kappa(\theta)$ from weak lensing
2056: observations.
2057:
2058:
2059:
2060:
2061: \subsection{Method (I): 2D Convergence Map}
2062:
2063:
2064: The first method makes a direct use of the 2D $\kappa$ map
2065: reconstructed by the entropy-regularized maximum likelihood method
2066: (\S \ref{sec:map}).
2067: The $\kappa$ map is directly compared with the model convergence field
2068: for an NFW spherical halo specified by two model parameters.
2069: %%%%
2070: We take
2071: the virial mass $M_{\rm vir}$ and the concentration parameter
2072: $c_{\rm vir}$
2073: %$c_{\rm vir}=r_{\rm vir}/r_s$, the ratio of the virial radius
2074: %$r_{\rm vir}$ to the scale radius $r_s$
2075: %evaluated at the clutser redshift
2076: %($z_d=0.183$)
2077: for describing an NFW halo.
2078: We employ the radial dependence of the
2079: convergence profile for the NFW model given by Bartelmann (1996).
2080: Note that the Bartelmann's
2081: formulae for the NFW convergence and shear profiles are obtained
2082: assuming the projection integral to extend to infinity.
2083: Alternatively, a truncated NFW profile can be used to model the
2084: convergence profile (Takada \& Jain 2003).
2085: We have confirmed that
2086: the best-fitting NFW parameters obtained using
2087: the above two different models agree to within $1\%$
2088: for the case of A1689 lensing
2089: (for detailed discussions,
2090: see Baltz, Marshall, \& Oguri 2007 and
2091: Hennawi et al. 2007). We thus simply use
2092: Bartlemann's formulae to calculate the relevant
2093: lensing fields for an NFW halo
2094: as done in B05a and B05b.
2095:
2096:
2097: We constrain the two NFW parameters from $\chi^2$ fitting to the
2098: 2D convergence map $\kappa(\btheta)$ derived from
2099: Subaru weak lensing
2100: observations. We adopt a flat prior of $c_{\rm vir}\le 30$ for the
2101: halo concentration because the NFW profiles with $c_{\rm vir} \simgt 20$
2102: cannot be distinguished by the Subaru data alone due to lack of
2103: information on the inner density profile (B05a).
2104: The $\chi^2$-function for the Subaru weak lensing observations,
2105: $\chi^2=\chi^2_{\rm WL}$,
2106: can be expressed as
2107: (Oguri et al. 2005)
2108: \begin{equation}
2109: \label{eq:chi2_wl}
2110: \chi^2_{\rm WL} = \sum_{i,j}
2111: \left[\hat{\kappa}(\btheta_i)-\kappa(\btheta_i)\right]
2112: \left(C^{-1}\right)_{ij}
2113: \left[\hat{\kappa}(\btheta_j)-\kappa(\btheta_j)\right],
2114: \end{equation}
2115: where $\hat{\kappa}(\btheta_i)$ is the model prediction of the
2116: NFW halo for the $i$th bin ($i=1,2,...,N_{\rm SL}$),
2117: and $(C^{-1})_{ij}$ is the inverse of
2118: the pixel-pixel covariance matrix; $N_{\rm pix}=31\times 27=837$.
2119: In the model fitting
2120: we exclude the central pixel
2121: in the strong lensing region
2122: (see \S \ref{subsec:acs}).
2123: %%%%
2124:
2125:
2126: We have derived sets of best-fitting NFW parameters for
2127: the three different MEM reconstructions
2128: listed in Table \ref{tab:memmethod}.
2129: %based on different combinations of datasets and boundary conditions.
2130: %%%%
2131: Table \ref{tab:nfwfit} shows a summary
2132: of the best-fitting NFW parameters $(M_{\rm vir},c_{\rm vir})$
2133: %with $1\sigma$ uncertainties
2134: and the resulting
2135: $\chi^2$ value for our Subaru weak lensing observations;
2136: %%%%
2137: the errors quote $68\%$ confidence intervals
2138: estimated from $\Delta\chi^2\equiv
2139: \chi^2-\chi^2_{\rm min}=1$
2140: in the $(c_{\rm vir}, M_{\rm vir})$-plane.
2141: %%%%
2142: Combining the distortion and magnification measurements
2143: with the central ACS constraint (MEM+)
2144: yields
2145: the best-fitting NFW parameters,
2146: $M_{\rm vir}=(1.97\pm 0.20)\times 10^{15}M_{\odot}$ and
2147: $c_{\rm
2148: vir}= 13.4^{+5.4}_{-3.3}$,
2149: with $\chi^2_{\rm min}/{\rm dof}=332/834 (421)$,
2150: where the value in parentheses refers to an effective degrees of
2151: freedom excluding upper limit bins with $\kappa<0$.
2152: %%%
2153: As a test for the consistency, we compare
2154: best-fit NFW parameters for the three MEM reconstructions
2155: %based on different combinations of
2156: %datasets and boundary conditions
2157: described in \S \ref{subsec:3mem}. Firstly, all of the three MEM
2158: reconstructions yield
2159: consistent results on the concentration parameter in the range,
2160: $c_{\rm vir}= 12.6-13.6$, but with rather large uncertainties
2161: allowing a wide range of the concentration ($8\simlt c_{\rm vir} \simlt
2162: 20$).
2163: %%%
2164: On the other hand, the best-fit values for $M_{\rm vir}$ range from
2165: $1.47\times 10^{15}M_{\odot}$ for MEM-S, through $1.62\times
2166: 10^{15}M_{\odot}$ for MEM, to $1.97 \times 10^{15}M_{\odot}$ for MEM+,
2167: while the $1\sigma$ error level for each is $\sigma(M_{\rm vir}) \sim
2168: 0.2\times 10^{15}M_{\odot}$.
2169:
2170:
2171:
2172: The observed location of the Einstein radius can be used for a powerful,
2173: model independent test of the $\kappa$ profile (B05b). This is based on
2174: the fact that the enclosed mass interior to the Einstein radius is given
2175: by fundamental constants and with knowledge of distances involved,
2176: namely $\bar{\kappa}(\theta_{\rm E})=1$ or $M(<\theta_{\rm
2177: E})=\pi(D_d\theta_{\rm E})^2\Sigma_{\rm crit}$,
2178: provided that the critical curve is nearly circular; this is the case
2179: for A1689 (B05b).
2180: %%%%%
2181: Although our weak lensing measurements
2182: do not resolve such strong lensing phenomena,
2183: the observed Einstein radius of $\theta_{\rm E}=45''$
2184: %from the ACS observations
2185: (for $z_s=1$)
2186: %%%%%
2187: may be used to test the derived NFW models.
2188: %with the observed Einstein radius
2189: %of $\theta_{\rm E}=45''$ for $z_s=1$ (B05b),
2190: To do this we numerically solve the equation $1=\bar{\kappa}_{\rm
2191: NFW}(\theta_{\rm E})$ for the Einstein radius $\theta_{\rm E}$ by the
2192: Newton-Raphson method.
2193: We found
2194: $\theta_{\rm E}=36.7^{+22}_{-18}, 36.7^{+22}_{-20}, 45.4^{+17}_{-15}$
2195: for the MEM-S, MEM, and MEM+ reconstructions, respectively (Table \ref{tab:nfwfit}).
2196: All of the MEM reconstructions are roughly consistent with the ACS
2197: measurement of the Einstein radius; however,
2198: the constraints on
2199: the $\theta_{\rm E}$
2200: placed by the Subaru data alone
2201: %from the Subaru observations alone
2202: are still rather weak, naturally due to the lack of central mass
2203: distribution.
2204: %satisfy the Einstein radius
2205: %constraint by the ACS observations
2206:
2207:
2208: To make a direct comparison with model predictions,
2209: we compute the convergence profile
2210: $\kappa(\theta)$
2211: from a weighted radial projection of the 2D $\kappa$ map as:
2212: %Given a two-dimensional $\kappa(\btheta_i)$ map
2213: %and the covariance matrix $C_{ij}$ of the reconstruction errors
2214: %($i,j=1,2,...,N_{\rm pix}$),
2215: %we can calculate the convergence profile $\kappa(\theta)$
2216: %of the cluster
2217: %from a weighted radial projection of the pixelized
2218: % $\kappa(\btheta)$ map:
2219: \begin{equation}
2220: \label{eq:binkappa}
2221: \kappa_m\equiv
2222: \kappa(\bar\theta_m) = \sum_{i\in {\rm Bin} m} W_{im} \kappa(\btheta_i)
2223: \Big/
2224: \sum_{i\in {\rm Bin} m} W_{im},
2225: \end{equation}
2226: where $W_{im}$ ($0\le W_{im}\le 1$) is the fraction of the area of the
2227: $i$th pixel lying within the $m$th annular bin. We use Monte Carlo
2228: integration to calculate these area fractions for individual pixels
2229: (Marshall et al. 2002), and $\bar\theta_m$ is the
2230: area-weighted center of the $m$th radial bin:
2231: \begin{equation}
2232: \theta_m = \sum_{i\in {\rm Bin} m} W_{im} |\btheta_i|
2233: \Big/
2234: \sum_{i\in {\rm Bin} m} W_{im}.
2235: \end{equation}
2236: Since the $\kappa_m$ profile is
2237: expressed as
2238: a linear combination of $\kappa(\btheta_i)$ values, it is
2239: straightforward to calculate the covariance matrix
2240: $C_{mn} \equiv \langle
2241: \delta\kappa_m\delta\kappa_n
2242: %%\delta\kappa_(\bar\theta_m)\delta\kappa(\bar\theta_n)
2243: \rangle$
2244: of the reconstruction errors:
2245: \begin{eqnarray}
2246: \label{eq:bin2bincovar}
2247: C_{mn}
2248: &=& \sum_{i\in {\rm Bin} m} \sum_{j\in {\rm Bin} m}
2249: W_{im}W_{jn} C_{ij}
2250: \Big/
2251: \sum_{i\in {\rm Bin} m} W_{im}
2252: \sum_{j\in {\rm Bin} n} W_{jn}\nonumber\\
2253: &+&
2254: \delta_{mn}
2255: %\left(
2256: \sum_{i\in {\rm Bin} m}
2257: W_{im}^2 [\kappa_m-\kappa(\btheta_i)]^2
2258: %\right)
2259: \Big/
2260: \left(\sum_{i\in {\rm Bin} m} W_{im}
2261: \right)^2,
2262: \end{eqnarray}
2263: where the first term represents the errors propagated from the
2264: pixel-pixel covariance matrix of the 2D $\kappa$ reconstruction,
2265: and the second term is responsible for variations of
2266: the $\kappa(\btheta)$-field along the azimuthal direction, i.e.,
2267: departure from circular symmetry.
2268:
2269:
2270: Here we derive
2271: for each of the 2D $\kappa$ maps
2272: a discrete convergence profile
2273: (\ref{eq:binkappa})
2274: in $N_{\rm bin}$
2275: logarithmically spaced bins for $\theta =
2276: [\theta_{\rm min},\theta_{\rm max}]$; we adopt the same binning as in B05a:
2277: $N_{\rm bin}=10$, $\theta_{\rm min}=1'$,
2278: and $\theta_{\rm max}=18'$, as summarized in Table \ref{tab:mrec}.
2279: In Figure \ref{fig:kprof_mem} we compare the convergence profiles from the
2280: 2D MEM+ and MEM reconstructions,
2281: both of which are based on the
2282: lens distortion and magnification measurements,
2283: to clarify the effect of the ACS
2284: constraint on the central surface mass density.
2285: The error bars represent the $1\sigma$ uncertainties
2286: from the diagonal
2287: part ($C_{mm}$) of the bin-to-bin covariance matrix given by
2288: equation (\ref{eq:bin2bincovar}), and hence are correlated between
2289: the different bins.
2290: One can see that overall the convergence profile obtained with the
2291: central ACS
2292: constraint is slightly steeper than that without the ACS constraint,
2293: and has a slightly higher overall normalization;
2294: these features are well explained by
2295: %as explained by
2296: slightly higher NFW mass ($M_{\rm vir}$) and concentration ($c_{\rm vir}$)
2297: derived for the results with the central ACS constraint.
2298: Nonetheless, the two convergence profiles
2299: are overall in good agreement within the
2300: statistical uncertainties.
2301: Steep NFW profiles with
2302: $c_{\rm vir}\sim 13$ and
2303: $M_{\rm vir}\sim 2\times 10^{15}M_{\odot}$
2304: are well fitted to the reconstructed convergence profiles
2305: (solid and dashed curves).
2306: Figure \ref{fig:kprof_3mem} shows convergence profiles from the three
2307: MEM reconstructions: MEM+ (squares), MEM (triangles), and MEM-S (crosses).
2308: Note that the vertical axis is linear here rather than logarithmic.
2309: The convergence profile from the distortion data alone (MEM-S) shows a
2310: slight negative dip of $\kappa\sim -0.01$ at $\theta'\simgt 6'$ due to
2311: boundary effects.
2312:
2313:
2314:
2315:
2316: \subsection{Method (II): 1D Maximum Likelihood Analysis}
2317:
2318: %upporting the assumption of. quasi-circular symmetry in the projected
2319: %mass ... 2005, ApJ, 621, 53) and the Subaru weak lensing analysis with
2320: %the 1D reconstruction method ...
2321:
2322: In B05a we developed a
2323: %model-independent
2324: method for reconstructing the discrete convergence profile $\kappa_m$
2325: from a maximum likelihood combination of radial profiles of the lens
2326: distortion and magnification effects on red background galaxies (see
2327: Figure 3 of B05a).
2328: %%%
2329: With the assumption of quasi-circular symmetry in the
2330: projected mass distribution (or the projected potential field),
2331: one can express the lensing observables (i.e., tangential distortion and
2332: magnification bias) in terms of the binned convergence profile (see also
2333: \S \ref{subsec:zeta}). Owing to the nature of shear-based
2334: reconstruction,
2335: boundary conditions need to be specified (as discussed in \S
2336: \ref{subsec:3mem}).
2337: In B05a we set the inner boundary condition on the mass interior to
2338: $\theta_{\rm min}=1'$, which is readily obtained from the
2339: well-constrained ACS mass model.
2340: %this 1D method also requires to set inner or outer
2341: %boundary conditions
2342:
2343:
2344: One of the advantages of such a 1D method is that one can improve up on
2345: the statistical significance of each measurement pixel by azimuthal
2346: averaging, provided that the system is nearly symmetric.
2347: %%%
2348: B05a measured from the red background sample
2349: a tangential distortion
2350: and a radial depletion profile
2351: in 10 discrete radial bins,
2352: with total significance of ${\rm S/N}\simeq 14.2$
2353: and ${\rm S/N}=9.2$, respectively
2354: (see \S \ref{subsec:3mem}). Thus, the per-pixel S/N of each measurement
2355: is of the order of unity, which is optimal for an inversion problem.
2356: Accordingly,
2357: the 1D analysis does not require
2358: regularization techniques.
2359: The best-fitting NFW parameters for the Subaru 1D analysis
2360: are listed in
2361: Table \ref{tab:memmethod}.
2362: %%%
2363: %In \S \ref{subsec:comparison} we compare the convergence profile from
2364: %the Subaru 1D analysis by B05a with the results from the 2D MEM
2365: %reconstructions.
2366:
2367:
2368:
2369: \subsection{Method (III): Aperture Densitometry}
2370: \label{subsec:zeta}
2371:
2372: The spin-2 shape distortion of an object
2373: due to gravitational lensing
2374: is described by
2375: the complex reduced shear, $g=g_1+i g_2$ (see equation [\ref{eq:redshear}]),
2376: which is coordinate dependent.
2377: For a given reference point on the sky, one can instead
2378: form coordinate-independent
2379: quantities,
2380: the tangential distortion $g_+$ and the $45^\circ$ rotated component,
2381: from linear combinations of the distortion coefficients
2382: $g_1$ and $g_2$ as
2383: \begin{equation}
2384: g_+ = -(g_1 \cos 2\phi + g_2\sin 2\phi), \ \
2385: g_{\times} = -(g_2 \cos 2\phi - g_1\sin 2\phi),
2386: \end{equation}
2387: %For a given point on the sky,
2388: %the tangential component $g_+$ is used to obtain the azimuthally
2389: %averaged distortion
2390: %due to lensing and computed from the distortion
2391: %coefficients $g_1$ and $g_2$ as:
2392: %\begin{equation}
2393: %g_+= -(g_1 \cos 2\phi + g_2\sin 2\phi),
2394: %\end{equation}
2395: where $\phi$ is the position angle of an object with respect to
2396: the reference position, and the uncertainty in the $g_+$ and
2397: $g_{\times}$
2398: measurement
2399: is $\sigma_+ = \sigma_{\times } = \sigma_{g}/\sqrt{2}\equiv \sigma$
2400: in terms of the rms error $\sigma_{g}$ for the complex shear measurement.
2401: In practice, the reference point is taken to be the cluster center,
2402: which is well determined from symmetry of the strong lensing pattern.
2403: %%%%%
2404: To improve the statistical significance of the distortion measurement,
2405: we calculate the weighted average of the $g_+$'s and its weighted error
2406: as
2407: \begin{eqnarray}
2408: \label{eq:gt}
2409: \langle g_+(\theta)\rangle &=& \frac{\sum g_+/\sigma^2}{\sum
2410: 1/\sigma^2},\\
2411: \sigma_+(\theta) &=& \left(\sum 1/\sigma^2\right)^{-1/2}.
2412: \end{eqnarray}
2413:
2414:
2415:
2416: For a shear-based estimation of the cluster mass profile
2417: we use a variant of the weak lensing aperture densitometry, or
2418: so-called the $\zeta$-statistic (Fahlman et al. 1994; Clowe et
2419: al. 2000),
2420: of the form:
2421: \begin{eqnarray}
2422: \label{eq:zeta}
2423: \zeta_{\rm c}(\theta)
2424: &\equiv &
2425: 2\int_{\theta}^{\theta_{\rm inn}}\!d\ln\theta'
2426: \gamma_+(\theta')\nonumber\\
2427: &&+ \frac{2}{1-(\theta_{\rm inn}/\theta_{\rm out})^2}\int_{\theta_{\rm
2428: inn}}^{\theta_{\rm out}}\! d\ln\theta'
2429: \gamma_+(\theta')
2430: \nonumber\\
2431: &=&
2432: \bar{\kappa}(\theta) - \bar{\kappa}
2433: (\theta_{\rm inn}<\vartheta <\theta_{\rm out}),
2434: \end{eqnarray}
2435: where
2436: $\kappa(\theta)$ is the azimuthal average of
2437: the convergence field $\kappa(\btheta)$ at radius $\theta$,
2438: $\bar{\kappa}(\theta)$ is the average convergence interior to
2439: radius $\theta$,
2440: $\theta_{\rm inn}$ and $\theta_{\rm out}$ are the inner and
2441: outer radii of the annular background region in which the mean
2442: background contribution,
2443: $\bar{\kappa}_b\equiv
2444: \bar{\kappa}(\theta_{\rm inn}<\vartheta <\theta_{\rm out})$,
2445: is defined;
2446: the $\gamma_+(\theta) = \bar{\kappa}(\theta)-\kappa(\theta)$
2447: is an azimuthal average of the tangential
2448: component of the gravitational shear at radius $\theta$
2449: (Fahlman et al. 1994),
2450: which is observable in the weak lensing limit: $\gamma_+(\theta)\approx
2451: \langle g_+(\theta)\rangle$.
2452: %%%%%
2453: This cumulative mass estimator
2454: subtracts from the mean convergence
2455: $\bar{\kappa}( \theta)$
2456: a constant
2457: $ \bar{\kappa}_{\rm bg}$
2458: %=\bar{\kappa}(\theta_{\rm inn}<\vartheta <\theta_{\rm out})$
2459: %for fixed inner and outer radii,
2460: %$\theta_{\rm inn}$ and $\theta_{\rm out}$,
2461: for all apertures $\theta$ in the measurements,
2462: thus removing any DC component in the control
2463: region $\theta = [\theta_{\rm inn}, \theta_{\rm out}]$.
2464: We note that the $\bar{\kappa}_b$ is
2465: a non-observable free parameter, and we
2466: use this degree-of-freedom
2467: to fix the outer boundary condition,
2468: and hence to derive a convergence profile.
2469: %%%%%
2470: %%%%%
2471:
2472:
2473:
2474:
2475: We compute the aperture densitometry profile
2476: $\zeta_{\rm c}(\theta_m)$
2477: in $N_{\zeta}=10$ logarithmically spaced bins
2478: for $\theta = [\theta_{\rm min},\theta_{\rm max}]$,
2479: which we set to
2480: $\theta_{\rm min}=1'$ and $\theta_{\rm max}=16'$;
2481: the maximum radius $\theta_{\rm max}$ is comparable to
2482: the angular size of
2483: the cluster virial radius,
2484: $\theta_{\rm vir}=15\farcm 7$
2485: ($r_{\rm vir} \sim 2 {\rm Mpc}/h$),
2486: according to the ACS+Subaru-1D model (B05a).
2487: %%%%
2488: The inner and outer background radii
2489: %$(\theta_{\rm inn},\theta_{\rm out})$
2490: are set to $\theta_{\rm inn}=\theta_{\rm max}=16'$ and $\theta_{\rm
2491: out}=19'$, respectively.
2492: %%%%
2493: Using the ACS+Subaru-1D model by B05a,
2494: the mean background level, $\bar\kappa_b$, is calculated to be
2495: $\bar\kappa_b\sim 4.0 \times 10^{-3}$
2496: for our red background population with $\langle
2497: D_{ds}/D_s\rangle=0.693$, at an effective source redshift of $z_{s,D}=0.68$
2498: (see equation [\ref{eq:zD}]).
2499: Note that the current
2500: $1\sigma$ upper limit is
2501: $\kappa(\theta) \simlt 0.01$ at $\theta\simgt 8'$ (B05a).
2502: %%%
2503: For a given boundary condition $\bar{\kappa}_b$, the average
2504: convergence $\bar{\kappa}(\theta_m)$ is estimated as
2505: \begin{equation}
2506: \label{eq:z2avk}
2507: \bar{\kappa}(\theta_m) = \zeta_{\rm c}(\theta_m) + \bar{\kappa}_b.
2508: \end{equation}
2509: Then, we define a discretized estimator for $\kappa$ as
2510: \begin{equation}
2511: \label{eq:zeta2kappa}
2512: \kappa_m\equiv
2513: \kappa(\overline{\theta}_m)
2514: = \alpha_2^m \zeta(\theta_{m+1})
2515: - \alpha_1^m \zeta(\theta_{m})
2516: +(\alpha_2^m-\alpha_1^m)\bar{\kappa}_b,
2517: \end{equation}
2518: where
2519: \begin{equation}
2520: \alpha_1^m = \frac{1}{2\Delta\ln\theta_m}
2521: \left(
2522: \frac{\theta_{m}}{ \overline{\theta}_m }
2523: \right)^2, \, \,
2524: %%
2525: \alpha_2^m = \frac{1}{2\Delta\ln\theta_m}
2526: \left(\frac{\theta_{m+1}}{ \overline{\theta}_m }\right)^2,
2527: \end{equation}
2528: and
2529: $\bar\theta_m$ is the weighted center of the $m$th radial bin
2530: ($m=1,2,...,N_{\zeta}-1$; see Appendix \ref{app:kappad}).
2531: %%%
2532: It is worth noting that, unlike strong-lensing based boundary
2533: conditions (e.g., B05a),
2534: %where the location of Einstein rings or the model-predicted
2535: %interior mass is used to normalize $\kappa$,
2536: this method utilizes an outer boundary condition on the mean background
2537: density $\bar\kappa_b$ to derive a $\kappa$ profile (see Schneider \&
2538: Seitz 1995 for an alternative method for a direct inversion of the mass
2539: profile).
2540: The error covariance matrix $C_{mn}$ of $\kappa_m$ is expressed as
2541: \begin{eqnarray}
2542: C_{mn} \equiv \langle \delta\kappa_m \delta\kappa_n \rangle
2543: &=&
2544: \alpha_2^m \alpha_2^n C^{\zeta}_{m+1,n+1}
2545: +
2546: \alpha_1^m \alpha_1^n C^{\zeta}_{m,n}\nonumber\\
2547: &-&
2548: \alpha_1^m \alpha_2^n C^{\zeta}_{m,n+1}
2549: -
2550: \alpha_2^m \alpha_1^n C^{\zeta}_{m+1,n},
2551: \end{eqnarray}
2552: where $C^{\zeta}_{mn}\equiv \langle \delta\zeta_m\delta\zeta_n\rangle$
2553: is the bin-to-bin error covariance matrix of the aperture densitometry
2554: measurements which is calculated by propagating the rms errors
2555: $\sigma_+(\theta_m)$ for the tangential shear measurement (Okabe \&
2556: Umetsu 2008).
2557:
2558:
2559:
2560: In the non-linear regime, however,
2561: the $\gamma_+(\theta)$ is not a direct observable.
2562: Therefore, non-linear corrections need to be taken into account
2563: in the mass reconstruction process.
2564: %%%
2565: In the subcritical regime (i.e., outside the critical curves),
2566: the $\gamma_+(\theta)$
2567: can be expressed in terms of the
2568: the averaged tangential reduced shear as
2569: $\langle g_+(\theta) \rangle \approx
2570: \gamma_+(\theta)/[1-\kappa(\theta)]$
2571: assuming a quasi circular symmetry in the projected mass distribution
2572: (B05a; Umetsu et al. 2007).
2573: This non-linear equation (\ref{eq:zeta})
2574: for $\zeta_{\rm c}(\theta)$ can be solved by an iterative procedure:
2575: Since the weak lensing limit ($\kappa,|\gamma|,|g|\ll 1$) holds in the
2576: background region $\theta_{\rm inn}\le \theta \le \theta_{\rm
2577: max}$, we have the following iterative equation for $\zeta_{\rm
2578: c}(\theta)$:
2579: \begin{eqnarray}
2580: \zeta_{\rm c}^{(k+1)}(\theta)
2581: &\approx&
2582: 2\int_{\theta}^{\theta_{\rm inn}}\!d\ln\theta'
2583: \langle g_+(\theta)\rangle [1-\kappa^{(k)}(\theta)] \nonumber\\
2584: &+& \frac{2}{1-(\theta_{\rm inn}/\theta_{\rm out})^2}\int_{\theta_{\rm
2585: inn}}^{\theta_{\rm out}}\! d\ln\theta'
2586: \langle g_+(\theta')\rangle,
2587: \end{eqnarray}
2588: where $\zeta_{\rm c}^{(k+1)}$
2589: represents the
2590: aperture densitometry in the $(k+1)$th
2591: step of the iteration
2592: $(k=0,1,2,...,N_{\rm iter})$;
2593: the $\kappa^{(k+1)}$ is
2594: calculated from $\zeta_{\rm c}^{(k+1)}$ using equation (\ref{eq:zeta2kappa}).
2595: This iteration is preformed by
2596: starting with $\kappa^{(0)}=0$ for all radial bins, and repeated
2597: until convergence is reached at all radial bins. For a fractional
2598: tolerance of $1\times 10^{-5}$, this iteration procedure converges
2599: within $N_{\rm iter} \sim 10$ iterations.
2600: We compute errors for $\zeta_{\rm c}$ and $\kappa$
2601: with the linear approximation.
2602:
2603:
2604: Figure \ref{fig:zeta2kappa} shows the resulting model-independent mass
2605: profile (squares) of A1689 with decorrelated error bars, along with the
2606: results without the non-linear corrections (triangles).
2607: Without the non-linear corrections, central bins at $\theta \simlt 3'$
2608: are underestimated by $\sim 15\%$ at maximum.
2609: %%% revised version
2610: %The outer radial bins at $\theta \approx 7\arcmin$ and $12\arcmin$ are
2611: %consistent with a null signal to within $1\sigma$.
2612: %%% submitted version
2613: %The outer profile at $\theta \simgt 6'$ is consistent with a null signal
2614: %to within $1\sigma$.
2615: The reconstructed convergence in the first bin is
2616: %%% revised version :: kappa(1\farcm 2) = 0.322 \pm 0.15
2617: $\kappa(1\farcm 2)= 0.32 \pm 0.15$
2618: %%% submitted version :: wrong background aperture used
2619: %$\kappa(1\farcm 2)= 0.316 \pm 0.122$
2620: (decorrelated $1\sigma$ error),
2621: which is consistent at the $\sim 1\sigma$ level
2622: with the previous Subaru 1D results (B05a):
2623: $\kappa(1\farcm 2)=0.44^{+0.11}_{-0.12}$ ($1\sigma$) for the red
2624: background population with $\langle D_{ds}/D_s\rangle\approx 0.693$.
2625: %based on the combined Subaru distortion and depletion profiles.
2626: %%%%%
2627: We note that B05a utilized an inner boundary condition on the mean
2628: $\kappa$ interior to $\theta_{\rm min}=1'$ based on the ACS mass profile.
2629:
2630:
2631: Force fitting an NFW profile to the derived $\kappa$ profile yields
2632: %% @@submitted version
2633: %% $M_{\rm vir}=(1.51\pm 0.26)\times 10^{15} M_{\odot}$ ($r_{\rm vir}=
2634: %% 1.87\pm 0.10 {\rm Mpc}/h$) with the adopted prior $c_{\rm vir}\le
2635: %% 30$,
2636: %% @@revised version (2008/04/08)
2637: $M_{\rm vir}=(1.48\pm 0.27)\times 10^{15} M_{\odot}$ and $c_{\rm
2638: vir}=27.3^{+2.7}_{-19.3}$
2639: with the adopted prior $c_{\rm vir}\le 30$,
2640: %($r_{\rm vir}=
2641: % 1.87\pm 0.10 {\rm Mpc}/h$) with the adopted prior $c_{\rm vir}\le
2642: where $\chi^2_{\rm min}/{\rm dof}=5.2/7 (7)$.
2643: %%%
2644: It is interesting to compare the reconstructed $\kappa$ profile with
2645: the tangential shear profile, $\langle g_{+}(\theta)\rangle$,
2646: which is a direct observable with uncorrelated measurement errors
2647: (see equation [\ref{eq:gt}]). In B05a we show the tangential distortion
2648: profile for the same red background sample as used in the present study
2649: (see Figure 1 of B05a, or Figure \ref{fig:gt} below).
2650: %%%
2651: The best-fitting NFW model for the shear profile is
2652: given by $M_{\rm vir}=(1.51\pm 0.26)\times 10^{15}M_{\odot}$ and
2653: a high concentration,
2654: $c_{\rm vir}=20^{+8.8}_{-5.3}$ ($\chi^2_{\rm min}/{\rm dof}=5.0/8$),
2655: which is in good agreement with the results from the 1D reconstruction
2656: based on the aperture densitometry.
2657: %%%%
2658: Such a detailed agreement ensures the validity of the boundary condition
2659: for a shear-based mass reconstruction.
2660: %%%
2661: We note that simply assuming $\bar\kappa_b=0$ yields similar results,
2662: $M_{\rm vir}=(1.51\pm 0.3) \times 10^{15}M_{\odot}$
2663: ($c_{\rm vir}\le 30$)
2664: with $\chi^2_{\rm min}/{\rm dof}=5.5/7 (5)$,
2665: being consistent within the $1\sigma$ uncertainty.
2666:
2667:
2668:
2669:
2670: \subsection{Combining Strong and Weak Lensing Results}
2671: \label{subsec:acs+subaru}
2672:
2673: \subsubsection{ACS Constraints}
2674: \label{subsubsec:acs}
2675:
2676: As demonstrated in B05a and the previous sections,
2677: it is crucial to have information on the
2678: central mass distribution in order to
2679: derive useful constraints on the concentration parameter,
2680: and hence to distinguish the NFW profile from others.
2681: To do this, we constrain the two NFW parameters
2682: from $\chi^2$ fitting to the
2683: combined Subaru weak lensing and ACS
2684: strong lensing observations following B05a and Oguri et al. (2005):
2685: \begin{equation}
2686: \label{eq:chi2_combined}
2687: \chi^2=\chi^2_{\rm WL} + \chi^2_{\rm SL},
2688: \end{equation}
2689: where the $\chi^2_{\rm WL}$ for the Subaru observations is defined by
2690: equation (\ref{eq:chi2_wl}).
2691: For the ACS data we use an azimuthally averaged profile of the $\kappa$
2692: map well constrained by the strong lensing observations of B05b;
2693: this profile is given in $N_{\rm SL}=12$ bins linearly spacing
2694: over the range $\theta = [0\farcm 077, 0\farcm 97]$
2695: (see Figure 22 of B05b, or Figure 3 of B05a), and the amplitude
2696: is scaled to $\langle D_{ds}/D_s\rangle=0.693$ of our red background
2697: sample from $D_{ds}/D_s=0.881$ at $z_s=3$.
2698: It is important to note that the ACS strong lensing analysis of B05b
2699: unveiled the secondary mass clump associated with a small clump of
2700: galaxies (Teague et al. 1990; Czoske 2004). However, the mass
2701: contribution of this subclump is only a small fraction of
2702: the cluster mass component (see Figures 21 and 22 of B05b)
2703: and it only slightly perturbs the tangential critical curve
2704: (see B05b). The ACS mass profile is corrected for the subclump, where
2705: the secondary clump region is locally masked out when taking an azimuthal
2706: average (B05b).
2707: %%%
2708: By combining the Subaru and ACS lensing analyses,
2709: we can trace the cluster mass distribution over a
2710: large range in amplitude $\kappa\sim [10^{-3},1]$ and
2711: in projected radius $r=[10^{-2},2] {\rm Mpc}/h$.
2712: %%%
2713: The $\chi^2$ function for the ACS observations is expressed as
2714: \begin{equation}
2715: \chi^2_{\rm SL} = \sum_{i=1}^{N_{{\rm SL}}}\frac{
2716: (\kappa_i-\hat\kappa_i)^2
2717: }
2718: {\sigma_i^2}
2719: \end{equation}
2720: where $\hat{\kappa}(\btheta_i)$ is the model prediction of the
2721: NFW halo for the $i$th pixel ($i=1,2,...,N_{\rm pix}$),
2722: and $\sigma_i$ is the $1\sigma$ error for $\kappa_i$;
2723: the bin width of the
2724: ACS-derived convergence profile is sufficiently broad to
2725: ensure that the errors between different bins are independent (B05b)
2726:
2727: \subsubsection{Einstein-Radius Constraint}
2728: \label{subsubsec:ein}
2729:
2730: Alternatively, as a model independent constraint,\footnote{We
2731: note that this is based on the assumption of the circular symmetry in
2732: the projected lens mass distribution, and that a rather tight prior
2733: (estimated from data) is applied to one of the ``parameters'' (i.e.,
2734: $\theta_{\rm E}$). }
2735: we can utilize the observed location of tangential critical curves
2736: traced by giant arcs and multiply imaged background galaxies, defining
2737: an approximate Einstein radius, $\theta_{\rm E}$.
2738: This radius is determined from strong lensing modeling of many multiple
2739: images visible around these clusters in deep HST/ACS images (B05b),
2740: and corresponds to the theoretically extreme value of the ellipticity,
2741: $g_+(\theta_{\rm E})=1$. This Einstein-radius constraint
2742: complements in a model independent manner the weak lensing shape
2743: measurements.
2744:
2745: We constrain the NFW model parameters ($c_{\rm vir}, M_{\rm vir}$)
2746: by combining weak lensing profiles and the Einstein-radius
2747: constraint. We define the $\chi^2$ function for combined weak lensing
2748: distortion and Einstein-radius constraints by
2749: \begin{equation}
2750: \label{eq:chi2_g_rein}
2751: \chi^2 = \sum_m
2752: \frac{\left[ \langle g_{+}(\theta_m) \rangle - \hat{g}_{+}(\theta_m)
2753: \right]^2}
2754: {\sigma_{+}^2(\theta_m)} +
2755: \frac{\left[1-\hat{g}_+(\theta_{\rm E},z_{\rm E})\right]^2}
2756: {\sigma_{+,{\rm E}}^2},
2757: \end{equation}
2758: where the first term is the $\chi^2$-function for the Subaru
2759: tangential shear measurements (\S \ref{subsec:zeta}) and the second term
2760: is the $\chi^2$ function for the Einstein radius constraint;
2761: $\hat{g}_{+}(\theta_m)$ is the NFW model prediction
2762: for the reduced tangential shear at $\theta=\theta_m$ calculated for the
2763: red background sample (see \S \ref{subsec:red}),
2764: $\hat{g}_{+}(\theta_{\rm E}, z_{\rm E})$ is the
2765: model prediction for the reduced tangential shear at $\theta=\theta_{\rm
2766: E}$, evaluated at the arc redshift, $z_s=z_{\rm E}$, and $\sigma_{+,{\rm
2767: E}}$ is the rms error for $g_+(\theta_{\rm E}, z_{\rm E})$.
2768: Following B05b, we take $\theta_{\rm E}=45\arcsec$ at $z_E=1$, and
2769: assume conservatively a 10\% error for the Einstein radius:
2770: $\sigma_{\theta_{\rm E}}/\theta_{\rm E}=0.1$. We then
2771: propagate this error to $g_+$ assuming a singular isothermal sphere (SIS)
2772: model; At $r\simlt r_s$, the density slope of NFW is shallower than that
2773: of SIS (see Figure 1 of Wright \& Brainerd 2000), so that this gives a
2774: fairly conservative estimate of $\sigma_{+,{\rm E}}$.
2775: For the SIS model, $\partial\ln g_+/\partial\ln \theta_{\rm E}=2$
2776: %$\Delta\ln g_+ = 2\Delta\ln \theta_{\rm E}$
2777: at $\theta=\theta_{\rm E}$, so that $\sigma_{+,{\rm E}} =
2778: 2\sigma_{\theta_{\rm E}}/\theta_{\rm E}=0.2$.
2779: %for our assumed $10\%$ uncertainty in $\theta_{\rm E}$.
2780:
2781:
2782: Alternatively we can combine the $\kappa$ profile reconstructed from
2783: weak lensing distortion measurements (\S \ref{subsec:zeta}) with the
2784: Einstein-radius constraint. For an axially symmetric lens, the
2785: Einstein-radius constraint is simply written as
2786: $\bar{\kappa}(\theta_{\rm E},z_{\rm E})=1$. Thus, the $\chi^2$ function
2787: is now expressed as
2788: \begin{equation}
2789: \label{eq:chi2_k_rein}
2790: \chi^2 = \sum_{m,n}
2791: \left[\kappa(\theta_m) - \hat\kappa(\theta_m) \right]
2792: \left(C^{-1}\right)_{mn}
2793: \left[\kappa(\theta_n) - \hat\kappa_n(\theta_n) \right]
2794: +
2795: \frac{
2796: \left[ 1-\hat{\bar{\kappa}}(\theta_{\rm E},z_{\rm E})
2797: \right]^2
2798: }
2799: {\sigma_{\bar\kappa,{\rm E}}^2},
2800: \end{equation}
2801: where $\hat{\bar{\kappa}}(\theta_{\rm E}, z_{\rm E})$ is the model
2802: prediction for the average convergence interior to the Einstein radius
2803: $\theta_{\rm E}$ evaluated at the arc redshift of $z_s=z_{\rm E}$, and
2804: $\sigma_{\bar\kappa, {\rm E}}$ is the rms error for
2805: $\bar\kappa(\theta_{\rm E}, z_{\rm E})$,
2806: which we take as
2807: $\sigma_{\bar\kappa,{\rm E}} = \sigma_{\theta_{\rm E}}/\theta_{\rm
2808: E}=0.1$.
2809:
2810:
2811: %normalize but leave a degree of freedom corresponding to the
2812: %uncertainty in the Einstein radius.
2813:
2814:
2815: \subsubsection{Comparison}
2816: \label{subsubsec:comparison}
2817:
2818:
2819: Table \ref{tab:nfwfit} summarizes for each mass reconstruction
2820: the best-fitting NFW parameters
2821: $(M_{\rm vir},c_{\rm vir})$, the minimized $\chi^2$ value with respect
2822: to the degrees of freedom, and the predicted Einstein radius
2823: $\theta_{\rm E}$ for a fiducial source at $z_s=1$.
2824: Here for each of the Subaru reconstructions, we compare the
2825: best-fit NFW parameters with and without the ACS inner profile combined,
2826: as indicated in the third column of Table \ref{tab:nfwfit}.
2827: When combined with the inner ACS profile,
2828: all of the Subaru reconstructions yield a virial mass
2829: in the range $M_{\rm vir}\sim (1.8-2.1)\times 10^{15}M_{\odot}$
2830: and a high concentration in the range $c_{\rm vir}\sim 13-15$,
2831: properly reproducing the observed Einstein radius of $\sim 45\arcsec$
2832: at $z_s=1$ (B05b). In particular, fitting an NFW
2833: profile to the combined ACS and Subaru-2D data based on the MEM+ mass
2834: reconstruction (see Table \ref{tab:memmethod}) yields $M_{\rm
2835: vir}=(2.10\pm 0.17)\times 10^{15}M_{\odot}$ and $c_{\rm
2836: vir}=12.7^{+1.0}_{-0.9}$ ($\chi^2_{\rm min}/{\rm dof}=335/846$;
2837: effective dof of $433$ without including upper limit bins with
2838: $\kappa<0$), with the predicted Einstein radius of $\theta_{\rm
2839: E}=45.3^{+5.9}_{-6.2}$ arcsec.
2840: %%%%
2841:
2842:
2843: The 2D-based results here can be compared with the corresponding results
2844: from the Subaru 1D analysis of B05a (see Table \ref{tab:nfwfit}) based
2845: on the combined distortion and magnification measurements. The
2846: combined ACS and Subaru-1D convergence profile is well fitted by an NFW
2847: profile with $M_{\rm vir}=(1.9\pm 0.2)\times 10^{15}M_{\odot}$ and
2848: $c_{\rm vir}=13.7^{+1.4}_{-1.1}$ (B05a), with the predicted Einstein radius of
2849: $\theta_{\rm E}=45.4^{+7.6}_{-6.9}$ arcsec, in good agreement with
2850: the present full 2D analysis within the statistical uncertainties;
2851: this agreement between the 1D and 2D analyses supports
2852: the assumption of quasi-circular symmetry in the projected
2853: mass distribution.
2854: It is interesting to compare these results with different combinations
2855: of lensing measurements having different systematics.
2856: Another combination of the ACS and Subaru 1D convergence profile,
2857: derived from the shear-based $\zeta$-statistic measurements (\S
2858: \ref{subsubsec:ein}), yields fairly
2859: consistent results: $M_{\rm vir}=1.91^{+0.24}_{-0.20}\times
2860: 10^{15}M_{\odot}$ and $c_{\rm vir}=13.7^{+1.5}_{-1.3}$. This consistency
2861: clearly demonstrates that our results here are insensitive to systematic
2862: errors in the lensing measurements, such as the shear calibration
2863: error.
2864: %%%%
2865:
2866: In Figure \ref{fig:mass} we make a direct comparison
2867: between model-independent convergence profiles of A1689 from
2868: different weak lensing techniques, along with the ACS-derived
2869: inner profile obtained by B05b (filled triangles).
2870: The filled circles with error bars show
2871: the results based on the 2D $\kappa$ map reconstructed
2872: from an entropy-regularized maximum-likelihood combination
2873: of the lens magnification and distortion of red background galaxies
2874: (MEM+, Table \ref{tab:memmethod}).
2875: %the error bars are correlated,
2876: %and are given by the square root of the diagonal part of the
2877: %bin-bin covariance matrix (equation [\ref{eq:bin2bincovar}]).
2878: %%%%%%%%%%%%%%%%
2879: The open triangles represent the $\kappa$ profile
2880: from the non-linear $\zeta_{\rm c}$-statistic measurements.
2881: %where decorrelated error bars are shown.
2882: The filled circles show the results from the Subaru 1D analysis
2883: by B05a based on the combined magnification and distortion
2884: profiles of the same red background population as
2885: in the present study.
2886: %measured from the same red background sample as
2887: %in the prsent study.
2888: %The 1D- and 2D-based NFW models from the respective combined
2889: %ACS and Subaru data are also shown as solid and dashed curves,
2890: %respectively.
2891: It is clearly seen from Figure \ref{fig:mass} that
2892: the Subaru-derived $\kappa$ profiles are all in good agreement
2893: within the statistical uncertainties, over the full range of radii up to
2894: $\sim 2{\rm Mpc}/h$.
2895: %%%%
2896: The entire mass profile traced by the combined ACS and Subaru
2897: information is well described by a single NFW profile (solid and dashed)
2898: with a high concentration ($c_{\rm vir}\sim 13-14$).
2899: For comparison an NFW profile with a low concentration $c_{\rm vir}=5$
2900: normalized to the observed Einstein radius of $45\arcsec$ (at $z_s=1$)
2901: is shown as a dotted curve.
2902: Such a low concentration profile as favored by
2903: the standard $\Lambda$CDM model
2904: clearly overpredicts the outer
2905: profile constrained by the Subaru weak lensing observations.
2906:
2907: Figure \ref{fig:joint} shows the
2908: $68\%$, $95\%$, and $99.7\%$ confidence levels
2909: ($\Delta\chi^2=2.3, 6.17$, and $11.8$) in
2910: the $(c_{\rm vir},M_{\rm vir})$-plane for
2911: each of the three-different 2D MEM reconstructions
2912: (from top to bottom: MEM-S, MEM, and MEM+,
2913: see Table \ref{tab:memmethod})
2914: with (right) and without (left) the central ACS profile
2915: at $10{\rm kpc}/h \simlt r \simlt 180 {\rm kpc}/h$
2916: combined.
2917: As shown in the left panels, $M_{\rm vir}$ is well constrained
2918: by the Subaru data alone, while the Subaru constraint on
2919: $c_{\rm vir}$ is rather weak, allowing a wide range of
2920: the concentration parameter, $c_{\rm vir}$. The complementary
2921: ACS observations, when combined with the Subaru observations,
2922: significantly narrow down the uncertainties on $c_{\rm vir}$
2923: (right panels), placing stringent constraints on the inner mass
2924: profile.
2925: For each sub-panel, the observed constraints on
2926: the Einstein radius,
2927: $\theta_{\rm E}\simeq 45\arcsec$ at $z_s=1$,
2928: are shown as a dashed curve.
2929: This clearly demonstrates that the ACS-derived constraints
2930: on the $\kappa$-field ensure the correct size of the observed
2931: Einstein radius.
2932:
2933:
2934:
2935: Table \ref{tab:nfwref} lists the best-fitting NFW parameters
2936: obtained in different studies. Here we quote as the NFW parameters
2937: $(c_{200},M_{200})$ evaluated at a specific fractional overdensity of
2938: $\Delta_c=200$ with respect to the critical density $\rho_{\rm
2939: crit}(z_d)$ for closure of the universe at $z_d=0.183$ as well as those
2940: in terms of the virial properties, $(c_{\rm vir},M_{\rm vir})$.
2941: We convert a given set of $(c_{200},M_{200})$ into $(c_{\rm vir},
2942: M_{\rm vir})$ assuming a spherical NFW density profile (see, e.g.,
2943: Appendix A of Shimizu et al. 2003).
2944: %%%%%%%%%%%%%%%%%%%
2945: We show in Figure \ref{fig:joint_gt} the same confidence levels as in
2946: Figure \ref{fig:joint} but for the Subaru $g_+$ profile of B05a.
2947: Also shown are best-fit sets of $(c_{\rm vir}, M_{\rm vir})$
2948: taken from
2949: %% the combined strong and weak lensing analysis of
2950: %Halkola et al. (2006) and the CFHT weak lensing analysis of Limousin et
2951: %al. (2007)
2952: Halkola et al. (2006) and Limousin et al. (2007) as well as from our
2953: combined ACS and MEM+ results.
2954: %%%%
2955: %%%%
2956: The best-fitting NFW parameters for the Subaru $g_+$ profile (cross)
2957: are $M_{\rm vir}=1.51^{+0.27}_{-0.24}\times 10^{15}M_{\odot}$
2958: and $c_{\rm vir}=20.0^{+8.8}_{-5.3}$ (Table \ref{tab:nfwfit}),
2959: being marginally consistent with the very high concentration
2960: of $c_{\rm vir}=27.2^{+3.5}_{-5.7}$
2961: derived by Medezinski et al. (2007)
2962: \footnote{
2963: Unlike our analysis here, Medezinski et al. (2007) used
2964: the observed constraints on the Einstein radius
2965: when fitting an NFW profile to
2966: the $g_{+}$ profile measured from
2967: the combined red and blue
2968: sample of the background.}.
2969: The $\kappa$ field derived from the present full 2D analysis,
2970: in conjunction with the central ACS profile, favors
2971: a slightly lower, but still high, concentration of $c_{\rm vir}\sim 13$
2972: (triangle), reproducing the observed Einstein radius of
2973: $\sim 45\arcsec$ at $z_s=1$.
2974: %%%%
2975: On the other hand,
2976: Limousin et al. (2007) found from their CFHT weak lensing analysis
2977: a concentration of $c_{\rm vir} = 9.6\pm 2.0$
2978: %($c_{200}=7.6\pm 1.6$)
2979: and a virial mass of
2980: $M_{\rm vir} = 1.5^{+0.3}_{-0.2}\times 10^{15}M_{\odot}$
2981: (filled circle),
2982: corresponding to the Einstein radius of $24\pm 11$ arcsec (at $z_s=1$),
2983: %%completely
2984: inconsistent with the observed Einstein radius.
2985: %%%%
2986: The results from Halkola et al. (2006) with
2987: $c_{\rm vir} = 9.6\pm 0.4$
2988: and $M_{\rm vir} = 2.6^{+0.2}_{-0.1}\times 10^{15}M_{\odot}$
2989: (square)
2990: come closer to the observed Einstein radius,
2991: being marginally consistent within the $1\sigma$ uncertainty
2992: ($\theta_{\rm E}=39^{+6}_{-3}$ arcsec at $z_s=1$),
2993: based on almost exactly the same shear data used in B05a but with a
2994: different weighting which prefers their inner strong-lensing based
2995: profile where the ACS data imply a shallower
2996: slope, as discussed in Medezinski et al. (2007) and Limousin et
2997: al. (2007).
2998:
2999:
3000:
3001: It is useful to compare the results from different lensing studies
3002: %%(see Table \ref{tab:nfwref})
3003: %%model predictions
3004: in terms of the tangential distortion, $g_+$, that is directly
3005: observable in weak lensing.
3006: Figure \ref{fig:gt} shows the radial profiles of $g_+$ and $g_\times$
3007: %the tangential
3008: %distortion, $g_+$, and the $45^\circ$-rotated component, $g_\times$,
3009: derived in B05a
3010: from the same red background sample as used in this work,
3011: where we have added the observed Einstein-radius constraint of
3012: $\theta_{\rm E}=45\pm 5$ arcsec ($z_s=1$), translated to the mean depth
3013: of the red background sample (\S \ref{subsec:red}), marking the point of
3014: maximum distortion, $g_+=1$ ($\theta_{\rm E}=39\pm 4$ arcsec). The
3015: ACS+Subaru-2D NFW model (solid curve)
3016: fits reasonably well with the entire distortion profile,
3017: $r\sim [80,2000] {\rm kpc}/h$, although it slightly overpredicts the
3018: outer profile at $\theta \simgt 9\arcmin$, meaning that the observed
3019: $g_+$ profile is steeper at large radii, as pointed out in
3020: Medezinski et al. (2007).
3021: %%%%
3022: In contrast the best-fit NFW model of Limousin et al. (2007),
3023: shown with the dashed curve, is in agreement
3024: with the Subaru outer profile, particularly
3025: at $10\arcmin \simlt \theta \simlt 18\arcmin$,
3026: but underpredicts significantly the inner $g_+$ profile,
3027: leading to a significant underprediction of the Einstein radius.
3028: %%($\theta_{\rm E}=24\pm 11$ arcsec for $z_s=1$).
3029: On the other hand,
3030: the dotted curve in Figure \ref{fig:gt} shows the NFW profile of
3031: Halkola et al. (2006),
3032: %for a simultaneous fit to their ACS inner mass
3033: %profile and the Subaru distortion profile of B05a.
3034: which is overall in good agreement with the observed $g_+$ profile,
3035: and with the NFW prediction of this work,
3036: but increasingly overestimates the distortion signal with radius,
3037: %being too shallow at large radii $\simgt 10\arcmin$,
3038: and slightly underpredicts the Einstein radius.
3039: This is again consistent with that the discrepancy between the derived
3040: NFW parameters of this work and Halkola et al. (2006) is due to
3041: the relative weights in the fitting procedure
3042: assigned differently to the strong and weak lensing
3043: measurements; that is, a relatively higher
3044: weight is given by Halkola et al. (2006)
3045: to the shallower inner profile at $r\simlt 40 {\rm kpc}/h$ constrained
3046: by the radial arcs,\footnote{The location of the observed radial critical
3047: curve is at $\theta\approx 17\arcsec$ as found in B05b.}
3048: and consequently, the location of the outer critical curve is slightly
3049: underpredicted. However, we argue this does not indicate a discrepancy
3050: between our strong and weak lensing results, but is simply that the form
3051: of the NFW profile is not entirely consistent with our data ranging from
3052: $10$ to $2000 {\rm kpc}/h$; the best-fit NFW profile is either too
3053: shallow at large radii or too steep at small radii, depending on the
3054: radial limits being examined.
3055:
3056:
3057:
3058:
3059:
3060: \subsection{Systematic Errors of the Concentration Parameter}
3061: \label{subsec:sys}
3062:
3063:
3064:
3065: In this subsection we address the issues of systematic errors on the
3066: halo concentration $c_{\rm vir}$ inherent in our lensing measurements and
3067: analysis methods. Here we discuss the following potential sources of
3068: systematic uncertainty: (i) selection criteria for the red
3069: background sample (\S \ref{subsec:red}), (ii) background redshift
3070: distribution (\S \ref{subsec:red}), (iii)
3071: magnification analysis (\S \ref{subsec:mag}, \S
3072: \ref{subsec:declustering}), (iv) inner boundary condition (\S
3073: \ref{subsec:acs}),
3074: (v) strong lensing model (\S \ref{subsec:acs} and \S \ref{subsubsec:ein}),
3075: (vi) entropic prior (\S \ref{subsec:massrec}), and
3076: (vii) shear calibration error.
3077: In Table \ref{tab:sys} we summarize the systematic errors of
3078: the halo concentration, $c_{\rm vir}$,
3079: given in fraction of the best-fit value of
3080: $c_{\rm vir}=12.7$ obtained from the combined ACS and Subaru-2D (MEM+)
3081: data. Adding all potential sources of error
3082: in quadrature, the total uncertainty in our determination of
3083: the halo concentration is $c_{\rm vir}=12.7 \pm 1 ({\rm statistical})
3084: \pm 2.8 ({\rm systematic})$.
3085: %a joint fit to the ACS and Subaru-2D (MEM+) data.
3086: %e use the 2D MEM+ method for reconstructing the
3087: %$\kappa$ map, and combine it with the ACS inner mass profile to deduce
3088: %the best-fit NFW parameters, unless otherwise noted.
3089:
3090:
3091:
3092:
3093:
3094: \subsubsection{Background Selection Criteria}
3095: \label{subsubsec:sys_zs}
3096:
3097: As clearly demonstrated by B05a and Medezinski et al. (2007), a secure
3098: background selection is critical in the cluster weak lensing
3099: analysis, in order to avoid dilution of the distortion signal by
3100: contamination of unlensed cluster member galaxies. The degree of
3101: dilution, which is proportional to the fraction of cluster membership,
3102: is particularly prominent in the central region of rich clusters,
3103: as in the case of A1689 (see Medezinski et al. 2007). Practically, a
3104: reliable background sample can be defined by selecting galaxies with
3105: colors redder than the cluster E/S0 sequence (see \S \ref{subsec:red}).
3106: Figure \ref{fig:dilution} shows that the dilution effect becomes
3107: significant when the lower (bluer) color limit of the entire red sample
3108: is decreased below a color of $\Delta(V-i')\sim 0.1$, while no signature
3109: of systematic variations is seen when the lower color limit is increased
3110: above $\Delta(V-i')\sim 0.1$, ensuring that the dilution effect is
3111: almost negligible there. Based on this, we defined our red background
3112: sample by choosing a conservative color limit as $\Delta(V-i')>0.22$.
3113:
3114: Here we vary the lower color limit of the red sample in the interval
3115: of $[0.1,0.5]$ where the dilution effect is negligible. At the lower
3116: color limit of $\sim 0.5$, however, the number of red galaxies is
3117: decreased by about $40\%$ (see Figure \ref{fig:dilution}), when compared
3118: with our fiducial sample of $\Delta(V-i')>0.22$, leading to noisier
3119: results.
3120: We generate red samples at the lower color limit of $0.1$,
3121: $0.15$, $0.2$, ..., $0.5$, and assess the scatter in the best-fit
3122: concentration parameter, $c_{\rm vir}$.
3123: We find the error distribution is fairly random, with a small spread of
3124: $\Delta c_{\rm vir}\sim \pm 0.5$, corresponding to a $\sim 4\%$
3125: fractional systematic error.
3126:
3127: \subsubsection{Background Redshift Distribution}
3128: \label{subsubsec:sys_pz}
3129:
3130: % redshift distribution plays a crucial role in
3131:
3132: Redshift information of background galaxies plays a crucial role in the
3133: determination of cluster mass profiles
3134: In order to convert the observed lensing signal into physical mass
3135: units, one needs to evaluate the mean distance ratio
3136: $\langle D_{ds}/D_s\rangle$ over the redshift distribution of background
3137: galaxies (\S \ref{subsec:red}). An overestimate of the source redshift
3138: will systematically lead to an underestimate of the cluster mass, and
3139: vice versa.
3140: In this way, the uncertainty in the
3141: background redshift distribution will lead to systematic errors of the
3142: cluster mass determination. The level of uncertainty depends on the
3143: %%relative distance between the lens and the source,
3144: lensing geometry, and is less
3145: significant for low-$z$ clusters (say, $z_d\simlt 0.2$).
3146: For purely weak lensing based data, this effect is less important for
3147: the determination of the halo concentration, because the identification
3148: of the inner characteristic radius ($r_s$) is basically independent of
3149: the background redshift. However, when the weak lensing measurements are
3150: combined with inner strong lensing information,
3151: this depth information becomes crucial
3152: for the determination of halo concentration as well, because it
3153: determines the relative normalization between the inner and outer
3154: profiles. That is, an overestimate of the background redshift will cause
3155: an underestimate of the surface mass density, which will increase the
3156: difference between the inner profile derived from strong lensing and the
3157: outer profile from weak lensing,
3158: leading to
3159: an overprediction of the halo concentration.
3160: %This degeneracy of the NFW
3161: %parameters ($c_{\rm vir}$, $M_{\rm vir}$) is clearly demonstrated in
3162: %Figure \ref{fig:joint}.
3163:
3164:
3165:
3166: Here we turn to assess the level of systematic error arising from the
3167: uncertainty in the background redshift distribution. Based on the
3168: multicolor photometry of Capak et al. (2004) in the HDF-N, we obtain
3169: a mean distance ratio of $\langle
3170: D_{ds}/D_s\rangle = 0.693 \pm \pm 0.02$, or a distance-equivalent
3171: redshift of $z_{s,D}=0.68\pm 0.05$,
3172: for our color-magnitude selected red background sample
3173: (\S \ref{subsec:red}; see also Medezinski et al. 2007).
3174: A good agreement has been also found using the COSMOS deep multicolor
3175: photometric catalog of Capak et al. (2007), yielding a similar depth of
3176: $\langle D_{ds}/D_s\rangle \approx 0.703$ (Medezinski et al. 2007 in
3177: preparation), suggesting that the field-to-field variance (cosmic
3178: variance) is not significant, and as small as the statistical uncertainty
3179: obtained here.
3180: We find this level of depth uncertainty will lead to
3181: only a $\sim 1\%$ fractional systematic error of $c_{\rm vir}$.
3182: %%%
3183: Further,
3184: %for a more conservative error estimate of the concentration,
3185: we assign a conservative uncertainty in the
3186: % a generous range of systematic errors in the
3187: (distance-equivalent) source redshift, $z_{s,D}=[0.7,1.0]$, and find a
3188: fractional systematic error of about $10\%$ in the concentration
3189: parameter. It is interesting to note that B05a obtained the best-fit
3190: concentration parameter
3191: of $c_{\rm vir}=13.7^{+1.4}_{-1.1}$ assuming a source redshift of
3192: $z_s(=z_{s,D})=1$. This concentration is slightly higher ($\sim +8\%$) than
3193: the best-fit value of $c_{\rm vir}=12.7\pm 1$ found in this work, but
3194: this level of discrepancy can be easily reconciled by the systematic
3195: error in the assumed depth of B05a: Applying this bias correction to
3196: the results of B05a yields $c_{\rm vir} \approx 12.3$,
3197: which agrees quite well with the present results of $c_{\rm vir}=12.7\pm
3198: 1$.
3199:
3200:
3201: Lensing magnification influences the observed
3202: surface density of background galaxies (Broadhurst et al. 1995), by
3203: expanding the
3204: observed solid angle of the background (area distortion),
3205: and decreasing the effective flux limit of the survey (flux
3206: amplification). Thus the latter effect of magnification bias may change
3207: the redshift distribution of background galaxies depending on the
3208: distance from the cluster center,
3209: which could be a potential source of the
3210: systematic errors in the determination of the halo concentration.
3211: Our red background sample, however, is highly depleted (see Figure
3212: \ref{fig:magbias}), meaning that the area distortion effect is
3213: dominating over the flux amplification of fainter, distant background
3214: galaxies. Indeed, the unlensed count slope measured at our magnitude
3215: limit $i'=25.5$ is $s=d\log N_0(m)/dm\approx 0.22$, being fairly flat as
3216: compared to the blue sample with $s\approx 0.4$, close to the lensing
3217: invariant slope (see equation [\ref{eq:magbias}]).
3218: Consequently, relatively few fainter objects are magnified into the
3219: sample even in the central cluster region, so that
3220: the magnification effect on the source redshift distribution
3221: is fairly negligible for the red background defined
3222: at fainter magnitude limits
3223: For our sample,
3224: magnification at $\theta\sim 2\arcmin$ is $\mu\simeq 1+2\kappa \approx
3225: 1.2$, or about $0.2 {\rm mag}$ of increased depth, corresponding to
3226: the fractional increase of only $\sim 5\%$ in the number of red
3227: background galaxies.
3228: Further, given
3229: the weak dependence of the redshift distribution of faint galaxies on
3230: apparent magnitude (e.g., Medezinski et al. 2007), we do not need to
3231: take this effect into serious consideration for our red background
3232: sample.
3233:
3234: %increase of only $\sim 5\%$
3235: %The magnification bias arises because gravitational lensing changes the
3236: %apparent solid angle of the background but conserves the surface
3237: %brightness, leading to an increase or decrease of the observed
3238: %%%Gravitational deflection of the light ray will
3239: %Magnification bias arises because gravitational lensing changes the
3240: %solid angle of a source but conserves its surface brightness
3241: %The magnification effect will enlarge the sky solid angle, thus
3242: %modifying the source density by a factor 1/mu, .
3243:
3244: Still, it is instructive to consider the potential systematic bias
3245: caused by the magnification effect on the background redshift
3246: distribution. The net effect of depth correction is to reduce the
3247: central surface mass density of the lens, since we attribute the high
3248: lensing signal to the geometric information of the background.
3249: For purely weak-lensing based data, this will lead to a lower
3250: concentration. However, when the strong lensing information in the inner
3251: region is taken into account as well, this depth correction will further
3252: increase the difference between the inner and outer profiles derived
3253: from strong and weak lensing, respectively, thereby enhancing the
3254: concentration. However, the amount of correction is negligible in
3255: practice when the weak lensing analysis is based on the red background
3256: galaxies as discussed above.
3257:
3258:
3259:
3260:
3261: \subsubsection{Magnification Analysis}
3262:
3263: Here we address the systematic uncertainties arising from a particular
3264: treatment and various cuts in the weak lensing magnification analysis.
3265:
3266: \begin{itemize}
3267: \item[(1)]
3268: {Mask area correction} % 7\%
3269: \end{itemize}
3270:
3271: The masking effect due to cluster member galaxies and bright foreground
3272: objects acts to reduce the
3273: apparent number of background galaxies, and this reduction increases
3274: towards the cluster center, leading to an overestimate of the central
3275: depletion signal without the masking correction.
3276: %taken into account.
3277: %%%%%%
3278: In the present study the mask area
3279: of bright objects
3280: is evaluated as the area inside the ellipse of
3281: $\nu_{\rm mask}$-times the major ({\tt A\_IMAGE}) and minor ({\tt
3282: B\_IMAGE}) axes
3283: computed from SExtractor (see Cobb et al. 2006 for a similar
3284: discussion),
3285: where the multiplier is chosen as $\nu_{\rm mask}=3$ so that the
3286: ellipse is visually
3287: consistent with the isophotal detection limit in our SExtractor
3288: configuration (\S \ref{subsec:data}).
3289: Here we adopt a conservative uncertainty of $\nu_{\rm mask}=3\pm 1$ on the
3290: masking factor, and find the corresponding systematic uncertainty of
3291: $\pm 7\%$ in
3292: the halo concentration. The lower limit on $c_{\rm vir}$ is obtained for
3293: $\nu_{\rm mask}=2$, when the mask area correction is underestimated and
3294: hence the depletion signal is overestimated.
3295: %Note that the direction of the correction for $c_{\rm vir}$ is opposite
3296: %when only weak lensing data are used (i.e., wihtout the inner strong
3297: %lensing information; see the discussion in \S \ref{subsubsec:sys_pz}).
3298: %%%%
3299: %Here we vary the value of $\nu_{\rm mask}$ in the range of 2 to 4,
3300: %and assess the corresponding uncertainties in $c_{\rm vir}$.
3301: %For $\nu_{\rm mask}=3\pm 1$,
3302: %As found in B05a, the mask area correction is small (less than several
3303: %percent) and negligible at large radii, but becomes
3304: %increasingly significant at $\theta\simlt 4'$.
3305: %%When $\nu_{\rm mask}=2$, the mask area
3306: %Here we assess the uncertainties in $c_{\rm vir}$ due to a particular
3307: %choice of the mask correction multiplier.
3308:
3309:
3310:
3311: \begin{itemize}
3312: \item[(2)]
3313: {Clustering rejection}
3314: %%: 1.5\%
3315: \end{itemize}
3316:
3317: Similarly, we adopt a conservative uncertainty of $\nu_{\rm clust}=4\pm 1$
3318: in the rejection
3319: threshold $\nu_{\rm clust}$, which has been introduced to downweight
3320: locally the
3321: intrinsic clustering noise
3322: which otherwise perturbs the depletion signal (\S
3323: \ref{subsec:declustering}).
3324: This yields a fractional uncertainty of only $\pm 1.5\%$ in $c_{\rm
3325: vir}$.
3326: We note that the intrinsic clustering of background galaxies is a local
3327: effect, and
3328: hence it does not affect significantly the radial profile fitting.
3329:
3330:
3331:
3332:
3333:
3334: \begin{itemize}
3335: \item[(3)]
3336: {Unlensed count slope}
3337: %%: 3.5\%
3338: \end{itemize}
3339:
3340: The conversion from the observed counts of the red background sample
3341: into the magnification bias $\delta_\mu$ (equation [\ref{eq:magbias}])
3342: depends on the slope parameter $s=d\log{N_0(m)}/dm$ of the unlensed
3343: number counts (\S \ref{subsec:mag}),
3344: which was estimated as $s=0.22\pm 0.03$ from the outer region $\simgt
3345: 10\arcmin$ where the magnification effect is negligibly small
3346: ($\kappa, \delta_\mu \simlt 0.01$). We find that this level of
3347: uncertainty in $s$ will cause an uncertainty in $c_{\rm vir}$ of about
3348: $3.5\%$.
3349:
3350:
3351: \subsubsection{Inner Boundary Condition}
3352: %% 10\%
3353:
3354: Any mass reconstruction technique based on the gravitational shear field
3355: involves a non local process (see equation [\ref{eq:local}]), and hence
3356: it is crucial to have a proper boundary condition for an accurate
3357: determination of the cluster mass profile.
3358: In particular, our MEM+ method is based on the ACS strong lensing
3359: information, which is incorporated as an inner boundary condition
3360: on the central pixel, $\kappa_{\rm c}$ (\S \ref{subsec:acs}).
3361: %It is threfore important to assess the systematic errors in our mass
3362: %reconstruction introduced by uncertainties in the adopted strong lensing
3363: %mass model.
3364: B05a showed that the combined ACS and Subaru mass profile can be well
3365: fitted by a high concentration NFW profile ($c_{\rm vir}\sim 14$) over
3366: the full range of ACS and Subaru data, $r=[10^{-2},2]$ Mpc$/h$. However,
3367: it is also found in B05a that this high-concentration model somewhat
3368: overestimates the inner slope at $r\simlt 40$ kpc$/h$ (see Figure
3369: \ref{fig:mass}).
3370: This model yields a central surface mass density of
3371: $\kappa_{\rm c}(z_s=1)\approx 0.810$, which is
3372: slightly higher than, but still consistent with, the prediction based on
3373: B05b adopted in the present work, $\kappa_{\rm c}(z_s=1)=0.781\pm 0.1$,
3374: within the $1\sigma$ statistical uncertainty.
3375: %%%%%
3376: The mass profile of A1689 has also been examined by Limousin et
3377: al. (2007) using independent weak lensing shape measurements from
3378: CFHT12K data.
3379: %From this Limousin et al. (2007) obtained the best-fit
3380: %parameters of $c_{\rm vir}\approx 10$ and $M_{\rm vir}\approx 1.5\times
3381: %10^{15}M_{\odot}$,
3382: Their best-fit NFW model, however, underpredicts the observed Einstein
3383: radius (\S \ref{subsubsec:comparison}), and accordingly yields a much
3384: lower central value of
3385: $\kappa_{\rm c}(z_s=1)\approx 0.527$, which we take as the lower limit
3386: on $\kappa_{\rm c}(z_s=1)$.
3387: %%%
3388: Allowing for a conservative uncertainty of
3389: $\kappa_{\rm c}(z_s=1)=0.78^{+0.1}_{-0.25}$,
3390: corresponding to $\kappa_{\rm c}(z_{s,D}=0.68) = 0.70^{+0.09}_{-0.22}$
3391: for the effective depth of our red sample,
3392: we find a $\pm 10\%$ fractional systematic uncertainty in the
3393: halo concentration, $c_{\rm vir}$.
3394:
3395: \subsubsection{Inner Strong Lensing Information}
3396: \label{subsubsec:inner}
3397:
3398:
3399: %% 10%
3400:
3401: The inner strong lensing information
3402: %plays a crucial role in the determination of the halo
3403: %concentration parameter
3404: provides strong constraints on the halo concentration parameter
3405: as demonstrated in Figure \ref{fig:joint}.
3406: In the present work, the ACS-derived inner mass profile of B05b
3407: is specifically used to determine the NFW halo parameters
3408: of A1689 in the strongly lensed region, $r=[10,130]$ ${\rm kpc}/h$.
3409: It is practically difficult to assess potential systematic errors
3410: introduced in strong lensing modeling because of the complex, non-linear
3411: error propagation (B05b).
3412: Instead, here we simply estimate the level of uncertainty
3413: %%due to the inner strong lensing model,
3414: by a comparison with
3415: the result obtained with the model-independent constraint on the
3416: Einstein radius $\theta_{\rm E}$ (see \S \ref{subsubsec:ein})
3417: %$\theta_{\rm E}=45\arcsec$ for $z_s=1$,
3418: based on multiply lensed images identified
3419: in the ACS observations of B05b.
3420: Both the strong lensing analysis of B05b and Limousin et al. (2007)
3421: yield a consistent value for the projected mass interior to
3422: $45\arcsec$ of $M_{2{\rm D}}(45\arcsec)\approx 2\times 10^{14} M_\odot$,
3423: or equivalently, $\bar\kappa(45\arcsec) = 1$ for a
3424: source redshift of $z_s=1$: i.e., $\theta_{\rm E}=45\arcsec$ at
3425: $z_s=1$.
3426: In contrast to the fit to the inner mass profile, this Einstein-radius
3427: information provides an integrated constraint on the inner mass profile
3428: interior to $45\arcsec$, or $r\approx 100{\rm kpc}/h$ in projected
3429: radius.
3430: %%%
3431: We find that a joint fit of the Subaru $\kappa$ map and the
3432: Einstein-radius constraint yields a slightly higher concentration of
3433: $c_{\rm vir}=14.0^{+2.5}_{-2.1}$ (see Table \ref{tab:nfwfit}),
3434: corresponding to a fractional increase of about $10\%$.
3435: This tendency is also found for the
3436: results with the shear-based 1D mass reconstruction from the
3437: $\zeta_{\rm c}$-statistic measurements (see Table \ref{tab:nfwfit}).
3438: %in which case a fractional increase of $20\%$ is found. We thus adopt a
3439: %conservative systematic error of $20\%$ in $c_{\rm vir}$ associated with
3440: %the strong lensing model.
3441: %%We note again this does not simply reflect the actual uncertainty
3442: %%in the strong lensing mass model, but is likely caused by
3443: %% error arising from the way in which the model is specified.
3444: %% error arising from the simplification
3445:
3446: \subsubsection{Entropic Prior}
3447:
3448: % 10\%
3449:
3450: A particular choice of the regularization could be a
3451: potential source of the systematic errors in the cluster mass
3452: reconstruction. In our mass reconstructions the model parameter
3453: $m$ of the entropy prior is
3454: fixed, but the Bayesian value of the
3455: regularization parameter $\alpha$ that satisfies equation
3456: (\ref{eq:alpha}) is obtained for a given value of $m$.
3457: When we vary the value of $m$ over a
3458: relevant range of the cluster lensing signal, $m=[0.1,0.9]$ ($m=0.5\pm
3459: 0.4$), we find the error distribution in the resulting value of $c_{\rm
3460: vir}$ is almost random with a small spread of $\pm 10\%$.
3461:
3462: Furthermore, a particular choice of the entropy function may lead to
3463: some systematic bias in the determination of cluster mass profiles.
3464: To check this possibility,
3465: %here we simply compare the present results
3466: %obtained with the entropy regularization and those from other standard
3467: %reconstructions (\S \ref{sec:massprofile}).
3468: here we simply compare
3469: the present results from the MEM+ reconstruction
3470: with earlier 1D maximum-likelihood results of B05a,
3471: both of which are based on the same distortion and magnification data,
3472: and adopt the ACS-based inner boundary conditions.
3473: After the correction for the systematic bias (\S
3474: \ref{subsubsec:sys_zs}) the best-fit concentration of B05a is $c_{\rm
3475: vir}\approx 12.3$, which is in good agreement with $c_{\rm vir}=12.7\pm
3476: 1$ obtained with the entropy regularization. Thus, it is likely that
3477: the level of systematic uncertainty due to the particular choice of the
3478: entropy regularization is negligibly small as compared to other sources
3479: of the systematic errors.
3480:
3481: Our use of the entropy prior in the low S/N regime might potentially induce
3482: some slight bias in the
3483: regularized maximum likelihood solution, and/or some slight
3484: non-Gaussianity in the error distribution, underestimating the actual
3485: error bars for the mass reconstruction and the NFW halo
3486: parameters. However, our results show good consistency between the
3487: entropy-regularized reconstructions and other standard reconstructions
3488: within the statistical uncertainties. Thus, it seems this bias is not
3489: significant for this work.
3490:
3491:
3492: \subsubsection{Shear Calibration Error}
3493: \label{subsubsec:calib}
3494:
3495: A shear calibration error is one of the systematic errors that could
3496: bias the weak lensing shape measurements (Haymans et al. 2006; Massey et
3497: al. 2007) and
3498: potentially have some influence on recovered mass profiles. To assess
3499: this possibility we have measured the strength
3500: of the weak lensing signal as a function of magnitude limit for our
3501: background galaxy sample.
3502: We found no particular tendency towards a
3503: loss of the weak lensing signal with apparent magnitude
3504: for red background galaxies, which span over a wide range of sizes.
3505: This is comforting and
3506: consistent with the expectations of the model-independent
3507: KSB+ based technique for which the recovered signal should match reality
3508: within the noise.
3509: We note that at the very weak lensing limit, our distortion
3510: measurements are quite consistent with an independent estimate of the
3511: weak lensing signal by Limousin et al. (2007).
3512:
3513: However we have found that for blue background galaxies there is a
3514: significant loss of the signal at faint magnitudes and this raises the
3515: worrying question of unresolvable HII regions which we know do become
3516: prevalent at faint blue magnitudes, acting effectively as point
3517: sources and hence reducing the weak lensing signal. Such objects are
3518: not included in our analysis, so as not to bias our lensing measurements.
3519: The STEP project, aimed at assessing signal which may be lost in
3520: ground based data, described in Heymans et al (2006) and Massey et
3521: al. (2007), does not allow
3522: for unresolvable sources within galaxies, as it is
3523: inherently assumed that galaxies are continuously resolvable, so that
3524: stretched HST/ACS images of faint galaxies are taken to be perfectly
3525: empirical representations of reality for the purpose of calibrating
3526: galaxies dominated by bright HII regions.
3527:
3528:
3529: Furthermore, we have found a good consistency between the purely
3530: shear-based results (e.g., $\zeta$-statistic based 1D reconstruction)
3531: and the results based on the combined distortion and magnification data
3532: (e.g., MEM+ results, B05a results), implying that any shear calibration
3533: error is not noticeable at the level of our analysis, otherwise we would
3534: see inconsistency with our magnification analysis.
3535:
3536:
3537: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3538:
3539: \section{Summary and Discussion}\label{con}
3540:
3541: In this paper, we have derived a projected 2D mass map of the well
3542: studied lensing cluster A1689 ($z=0.183$) based on an
3543: entropy-regularized maximum-likelihood combination of the lens
3544: magnification and distortion of red background galaxies registered in
3545: deep Subaru images. The combination of distortion and
3546: magnification data breaks the mass sheet degeneracy inherent in all
3547: reconstruction methods based on distortion information alone.
3548: %%%
3549: The method is not restricted to the weak lensing regime but applies to
3550: the whole area outside the tangential critical curve, where
3551: non-linearity between the surface mass-density and the observables
3552: extends to a radius of a few arcminutes.
3553: %%%
3554: The strong lensing information from ACS observations was also readily
3555: incorporated in this maximum likelihood approach (\S
3556: \ref{subsec:acs}).
3557: We also utilized the distortion measurements to locally downweight
3558: the intrinsic clustering noise in the magnification measurements,
3559: which otherwise perturbs the depletion signal (\S
3560: \ref{subsec:declustering}).
3561: %%%%
3562: The resulting 2D map showed that the projected surface density of A1689
3563: is smoothly varying and symmetric, similar to the distribution of
3564: cluster members.
3565: %%%%
3566: The 2D mass map is well fitted by the Navarro-Frenk-White model, with a
3567: continuously steepening profile, but the concentration parameter much
3568: higher than expected for its virial mass ($\sim 2\times
3569: 10^{15}M_{\odot}$), according to the clear predictions of standard
3570: $\Lambda$CDM (Bullock et al. 2001; Neto et al. 2007).
3571: %%%%
3572: For consistency we have compared the best-fitting
3573: NFW parameters obtained from different combinations of datasets,
3574: boundary conditions, and weak lensing techniques (\S
3575: \ref{subsubsec:comparison}).
3576: %%%
3577: We find that all of the reconstructions tested here are consistent with
3578: a virial mass in the range, $M_{\rm vir}=(1.5-2.1)\times
3579: 10^{15}M_{\odot}$, and the combined ACS and Subaru-2D mass
3580: reconstruction yields a tight constraint on the concentration parameter,
3581: $c_{\rm vir}=12.7^{+1.0}_{-0.9}$ ($c_{200}\sim 10$), improving upon the
3582: statistical accuracy of our earlier 1D analysis (B05a).
3583: %%%
3584: Very good agreement is found between the present full 2D reconstruction
3585: and the earlier 1D reconstruction (B05a), both of which are based on the
3586: same distortion and magnification measurements, supporting the
3587: assumption of quasi-circular symmetry in the projected mass distribution.
3588: We have also explored potential sources of systematic error on the
3589: concentration parameter, such as the uncertainties in background
3590: redshift distribution, selection criteria for the red background sample,
3591: and strong lensing modeling. Taking into account all of the systematic
3592: errors as well as the statistical uncertainty, our constraint on
3593: the concentration is $c_{\rm vir}=12.7 \pm 1 ({\rm stat.}) \pm 2.8 ({\rm
3594: systematic})$.
3595: %% where the second set of errors include the systematic uncertainties
3596: %% Spergel et al. 2003
3597: %% where the error includes random and systematic uncertainties
3598: % we assess carefully various
3599: %sources of potential systematic error in the halo concentration
3600: %parameter derived from the lensing observations.
3601: %We further explored an alternative combination of the ACS
3602: %profile with the 1D convergence profile based solely on the distortion
3603: %measurement, and obtained
3604:
3605:
3606: %%% Comparison with Limousin et al. (2007) etc.
3607: For clusters well fitted by an NFW profile, the derived virial mass of
3608: a cluster and the concentration parameter can be used to find the
3609: Einstein radius $\theta_{\rm E}$, through the simple relationship
3610: $1=\bar\kappa_{\rm NFW}(\theta_{\rm E})$ (see Appendix \ref{app:nfw}).
3611: %%%%
3612: For A1689 with $M_{\rm vir}=(2.1\pm 0.2)\times 10^{15}_{\odot}$ and
3613: $c_{\rm vir}=12.7^{+1.0}_{-0.9}$ (only statistical errors quoted),
3614: this yields an Einstein radius of
3615: $45 \pm 6$ arcsec at $z_s=1$, or $53 \pm 7$ arcsec at $z_s=3$,
3616: in very good agreement with the mean estimated radius of $\sim
3617: 50\arcsec$, based on the locations of the multiple images (Table 2 of
3618: B05b).
3619: %%%%
3620: The strong lensing mass model of Limousin et al. (2007), based on the
3621: multiple images identified by B05b, properly reproduces the observed
3622: Einstein radius of $45\arcsec$ at $z_s=1$,
3623: %($\bar\kappa[45\arcsec]=1$ at
3624: %$z_s=1$), or equivalently,
3625: %$M(45\arcsec)\approx 2\times 10^{14} M_\odot$ in terms of the
3626: %projected mass interior to $\theta_{\rm E}=45\arcsec$,
3627: consistent with the strong lensing mass model of B05b (see \S
3628: \ref{subsubsec:inner}).
3629: %%%%
3630: In contrast, an Einstein radius of
3631: $24\pm 11$ arcsec at $z_s=1$ is implied by the NFW fit to this cluster
3632: by Limousin et al. (2007), to independent weak lensing data from CFHT
3633: ($c_{\rm vir}\sim 9.6$, $M_{\rm vir}\sim 1.5\times 10^{15}M_\odot$; see
3634: Table \ref{tab:nfwref}, Figures \ref{fig:joint_gt} and \ref{fig:gt}).
3635: This discrepancy may be attributed to a degree of contamination of the
3636: sample of galaxies used to define the lensed background, which,
3637: as pointed out in B05a, can drag down the weak lensing signal
3638: if accidentally included in the background sample, and preferentially
3639: so at small radius (see Figure \ref{fig:gt}) where the
3640: ratio of cluster members compared with background is much higher,
3641: resulting in a shallow $g_+$ profile and hence a lower concentration
3642: fit.
3643:
3644:
3645: %%% Comparison with Limousin and Halkola results
3646: Taking into account the systematic errors (\S \ref{subsec:sys}),
3647: combined ACS and Subaru constraints allow a shallower (but still steeper
3648: than theoretically expected) mass profile
3649: with $c_{\rm vir}= 9-10$, similar to the values found in Halkola
3650: et al. (2006) and Limousin et al. (2007).
3651: This, however, does not simply mean that the discrepancy
3652: between different lensing studies
3653: %%% in the derived NFW halo parameters
3654: %%% between the models
3655: is fully solved: When the NFW model is normalized to reproduce the
3656: observed Einstein radius ($45\arcsec$ at $z_s=1$), then this
3657: concentration would indicate a large virial mass of
3658: $M_{\rm vir}=(3-3.3)\times 10^{15}M_\odot$ (see Figures \ref{fig:joint}
3659: and \ref{fig:joint_gt}), which however is considerably higher than the
3660: mass estimates derived from the X-ray observations ($M_{\rm vir}\approx
3661: 10^{15}M_\odot$ in Andersson \& Madejski 2004, $M_{\rm vir}\approx
3662: 1.4\times 10^{15}M_\odot$ in Lemze et al. 2008).
3663: The discrepancy between the present results and the results by Halkola
3664: et al. (2006) can be explained by the relative weights in the least
3665: $\chi^2$ fitting,
3666: assigned differently to the ACS- and Subaru-based measurements,
3667: indicating slight deviation of the observed profile from the NFW
3668: predictions (see discussion in \S \ref{subsubsec:comparison}; also see
3669: discussion in Medezinski et al. 2007).
3670: This is seen at
3671: the innermost radii $r\simlt 40{\rm kpc}/h$ (Figure
3672: \ref{fig:mass}), where the ACS data indicate a shallower profile.
3673: Consequently, when one prefers the innermost region to fit the data,
3674: this could lead to a lower concentration ($c_{\rm vir}=9-10$) as favored
3675: by the shallower slope in the innermost region, and to a lower value for
3676: the Einstein radius when the data at outer critical radius are less
3677: weighted. Nonetheless, our best-fitting NFW model provides a good
3678: approximation to our data over the radii we have considered, $r=[10^{-2},
3679: 2]$ ${\rm Mpc}/h$.
3680:
3681:
3682:
3683: %%% Importance of dilution
3684: B05a demonstrate that dilution of the lensing signal is certainly the
3685: cause of the very low concentration fit ($c_{\rm vir}\sim 4.5$, Table
3686: \ref{tab:nfwref}) obtained by Bardeau et al. (2005), due to the
3687: inclusion of relatively blue cluster members in the definition of the
3688: background sample of the same
3689: CFHT weak lensing data as Limousin et al. (2007), and for which the
3690: equivalent Einstein radius is only $\sim 4\arcsec$
3691: (at $z_s=1$, see Table \ref{tab:nfwref}).
3692: %The study by Halkola et al. (2006) comes closer to the observed radius
3693: %with a value of $39^{-3}_{+6}$ arcsec (at $z_s=1$, see Table
3694: %\ref{tab:nfwref}), as shown in
3695: %Figure \ref{fig:joint_gt}, implied by their published best-fit NFW
3696: %parameters for the joint combination of the weak lensing results of B05a
3697: %and a parametric model of the strong lensing region based on the
3698: %multiple images identified in B05b. The difference here seems to be due
3699: %to the relatively higher weighting given to the strong lensing region by
3700: %Halkola et al. (2006) where the slope of the projected mass profile is
3701: %shallower than at large radius (see also Figure \ref{fig:gt}).
3702: %%%% Application of dilution
3703: On the other hand,
3704: the dilution of the lensing signal caused by cluster members can be
3705: used to derive the proportion of galaxies statistically belonging
3706: to the cluster by comparing the undiluted red background distortion
3707: signal with the radial distortion profile of color-magnitude space
3708: occupied by the cluster members, but including inevitable background
3709: galaxies falling in the same space (Medezinski et al. 2007).
3710: This technique allows the light profile of the cluster to be determined in
3711: a way which is independent of the number density fluctuations
3712: in the background population, which otherwise limit the
3713: calculation of the cluster light profiles and luminosity functions
3714: from counts of member galaxies.
3715: The resulting light profile can be compared with the mass profile
3716: to examine the radial behavior of $M/L$ (Medezinski et al. 2007).
3717:
3718:
3719:
3720: %%% Lemze et al.'s X-ray analysis
3721: A recent joint X-ray and lensing analysis of A1689 by Lemze et al.
3722: (2008) also produces very similar concentration,
3723: $c_{\rm vir}=12.2^{+0.9}_{-1}$, and virial mass,
3724: $M_{\rm vir}\sim (1.4\pm 0.4)\times 10^{15}M_{\odot}$,
3725: to that derived here in our analysis (Table \ref{tab:nfwref}).
3726: This is derived from a model-independent approach to the X-ray emission
3727: profile and the projected lensing mass profile, assuming hydrostatic
3728: equilibrium, utilizing the mass profile derived in B05a.
3729: Interestingly, the observed temperature profile falls
3730: short of the predicted temperature profile derived from the joint fit
3731: and this may imply that the gas distribution is clumpy on small scales,
3732: in the form of a higher density cold gas phase (Lemze et al. 2008).
3733: %%%%%%
3734: Moreover,
3735: this anomaly seems to be consistent with conclusions regarding
3736: gas substructure in the recent detailed hydrodynamical simulations
3737: of cluster gas by Kawahara et al. (2007),
3738: implying that other similar detailed lensing-X-ray
3739: studies should also show a similar temperature discrepancy.
3740:
3741:
3742: Great progress continues to be made in the detailed predictions of
3743: $\Lambda$CDM, particularly on cluster scales where gas cooling is not a
3744: worry. Recently the whole Millennium survey (Springel et al. 2005)
3745: has been converted into the
3746: observer's frame following the full geodesics through the volume to
3747: simulate the effect of structure on the light received by an observer
3748: (Hilbert et al. 2007). This work has shown that although in general
3749: clusters form in overdense regions, the material associated with a
3750: given cluster in the form of extended groups and filaments outside the
3751: virial radius of the cluster is of relatively low mass contrast
3752: compared to the projected mass due to the cluster itself, and
3753: therefore lensing based projected masses of clusters are not
3754: overestimated by more that a few percent (Hilbert et al. 2007).
3755:
3756:
3757:
3758:
3759: This simulation has also been used to better define the relationship
3760: between the concentration parameter and the virial mass of halos,
3761: over the full range of mass from galaxies up to the most massive
3762: cluster-sized halos, in the context of standard $\Lambda$CDM (Neto et al
3763: 2007). A clear prediction has emerged that most massive halos generated
3764: in these simulations have the lower concentration ($c_{\rm vir}\sim 5$),
3765: and the cause of this is in part attributed to the generally later
3766: collapse of the more massive halos reflecting the lower mean density of
3767: the universe. For example, at $z_{\rm vir}=0.183$, for the standard
3768: choice of cosmological parameters the criterion for virialization is
3769: $\bar\rho(<r_{\rm vir}) \sim 115 \rho_{\rm crit}(z_{\rm vir})
3770: \sim 277 \bar\rho(z_{\rm vir})$.
3771: One possibility to achieve earlier formation of massive clusters is
3772: to allow deviation from Gaussianity of the primordial density
3773: fluctuations in the early universe (e.g., Grossi et al. 2006; Sadeh,
3774: Rephaeli, \& Silk 2007).
3775:
3776:
3777: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3778: %%% Halo triaxiality in general
3779: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3780: A degree of triaxiality is inevitable for collisionless
3781: gravitationally collapsed structures. Discussion of the likely effect
3782: of triaxiality on the measurements of lensing properties has been
3783: examined analytically (Oguri et al. 2005; Sereno 2007;
3784: Corless \& King 2007) and in numerical investigations (Jing \& Suto
3785: 2002; Hennawi et al. 2007).
3786: A bias in favor of prolate structure pointed to the observer is
3787: unavoidable at some level, as this orientation boosts the projected
3788: surface mass density and hence the lensing signal. This effect
3789: has been evaluated in the context of the CDM model and serves as a
3790: guide to the likely degree of bias which may affect lensing work.
3791: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3792: %%% Theoretical efforts to examine the triaxiality effect
3793: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3794: Hennawi et al. (2007) conducted a detailed study of the properties of
3795: lensing cluster population identified in $\Lambda$CDM cosmological
3796: $N$-body simulations.
3797: The level of bias in terms of the concentration
3798: %The effect of this bias in terms of the concentration
3799: parameter derived from 2D lensing measurements
3800: was explicitly estimated and found to amount to an $\sim 34\%$
3801: increase, which results from a combination of two effects, namely
3802: the orientation bias ($\sim 19\%$) due to halo triaxiality
3803: and
3804: the selection effect towards higher 3D concentrations ($\sim 18\%$).
3805: %that favors projections along the major axis of triaxial CDM halos.
3806: The level of correction for the orientation bias is also derived from
3807: semi-analytical
3808: representation of simulated CDM triaxial halos by Oguri et al. (2005).
3809: The anomalously high concentrations of $c_{\rm vir}\simgt 13$, such as
3810: found for A1689, CL0024+1654 (Kneib et al. 2003), and MS2137-23 (Gavazzi
3811: et al. 2003), appear inconsistent with the distribution of
3812: concentrations found in detailed simulations of Hennawi et al. (2007),
3813: which predict that only $<2\%$ of lensing clusters should have such high
3814: concentrations.
3815: It is also unlikely that the baryonic component in these massive
3816: clusters increases the concentrations over the $\Lambda$CDM prediction
3817: for dark matter halos (see discussion in Hennawi et al. 2007, B05b, and
3818: Broadhurst \& Barkanna 2008).
3819:
3820:
3821:
3822: A chance projection of foreground/background structure along the
3823: line-of-sight can potentially influences projected lensing observations,
3824: boosting the surface mass density locally and hence the
3825: concentration. The ACS strong lensing analysis of B05b revealed the
3826: secondary mass clump in the central region of A1689
3827: associated with a small clump of galaxies. The existence of this
3828: subclump has been suggested by earlier observations, such as
3829: the spectroscopic study of Teague et al. (1990) and Czoske (2004),
3830: and the X-ray study of Andersson \& Madejski (2004).
3831: No obvious substructure is visible in a large spectroscopic sample of
3832: 525 cluster members identified in Czoske (2004), in contrast to
3833: earlier work of Teague et al. (1990) based on a smaller sample of 176
3834: identified cluster members.
3835: %who suggested a complex dynamical
3836: %state of A1689 based on 176 identified cluster members.
3837: %%%
3838: %%% no obvious substructure visible in
3839: %%% large sample of czoske etal 2004, something like that.
3840: Recently, this
3841: secondary mass clump has also been directly detected by the weak lensing
3842: flexion analyses by Leonard et al. (2007) and Okura et al.
3843: (2008) based on the ACS and Subaru data, respectively. However, the
3844: detailed ACS strong lensing analyses based on multiply imaged background
3845: galaxies showed that the mass contribution of the secondary mass clump
3846: is only a small fraction of the main cluster component (Figures 21 and
3847: 22 of B05b),
3848: implying a lower $M/L$ for the subclump than for the main cluster,
3849: which is tightly constrained by the geometrical positions of sets of
3850: multiply lensed images, or the location of critical curves.
3851:
3852:
3853:
3854: %%%%%%%%% Practical method for examining the triaxiality
3855: In the near future,
3856: the question of the effect of triaxiality on lensing based cluster
3857: mass profiles may be examined empirically. For example, the total
3858: X-ray luminosity or the total lensing based mass of a cluster should
3859: not depend on the orientation with respect to the line of sight,
3860: whereas the concentration parameter and the Einstein radius will be
3861: affected and hence expanded lensing studies could in principle reveal
3862: whether relaxed clusters of fixed mass or fixed X-ray luminosity tend
3863: to have consistent lensing based concentrations, or instead a broader
3864: distribution may be uncovered, with a mean concentration smaller than
3865: derived for A1689, indicating triaxiality produces a significant
3866: bias. Current indications based on several massive clusters favor the
3867: NFW profile, but with consistent concentrations
3868: (Medezinski et al. 2007, in preparation),
3869: similar in value to A1689, underscoring the tension between
3870: detailed lensing based mass profiles and the predictions of standard
3871: $\Lambda$CDM.
3872:
3873:
3874:
3875: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3876: %%%
3877: %%% Acknowledgments
3878: %%%
3879: %%%
3880: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3881:
3882: \acknowledgments
3883:
3884: We thank the anonymous referee for a careful reading of the manuscript
3885: and and for providing invaluable comments.
3886: We are grateful to Masahiro Takada for valuable discussions and
3887: comments.
3888: %We thank Masahiro Takada for valuable discussions and suggestions.
3889: We thank
3890: J.-H. Proty Wu,
3891: Elinor Medezinski,
3892: Guo-Chin Liu,
3893: Ue-Li Pen,
3894: and Sandor Molnar for fruitful discussions.
3895: We thank Nick Kaiser for making the IMCAT package publicly available.
3896: KU thanks Ludovic Van Waerbeke for kindly providing
3897: his code for a maximum likelihood mass reconstruction.
3898: %We thank the Suprime-Cam team for their support during
3899: %the observation.
3900: %KU acknowledges fruitful discussions with xxx.
3901: %KU thanks Patrick Koch for providing helpful comments.
3902: %valuable discussions.
3903: % We thank the annonymous referee for invaluable comments and suggestions
3904: % which improved the paper significantly,
3905: Part of this work is based on
3906: data collected at the Subaru Telescope,
3907: which is operated by the National Astronomical Society of Japan.
3908: %%%
3909: %The work is partially supported by the COE program at Tohoku University.
3910: This work in part supported by
3911: the National Science Council of Taiwan
3912: under the grant NSC95-2112-M-001-074-MY2.
3913:
3914:
3915:
3916: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3917: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3918: %%%
3919: %%% Tables
3920: %%%
3921: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3922: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3923:
3924: %%\input{tab.tex}
3925:
3926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3927: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3928: %%%
3929: %%% Tables
3930: %%%
3931: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3932: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3933:
3934:
3935: %%% Table 1
3936:
3937: \begin{deluxetable*}{lllll}
3938: \tabletypesize{\scriptsize}
3939: \tablecaption{
3940: \label{tab:memparam}
3941: Parameters used in the 2D MEM reconstruction}
3942: \tablewidth{0pt}
3943: \tablehead{
3944: \colhead{$m$ \tablenotemark{a}} &
3945: \colhead{$\alpha$ \tablenotemark{b}} &
3946: \colhead{$N_{\rm pix}$ \tablenotemark{c}}&
3947: \colhead{$N_{\rm data}$ \tablenotemark{d}} &
3948: \colhead{$\kappa_{\rm c}$ \tablenotemark{e}}
3949: }
3950: \startdata
3951: %% ** Submitted Version **
3952: %% $0.5$ & $96.6$ & $31\times 27$ & 1012 & 0.700 \\
3953: $0.5$ & $96.2$ & $31\times 27$ & 1012 & 0.700 \\
3954: \enddata
3955: \tablenotetext{a}{MEM model parameter.}
3956: \tablenotetext{b}{Bayesian value of $\alpha$.}
3957: \tablenotetext{c}{This includes the ACS-constrained central pixel.}
3958: \tablenotetext{d}{Number of usable measurements.}
3959: \tablenotetext{e}{ACS constraint on the central $\kappa$ pixel assuming
3960: $\langle D_{ds}/D_s\rangle=0.693$.}
3961: \end{deluxetable*}
3962:
3963:
3964: %%% Table 2
3965:
3966: \begin{deluxetable*}{lllrrrrrr}
3967: \tabletypesize{\scriptsize}
3968: \tablecaption{
3969: \label{tab:memmethod}
3970: 2D MEM reconstructions
3971: with different datasets and boundary conditions}
3972: \tablewidth{0pt}
3973: \tablehead{
3974: \colhead{Method \tablenotemark} &
3975: \colhead{Dataset \tablenotemark{a}} &
3976: \colhead{Boundary Conditions \tablenotemark{b}} &
3977: \colhead{$N_{\rm data}${c}} &
3978: \colhead{$\alpha$\tablenotemark{d}} &
3979: %% ** Submitted Version **
3980: %%\colhead{$F_{\rm min}$ \tablenotemark{e}} &
3981: %% ** Version 2 (revised) **
3982: \colhead{NDF \tablenotemark{e}} &
3983: \colhead{$\chi^2_{\rm min}$ \tablenotemark{f}} &
3984: \colhead{RMS \tablenotemark{g}} &
3985: \colhead{S/N \tablenotemark{h}}
3986: }
3987: \startdata
3988: %% ** Submitted **
3989: %2D MEM-S & shear & MEM + ACS & 710 & 88.7 & 322 & 448 &
3990: %0.082 & 14.7\\
3991: %2D MEM & shear + magbias & MEM & 1012 & 100.1 & 625 & 1039 &
3992: % 0.075 & 18.3 \\
3993: %2D MEM+ & shear + magbias & MEM + ACS & 1012 & 96.6 & 651 & 1089 &
3994: % 0.076 & 19.4\\
3995: 2D MEM-S & shear & MEM + ACS & 710 & 91.2 & 515 & 448 & 0.081
3996: & 14.4\\
3997: 2D MEM & shear + magbias & MEM & 1012 & 97.3 & 785 & 1014 & 0.076
3998: & 18.2 \\
3999: 2D MEM+ & shear + magbias & MEM + ACS & 1012 & 96.2 & 784 & 1065 & 0.077
4000: & 19.4\\
4001: \enddata
4002: \tablecomments{
4003: The following three sets of combinations of datasets and
4004: boundary conditions are considered:
4005: (i) 2D MEM+ method using shear and magnification
4006: data with the ACS constraint on the central pixel,
4007: (ii) 2D MEM method using shear
4008: and magnification data without the central ACS constraint,
4009: and (iii) 2D MEM-S method using shear data with the central ACS constraint.
4010: }
4011: \tablenotetext{a}{Dataset used for weak lensing analysis.}
4012: \tablenotetext{b}{With or without the ACS constraint on
4013: the central pixel in the strong lensing region.}
4014: \tablenotetext{c}{Number of usable measurements.}
4015: \tablenotetext{d}{Bayesian value of $\alpha$ ($m=0.5$).}
4016: %\tablenotetext{e}{Minimum functional value of $F$.}
4017: \tablenotetext{e}{Classical number of degrees of freedom, ${\rm NDF} \equiv N_{\rm data}-N_{\rm good}$ (see equation [\ref{eq:alpha}]).}
4018: \tablenotetext{f}{Minimum functional value of $\chi^2$.}
4019: \tablenotetext{g}{Average rms error of $\kappa$ in the observed region
4020: $(30'\times 24')$.}
4021: \tablenotetext{h}{Detection significance of the convergence signal
4022: defined by equation (\ref{eq:sn}).}
4023: \end{deluxetable*}
4024:
4025:
4026: %%% Table 3
4027:
4028: \begin{deluxetable*}{lllrr}
4029: \tabletypesize{\scriptsize}
4030: \tablecaption{
4031: \label{tab:mrec}
4032: Summary of the Methods for Mass Profile Reconstructions
4033: }
4034: \tablewidth{0pt}
4035: \tablehead{
4036: \colhead{Method \tablenotemark{a}} &
4037: \colhead{Dataset \tablenotemark{b}} &
4038: \colhead{Boundary Conditions \tablenotemark{c}} &
4039: \colhead{$(\theta_{\rm min},\theta_{\rm max})$ \tablenotemark{d}} &
4040: \colhead{$N_{\rm bin}$ \tablenotemark{e}}
4041: }
4042: \startdata
4043: $\zeta_{\rm c}$-statistic & tangential shear & $\bar{\kappa}_b=4\times
4044: 10^{-3}$\tablenotemark{\dag} & ($1',16'$) & 9 \\%& 6.6\\
4045: 2D MEM-S & shear & MEM + ACS & ($1',18'$) & 10 \\
4046: 2D MEM & shear + magbias & MEM & ($1',18'$) & 10 \\
4047: 2D MEM+ & shear + magbias & MEM + ACS & ($1',18'$) & 10 \\
4048: Subaru 1D\tablenotemark{*} & tangential shear + magbias & ACS & ($1',18'$) & 10 \\% & 6.9\\
4049: \enddata
4050: \tablenotetext{a}{Weak lensing mass reconstruction method. All methods
4051: apply to the non-linear but subcritical regime.}
4052: \tablenotetext{b}{Dataset used for weak lensing analysis.}
4053: \tablenotetext{c}{With or without the ACS constraint on the central
4054: pixel in the strong lensing region.}
4055: \tablenotetext{d}{Lower and upper radial limits.}
4056: \tablenotetext{e}{Number of radial bins in the range of $(\theta_{\rm
4057: min},\theta_{\rm max})$.}
4058: \tablenotetext{\dag}{This employs an outer boundary condition
4059: %that the
4060: on the
4061: mean convergence $\bar\kappa_b$ within
4062: $16'<\theta<19'$. The mean background level $\bar\kappa_b$ is calculated
4063: to be $\bar\kappa_b=4\times 10^{-3}$ using the ACS+Subaru-1D
4064: best-fit NFW model by B05a.
4065: }
4066: \tablenotetext{*}{1D maximum likelihood analysis by B05a
4067: based on the joint measurements of weak lensing distortion and depletion profiles.}
4068: \end{deluxetable*}
4069:
4070: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4071: %% Table 4
4072: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4073:
4074: \begin{deluxetable*}{llllllll}
4075: \tabletypesize{\scriptsize}
4076: \tablecaption{
4077: \label{tab:nfwfit}
4078: Summary of the best-fitting NFW parameters
4079: for Subaru weak lensing observations
4080: }
4081: \tablewidth{0pt}
4082: \tablehead{
4083: \colhead{Method \tablenotemark{a}} &
4084: \colhead{Data} &
4085: \colhead{ACS \tablenotemark{b}} &
4086: \colhead{ER \tablenotemark{c}} &
4087: \colhead{$M_{\rm vir}$ \tablenotemark{d}} &
4088: \colhead{$c_{\rm vir}$\tablenotemark{e}} &
4089: \colhead{ $\chi^2_{\rm min}/{\rm dof}$\tablenotemark{f}} &
4090: \colhead{ $\theta_{\rm E}$\tablenotemark{g}}
4091: }
4092: \startdata
4093: %%%
4094: tangential shear & 1D $g_+$ &--- &--- & $1.51^{+0.27}_{-0.24}$ &
4095: $20.0^{+8.8}_{-5.3}$ & 5.0/8 (8) & 50.9 \\
4096: tangential shear & 1D $g_+$ &--- & yes & $1.59^{+0.24}_{-0.22}$ &
4097: $15.7^{+3.4}_{-2.5}$ & 9.1/9 (9) & 44.5 \\
4098: $\zeta_{\rm c}$-statistic & 1D $\kappa$ & --- & --- & $1.48 \pm 0.27 $ & $27.3^{+2.7}_{-19.3}$ & 5.2/7
4099: (7) & 59.1 \\
4100: $\zeta_{\rm c}$-statistic & 1D $\kappa$ &--- & yes & $1.51^{+0.25}_{-0.22}$ & $16.5^{+4.0}_{-3.1}$
4101: & 11.4/8 (8) & 44.8 \\
4102: $\zeta_{\rm c}$-statistic & 1D $\kappa$ &yes & --- & $1.91^{+0.24}_{-0.20}$ & $13.7^{+1.5}_{-1.3}$
4103: & 12.3/19 (19) & 45.3 \\
4104: 2D MEM-S & 2D $\kappa$ & --- & --- & $1.48^{+0.20}_{-0.18} $ & $14.1^{+10.3}_{-4.8}$ & 361/834 (406)
4105: & 38.2 \\
4106: 2D MEM-S & 2D $\kappa$ & yes & --- &$1.75^{+0.17}_{-0.16} $ & $14.6^{+1.3}_{-1.1}$ & 369/846 (418)
4107: & 45.1 \\
4108: 2D MEM & 2D $\kappa$ & --- &--- & $1.60^{+0.21}_{-0.17} $ &
4109: $14.9^{+11.9}_{-5.2}$ & 287/834 (478) & 42.7 \\
4110: 2D MEM & 2D $\kappa$ & yes & --- & $1.81^{+0.21}_{-0.14} $ &
4111: $14.3^{+1.2}_{-1.1}$ & 294/846 (490) & 45.1 \\
4112: 2D MEM+ & 2D $\kappa$ & --- & --- & $1.76^{+0.20}_{-0.17} $ & $15.5^{+7.4}_{-4.2}$
4113: & 323/836 (423) & 51.0 \\
4114: %% new:: aho-kprof.txt, aho-kprof.cov
4115: 2D MEM+ & 2D $\kappa$ & --- & yes & $1.93^{+0.22}_{-0.19} $ & $14.0^{+2.5}_{-2.1}$
4116: & 387/837 (424) & 46.4 \\
4117: 2D MEM+ & 2D $\kappa$ & yes & --- & $2.10 \pm 0.17 $ & $12.7^{+1.0}_{-0.9}$
4118: & 327/848 (435) & 45.3 \\
4119: Subaru 1D\tablenotemark{*} & 1D $\kappa$ & --- & --- &
4120: $1.69^{+0.30}_{-0.28}$ & $\le 30$ & $5.36/8$ (6) & 66.9 \\
4121: Subaru 1D\tablenotemark{*} & 1D $\kappa$ &yes & --- & $1.93\pm 0.20$ &
4122: $13.7^{+1.4}_{-1.1}$ & $13.3/20$ (18) & 45.4 \\
4123: \enddata
4124: %%%%
4125: \tablecomments{
4126: A flat prior of $c_{\rm vir}\le 30$ is adopted in the model fitting.
4127: }
4128: \tablenotetext{a}{Mass reconstruction method (see Table
4129: \ref{tab:mrec}).}
4130: \tablenotetext{b}{Fitting with or without the ACS-derived inner mass
4131: profile (\S \ref{subsubsec:acs}).}
4132: \tablenotetext{c}{Fitting with or without the strong lensing
4133: Einstein-radius (ER) constraint, $\theta_{\rm E}=45\arcsec$ at $z_s=1$
4134: (\S \ref{subsubsec:ein}).}
4135: \tablenotetext{d}{Virial mass and $1\sigma$ error in units of $10^{15}M_{\odot}$.}
4136: \tablenotetext{e}{Concentration parameter, $c_{\rm vir}=r_{\rm
4137: vir}/r_s$, and $1\sigma$ error.}
4138: \tablenotetext{f}{Values in parentheses refer to effective degrees of
4139: freedom excluding upper limit bins with $\kappa<0$.}
4140: \tablenotetext{g}{Einstein radius in units of arcsec for a fiducial
4141: source at $z_s=1$, defined as $1=\bar\kappa(\theta_{\rm E})$.}
4142: \tablenotetext{*}{Taken from B05a. A fiducial source redshift of $z_s=1$
4143: is assumed in B05a.}
4144: \end{deluxetable*}
4145:
4146:
4147: %% Table 5 :: Mvir-Cvir constraints from the literature
4148:
4149:
4150: \begin{deluxetable*}{llllllll}
4151: \tabletypesize{\scriptsize}
4152: \tablecaption{
4153: \label{tab:nfwref}
4154: Comparison between best-fitting NFW parameters
4155: for A1689 from different observations and methods
4156: %by different authors
4157: %taken from the literature
4158: }
4159: \tablewidth{0pt}
4160: \tablehead{
4161: \colhead{Reference} &
4162: \colhead{Method \tablenotemark{a}} &
4163: \colhead{$M_{\rm vir}$ \tablenotemark{b}} &
4164: \colhead{$c_{\rm vir}$ \tablenotemark{c}} &
4165: \colhead{$M_{200}$ \tablenotemark{d}} &
4166: \colhead{$c_{200}$\tablenotemark{e}} &
4167: \colhead{ $\theta_{\rm E}$\tablenotemark{f}} &
4168: \colhead{ Remarks}
4169: }
4170: \startdata
4171: King, Clowe, \& Schneider 2002 & WL & $1.0$ & $6.1$ & $0.84$ & $4.8$ &
4172: 11 & ESO/MPG\\%% (fit to the shear pattern)\\
4173: %%% Clowe 2003
4174: Clowe 2003 & WL & $1.3$ & $9.9$ & $1.1 $ & $ 7.9$ &
4175: 22 & ESO/MPG\\%% (shear profile)\\
4176: %%% Bardeau 2005
4177: Bardeau et al. 2005 & WL & $ 1.72^{+0.8}_{-0.6} $ & $ 4.5^{+0.6}_{-0.4} $
4178: & $ 1.41^{+0.63}_{-0.47}$ & $3.5^{+0.5}_{-0.3}$
4179: & 3.5 & CFH12K\\
4180: %%% ACS strong lensing
4181: %%% Mvir=2.62e15Msun/h
4182: B05b & SL & $3.7$ & $8.2^{+2.1}_{-1.8}$ & $3.2$ & $6.5^{+1.9}_{-1.6}$ &
4183: 44 & ACS\\
4184: %%% BTU05
4185: B05a & SL+WL & $1.93\pm 0.20$ & $13.7^{+1.4}_{-1.1}$ & $1.72\pm
4186: 0.19 $ & $ 10.9^{+1.1}_{-0.9}$ &
4187: 45 & ACS + Subaru ($\kappa$ profile)\\
4188: %%% Halkola+ 2006 (SL)::
4189: %%% r200 = 2.82Mpc/h70 +/- 0.11
4190: %%% C200 = 6.0 \pm 0.5
4191: %%% M200 = 2.13954868143346 [1.89881528994971 -2.39981469631041]1e15Msun/h
4192: %%% M200 = 3.05e15Msun/h70
4193: %%% Mvir = 2.4902 [2.1973--2.8116] = 3.55e15Msun/h70
4194: %%% Cvir = 7.592 [6.9746 -- 8.2089]
4195: Halkola et al. 2006 & SL
4196: & $3.55\pm 0.4$ & $7.6 \pm 0.6$
4197: & $3.05\pm 0.3$ & $6.0 \pm 0.5$ & 37 & ACS \\
4198: %% Halkola+ 2006 (SL + WL)::
4199: %% r200=2.55Mpc/h70 \pm 0.07
4200: %% C200=7.6 \pm 0.5
4201: %% M200(1e15Msun/h) = 1.5799 [1.4552 -- 1.71584]
4202: %% Mvir(1e15Msun/h) = 1.8093 [1.6599 -- 1.9736]
4203: %% Cvir = 9.5647 [8.9486 -- 10.181]
4204: %% rE: 38.889220 [35.474812 -- 44.9133]-3.414408 +6.024080
4205: Halkola et al. 2006 & SL+WL
4206: & $2.58 \pm 0.2$ & $9.6 \pm 0.6 $
4207: & $2.25 \pm 0.2$ & $7.6 \pm 0.5 $ & 39
4208: & ACS + Subaru ($g_+$ profile) \\
4209: %% Bardeau+ 2007
4210: %% M200 = 1.379e15Msun/h \pm 0.238 [1.141--1.617]
4211: %% C200 = 4.28 \pm 0.82
4212: %% Mvir = 1.65e15 Msun/h [1.3449 -- 1.9760]
4213: %% Cvir = 5.466 [4.449-6.480]
4214: Bardeau et al. 2007 & WL
4215: & $2.35\pm 0.4$ & $5.5 \pm 1.0$
4216: & $1.97\pm 0.3$ & $4.28 \pm 0.8$ &
4217: 12 & CFH12K \\%%(fit to the shear pattern) \\
4218: %%% Limousin et al. 2007
4219: %%% M200 = 0.924 e15Msun/h [0.784--1.064]
4220: %%% C200 = 7.6\pm 1.6 [6.0 -- 9.2]
4221: %%% Mvir = 1.51^{+0.25}_{-0.24}1e15Msun/h70
4222: %%% Mvir=1.057e15Msun/h [0.88768--1.2384E]
4223: %%% Cvir=9.6\pm 2.0 [7.6--11.6]]
4224: %%% rE = 24.087869 [12.741739 -- 35.662879]-11.34+11.574931
4225: Limousin et al. 2007 & WL
4226: & $1.51^{+0.3}_{-0.2}$ & $9.6 \pm 2.0$
4227: & $1.32\pm 0.2$ & $7.6 \pm 1.6$
4228: & 24 & CFH12K \\%%(shear profile) \\
4229: %%% Lemze et al. 2007 (X-ray)
4230: %%% Mvir = 1.5661e15Msun/h
4231: %%% Cvir = 12.2 [11.2--13.1]
4232: %%% M200 = 1.3875e15Msun/h
4233: %%% C200 = 9.741 [8.9282--10.554]
4234: Lemze et al. 2008\tablenotemark{*} & SL+WL+X
4235: & $2.23\pm 0.6^\dag$ & $12.2^{+0.9}_{-1}$
4236: & $1.98$ & $9.7 \pm 0.8$
4237: & 45 & ACS + Subaru + Chandra\\
4238: %%% This work
4239: %%% M200=1.086e15Msun/h [1.086-1.384]
4240: %%% C200=10.709 [7.96-15.17]]
4241: This work & WL & $ 1.97 \pm 0.20 $ & $13.4^{+5.4}_{-3.3}$ &
4242: $1.76 \pm 0.20$ & $10.7^{+4.5}_{-2.7}$ &
4243: 45 & Subaru ($\kappa$ map) \\
4244: %%%
4245: %%%
4246: This work & SL+WL & $2.10 \pm 0.17$ & $12.7^{+1.0}_{-0.9}$ &
4247: $ 1.86 \pm 0.16$ & $ 10.1^{+0.8}_{-0.7} $ &
4248: 45 & ACS + Subaru ($\kappa$ map) \\
4249: \enddata
4250: \tablecomments{
4251: A similar table of best-fitting NFW parameters
4252: is found in Comerford \& Natarajan 2007 (Table 1)
4253: which also include the results for other clusters as well as A1689.
4254: }
4255: \tablenotetext{a}{Analysis method.}
4256: \tablenotetext{b}{Virial mass $M_{\rm vir}$ and $1\sigma$ error in units
4257: of $10^{15}M_{\odot}$.}
4258: \tablenotetext{c}{Virial concentration $c_{\rm vir}=r_{\rm vir}/r_s$ and
4259: $1\sigma$ error.}
4260: \tablenotetext{d}{$M_{200}$ and $1\sigma$ error in units of
4261: $10^{15}M_{\odot}$.}
4262: \tablenotetext{e}{Specific concentration $c_{200}=r_{200}/r_s$ and
4263: $1\sigma$ error.}
4264: \tablenotetext{f}{Einstein radius in units of arcsec for a fiducial
4265: source at $z_s=1$, defined as $1=\bar\kappa(\theta_{\rm E})$.}
4266: \tablenotetext{*}{Based on the Chandra X-ray data and the
4267: projected mass profile from the joint ACS and Subaru-1D analysis by
4268: B05a. Hydrostatic equilibrium assumed.}
4269: \tablenotetext{\dag}{Note that the outermost radius point of Lemze et
4270: al. (2008) is at $1.5 {\rm Mpc}/h$, which is smaller than the virial
4271: radius of A1689, $r_{\rm vir}\approx 2 {\rm Mpc}/h$.
4272: As compared to the NFW-based prediction in the table,
4273: their model-independent
4274: reconstruction of the total mass density $\rho(r)$
4275: %%with an extrapolation index of $-3$
4276: yields a virial mass of
4277: $M_{\rm vir} = (1.4\pm 0.4)\times 10^{15} M_\odot$,
4278: assuming an extrapolation index of -3.}
4279: \end{deluxetable*}
4280:
4281:
4282: %% Table 6 :: systematic uncertainties
4283:
4284:
4285: \begin{deluxetable*}{llr}
4286: \tabletypesize{\scriptsize}
4287: \tablecaption{
4288: \label{tab:sys}
4289: Sources of systematic error and their effects on the determination of
4290: the halo concentration parameter $c_{\rm vir}$.
4291: }
4292: \tablewidth{0pt}
4293: \tablehead{
4294: \colhead{Source of error} &
4295: \colhead{Uncertainty range} &
4296: \colhead{Fractional error in $c_{\rm vir}$}
4297: }
4298: \startdata
4299: Lower color limit for the red sample & $[0.1,0.5]$ & 4\% \\
4300: Source redshift\tablenotemark{*} & $z_{s,D}=[0.7,1]$ & 10\% \\
4301: Mask area correction & $\nu_{\rm mask}=3\pm 1$ & 7\% \\
4302: Clustering noise rejection & $\nu_{\rm clust}=4\pm 1$ & 1.5\% \\
4303: Slope of unlensed number counts\tablenotemark{\dag} & $s=0.22\pm 0.03$ & 3.5\% \\
4304: Inner boundary condition & $\kappa_{\rm c}(z_s=1)=0.78^{+0.1}_{-0.25}$ &
4305: 10\% \\
4306: Strong lensing modeling\tablenotemark{\dag\dag} & B05b, ER & 10\% \\
4307: Entropy prior & $m=0.5\pm 0.4$ & 10\% \\
4308: \enddata
4309: \tablecomments{
4310: The systematic errors are presented in fraction of $c_{\rm vir}=12.7$
4311: derived from a joint fit to the ACS-based inner $\kappa$ profile of
4312: B05b and the Subaru-based $\kappa$ map reconstructed with the MEM+ method.
4313: }
4314: \tablenotetext{*}{Effective source redshift $z_{s,D}$ equivalent to the
4315: mean distance ratio $\langle D_{ds}/D_s\rangle$ defined by equation
4316: (\ref{eq:zD}).}
4317: \tablenotetext{\dag}{$s=d\log N(m)/dm$.}
4318: \tablenotetext{\dag\dag}{Einstein-radius constraint ($\theta_{\rm
4319: E}=45\arcsec$ for $z_s=1$) used instead of the ACS inner mass profile
4320: of B05b.}
4321: \end{deluxetable*}
4322:
4323:
4324: \clearpage
4325:
4326:
4327: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4328: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4329: %%%
4330: %%% Figures
4331: %%%
4332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4334:
4335:
4336: %% Use the figure environment and \plotone or \plottwo to include
4337: %% figures and captions in your electronic submission.
4338: %% To embed the sample graphics in
4339: %% the file, uncomment the \plotone, \plottwo, and
4340: %% \includegraphics commands
4341: %%
4342: %% If you need a layout that cannot be achieved with \plotone or
4343: %% \plottwo, you can invoke the graphicx package directly with the
4344: %% \includegraphics command or use \plotfiddle. For more information,
4345: %% please see the tutorial on "Using Electronic Art with AASTeX" in the
4346: %% documentation section at the AASTeX Web site,
4347: %% http://www.journals.uchicago.edu/AAS/AASTeX.
4348: %%
4349: %% The examples below also include sample markup for submission of
4350: %% supplemental electronic materials. As always, be sure to check
4351: %% the instructions to authors for the journal you are submitting to
4352: %% for specific submissions guidelines as they vary from
4353: %% journal to journal.
4354:
4355:
4356: %% This example uses \plotone to include an EPS file scaled to
4357: %% 80% of its natural size with \epsscale. Its caption
4358: %% has been written to indicate that additional figure parts will be
4359: %% available in the electronic journal.
4360:
4361:
4362:
4363: %%%%%%%%%%%%% Figure 1
4364:
4365: \begin{figure*}[!htb]
4366: \begin{center}
4367: \includegraphics[width=60mm, angle=270]{f1.ps}
4368: \end{center}
4369: \caption{
4370: The quadrupole PSF anisotropy field as measured from
4371: stellar ellipticities before and after the PSF anisotropy correction.
4372: The left panel shows the raw ellipticity field of stellar objects,
4373: and the right panel shows the residual ellipticity field after
4374: the PSF anisotropy correction.
4375: The orientation of the sticks indicates the position angle of
4376: the major axis of stellar ellipticity, whereas the length is
4377: proportional to the modulus of stellar ellipticity. A stick with the
4378: length of $5\%$ ellipticity is indicated in the top right of the right
4379: panel.
4380: }
4381: \label{fig:anisopsf1}
4382: \end{figure*}
4383:
4384:
4385:
4386:
4387:
4388: %%%%%%%%%%%%% Figure 2
4389:
4390: \begin{figure*}[!htb]
4391: \begin{center}
4392: \includegraphics[width=60mm, angle=270]{f2.ps}
4393: \end{center}
4394: \caption{
4395: Stellar ellipticity distributions before and after the PSF anisotropy
4396: correction. The left panel shows the raw ellipticity components
4397: $(e_1^*,e_2^*)$ of stellar objects, and the right panel shows
4398: the residual ellipticity components $(\delta e_1^*, \delta e_2^*)$
4399: after the PSF anisotropy correction.
4400: }
4401: \label{fig:anisopsf2}
4402: \end{figure*}
4403:
4404:
4405:
4406: %%%%%%%%%%%%% Figure 3
4407:
4408: \begin{figure*}[!htb]
4409: \begin{center}
4410: \includegraphics[width=80mm]{f3.ps}
4411: \end{center}
4412: \caption{
4413: Averaged
4414: shear correction factor, $\langle P_g^{\rm s} \rangle$,
4415: as a function of object size, $r_g$.
4416: The horizontal error bar represents the size of the bin ($\Delta r_g=1$
4417: pixel),
4418: and the vertical error bar represents the {\it rms} scatter in
4419: the smoothed scalar correction factor, $\langle P_g\rangle$.
4420: }
4421: \label{fig:Pg}
4422: \end{figure*}
4423:
4424:
4425:
4426: %%%%%%%%%%%%% Figure 4
4427:
4428: \begin{figure*}[!htb]
4429: \begin{center}
4430: \includegraphics[width=70mm,angle=270]{f4.ps}
4431: \end{center}
4432: \caption{
4433: \label{fig:dilution}
4434: Top panel:
4435: mean distortion strength averaged over a wide radial range of $1\arcmin
4436: < \theta < 18\arcmin$,
4437: done separately for the blue and red galaxy samples.
4438: No area weighting is used here to enhance the effect of dilution in the
4439: central cluster region.
4440: %in order to
4441: %establish the boundaries of the color distribution free of cluster members.
4442: Shown are the measurements of the tangential component ($g_+$)
4443: with open squares, and those of the $45{\rm deg}$-rotated component
4444: ($g_\times$).
4445: %%%
4446: On the right ({\it red}), the square symbols show
4447: that $g_+$ drops rapidly when the bluer limit of the entire red sample
4448: is decreased below a color indicated by the vertical dashed line
4449: which lies $+0.22$ mag redward of the cluster sequence.
4450: %%%%
4451: This sharp decline marks the point at which the red sample encroaches on
4452: the E/S0 sequence of the cluster.
4453: For galaxies with colors bluer than the cluster seqeuence, cluster
4454: members are present along with background galaxies. Consequently, the
4455: mean lensing strengh of the blue sample, as shown on the left ({\it
4456: blue}), is systematically lower than that of the red sample.
4457: Bottom panel: the respective numbers of galaxies as a function of color
4458: limit, contained in the range
4459: $1\arcmin <\theta < 18\arcmin$ in the red ({\it right}) and the blue
4460: ({\it left}) samples.
4461: }
4462: \end{figure*}
4463:
4464:
4465:
4466:
4467: %%%%%%%%%%%%% Figure 5
4468:
4469:
4470: \begin{figure*}[!htb]
4471: \begin{center}
4472: \includegraphics[width=70mm, angle=270]{f5.ps}
4473: \end{center}
4474: \caption{Gravitational reduced-shear field in A1689 obtained from shape
4475: distortions of the red background galaxies, smoothed with a
4476: Gaussian with ${\rm FWHM} = 2'$ for visualization purposes.
4477: A stick with the length of $20\%$ ellipticity is indicated in the top
4478: right corner.
4479: The shaded circle indicates the FWHM of the Gaussian.
4480: The coordinate origin is at the optical center (see Figure
4481: \ref{fig:image}).
4482: }
4483: \label{fig:shear}
4484: \end{figure*}
4485:
4486:
4487:
4488:
4489: %%%%%%%%%%%%% Figure 6
4490:
4491: \begin{figure*}[!htb]
4492: \begin{center}
4493: \includegraphics[width=60mm, angle=270]{f6.ps}
4494: \end{center}
4495: \caption{
4496: Distribution of the lensing magnification bias $n/n_0$
4497: measured from a red-galaxy sample in the background of A1689,
4498: smoothed with a Gaussian
4499: with ${\rm FWHM} = 2'$ for visualization purposes.
4500: The shaded circle indicates the FWHM of the Gaussian.}
4501: \label{fig:magbias}
4502: \end{figure*}
4503:
4504:
4505:
4506:
4507: %%%%%%%%%%%%% Figure 7
4508:
4509: \begin{figure*}[!htb]
4510: \begin{center}
4511: \begin{tabular}{cc}
4512: \includegraphics[width=65mm,angle=270]{f7a.ps} &
4513: \includegraphics[width=65mm,angle=270]{f7b.ps}\\
4514: \end{tabular}
4515: \end{center}
4516: \caption{
4517: Left panel:
4518: The projected mass distribution $\kappa$ of A1689
4519: on a grid of $31\times 27$ pixels
4520: reconstructed
4521: from an entropy-regularized, maximum-likelihood combination of
4522: %from a maximum entropy combination of
4523: Subaru shape distortion and
4524: number-count depletion
4525: data of red background galaxies.
4526: ACS strong lensing constraints are used to determine the
4527: convergence value
4528: at the central pixel
4529: that falls in the strong
4530: lensing region.
4531: %%%%
4532: The pixel width is $1.4$ arcmin.
4533: The observing region is limited to the central $30'\times 24'$,
4534: covering a projected area of $3.9 \times 3.1$ ${\rm Mpc}^2/h^2$,
4535: while the field size for the reconstruction is $43\farcm 4\times 37\farcm 8$.
4536: %%%
4537: Overlayed up on the image is the reconstructed spin-2 gravitational
4538: shear field. A stick with the length of $5\%$ shear is
4539: indicated in the top right corner.
4540: The north is to the top, and the east is to the left.
4541: Right panel:
4542: distribution of the reconstruction errors for the
4543: convergence $\kappa$
4544: calculated using the maximum entropy method.
4545: The uncertainties are given by the square root of the diagonal part
4546: of the covariance matrix.
4547: \label{fig:rawkappa}
4548: }
4549: \end{figure*}
4550:
4551:
4552:
4553:
4554:
4555: %%%%%%%%%%%%% Figure 8
4556:
4557: \begin{figure*}[!htb]
4558: \begin{center}
4559: \includegraphics[width=120mm]{f8_lower.ps}
4560: \end{center}
4561: \caption{
4562: Contours ({\it thick}) of the
4563: dimensionless surface mass density $\kappa$
4564: smoothed with a Gaussian of ${\rm FWHM=1\farcm 4'}$,
4565: superposed on the
4566: $V+i'$ pseudo-color image of A1689.
4567: The image size is $\approx 30'\times 25'$,
4568: covering a projected area of $3.9\times 3.2$
4569: ${\rm Mpc}^2/h^2$ at $z=0.183$.
4570: The lowest contour
4571: and the contour interval are $0.05$.
4572: North is to the top and east is to the left.
4573: }
4574: \label{fig:image}
4575: \end{figure*}
4576:
4577:
4578:
4579:
4580: %%%%%%%%%%%%% Figure 9
4581:
4582: \begin{figure*}[!htb]
4583: \begin{center}
4584: \includegraphics[width=150mm]{f9_lower.ps}
4585: \end{center}
4586: \caption{
4587: Comparison of the reconstructed lensing fields and the
4588: cluster galaxy distributions in A1689.
4589: Top: reconstructed convergence $
4590: \kappa$ ({\it left}) and magnification bias
4591: $\mu^{2.5s-1}$ ({\it right}) fields.
4592: %smoothed with a Gaussian of ${\rm FWHM}=2'$.
4593: Bottom:
4594: %distributions of
4595: observed $i'$-band surface luminosity ({\it left}) and number ({\it right})
4596: density distributions of $(V-i')$-selected
4597: member galaxies with $i'<23$ ABmag.
4598: % in A1689,
4599: %smoothed with a Gaussian of ${\rm FWHM}=2'$.
4600: All images are smoothed with a Gaussian of ${\rm FWHM}=2'$.
4601: The field size is $30'\times 24'$. North is to the top,
4602: east to the left.
4603: %%%
4604: The contours show the
4605: Gaussian smoothed convergence distribution
4606: shown in the top-left panel.
4607: %as the surface luminosity and number density maps.
4608: %smoothed with a Gaussian of ${\rm FWHM}=2'$.
4609: The lowest contour level and the contour interval are $0.05$.
4610: %$\Delta\Sigma_m \approx 2.4\times h 10^{14} M_{\odot} {\rm Mpc}^{-2}$.
4611: %with contours levels
4612: For each panel the color scale is linear, and
4613: ranges from $0\%$--$100\%$ of the peak value.
4614: }
4615: \label{fig:lmap}
4616: \end{figure*}
4617:
4618:
4619:
4620:
4621: %%%%%%%%%%%%% Figure 10
4622:
4623:
4624: \begin{figure*}[!htb]
4625: \begin{center}
4626: \includegraphics[width=100mm, angle=270]{f10.ps}
4627: \end{center}
4628: \caption{
4629: Mass profiles of A1689
4630: obtained by a radial projection of
4631: the 2D $\kappa$ map
4632: (see Figure \ref{fig:rawkappa})
4633: reconstructed from an entropy-regularized
4634: maximum likelihood combination of
4635: Subaru distortion and depletion measurements.
4636: All of the profiles are scaled to a fiducial source redshift of
4637: $z_s=1$.
4638: %%%%
4639: The square and triangle symbols
4640: represent the results with (2D MEM+) and without (2D MEM)
4641: ACS strong lensing constrains on the central mass density,
4642: respectively.
4643: %%
4644: The error bars are highly correlated in the different bins.
4645: %Downwards-pointing arrows
4646: %are used where the lower error bar drops below zero.
4647: %%%
4648: %When the ACS constraints are not incorporated
4649: %in the mapmaking,
4650: Without the central ACS constraint,
4651: central $\kappa$-bins at $\theta\simlt 4'$ are
4652: slightly underestimated ($\sim 10\%$), but the two profiles are
4653: overall in good agreement within the statistical uncertainties.
4654: %%%
4655: The solid curve shows the best-fitting NFW profile
4656: for the 2D $\kappa$ map (2D MEM+) combined with the ACS-derived
4657: inner mass profile.
4658: %the mass profile with the ACS constraints (2D MEM+).
4659: %%%%
4660: For comparison
4661: an NFW model based on the ACS+Subaru 1D analysis (B05a)
4662: %on the 1D ACS data (B05b) and
4663: %combined Subaru distortion/depletion data (B05a)
4664: is shown as a dashed curve.
4665: %The dashed curve shows
4666: %the best-fitting NFW model
4667: %from the combined ACS+Subaru-1D analysis (B05a).
4668: }
4669: \label{fig:kprof_mem}
4670: \end{figure*}
4671:
4672:
4673:
4674:
4675: %%%%%%%%%%%%% Figure 11
4676:
4677: \begin{figure*}[!htb]
4678: \begin{center}
4679: \includegraphics[width=100mm, angle=270]{f11.ps}
4680: \end{center}
4681: \caption{
4682: Comparison of
4683: mass profiles from MEM-reconstructed $\kappa$ maps
4684: based on different combinations of Subaru datasets and boundary conditions.
4685: %%%
4686: All of the profiles are scaled to a fiducial source redshift of $z_s=1$.
4687: The square and triangle symbols represent the results from the
4688: combined distortion and depletion measurements,
4689: with (MEM+) and without (MEM) the
4690: ACS constraint on the mean surface mass density
4691: in the central pixel, respectively.
4692: %%%
4693: The crosses show the results from the distortion data with
4694: the central ACS constraint (MEM-S).
4695: %%%
4696: The model curves are shown for comparison as in
4697: Figure \ref{fig:kprof_mem}.
4698: The mass profile from the distortion data alone ({\it crosses})
4699: shows a slight negative dip of $\kappa\sim -0.01$
4700: at $6'\simlt \theta \simlt 10$ due to spurious boundary effects.
4701: %due to incorrect
4702: %boundary condition in the biased rich cluster field.
4703: %The 1D $\zeta_{\rm c}$-based mass profile
4704: %increases continously
4705: %from the outer boundary towards the center, but lies
4706: %systematically below the MEM+ reconstruction with the
4707: %ACS inner boundary condition.
4708: }
4709: \label{fig:kprof_3mem}
4710: \end{figure*}
4711:
4712:
4713:
4714:
4715:
4716: %%%%%%%%%%%%% Figure 12
4717:
4718:
4719: \begin{figure*}[!htb]
4720: \begin{center}
4721: \includegraphics[width=100mm, angle=270]{f12.ps}
4722: \end{center}
4723: \caption{
4724: %Importance of non-linear corrections in
4725: %the
4726: Shear-based 1D mass reconstruction
4727: utilizing the $\zeta_{\rm c}$-statistic.
4728: As an outer boundary condition,
4729: the mean background density $\bar\kappa_b$
4730: in the range $16'\simlt \theta\simlt 19'$ is set to
4731: $\bar\kappa_b=4\times 10^{-3}$ according to the ACS+Subaru-1D best-fit
4732: NFW model (B05a).
4733: %%%
4734: The square symbols represent the results
4735: with the non-linear corrections. The triangle symbols
4736: show the reconstruction with linear approximation.
4737: The mass profiles are scaled to a fiducial source redshift
4738: of $z_s=1$.
4739: Decorrelated error bars are shown.
4740: Downwards-pointing arrows
4741: are used where the lower error bar drops below zero.
4742: %%%
4743: Without the non-linear corrections, central bins are underestimated
4744: by $\sim 15\%$ at maximum.
4745: %%%%
4746: The solid curve shows
4747: an NFW profile with a high concentration, $c_{\rm vir}=30$,
4748: matching well the overall profile obtained with the non-linear
4749: corrections.
4750: %the best-fitting NFW profile
4751: %with high concentration $c_{\rm vir}=30$
4752: %for the $\kappa$ profile obtained with the non-linear corrections.
4753: %%%%
4754: For comparison
4755: an NFW model
4756: %from on the 1D analysis (B05a)
4757: based on
4758: the combined
4759: ACS and Subaru distortion and depletion profiles (B05a)
4760: is shown as a dashed curve.
4761: }
4762: \label{fig:zeta2kappa}
4763: \end{figure*}
4764:
4765:
4766:
4767:
4768:
4769: %%%%%%%%%%%%% Figure 13
4770:
4771: \begin{figure*}[!htb]
4772: \begin{center}
4773: \includegraphics[width=100mm, angle=270]{f13.ps}
4774: \end{center}
4775: \caption{
4776: Comparison of
4777: model-independent
4778: mass profiles of A1689.
4779: All of the profiles are scaled to a fiducial source redshift of
4780: $z_s=1$.
4781: %%%%
4782: The filled circles represent the results
4783: based on the 2D $\kappa$ map
4784: reconstructed from an entropy-regularized maximum-likelihood combination
4785: of Subaru distortion and depletion data,
4786: with the ACS constraint on the mean surface mass density
4787: in the central pxiel (MEM+).
4788: %using the ACS constraint on the central mass density.
4789: %based on the MEM-reconstructed 2D $\kappa$ map using
4790: %the combined Subaru distortion and depletion data
4791: %and the ACS constraints on the central mass density.
4792: The error bars are correlated.
4793: %(shear)
4794: %(magnification bias)
4795: %measurements.
4796: %%%%
4797: The open triangles
4798: represent the mass profile from
4799: the non-linear $\zeta_{\rm c}$-statistic measurements
4800: based on averaged tangential distortion data.
4801: Decorrelated error bars are shown.
4802: %%%
4803: The filled triangles and circles show the results from the ACS strong
4804: lensing analysis (B05b) and
4805: %The filled circles with error bars
4806: %represent the mass profile from
4807: %the Subaru weak lensing analysis
4808: %with the 1D reconstruction method
4809: from the Subaru 1D weak lensing analysis
4810: based on the combined distortion and depletion profiles (B05a).
4811: %tangential distortion and depletion measurements (B05a).
4812: %Downwards-pointing arrows
4813: %are used where the lower error bar drops below zero.
4814: %%%%
4815: %The solid curve shows
4816: %the best-fitting NFW profile for the combined ACS+Subaru data
4817: %by BTU05 ($M_{\rm vir}=1.93\times 10^{15} M_{\odot}, c_{\rm
4818: % vir}=13.7$).
4819: The solid curve shows
4820: the best-fitting NFW profile for the MEM-reconstructed
4821: 2D $\kappa$ map (MEM+).
4822: The 1D- and 2D-based NFW models
4823: from the respective combined ACS+Subaru data
4824: (B05a) are also shown as solid and dashed curves, respectively.
4825: For comparison
4826: an NFW profile with a low concentration, $c_{\rm vir}=5$,
4827: normalized to the observed Einstein radius ($\theta_{\rm E}=45''$),
4828: is shown as a dotted curve.
4829: The low concentration model ({\it dotted}) predicted for $\Lambda$CDM
4830: clearly overestimates the outer profile constrained by
4831: the Subaru weak lensing observations.
4832: %%%
4833: The mass profiles are all
4834: in remarkable agreement
4835: over the full range of radii up to $\sim 2 {\rm Mpc}/h$.
4836: }
4837: \label{fig:mass}
4838: \end{figure*}
4839:
4840:
4841:
4842:
4843: %%%%%%%%%%%%% Figure 14
4844:
4845: \begin{figure*}[!htb]
4846: \begin{center}
4847: \includegraphics[width=30mm, angle=270]{f14a.ps}\\
4848: \includegraphics[width=30mm, angle=270]{f14b.ps}\\
4849: \includegraphics[width=30mm, angle=270]{f14c.ps}
4850: \end{center}
4851: \caption{
4852: Joint constraints on the NFW model parameters,
4853: $(c_{\rm vir}, M_{\rm vir})$, derived from
4854: gravitational lensing observations of A1689.
4855: %Subaru weak lensing
4856: %observations of A1689.
4857: %%%
4858: %The left panel shows
4859: Left panels show
4860: the $68\%$, $95\%$,
4861: and $99.7\%$ confidence levels
4862: ($\Delta\chi^2=2.3$, $6.17$, and $11.8$)
4863: in the $(c_{\rm vir}, M_{\rm vir})$-plane
4864: for the 2D $\kappa$ map reconstructed from
4865: Subaru weak lensing observations.
4866: %for the 2D $\kappa$ map reconstructed from an entropy-regularized
4867: %maximum likelihood combination of Subaru distortion and depletion data
4868: %(MEM+).
4869: Right panels show the same confidence levels but for the joint
4870: ACS+Subaru-2D NFW fitting, incorporating the inner mass profile
4871: ($10{\rm kpc}/h \simlt r \simlt 180 {\rm kpc}/h$)
4872: constrained by ACS strong-lensing observations by B05b.
4873: The cross in each panel shows
4874: the best-fit set of the NFW model parameters.
4875: %the best-fitting model parameters.
4876: The top, middle, and bottom panels
4877: correspond to the results
4878: based on the 2D MEM+, MEM, and MEM-S
4879: reconstructions,
4880: %from Subaru weak lensing observations,
4881: respectively (see Table \ref{tab:memmethod}).
4882: %%%%
4883: The virial mass $M_{\rm vir}$ is well constrained by the Subaru data
4884: alone, while the Subaru constraint on the concentration
4885: $c_{\rm vir}$ is rather weak.
4886: The complementary ACS observations, when combined with the Subaru
4887: observations, significantly narrow down the uncertainties on $c_{\rm
4888: vir}$, placing stringent constraints on the inner mass profile.
4889: In each panel the observed constraints on the Einstein radius
4890: ($\theta_{\rm E}\simeq 45$\arcsec at $z_s= 1$)
4891: are shown as a dashed curve.
4892: }
4893: \label{fig:joint}
4894: \end{figure*}
4895:
4896:
4897:
4898: %%%%%%%%%%%%% Figure 15
4899:
4900: \begin{figure*}[!htb]
4901: \begin{center}
4902: \includegraphics[width=100mm, angle=270]{f15.ps}
4903: \end{center}
4904: \caption{
4905: Joint constraints on the NFW model parameters,
4906: $(c_{\rm vir}, M_{\rm vir})$
4907: obtained from the Subaru tangential shear ($g_+$)
4908: profile of A1689 (see Figure 1 of B05a).
4909: %%%
4910: The cross shows the best-fitting set of the NFW parameters, and
4911: the contours show the $68\%$, $95\%$, and $99.7\%$ confidence levels
4912: ($\Delta\chi^2=2.3$, $6.17$, and $11.8$) in the $(c_{\rm vir}, M_{\rm
4913: vir})$-plane.
4914: The observed constraints on the Einstein radius,
4915: $\theta_{\rm E}\simeq 45$\arcsec at $z_s = 1$, are shown as a dashed
4916: curve. The dotted curve shows the $c_{\rm vir}-M_{\rm vir}$
4917: relation for $\theta_{\rm E}=24\arcsec$ at $z_s=1$.
4918: %%%
4919: The triangle symbol shows the best-fit set of
4920: $(c_{\rm vir}, M_{\rm vir})$ for the combined ACS and Subaru-2D (MEM+)
4921: results. The square and circle show the best-fit sets of $(c_{\rm vir},
4922: M_{\rm vir})$
4923: from the combined strong and weak lensing analysis of Halkola et
4924: al. (2006) and the weak lensing analysis of Limousin et al. (2007),
4925: respectively.
4926: }
4927: \label{fig:joint_gt}
4928: \end{figure*}
4929:
4930:
4931:
4932: %%%%%%%%%%%%% Figure 16
4933:
4934: \begin{figure*}[!htb]
4935: \begin{center}
4936: \includegraphics[width=.7\textwidth, angle=270]{f16.ps}
4937: \end{center}
4938: \caption{
4939: \label{fig:gt}
4940: Tangential distortion profile $g_+(\theta)$ ({\it square}, upper
4941: panel) from the Subaru weak lensing analysis of the red background sample
4942: (B05a). The solid curve shows
4943: the best-fit NFW profile derived from the joint strong and weak lensing
4944: analysis of ACS and Subaru observations (this work), incorporating full
4945: distortion and magnification information.
4946: The Einstein radius constraint ({\it triangle}) of
4947: $\theta_{\rm E}=45\arcsec$ ($z_s=1$), determined from multiply lensed
4948: images in ACS observations (B05b), is translated to the corresponding
4949: depth
4950: %%$\langle D_{ds}/D_s\rangle\approx 0.693$
4951: of the Subaru red background sample using the ACS+Subaru-2D NFW
4952: model ({\it solid}), and added to the distortion profile ($g_+=1$),
4953: marking the point of maximum distortion.
4954: The ACS+Subaru-2D NFW model ({\it solid}) fits well with the combined
4955: ACS and Subaru distortion information over the full range of data,
4956: $r=[80,2000] {\rm kpc}/h$, but somewhat overpredicts the outer
4957: distortion profile at $\theta \simgt 9\arcmin$ ($r\simgt 1.2{\rm
4958: Mpc}/h$).
4959: %%%%
4960: Also shown with the dashed curve is the best-fit NFW profile from the
4961: CFHT weak lensing analysis of Limousin et al. (2007),
4962: which, in contrast, is in good agreement with the Subaru outer profile
4963: at $\theta\simgt 4\arcmin$,
4964: %%($r\simgt 500 {\rm kpc}/h$),
4965: but underpredicts
4966: significantly the inner distortion profile and hence the Einstein radius.
4967: The dotted curve shows an NFW profile of Halkola et al. (2006)
4968: for a simultaneous fit to their ACS inner mass profile and the Subaru
4969: distortion profile of B05a shown here, but with a different weighting
4970: that prefers the inner strong-lensing based profile where the data
4971: imply a shallower slope (see Figure \ref{fig:mass}).
4972: %%%
4973: The lower panel shows
4974: the radial profile of the $45^\circ$ rotated component
4975: $g_\times(\theta)$ for the same Subaru red background sample
4976: (B05a). The $\times$-component of the red galaxy sample is consistent
4977: with a null signal at all radii, indicating the reliability the Subaru
4978: distortion analysis.
4979: }
4980: \end{figure*}
4981:
4982:
4983: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4985: %%%
4986: %%% Appendix
4987: %%%
4988: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4989: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4990:
4991: \appendix
4992:
4993:
4994:
4995: \section{Discretized Estimator for the Lensing Convergence}
4996: \label{app:kappad}
4997:
4998: In this Appendix, we aim to derive an expression
4999: for the discrete convergence profile
5000: %in terms of the
5001: using the
5002: weak lensing
5003: aperture densitometry
5004: $\zeta_{\rm c}(\theta)$
5005: given by
5006: equation (\ref{eq:zeta}).
5007: %We then derive the azimuthally averaged convergence,
5008: %$\kappa(\theta)$,
5009: %from the cumulative mass estimator,
5010: %$\zeta_{\rm c}(\theta)$.
5011: %%%%
5012: In the continuous limit,
5013: the averaged convergence $\bar\kappa(\theta)$
5014: and the convergence $\kappa(\theta)$
5015: are related by
5016: \begin{eqnarray}
5017: \bar{\kappa}(\theta)&=&\frac{2}{\theta^2}\int_0^{\theta}
5018: \!d\ln\theta'\theta'^2\kappa(\theta'),\\
5019: \kappa(\theta) &=&
5020: \frac{1}{2\theta^2}\frac{d(\theta^2\bar{\kappa})}{d\ln\theta}.
5021: \end{eqnarray}
5022: For a given set of annular radii $\theta_m$
5023: $(m=1,2,...,N)$,
5024: discretized estimators can be written in the following way:
5025: \begin{eqnarray}
5026: \label{eq:avkappa_d}
5027: \bar\kappa_m &\equiv&
5028: \bar{\kappa}(\theta_m)=
5029: \frac{2}{\theta_m^2}\sum_{l=1}^{m-1}
5030: \Delta\ln\theta_l
5031: %\frac{\Delta\theta_m}{\left<\theta_m\right>}
5032: %%%
5033: \bar\theta_l^2
5034: \kappa(\bar\theta_l),\\
5035: %
5036: \label{eq:kappa_d}
5037: \kappa_l&\equiv&
5038: \kappa(\bar\theta_l) =
5039: % \frac{
5040: %\theta^2_{l+1}\bar\kappa_{l+1}-
5041: %\theta^2_{l} \bar\kappa_l
5042: %}
5043: %{2\bar\theta_l^2 \Delta\ln\theta_l}
5044: %\equiv \alpha^{l}_2\bar\kappa_{l+1}
5045: %-\alpha^l_1\bar\kappa_l
5046: \alpha^l_2\bar\kappa_{l+1}-\alpha^l_1\bar\kappa_l
5047: \ \ \ \ \ (l=1,2,...,N-1),
5048: \end{eqnarray}
5049: where
5050: \begin{equation}
5051: \alpha_1^l = \frac{1}{2\Delta\ln\theta_l}
5052: \left(
5053: \frac{\theta_{l}}{ \overline{\theta}_l }
5054: \right)^2, \, \,
5055: %%
5056: \alpha_2^l = \frac{1}{2\Delta\ln\theta_l}
5057: \left(\frac{\theta_{l+1}}{ \overline{\theta}_l }\right)^2,
5058: \end{equation}
5059: with
5060: $\Delta\ln\theta_l \equiv (\theta_{l+1}-\theta_l)/\bar\theta_l$
5061: and $\bar\theta_l$
5062: being the area-weighted center of the $l$th
5063: annulus defined by $\theta_l$ and $\theta_{l+1}$;
5064: in the continuous limit, we have
5065: \begin{eqnarray}
5066: \bar\theta_l
5067: &\equiv&
5068: 2\int_{\theta_l}^{\theta_{l+1}}\!d\theta'\theta'^2/
5069: (\theta_{l+1}^2-\theta_{l}^2)\nonumber\\
5070: &=&
5071: \frac{2}{3}
5072: \frac{\theta_{l}^2+\theta_{l+1}^2+\theta_{l}\theta_{l+1}}
5073: { \theta_{l}+\theta_{l+1} }.
5074: \end{eqnarray}
5075:
5076: The technique of the aperture densitometry
5077: (Fahlman et al. 1994; Clowe et al. 2000) allows us
5078: to measure the azimuthally averaged convergence $\bar\kappa(\theta)$
5079: up to an additive constant $\bar{\kappa}_b$, corresponding to
5080: the mean convergence in the outer background annulus
5081: with inner and outer radii of $\theta_{\rm inn}$ and $\theta_{\rm out}$,
5082: respectively (\S \ref{subsec:zeta}):
5083: \begin{equation}
5084: \label{eq:z2k}
5085: \bar\kappa(\theta)=
5086: \zeta_{\rm c}(\theta)
5087: +\bar\kappa_b.
5088: \end{equation}
5089: Substituting equation (\ref{eq:z2k}) into equation (\ref{eq:kappa_d})
5090: yields the desired expression
5091: as
5092: \begin{equation}
5093: \kappa(\theta_l)=\alpha^{l}_2\zeta_{\rm c}(\theta_{l+1})
5094: -\alpha^l_1\zeta_{\rm c}(\theta_l)
5095: + (\alpha_2^l-\alpha_1^l)\bar\kappa_b.
5096: \end{equation}
5097:
5098:
5099: \section{The NFW Lens Model}
5100: \label{app:nfw}
5101:
5102: The NFW universal density profile
5103: has a two-parameter functional form as
5104: \begin{eqnarray}
5105: \rho_{\rm NFW}(r)= \frac{\rho_s}{(r/r_s)(1+r/r_s)^2} \label{eq:nfw}
5106: \end{eqnarray}
5107: where $\rho_s$ is a characteristic inner density, and $r_s$ is a
5108: characteristic inner radius.
5109: The virial properties are related
5110: %%%
5111: In stead of using $\rho_s$ and $r_s$,
5112: we introduce for an NFW halo
5113: the virial mass $M_{\rm vir}$
5114: and the concentration parameter,
5115: $c_{\rm vir}\equiv r_{\rm vir}/r_s$, defined as the ratio of the
5116: virial radius $r_{\rm vir}$ to the scale radius.
5117: The virial mass and virial radius are related through the
5118: following equation:
5119: \begin{eqnarray}
5120: \label{eq:mvirrvir}
5121: M_{\rm vir} = \frac{4\pi}{3}\bar{\rho}(z_{\rm vir})\Delta_{\rm
5122: vir}r_{\rm vir}^3,
5123: \end{eqnarray}
5124: where
5125: $\Delta_{\rm vir}$
5126: is the mean overdensity with respect to the mean cosmic density
5127: $\bar\rho(z_{\rm vir})$
5128: %$\bar{\rho}(z_{\rm vir}) = \Omega_m(z_{\rm vir})\rho_{\rm crit}(z_{\rm vir})$
5129: at the virialization epoch $z_{\rm vir}$,
5130: predicted by the dissipationless spherical
5131: tophat collapse model (Peeebles 1980; Eke, Cole, \& Frenk 1996).
5132: We assume the cluster redshift $z_d$ is
5133: equal to the cluster virial redshift $z_{\rm vir}$.
5134: We use the following fitting formula in a flat 3-space with cosmological
5135: constant (see Oguri, Taruya, \& Suto 2001):
5136: \begin{eqnarray}
5137: \Delta_{\rm vir} &=& 18 \pi^2 (1+0.4093 \omega_{\rm vir}^{0.9052}),
5138: \end{eqnarray}
5139: where $\omega_{\rm vir}\equiv 1/\Omega_m(z_{\rm vir})-1$.
5140:
5141:
5142:
5143: %%%
5144: The inner density $\rho_s$ can be then
5145: expressed in terms of other virial properties of the NFW halo:
5146: \begin{equation}
5147: \rho_s = \bar{\rho}(z_{\rm vir})
5148: \frac{\Delta_{\rm vir}}{3}\frac{c_{\rm vir}^3}{\ln(1+c_{\rm vir})-
5149: c_{\rm vir}/(1+c_{\rm vir})}.
5150: \end{equation}
5151: %where $\rho_{\rm cr}$ is the
5152: %critical density of the universe at the cluster redshift $z_d$,
5153: %and $r_s$ is the NFW scale radius.
5154: %%%
5155: Hence, for a given cosmological model and a halo virial redshift,
5156: we can specify the NFW model with the halo virial mass
5157: $M_{\rm vir}$ and the halo concentration parameter $c_{\rm vir}$.
5158:
5159: For an NFW profile, it is useful to decompose the convergence
5160: $\kappa(\theta)$ and the
5161: averaged convergence $\bar{\kappa}(\theta)$ as
5162: \begin{eqnarray}
5163: \kappa_{\rm NFW}(x) &=& \frac{b}{2} f(x),\\
5164: \bar{\kappa}_{\rm NFW}(x) &=& \frac{b}{x^2}g(x),
5165: \end{eqnarray}
5166: where $b=4\rho_s r_s/\Sigma_{\rm crit}(z_d,z_s)$ is the dimensionless
5167: scaling convergence,
5168: $x=\theta/(r_s/D_d)$ is the dimensionless angular radius,
5169: and $f(x)$ and $g(x)$ are dimensionless functions.
5170: We have analytic expressions for $f(x)$ and $g(x)$
5171: as (Bartelmann 1996):
5172: \begin{eqnarray}
5173: \label{eq:nfw_f}
5174: f(x) &=&
5175: \left\{
5176: \begin{array}{ll}
5177: \frac{1}{1-x^2}
5178: \left(
5179: -1+ \frac{2}{\sqrt{1-x^2}} {\rm arctanh}\sqrt{\frac{1-x}{1+x}}
5180: \right) & \ \ \ (x<1), \\
5181: \frac{1}{3} & \ \ \ (x=1), \\
5182: %%
5183: \frac{1}{x^2-1}
5184: \left(
5185: +1- \frac{2}{\sqrt{x^2-1}} \arctan\sqrt{\frac{x-1}{x+1}}
5186: \right) & \ \ \ (x>1),
5187: \end{array}
5188: \right. \\
5189: %%%
5190: \label{eq:nfw_g}
5191: g(x) &=& \ln\left(\frac{x}{2}\right)
5192: +
5193: \left\{
5194: \begin{array}{ll}
5195: \frac{2}{\sqrt{1-x^2}} {\rm arctanh}\sqrt{\frac{1-x}{1+x}}
5196: & \ \ \ (x<1), \\
5197: 1 & \ \ \ (x=1),\\
5198: \frac{2}{\sqrt{x^2-1}} \arctan \sqrt{\frac{x-1}{x+1}}
5199: & \ \ \ (x>1).
5200: \end{array}
5201: \right.
5202: \end{eqnarray}
5203: The tangential shear $\gamma_{+, {\rm NFW}}(\theta)$ is then
5204: evaluated by
5205: \begin{equation}
5206: \gamma_{+,{\rm NFW}}(\theta) = \bar\kappa_{\rm NFW}(\theta)-\kappa_{\rm
5207: NFW}(\theta).
5208: \end{equation}
5209:
5210: For a given source redshift $z_s$, the Einstein radius
5211: is then readily calculated by $1=\kappa_{\rm NFW}(\theta)$;
5212: or more explicitly,
5213: using equation (\ref{eq:nfw_g}) we have
5214: \begin{equation}
5215: \theta_{\rm E}^2 = b \theta_s^2
5216: g(\theta_{\rm E}/\theta_s),
5217: %\theta_{\rm E}^2 = b\left(\frac{\theta_{\rm vir}}{c_{\rm vir}}\right)^2
5218: %g(\frac{c_{\rm vir}\theta_{\rm E}}{\theta_{\rm vir}})
5219: \end{equation}
5220: where $\theta_s\equiv r_s/D_d=r_{\rm vir}/(c_{\rm vir}D_d)$
5221: is the angular size of the
5222: NFW scale radius.
5223: This equation for $\theta_{\rm E}$ can be solved
5224: numerically, for example, by the Newton-Raphson method.
5225:
5226:
5227:
5228:
5229: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5230: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5231: %%%
5232: %%% References
5233: %%%
5234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5235: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5236:
5237:
5238:
5239: \begin{thebibliography}{99}
5240:
5241: % X-ray
5242: \bibitem{andersson}
5243: Andersson, K. E. \& Madejski, G. M. 2004,
5244: \apj, 607, 190
5245:
5246: \bibitem{Bacon00}
5247: Bacon, D.~J., Refregier, A.~R., Ellis, R.~S. 2000, MNRAS, 318, 625
5248:
5249: \bibitem{baltz07}
5250: Baltz, E.~A., Marshall, P., Oguri, M. 2007, astro-ph/0705.0682
5251:
5252: \bibitem{BARD1689}
5253: Bardeau, S. et al. 2005,
5254: \aap, 434, 433
5255:
5256: \bibitem{bardeau07}
5257: Bardeau, S.,
5258: Soucail, G.,
5259: Kneib, J.-P.,
5260: Czoske, O.,
5261: Ebeling, H.,
5262: Hudelot, P.,
5263: Smail, I.,
5264: Smith, G.~P.,
5265: 2007, \aap, 470, 449
5266:
5267: \bibitem[Bartelmann(1995)]{bartelmann95}
5268: Bartelmann, M. 1995, \aap, 299, 11
5269:
5270: \bibitem{review}
5271: Bartelmann, M., \& Schneider, P. 2001, Phys.Rep., 340, 291
5272:
5273: \bibitem[Bertin, Arnouts (1996)]{bertin96}
5274: Bertin, E., \& Arnouts, S. \ 1998, A\&AS, 117, 393
5275:
5276:
5277: \bibitem{BridleMEM}
5278: Bridle, S.~L., Hobson, M.~P., Lasenby, A.~N., \& Saunders,
5279: R. 1998, MNRAS, 299, 895
5280:
5281: %1e0657
5282: \bibitem[Brada${\check {\rm c}}$ et al. 2006]{bra06} Brada${\check
5283: {\rm c}}$, M. et al. 2006, ApJ,
5284: 652, 937.
5285:
5286: \bibitem{BTP}
5287: Broadhurst, T., Taylor, A.~N., \&
5288: Peacock, J.~A. 1995, \apj, 438, 49
5289:
5290: \bibitem{BA1689}
5291: Broadhurst, T., et al. 2005, \apj, 621, 53
5292: (Broadhurst et al. 2005b, B05b)
5293:
5294: \bibitem{BA1689wl}
5295: Broadhurst, T., Takada, M., Umetsu, K.,
5296: Kong, X., Arimoto, N., Chiba, M., \& Futamase, T. 2005, 619,
5297: L143 (Broadhurst et al. 2005a, B05a)
5298:
5299: \bibitem{bb2008}
5300: Broadhurst, T. \& Barkana, R. 2008, submitted to \mnras
5301: (arXiv:astro-ph/0801.1875)
5302:
5303:
5304: \bibitem[]{Bullock}
5305: Bullock, J.~S., Kolatt, T.~S., Sigad, Y., Somerville, R.~S.,
5306: Kravtsov, A.~V., Klypin, A.~A., Primack, J.~R., Dekel, A. 2001,
5307: \mnras, 321, 559
5308:
5309: \bibitem[{{Capak} {et~al.}(2004)
5310: {Capak}, {Cowie}, {Hu}, {Barger}, {Dickinson},
5311: {Fernandez}, {Giavalisco}, {Komiyama},
5312: {Kretchmer}, {McNally}, {Miyazaki},
5313: {Okamura}, \& {Stern}}]{2004AJ....127..180C}
5314: {Capak}, P. {et~al.} 2004, \aj, 127, 180
5315:
5316: \bibitem[]{capak2007}
5317: Capak, P. et al. 2007, \apjs, 172, 99
5318:
5319:
5320: \bibitem{Clowe00}
5321: Clowe, D.,
5322: Luppino, G.~A.,
5323: Kaiser, N., \&
5324: Gioia, I.~M. 2000, \apj, 539, 540
5325:
5326: \bibitem{Clowe01}
5327: Clowe, D. \& Schneider, P. 2001, \aap, 379, 384
5328:
5329: %% Masking area
5330: \bibitem{Cobb}
5331: Cobb, B. E., Bailyn, C. D., van Dokkum, P. G.,
5332: \& Natarajan, P. 2006, \apjl, 651, L85
5333:
5334:
5335: \bibitem{comeford07}
5336: Comerford, J.~M. \& Natarajan, P. 2007, \mnras, 379, 190
5337:
5338: \bibitem{corless}
5339: Corless, V.~L. \& King, L.~J. 2007,
5340: \mnras, 380, 149
5341:
5342: \bibitem{crittenden02}
5343: Crittenden, R.~G., atarajan, P., Pen, U.-L., Theuns, T. 2002,
5344: \apj, 568, 20
5345:
5346:
5347: %%dm
5348: \bibitem[Eke, Cole, \& Frenk 1996]{eke96}
5349: Eke, V. R., Cole, S. \& Frenk,
5350: C. S. 1996, \mnras, 282, 263.
5351:
5352:
5353: %%wl
5354: \bibitem[]{Erben et al. (2001)}
5355: Erben, T., Van Waerbeke, L., Bertin, E., Mellier, Y., Schneider, P.
5356: 2001, A\&A, 366, 717
5357:
5358: \bibitem[]{Fahlman et al. (1994)}
5359: Fahlman, G., Kaiser, N., Squires, G., Woods, D., 1994, \apj, 437, 56
5360:
5361: \bibitem[Gavazzi et al.(2003)]{gavazzi03}
5362: Gavazzi, R., Fort, B., Mellier, Y., Pello, R.,
5363: \& Dantel-Fort, M.
5364: 2003, \aap, 403, 11
5365:
5366: % substructure
5367: \bibitem{girardi1997}
5368: Girardi, M.,
5369: Fadda, D.,
5370: Escalera, E.,
5371: Giuricin, G.,
5372: Mardirossian, F.,
5373: \& Mezzetti, M. 1997, \apj, 490, 56
5374:
5375: % non gaussian
5376: \bibitem{grossi}
5377: Grossi, M.,
5378: Dolag, K.,
5379: Branchini, E.,
5380: Matarrese, S. \& Moscardini, L. 2006, \mnras, 382, 1261
5381:
5382:
5383: % CFH12K
5384: %\bibitem{gavazzi04}
5385: % Gavazzi, R., Mellier, Y., Fort, B.,
5386: % Cuillandre, J.-C., Dantel-Fort, M. 2004,
5387: % \aap, 422, 407
5388:
5389: % N-body
5390: %\bibitem{ghigna2000}
5391: % Ghigna, J. et al. 2000, \apj, 544, 616
5392:
5393:
5394: % cosmic shear
5395: %\bibitem{Hamana03}
5396: % Hamana, T. et al. 2003, \apj, 597, 98
5397:
5398: \bibitem{STEP1}
5399: Heymans, C. et al. 2006, \mnras, 368, 1323
5400:
5401: \bibitem{halkola07}
5402: Halkola, A.
5403: Seitz, S.,
5404: \& Pannella, M. 2006, \mnras, 372, 1425
5405:
5406: \bibitem{hennawi2007}
5407: Hennawi, J. F.,
5408: Dalal, N., Bode, P., Ostriker, J. P. 2007, \apj, 654, 714
5409:
5410: \bibitem{hilbert07}
5411: Hilbert, S., White, S.~D.~M.,
5412: Hartlap, J., Schneider, P. 2007, \mnras,
5413: 382,121
5414:
5415: \bibitem{hobson98}
5416: Hobson, M.~P., Lasenby, A.XN. 1998,
5417: \mnras, 298, 905
5418:
5419: % HST CL 1358+62
5420: \bibitem{hoekstra98}
5421: Hoekstra, H., Franx, M., Kuijken, K., \& Squires, G. 1998,
5422: \apj, 504, 636
5423:
5424: % CFHTLS cosmic shear
5425: %\bibitem{hoekstra07}
5426: % Hoekstra, H. et al. 2007, \apj, 647, 116
5427:
5428: \bibitem[Hudson, Gwyn, Dahle \& Kaiser]{hud98}
5429: Hudson, M. J., Gwyn, S. D. J., Dahle, H. \& Kaiser, N., 1998,
5430: \apj, 503, 531.
5431:
5432:
5433: \bibitem{jingsuto2000}
5434: Jing, Y. P., \& Suto, Y. 2000, \apjl, 529, L6
5435:
5436: \bibitem{johnston}
5437: Johnston, D. E. et al. 2007, arxiv:astro-ph/0709.1159
5438:
5439:
5440: \bibitem{KS93}
5441: Kaiser, N., \& Squires, G. 1993, ApJ, 404, 441
5442:
5443: \bibitem[Kaiser 1995]{kai95a}
5444: Kaiser, N. 1995, \apj, 439, L1.
5445:
5446: \bibitem[]{KSB}
5447: Kaiser, N., Squires, G., Broadhurst, T., 1995, \apj, 449, 460
5448:
5449: \bibitem[Kawahara et al.(2007)]{2007ApJ...659..257K}
5450: Kawahara, H., Suto,
5451: Y., Kitayama, T., Sasaki, S.,
5452: Shimizu, M., Rasia, E.,
5453: \& Dolag, K.\ 2007, \apj, 659, 257
5454:
5455: \bibitem{KA1689}
5456: King, L. J., Clowe, D. I., Schneider, P. 2002 A\&A, 383, 118
5457:
5458: % projection
5459: \bibitem{kingcorless}
5460: King, L. \& Corless, V. 2007, \mnras, 374, 37
5461:
5462:
5463: \bibitem[Kneib et al.(2003)]{kneib03}
5464: Kneib, J.-P. et al. 2003, \apj, 598, 804
5465:
5466: \bibitem{lemze08}
5467: Lemze, D., Barkana, L., Broadhurst, T., Rephaeli, Y. 2008,
5468: \mnras in press (astro-ph/0711.3908)
5469:
5470: \bibitem{GA1689}
5471: Leonard, A., Goldberg, D.~M., Haaga, J.~L., Massey, R. 2007,
5472: \apj, 666, L51
5473:
5474:
5475: \bibitem{SW-A1689}
5476: Limousin, M. et al. 2007, ApJ, 668, 643
5477:
5478:
5479: %\bibitem[Mahdavi et al. 2007]{mah07}
5480: % Mahdavi, A., Hoekstra, H., Babul,
5481: % A., Balam, D.\& Capak, P. 2007,
5482: % \apj, 668., 806
5483:
5484:
5485: \bibitem{marshall02}
5486: Marshall, P.~J., Hobson, M.~P.,
5487: Gull, S.~F., Bridle, S.~L. 2002,
5488: \mnras, 335, 1037
5489:
5490: \bibitem{STEP2}
5491: Massey, R. et al. 2007, \mnras, 376, 13
5492:
5493: %\bibitem[Miyazaki et al. (2002)]{miya02}
5494: % Miyazaki, S., et al. 2002, PASJ, 54, 833
5495:
5496: \bibitem{elinor}
5497: Medezinski, E., Broadhurst, T., Umetsu, K.,
5498: Coe, D., Benitez, N., Ford, H., Rephaeli, Y.,
5499: Arimoto, N., \& Kong, X. 2007, \apj, 663, 717
5500:
5501: % WL DM search
5502: %\bibitem{miyazaki02}
5503: % Miyazaki, S. et al. 2002, \apj, 580, L97
5504:
5505:
5506: \bibitem{maisinger98}
5507: Maisinger, K., Hobson, M.~P., Lasenby, A.~N. 1998,
5508: \mnras, 290, 313
5509:
5510: \bibitem{Navarro et al. (1996)}
5511: Navarro, J.~F., Frenk, C.~S., White, S.~D.~M., 1996, \apj, 462, 563
5512:
5513: \bibitem{NFW}
5514: Navarro, J.~F., Frenk, C.~S., White, S.~D.~M., 1997, \apj, 490, 493
5515:
5516: \bibitem{Neto07}
5517: Neto, A.~F. et al. 2007, \mnras, 381, 1450
5518:
5519:
5520: \bibitem[Oguri, Taruya, \& Suto 2001]{4407}
5521: Oguri, M., Taruya, A., Suto, Y. 2001, \apj, 559, 572
5522:
5523:
5524: \bibitem{Oguri et al. (2005)}
5525: Oguri, M., Takada, M., Umetsu, K., \& Broadhurst, T. 2005,
5526: \apj, 632, 841
5527:
5528:
5529: \bibitem{OU07}
5530: Okabe, N. \& Umetsu, K. 2008, PASJ in press (arXiv:astro-ph/0702649)
5531:
5532: %\bibitem{HOLICs}
5533: % Okura, Y., Umetsu, K., \& Futamase, T., 2007, ApJ, 660, 995
5534:
5535:
5536: \bibitem{HOLICs2}
5537: Okura, Y., Umetsu, K., \& Futamase, T. 2008, ApJ in press
5538: (arXiv:astro-ph/0710.2262)
5539:
5540: %%dm halo
5541: \bibitem[Peebles 1980]{4428}
5542: Peebles, P. J. E., 1980 The large-scale structure of
5543: the universe (Princeton University Press)
5544:
5545: %% N-body
5546: %\bibitem[Power2003]{4433}
5547: % Power et al. 2003, \mnras, 338, 14
5548:
5549:
5550: \bibitem{Press}
5551: Press, W.~H., Teukolsky, S.,
5552: Vetterling, T.~W., and Flannery, B.~P. 1992,
5553: {\it Numerical Recipes in Fortran}, Cambidge University Press,
5554: 2nd edition
5555:
5556: %% non gaussian
5557: \bibitem{sadeh}
5558: Sadeh, S., Rephaeli, Y., \& Silk, J. 2007, \mnras, 380, 637
5559:
5560:
5561: \bibitem{sand04}
5562: Sand, D.~J., Treu, T., Smith, G.~P., \& Ellis, R.~S. 2004,
5563: \apj, 604, 88
5564:
5565:
5566: \bibitem[Sato et al. 2003]{sat03}
5567: Sato, J., Umetsu, K., Futamase, T. \&
5568: Yamada, T. 2003, \apj, 582, L67.
5569:
5570:
5571: \bibitem[Seitz et al. (1998)]{seitz98}
5572: Seitz, S., Schneider, P., \& Bartelmann, M. 1998, \aap, 337, 325
5573:
5574: \bibitem{seljak98}
5575: Seljak, U. 1998, \apj, 506, 64
5576:
5577: \bibitem{sereno07}
5578: Sereno, M. 2007, \mnras, 380, 1207
5579:
5580:
5581: \bibitem{ss95}
5582: Schneider, P. \& Seitz, C. 1995, \aap, 294, 411
5583:
5584:
5585: \bibitem[Schneider et al. (2000)]{schneider00}
5586: Schneider, P., King, L., \& Erben, T. 2000, \aap, 353, 41
5587:
5588:
5589: \bibitem{shimizu03}
5590: Shimizu, M.,
5591: Kitayama, T.,
5592: Sasaki, S.,
5593: \& Suto, Y. 2003, \apj, 590, 197
5594:
5595: \bibitem{Smith05}
5596: Smith, G.~P. et al. 2005, \mnras, 359, 417
5597:
5598: \bibitem{wmap}
5599: Spergel, D.~N. et al. 2003, \apjs, 148, 175
5600:
5601: \bibitem{wmap3}
5602: Spergel, D.~N. et al. 2007, \apjs, 170, 377
5603:
5604: \bibitem{Springel05}
5605: Springel, V. et al. 2005, Nature, 435, 629
5606:
5607: % mem
5608: \bibitem{Suyu}
5609: Suyu, S. H., Marshall, P. J., Hobson, M. P., Blandford,
5610: R. D. 2006, \mnras, 371, 983
5611:
5612: \bibitem{takada05}
5613: Takada, M. \& Jain, B. 2003, \mnras, 340, 580
5614:
5615: \bibitem{tasitsiomi2004}
5616: Tasitsiomi, A. et al. 2004, \apj, 607, 125
5617:
5618: \bibitem{Taylor98}
5619: Taylor, A.~N., Dye, S., Broadhurst, T.,
5620: Benitez, N., van Kampen, E. 1998, \apj, 501, 539
5621:
5622: \bibitem{Taylor07}
5623: Taylor, A.~N., Kitching, T.~D., Bacon, D.~J., \& Heavens, A.~F.
5624: 2007, \mnras, 374, 1377
5625:
5626: \bibitem{Tegmark04}
5627: Tegmark, M. et al. 2004, \prd, 69, 3501
5628:
5629: \bibitem{Tyson90}
5630: Tyson, J.~A., Wenk, R.~A., \& Valdes, F. 1990, \apj, 349, L1
5631:
5632: \bibitem{TA1689}
5633: Tyson, J.~A. \& Fisher, P. 1995, \apj, 446, L55
5634:
5635:
5636: \bibitem{UTF1999}
5637: Umetsu, K, Tada, M., \& Futamase, T. 1999,
5638: Prog.~Theor.~Phys.~Suppl., 133, 53
5639:
5640: \bibitem{UF2000}
5641: Umetsu, K, \& Futamase, T. 2000, \apj, 539, L5
5642:
5643: \bibitem{UA1689}
5644: Umetsu, K., Takada, M., Broadhurst, T. 2007,
5645: Mod.~Phys.~Lett.~A, 22, 2099 (astro-ph/0702096)
5646:
5647: % cosmic shear
5648: %\bibitem{vanwaerbeke00}
5649: % Van Waerbeke, L. et al. 2000, \aap, 358, 30
5650:
5651: \bibitem{wechsler2002}
5652: Wechsler, R. H. et al. 2002, \apj, 568, 52
5653:
5654: \bibitem{wechsler2006}
5655: Wechsler, R. H.,
5656: Zentner, A. R.,
5657: Bullock, J. S.,
5658: Kravtsov, A. V., \&
5659: Allgood, B. 2006, \apj, 652, 71
5660:
5661: \bibitem{wright}
5662: Wright, C. O. \& Brainerd, T. G. 2000, \apj, 534, 34
5663:
5664:
5665: \bibitem{zhangpen}
5666: Zhang, P. \& Pen, U.-L. 2005, \prl, 95, 1302
5667:
5668: %\bibitem{vanWaerbeke}
5669: % Van Waerbeke, L. et al. 2001, A\&A, 374, 757
5670:
5671:
5672:
5673:
5674: \end{thebibliography}
5675:
5676: \end{document}
5677:
5678: %%
5679: %% End of file `sample.tex'.
5680:
5681: