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: %\documentclass[12pt]{emulateapj}
19: %\usepackage{apjfonts,rotating,graphicx,graphics}
20: %\usepackage{apjfonts}
21:
22: %% ApJ preprint
23: \documentclass[a4paper,12pt,preprint]{aastex}
24: \usepackage{graphics,graphicx,rotating}
25:
26: %% manuscript produces a one-column, double-spaced document:
27:
28: %% \documentclass[manuscript]{aastex}
29:
30: %% preprint2 produces a double-column, single-spaced document:
31:
32: %% \documentclass[preprint2]{aastex}
33:
34: %% Sometimes a paper's abstract is too long to fit on the
35: %% title page in preprint2 mode. When that is the case,
36: %% use the longabstract style option.
37:
38: %% \documentclass[preprint2,longabstract]{aastex}
39:
40: %% If you want to create your own macros, you can do so
41: %% using \newcommand. Your macros should appear before
42: %% the \begin{document} command.
43: %%
44: %% If you are submitting to a journal that translates manuscripts
45: %% into SGML, you need to follow certain guidelines when preparing
46: %% your macros. See the AASTeX v5.x Author Guide
47: %% for information.
48:
49: \newcommand{\tp}{\hspace{-1mm}+\hspace{-1mm}}
50: \newcommand{\tm}{\hspace{-1mm}-\hspace{-1mm}}
51: \newcommand{\cF}{{\cal F}}
52: \newcommand{\cG}{{\cal G}}
53: \newcommand{\cA}{{\cal A}}
54: \newcommand{\cD}{{\cal D}}
55: \newcommand{\dt}{\Delta\theta}
56: \newcommand{\db}{\Delta\beta}
57: \newcommand{\trQ}{{\rm tr}Q}
58: \newcommand{\paren}[1]{\left( #1 \right)}
59: \newcommand{\astral}[1]{#1^*}
60: \newcommand{\iotaIII}{\iota_{I\hspace{-.1em}I\hspace{-.1em}I}}
61: \newcommand{\upsilonII}{\upsilon_{I\hspace{-.1em}I}}
62: \newcommand{\upsilonIV}{\upsilon_{I\hspace{-.1em}V}}
63: \newcommand{\upsilonVI}{\upsilon_{V\hspace{-.1em}I}}
64: \newcommand{\tauIII}{\tau_{I\hspace{-.1em}I\hspace{-.1em}I}}
65: \newcommand{\tauVII}{\tau_{V\hspace{-.1em}I\hspace{-.1em}I}}
66: \newcommand{\Real}[1]{{\rm Re}\left[ #1 \right]}
67: \newcommand{\dtrQ}[1]{\frac{#1}{\trQ}}
68: \newcommand{\dxi}[1]{\frac{#1}{\xi}}
69: \newcommand{\dsig}[2]{\frac{#1}{\sigma^{#2}}}
70:
71: \newcommand{\simgt}{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}}
72: \newcommand{\simlt}{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}}
73: \newcommand{\gIII}{I\hspace{-.3mm}I\hspace{-.3mm}I}
74:
75: \def\bbeta{\mbox{\boldmath $\beta$}}
76: \def\btheta{\mbox{\boldmath $\theta$}}
77: \def\bnabla{\mbox{\boldmath $\nabla$}}
78: \def\bk{\mbox{\boldmath $k$}}
79: \def\bvtheta{\mbox{\boldmath $\vartheta$}}
80: \def\bdelta{\mbox{\boldmath $\Delta$}}
81: \def\bphi{\mbox{\boldmath $\phi$}}
82:
83: %% You can insert a short comment on the title page using the command below.
84:
85: %\slugcomment{Not to appear in Nonlearned J., 45.}
86:
87: %% If you wish, you may supply running head information, although
88: %% this information may be modified by the editorial offices.
89: %% The left head contains a list of authors,
90: %% usually a maximum of three (otherwise use et al.). The right
91: %% head is a modified title of up to roughly 44 characters.
92: %% Running heads will not print in the manuscript style.
93:
94: \shorttitle{A Method for Weak Lensing Flexion Analysis}
95: \shortauthors{Okura, Umetsu, \& Futamase}
96:
97: %% This is the end of the preamble. Indicate the beginning of the
98: %% paper itself with \begin{document}.
99:
100: \begin{document}
101:
102: %% LaTeX will automatically break titles if they run longer than
103: %% one line. However, you may use \\ to force a line break if
104: %% you desire.
105:
106:
107: \title{A Method for Weak Lensing Flexion Analysis by the HOLICs Moment
108: Approach\altaffilmark{1}}
109: %using Higher Order
110: %Lensing Image Characteristics}
111: %\title{A Method for Weak Lensing Flexion Analysis using HOLICs}
112: %\title{Generalized Weak Lensing Analysis: \ \ HOLICs to Gravitational Flexions}
113:
114: %\title{A Generalized Method for Weak Lensing Analysis: \ \ HOLICs to Gravitational Flexions}
115: %\title{Measuring mass distribution of A1689 using HOLICs}
116:
117:
118: %% Use \author, \affil, and the \and command to format
119: %% author and affiliation information.
120: %% Note that \email has replaced the old \authoremail command
121: %% from AASTeX v4.0. You can use \email to mark an email address
122: %% anywhere in the paper, not just in the front matter.
123: %% As in the title, use \\ to force line breaks.
124:
125:
126: \author{Yuki Okura\altaffilmark{2}}
127: \email{aepgstrx@astr.tohoku.ac.jp}
128: \author{Keiichi Umetsu\altaffilmark{3}}
129: \email{keiichi@asiaa.sinica.edu.tw}
130: \and
131: \author{Toshifumi Futamase\altaffilmark{2}}
132: \email{tof@astr.tohoku.ac.jp}
133: %% Notice that each of these authors has alternate affiliations, which
134: %% are identified by the \altaffilmark after each name. Specify alternate
135: %% affiliation information with \altaffiltext, with one command per each
136: %% affiliation.
137:
138: \altaffiltext{1}{Based in part on data collected at the Subaru Telescope,
139: which is operated by the National Astronomical Society of Japan}
140: \altaffiltext{2}
141: {Astronomical Institute, Tohoku University, Sendai 980-8578, Japan}
142: \altaffiltext{3}
143: {Institute of Astronomy and Astrophysics, Academia Sinica, P.~O. Box 23-141, Taipei 106, Taiwan, Republic of China}
144:
145: %% Mark off your abstract in the ``abstract'' environment. In the manuscript
146: %% style, abstract will output a Received/Accepted line after the
147: %% title and affiliation information. No date will appear since the author
148: %% does not have this information. The dates will be filled in by the
149: %% editorial office after submission.
150:
151:
152: \begin{abstract}
153: We have developed a method for measuring
154: higher-order weak lensing distortions of faint background galaxies,
155: namely the weak gravitational flexion,
156: by fully extending the Kaiser, Squires \& Broadhurst method to include
157: higher-order lensing image characteristics (HOLICs) introduced by Okura,
158: Umetsu, \& Futamase.
159: We take into account explicitly the weight function in
160: calculations of noisy shape moments and the effect of
161: higher-order
162: PSF anisotropy, as well as isotropic PSF smearing.
163: Our HOLICs formalism allows accurate measurements of flexion from
164: practical observational data in the presence of non-circular,
165: anisotropic PSF.
166: We test our method using mock observations of simulated galaxy images
167: and actual, ground-based
168: Subaru observations of the massive galaxy cluster A1689
169: ($z=0.183$). From the high-precision
170: measurements of spin-1 first flexion,
171: we obtain a high-resolution mass map in the central region
172: of A1689.
173: The reconstructed mass map shows a bimodal feature in the central
174: $4'\times 4'$ region of the cluster.
175: The major, pronounced peak is
176: associated
177: with the brightest cluster galaxy and central cluster members,
178: while the secondary mass peak is associated with a local concentration
179: of bright galaxies.
180: The refined, high-resolution mass map of A1689
181: demonstrates the power of the generalized weak lensing analysis
182: techniques for quantitative and accurate measurements of the
183: weak gravitational lensing signal.
184: \end{abstract}
185:
186: %% Keywords should appear after the \end{abstract} command. The uncommented
187: %% example has been keyed in ApJ style. See the instructions to authors
188: %% for the journal to which you are submitting your paper to determine
189: %% what keyword punctuation is appropriate.
190:
191: %% Authors who wish to have the most important objects in their paper
192: %% linked in the electronic edition to a data center may do so in the
193: %% subject header. Objects should be in the appropriate "individual"
194: %% headers (e.g. quasars: individual, stars: individual, etc.) with the
195: %% additional provision that the total number of headers, including each
196: %% individual object, not exceed six. The \objectname{} macro, and its
197: %% alias \object{}, is used to mark each object. The macro takes the object
198: %% name as its primary argument. This name will appear in the paper
199: %% and serve as the link's anchor in the electronic edition if the name
200: %% is recognized by the data centers. The macro also takes an optional
201: %% argument in parentheses in cases where the data center identification
202: %% differs from what is to be printed in the paper.
203:
204: %\keywords{globular clusters: general ---
205: %globular clusters: individual(\objectname{NGC 6397},
206: %\object{NGC 6624}, \objectname[M 15]{NGC 7078},
207: %\object[Cl 1938-341]{Terzan 8})}
208:
209:
210: %% Modified by KU (2007/07/21)
211: \keywords{cosmology: theory --- dark matter --- galaxies: clusters:
212: individual (A1689) --- gravitational lensing}
213:
214:
215: %% From the front matter, we move on to the body of the paper.
216: %% In the first two sections, notice the use of the natbib \citep
217: %% and \citet commands to identify citations. The citations are
218: %% tied to the reference list via symbolic KEYs. The KEY corresponds
219: %% to the KEY in the \bibitem in the reference list below. We have
220: %% chosen the first three characters of the first author's name plus
221: %% the last two numeral of the year of publication as our KEY for
222: %% each reference.
223:
224:
225: \section{Introduction}
226:
227:
228: Propagation of light rays from a distant source to the observer is
229: governed by the gravitational field of intervening mass fluctuations
230: as well as by the global geometry of the universe.
231: %Observations of the properties of
232: The images of background sources
233: %are observed as gravitational lensing phonomena, which
234: hence carry the imprint of the gravitational potential
235: of intervening cosmic structures, and
236: %%%
237: %The statistical properties of gravitational lensing
238: their statistical properties
239: can be used to test the background cosmological models.
240: %%%
241:
242:
243: Weak gravitational lensing is responsible for the weak
244: %shearing
245: shape-distortion
246: and
247: magnification of the images of background sources due to the
248: gravitational field of intervening matter
249: (e.g., Bartelmann \& Schneider 2001; Umetsu, Futamase, \& Tada 1999).
250: %%%%%
251: To the first order, weak lensing gives rise to
252: a few -- $10\%$ levels of
253: elliptical distortions
254: in images of background sources, responsible for the second-order
255: derivatives of the gravitational lensing potential.
256: Thus, the weak lensing signal, measured from tiny but coherent
257: quadrupole distortions
258: in galaxy shapes, can provide a direct measure of the projected mass
259: distribution of cosmic structures.
260: %%%%%%%%
261: %In the past decade,
262: However, practical weak lensing observations subject to
263: the effects of atmospheric seeing,
264: isotropic/anisotropic PSF, and (residual) camera distortion across the
265: field of view,
266: which must be examined from the stellar shape measurements
267: and corrected for in the weak lensing analysis.
268: Practical methods for PSF corrections and shear measurements/calibrations
269: have been studied and developed by many authors,
270: such as
271: pioneering work of Kaiser, Squires, \& Broadhurst (1995, hereafter
272: KSB), the Shapelets technique which describes PSF and object images in
273: terms of Gaussian-Hermite expansions (Refregier 2003),
274: and recent systematic, collaborative efforts by
275: The Shear TEsting Programme (Heymans et al. 2006; Massey et al. 2007).
276:
277: %In the past decade,
278: Thanks to these successful developments in
279: weak lensing techniques as well as in instrument technology,
280: the quadrupole weak lensing has become
281: one of the most important tools in observational cosmology
282: to map the mass distribution
283: in individual clusters of galaxies
284: %cosmic structures of galaxy-cluster sized halos
285: (e.g.,
286: Kaiser \& Squires 1993;
287: Broadhurst et al. 2005a;
288: Okabe \& Umetsu 2007;
289: Umetsu \& Broadhurst 2007),
290: measure ensemble-averaged mass profiles of galaxy-group sized halos
291: from the galaxy-galaxy lensing signal (e.g., Hoekstra et al. 2001; Hoekstra et
292: al. 2004; Parker et al. 2005; Mandelbaum et al. 2006),
293: study the statistical properties of the
294: large scale structure of the universe from the cosmic shear statistics
295: (e.g., Bacon et al. 2000;
296: van Waerbeke et al. 2001; Hamana et al. 2003), and search for
297: galaxy clusters by their mass properties
298: (e.g., Schneider 1996; Erben et
299: al. 2000; Umetsu \& Futamase 2000; Wittman et al. 2001).
300:
301:
302: In recent years, there have been theoretical efforts to include
303: the next higher order distortion effects
304: as well as the usual quadrupole
305: distortion effect in the weak lensing analysis
306: (Goldberg \& Natarajan 2002;
307: Goldberg \& Bacon 2005;
308: Bacon et al. 2006;
309: Irwin \& Shmakova 2006;
310: Goldberg \& Leonard 2007;
311: Okura, Umetsu, \& Futamase 2007).
312: %%%
313: We have proposed in Okura, Umetsu, \& Futamase (2007, hereafter OUF)
314: to use
315: certain convenient combinations of octopole/higher multipole moments of
316: background images which we call the Higher Order Lensing Image's
317: Characteristics (HOLICs), and have shown that
318: HOLICs serve as a direct measure for the next higher-order weak lensing
319: effect, or the gravitational flexion (Goldberg \& Bacon 2005)
320: and that the use of HOLICs in addition to the quadrupole shape
321: distortions can improve the accuracy and resolution of weak lensing
322: mass reconstructions based on simulated observations.
323: %Okura, Umetsu, \& Futamase (2007) have shown
324: %We have shown in Okura, Umetsu, \& Futamase (2007, hereafter OUF07)
325: %that there is a natural
326: %relation between flexion and certain combinations of octopole/higher
327: %multipole moments of background images that we call the higher order
328: %lensing image's characteristics (HOLICs).
329: % and developed a HOLICs
330: %approach
331: %to measure weak lensing flexion from weak lensing observations of galaxy
332: %shapes. OUF07 have
333:
334:
335: Recently, Goldberg \& Leonard (2007)
336: %have made an attempt to
337: extended the HOLICs approach for flexion measurements to
338: %take into account the correction for the isotropic PSF effect
339: include observational effects, namely
340: the Gaussian weighting in
341: shape-moment calculations (see the appendix therein) and
342: the isotropic PSF effect,
343: %the isotropic PSF correction in flexion measurements
344: under the assumption that PSF is nearly circular,
345: and tested their extended HOLICs approach with simulated and HST/ACS
346: observations. Leonard et al. (2007) have applied the extended HOLICs
347: method to reconstruct the projected mass distribution in the central
348: region of the
349: massive galaxy cluster A1689 at $z=0.183$, and revealed
350: substructures associated with small clumps of galaxies.
351: Further
352: Leonard et al. (2007) found that in dense systems such as galaxy clusters
353: the HOLICs technique is robust and less sensitive than the Shapelet technique
354: to contamination by light
355: from the extended wings of lens/foreground galaxies.
356:
357:
358: In the present paper we develop a method for measuring flexion
359: by the HOLICs approach
360: by fully extending the KSB formalism;
361: We take into account explicitly the effects of
362: Gaussian weighting in calculation of noisy shape moments and
363: higher-order PSF anisotropy as well as isotropic PSF
364: smearing.
365: We then apply our method to actual, ground-based Subaru observations
366: of A1689, and perform a mass reconstruction in the central region of
367: A1689.
368:
369: The paper is organized as follows. We first summarize in \S 2 the basis
370: of weak gravitational lensing and the flexion formalism.
371: In \S 3, we derive the relationship between HOLICs and flexion
372: by incorporating Gaussian smoothing in shape measurements
373: in the presence of isotropic and anisotropic PSF.
374: The practical method to correct the isotropic/anisotropic PSF effects
375: will be presented in \S 4.
376: In Section 5 we use simulations to test our flexion analysis method
377: based on the HOLICs moment approach.
378: We then perform
379: a weak lensing flexion
380: analysis of A1689 by our fully-extended HOLICs approach, and perform a
381: mass reconstruction of A1689 from the HOLICs estimates of flexion.
382: Finally summary and discussions are given in \S 6.
383: %%%
384: %%% @@ edited by KU (2008/2/17)
385: We refer interested readers
386: to a complete appendix\footnote
387: {Full appendix is available in electronic form at
388: http://www.asiaa.sinica.edu.tw/keiichi/OUF2/appendix.pdf.}
389: %http://www.asiaa.sinica.edu.tw/keiichi/pub_files/OUF2/appendix.pdf.}
390: for details of the derivation of flexion-observable relationships
391: in practical observations.
392:
393:
394: \section{Basis of Weak Lensing and Flexion}
395:
396: %% a complete list of all centers and
397: %% participants can be found in the Online Appendix
398: %% (www.circulationaha.org). ...
399: %%Complete Appendix A is only available in electronic form at.
400:
401: In this section,
402: we summarize general aspects of weak gravitational lensing and
403: flexion formalism, following the complex derivative notation
404: developed by Bacon et al. (2006).
405: A general review of quadrupole weak lensing can be found in
406: Bartelmann \& Schneider (2001).
407: %%%%%
408:
409:
410: \subsection{Spin Properties}
411:
412: We define the spin for weak-lensing quantities in the following way:
413: A quantity is said to have spin $N$ if it has the same value after
414: rotation by $2\pi/N$. The product of spin-$A$ and spin-$B$ quantities
415: has spin $(A+B)$, and the product of spin-$A$ and spin-$B^*$ quantities
416: has spin $(A-B)$. Then, as we shall see in the next subsection,
417: the lensing convergence $\kappa$ is a spin-0 (scalar) quantity.
418: The complex shear $\gamma$ and the reduced shear $g=\gamma/(1-\kappa)$
419: are spin-2 quantities. The first and second flexion fields, $F$ and $G$,
420: are a spin-1 and a spin-3 quantity, respectively.
421:
422:
423: \subsection{Weak Lensing and Flexion Formalism}
424:
425: The gravitational deflection of light rays can be described by the
426: lens equation,
427: \begin{equation}
428: \label{eq:lenseq}
429: \bbeta = \btheta - \bnabla \psi(\btheta),
430: \end{equation}
431: where $\psi(\btheta)$ is the effective lensing potential, which
432: is defined by the two-dimensional Poisson equation as
433: $\nabla^2\psi(\btheta)=2 \kappa(\btheta)$, with the lensing convergence.
434: Here the convergence $\kappa=\int\! d\Sigma_m \Sigma_{\rm crit}^{-1}$ is the
435: dimensionless surface mass density projected on the sky, normalized with
436: respect to the critical surface mass density of gravitational lensing,
437: \begin{equation}
438: \label{eq:sigma_crit}
439: \Sigma_{\rm crit} = \frac{c^2}{4\pi G}\frac{D_s}{D_d D_{ds}},
440: \end{equation}
441: where $D_d$, $D_s$, and $D_{ds}$ are the angular diameter distances
442: from the observer to the deflector, from the observer to the source,
443: and from the deflector to the source, respectively.
444: By introducing the complex gradient operator, $\partial = \partial_1 +
445: i\partial_2$ that transforms as a vector,
446: $\partial'=\partial e^{i\phi}$, with $\phi$ being the angle of rotation,
447: the lensing convergence $\kappa$ is expressed as
448: \begin{equation}
449: \label{eq:kappa}
450: \kappa = \frac{1}{2}\partial\partial^* \psi,
451: \end{equation}
452: where $^*$ denotes the complex conjugate.
453: Similarly, the complex gravitational shear of spin-2 is defined as
454: \begin{equation}
455: \gamma \equiv \gamma_1+i\gamma_2= \frac{1}{2}\partial\partial \psi.
456: \end{equation}
457: The third-order derivatives of $\psi(\btheta)$ can be combined to
458: from a pair of the complex flexion fields as
459: (Bacon et al. 2006):
460: \begin{eqnarray}
461: \label{eq:lpd3rd}
462: \cF &\equiv& \cF_1+i\cF_2 = \frac{1}{2}\partial\partial\partial^*\psi,\\
463: \cG &\equiv& \cG_1+i\cG_2 = \frac{1}{2}\partial\partial\partial\psi.
464: \end{eqnarray}
465: If the angular size of an image is small compared to the
466: scale over which the
467: lens potential $\psi$ varies, then we can locally expand the lens
468: equation (\ref{eq:lenseq}) to have:
469: \begin{eqnarray}
470: \label{eq:dbetaij}
471: d\beta_i = \cA_{ij}d\theta_j + \frac{1}{2}\cD_{ijk}d\theta_j d\theta_k
472: \end{eqnarray}
473: to the second order, where $\cA_{ij}$ is the Jacobian matrix of the lens
474: equation and $\cD_{ijk}=\cA_{ij,k}=-\psi_{,ijk}$ is the third-order
475: lensing tensor,
476: \begin{eqnarray}
477: \cA_{ij}&=&
478: \left(
479: \begin{array}{cc}
480: 1-{\kappa}-{\gamma}_1 & -{\gamma}_2 \\
481: -{\gamma}_2 & 1-{\kappa}+{\gamma}_1
482: \end{array}
483: \right),\\
484: \cD_{ijk}&=&\cF_{ijk}+\cG_{ijk}.
485: \end{eqnarray}
486: %%%%
487: The third-order tensor $\cD_{ijk}$ can be expressed with the sum of the
488: two terms, $\cD_{ijk}=\cF_{ijk}+\cG_{ijk}$, with the spin-1 part
489: $\cF_{ijk}$ and the spin-3 part $\cG_{ijk}$, composed of the
490: real/imaginary part of the flexion fields:
491: \begin{eqnarray}
492: {\cal F}_{ij1} = -\frac{1}{2}\left(
493: \begin{array}{@{\,}cc@{\,}}
494: 3{\cal F}_1 & {\cal F}_2 \\
495: {\cal F}_2 & {\cal F}_1
496: \end{array}
497: \right) &,& \ \
498: {\cal F}_{ij2} = -\frac{1}{2}\left(
499: \begin{array}{@{\,}cc@{\,}}
500: {\cal F}_2 & {\cal F}_1 \\
501: {\cal F}_1 & 3{\cal F}_2
502: \end{array}
503: \right), \\
504: {\cal G}_{ij1} = -\frac{1}{2}\left(
505: \begin{array}{@{\,}cc@{\,}}
506: {\cal G}_1 & {\cal G}_2 \\
507: {\cal G}_2 & -{\cal G}_1
508: \end{array}
509: \right)&,& \ \
510: %\\
511: {\cal G}_{ij2} = -\frac{1}{2}\left(
512: \begin{array}{@{\,}cc@{\,}}
513: {\cal G}_2 & -{\cal G}_1 \\
514: -{\cal G}_1 & -{\cal G}_2
515: \end{array}
516: \right).
517: \end{eqnarray}
518: Note that flexion has a dimension of inverse length (or inverse angle),
519: meaning that the flexion effect depends on the angular size of the
520: source.
521: The shape quantities affected by the first flexion $\cF$ alone have
522: spin-1 properties, while those affected by the second flexion $\cG$
523: alone have spin-3 properties. These third-order lensing fields naturally
524: appear in the transformation equations of HOLICs between the lens and
525: source planes.
526:
527:
528: \subsection{Quadrupole Lensing Observable -- Complex Ellipticity}
529:
530: %In KSB methods, We use quadrupole moment defined as
531:
532: In the KSB approach, we use quadrupole moments $Q_{ij}$
533: of the surface brightness distribution $I(\btheta)$ of background images
534: for quantifying the shape of the images:
535: \begin{equation}
536: \label{eq:Qij}
537: Q_{ij} \equiv \frac{\int\! d^2\theta\,
538: q_I[I(\btheta)]\dt_i \dt_j}{\int d^2\theta\,q_I[I(\btheta)]},
539: \end{equation}
540: where $q_I[I(\btheta)]$ denotes the weight function used in
541: noisy shape measurements and $\Delta\theta_i = \theta_i-\bar{\theta}_i$
542: is the offset vector from the image centroid.
543: The complex ellipticity $\chi$ is then defined as
544: \begin{equation}
545: \label{eq:cellip}
546: \chi \equiv \frac{Q_{11} - Q_{22} + 2iQ_{12}}{Q_{11} + Q_{22}},
547: \end{equation}
548: The $\chi$ transforms under the lens mapping as
549: \begin{equation}
550: \label{eq:chis2chi}
551: \chi^{(s)}=\frac{\chi-2g+g^2\chi^\ast}{1+|g|^2-2\Real{g\chi^\ast}},
552: \end{equation}
553: where $g = \gamma/(1 - \kappa)$ is the spin-2 reduced shear.
554: In the weak lensing limit ($\kappa, |\gamma|\ll 1$),
555: equation (\ref{eq:chis2chi}) reduces to
556: $\chi^{(s)} \approx \chi-2\gamma$.
557: Assuming the random orientation of the background sources,
558: we average observed ellipticities over a sufficient number of
559: images to obtain
560: \begin{equation}
561: \label{eq:chis2chiap}
562: \langle\chi\rangle\approx 2g \approx 2\gamma.
563: \end{equation}
564:
565: \subsection{Centroid Shift due to Lensing}
566: \label{subsec:csfgl}
567: %due to the First Flexion Effect}
568:
569: In the moment methods such as KSB, we quantify the shape of
570: an image by measuring various moments of $I(\btheta)$, in which
571: the moments are calculated with respect to the observable
572: centroid of the image $\bar{\btheta}$,
573: or the center of the light, defined by the
574: first moment of $I(\btheta)$.
575: However, in the presence of gravitational lensing, this apparent center
576: can be different from the point $\btheta(\bar{\bbeta})$
577: that is mapped using the lens
578: equation (\ref{eq:lenseq})
579: from the center of the unlensed light, $\bar{\bbeta}$.
580: %%%
581: We refer to this point
582: as the ``true'' center of the image. The difference between these two
583: centers, namely, the apparent and the true centers, causes a significant
584: effect in evaluating the first flexion, as pointed out by Goldberg \&
585: Bacon (2005). The relationship between the two centers is
586: \begin{equation}
587: \label{eq:csft}
588: \theta_i(\bar{\bbeta})
589: \approx \bar{\theta}_i - \trQ
590: \paren{\frac{3}{2} F_i + \frac{5}{4}[F^*\chi]_i +
591: \frac{1}{4}[G\chi^*]_i}
592: \equiv \bar{\theta}_i - \Delta_{L,i},
593: \end{equation}
594: where $\trQ=Q_{11}+Q_{22}$ is the trace of $Q_{ij}$,
595: $F$ and $G$ are the
596: reduce Flexion, defined by
597: $F=\cF/(1-\kappa)$ and $G=\cG/(1-\kappa)$, respectively,
598: and $\Delta_{L,i}$ is the displacement vector from the true to the
599: apparent center due to gravitational lensing.
600: %%%
601: Therefore, by taking into account the centroid shift in shape
602: measurements,
603: we have the relation between $\Delta\beta_i=\beta_i-\bar{\beta}_i$ and
604: $\Delta\theta_i = \theta_i -\bar{\theta}_i$ as
605: \begin{eqnarray}
606: \Delta\beta_i &\approx&
607: \cA_{ij}\Delta\theta_j
608: + \frac{1}{2}\cD_{ijk} \Delta\theta_j \Delta\theta_k
609: + \trQ \paren{\frac{3}{2} F_i + \frac{5}{4}[F^*\chi]_i
610: + \frac{1}{4}[G\chi^*]_i}\nonumber\\
611: &=&
612: \cA_{ij}\Delta{\theta}_j + \frac{1}{2}\cD_{ijk}\Delta\theta_j
613: \Delta\theta_k
614: + \Delta_{L,i}.
615: \end{eqnarray}
616: A detailed derivation of the above relation is given in
617: Appendix B of OUF.
618: %See the derivation in more detail given in Appendix B of OUF.
619:
620:
621: \subsection{Flexion Observable -- HOLICs}
622:
623: The flexion fields, $\cF$ and $\cG$, can be measured from
624: proper combinations of higher-order shape moments with the
625: corresponding spin properties and the dimension,
626: as explicitly shown by OUF.
627: Higher-order moments of images are defined as a straightforward
628: extension of the quadrupole moment.
629: %%%
630: The octopole moment $Q_{ijk}$
631: and the 16pole moment $Q_{ijkl}$ are defined as follows:
632: \begin{eqnarray}
633: \label{eq:Qijk}
634: Q_{ijk}&\equiv& \frac{\int\! d^2\theta\, q_I[I(\btheta)]\dt_i \dt_j \dt_k}{\int d^2\theta\,q_I[I(\btheta)]} \\
635: Q_{ijkl}&\equiv& \frac{\int\! d^2\theta\, q_I[I(\btheta)]\dt_i \dt_j \dt_k \dt_l}{\int d^2\theta\,q_I[I(\btheta)]}.
636: \end{eqnarray}
637: Then, $\zeta$ and $\delta$ of the spin-1 and spin-3 HOLICs,
638: respectively,
639: are defined by
640: \begin{eqnarray}
641: \label{eq:HOLICs}
642: \zeta&\equiv&\frac{Q_{111}+Q_{122}+i\left(Q_{112}+Q_{222}\right)}{\xi}\nonumber\\
643: \delta&\equiv&\frac{Q_{111}-3Q_{122}+i\left(3Q_{112}-Q_{222}\right)}{\xi},
644: \end{eqnarray}
645: where $\xi$ is the spin-0 normalization factor,
646: \begin{equation}
647: \xi = Q_{1111}+2Q_{1122}+Q_{2222}.
648: \end{equation}
649:
650: Finally, the transformation equations between unlensed and lensed HOLICs
651: are obtained as (see OUF)
652: \begin{eqnarray}
653: \label{eq:zeta}
654: \zeta^{(s)}
655: &=&
656: \frac{\zeta-2g\zeta^*-g^*\delta -\frac{1}{4}(8F^*\eta - 16\frac{\paren{\trQ}^2}{\xi}F^*\chi +9F - 12\frac{\paren{\trQ}^2}{\xi}F +2G\eta^* - 2\frac{\paren{\trQ}^2}{\xi}G\chi^*+G^*\lambda)}{(1-\kappa)(1-4\Real{g^*\eta}-5\Real{F\iota_I^*}-\Real{G\iotaIII^*})},\nonumber\\
657: \label{eq:delta}
658: \delta^{(s)}
659: &=
660: &\frac{\delta-3g\zeta -\frac{1}{4}(10F\eta+7F^*\lambda - 18\frac{\paren{\trQ}^2}{\xi}F\chi +3G)}{(1-\kappa)(1-4\Real{g^*\eta}-5\Real{F\iota_I^*}-\Real{G\iotaIII^*})},
661: \end{eqnarray}
662: where $\eta$ and $\lambda$ are dimensionless spin-2 and spin-4
663: quantities, respectively,
664: defined with 16-pole moments,
665: and $\iota_{I}$, $\iota_{\gIII}$ and $\iota_{V}$ are
666: spin-1, spin-3, and spin-5 quantities, respectively,
667: %%% @@ edited by KU (2008/02/13)
668: %defined with $32$-pole moments (see Appendix \ref{sec:basis} and OUF).
669: defined with $32$-pole moments (see Appendix and OUF).
670: We note that, the above equations (\ref{eq:zeta}) and (\ref{eq:delta})
671: are obtained under the sub-critical lensing condition, i.e., ${\rm
672: det}{\cal A}>0$ (see Schneider \& Er 2007).
673:
674:
675: Since the HOLICs $\zeta$ and $\delta$ are non-zero spin quantities
676: with a direction dependence, the expectation value of the intrinsic
677: $\zeta$ and $\delta$ are assumed to vanish. To the first order in
678: flexion,
679: we have the linear relations between the HOLICs and flexion fields as
680: \begin{eqnarray}
681: \label{eq:HOLICFL}
682: F &\approx&
683: \left<\frac{\zeta}{\frac{9}{4} - 3\frac{\paren{\trQ}^2}{\xi}}\right>\\
684: G &\approx& \frac{4}{3}\left<\delta\right>.
685: \end{eqnarray}
686:
687:
688: \section{HOLICs and Flexion in Practical Applications}
689: \label{sec:ksb+}
690:
691: For a practical application of the HOLICs approach,
692: we must take into account various observational effects
693: such as noise in the shape measurement due to readout
694: and/or sky background
695: and the dilution of the lensing signal due to the isotropic/anisotropic
696: PSF effects.
697: Thus, one cannot simply use equations
698: (\ref{eq:zeta}) and (\ref{eq:delta})
699: to measure the flexion fields.
700: In this section, we introduce the Gaussian weighting in moment
701: calculations, as done in the KSB formalism for quadrupole weak lensing,
702: and derive the relevant
703: transformation equations between unlensed and lensed HOLICs
704: by taking into account
705: explicitly the effect of Gaussian smoothing.
706: %%For details of the derivation,
707: %%%%
708:
709:
710: \subsection{Redefining HOLICs for Noisy Observations}
711:
712: Now we introduce a weight function $W(|\dt|^2/\sigma^2)$
713: having a characteristic width $\sigma$ for practical, noisy
714: moment measurements,
715: redefining the octopole and 16-pole moments of the
716: brightness distribution $I(\btheta)$
717: as
718: \begin{eqnarray}
719: \label{octW}
720: Q_{ijk}&\equiv&
721: \int d^2\theta I(\theta)\dt_i\dt_j\dt_kW(|\dt|^2/\sigma^2),\\
722: Q_{ijkl}&\equiv&
723: \int d^2\theta I(\theta)\dt_i\dt_j\dt_k\dt_lW(|\dt|^2/\sigma^2),
724: \end{eqnarray}
725: where $Q$s here are no longer normalized with the corresponding {\it flux}
726: of the image (see equation [\ref{eq:Qij}]).
727: The redefined shape moments enter equation (\ref{eq:HOLICs}).
728: %% @@ edited by KU (2008/02/13)
729: We provide in Appendix
730: %%We provide in Appendix \ref{sec:basis}
731: %%We provide in Appendix
732: detailed definitions of HOLICs in practical applications.
733:
734:
735: \subsection{Lensing-Induced Centroid Shift in Weighted Moment Calculations}
736:
737: When weighted moments are used for calculating HOLICs,
738: the centroid shift due to lensing (\ref{eq:csft}), to the first order,
739: is changed in the following way:
740: \begin{eqnarray}
741: \label{SCL}
742: \Delta_{L} = \bar{\theta} - \theta(\bar{\beta})
743: \approx
744: \frac{\frac{3\trQ^a}{2M} + \frac{3{\xi^a}'}{4M\sigma^2}}{1+\frac{{\trQ^a}'}
745: {M\sigma^2}}F \equiv \Delta^0_L F,
746: \end{eqnarray}
747: where
748: $\bar{\theta} = \bar{\theta}_1+i\bar{\theta}_2$
749: and $\theta(\bar\beta) = \theta_1(\bar\beta)+i\theta_2(\bar\beta)$
750: are the apparent and the true centers of the image
751: (see \S \ref{subsec:csfgl}), respectively,
752: in the complex form calculated using the weight function $W(x)$,
753: $M\equiv \int\!d^2\theta\,I(\theta)W(|\Delta\theta|^2/\sigma^2)$
754: is the monopole shape moment (or {\it flux}),
755: quantities with subscript ``a'' represent those calculated
756: with respect to the apparent center,
757: and quantities with prime represent those measured with
758: $W'(x)=\partial W(x)/\partial x$ as the weight function
759: instead of $W(x)$; $\Delta_L^0$ is the spin-0 coefficient in $\Delta_L$.
760: %%%%
761: The deviation from unity in
762: the denominator of $\Delta_L^0$ is obtained by properly expanding
763: the weight function W(x) in moment calculations, and this term does not
764: appear in Goldberg \& Leonard (2006)'s formulation.
765: %%%%
766: Hence,
767: the complex displacement from the true image center, $\dt^t$,
768: can be expressed in terms of that from the apparent image center,
769: $\dt^a$, and the complex centroid shift, $\Delta_L$, as
770: \begin{eqnarray}
771: \label{eq:csf_complex}
772: \dt^t = \dt^a + \Delta_L.
773: \end{eqnarray}
774: %% @@ edited by KU (2008/02/13)
775: %%We show in Appendix \ref{sec:CSL} detailed calculations of the
776: %%We show in Full Appendix B.1 detailed calculations of the
777: For interested readers, we refer to Full Appendix B.1 for detailed
778: calculations of the
779: lensing-induced centroid shift with a weight function.
780:
781: \subsection{Relation between the Weighted HOLICs and Flexion}
782:
783: %In practice, all of the shape moments are calculated with respect to
784: %the weighted first moment
785:
786: In weighted moment calculations, the transformation equations
787: between HOLICs and flexion must be modified accordingly.
788: To the first order, we have
789: \begin{eqnarray}
790: \label{eq:Lbt}
791: \zeta^{(s)}
792: &\approx&
793: \frac{1}{1-\kappa}\left(\zeta^t-\frac{9}{4}F-\frac{3{\upsilon^t_0}'}
794: {4\xi^t\sigma^2}F\right),\\
795: %%%
796: \delta^{(s)}&\approx&\frac{1}{1-\kappa}\left(\delta^t-
797: \frac{3}{4}G-\frac{{\upsilon^t_0}'}{4\xi^t\sigma^2}G\right),\\
798: \upsilon_0&\equiv&Q_{111111}+3Q_{112222}+3Q_{111122}+Q_{222222},
799: \end{eqnarray}
800: where quantities with subscript ``t'' represent those calculated
801: with respect to the true center, or using $\Delta\theta^t$.
802: By using equation (\ref{eq:csf_complex}),
803: we can express $\zeta^t$ and $\delta^t$ in terms of
804: practically observable quantities
805: (with subscript ``a'')
806: as
807: \begin{eqnarray}
808: \label{eq:zeta_t}
809: \zeta^t &\approx&
810: \zeta^a + 2\frac{\trQ^a}{\xi^a}\Delta_L+\frac{{\xi^a}'}{\xi^a\sigma^2}\Delta_L
811: = \zeta^a + \left(2\frac{\trQ^a}{\xi^a}\Delta^0_L+\frac{{\xi^a}'}
812: {\xi^a \sigma^2}\Delta^0_L\right)F,\\
813: %%%%
814: \label{eq:delta_t}
815: \delta^t&\approx&\delta^a
816: \end{eqnarray}
817: to the first order.
818: Here the term $\frac{{\xi^a}'}{\xi^a\sigma^2}\Delta_L$
819: in equation (\ref{eq:zeta_t}) is again caused by the centroid shift
820: in the weight function
821: (a similar, but different, expression was
822: obtained by Goldberg \& Leonard 2007).
823: %and is not taken into account in
824: %Goldberg \& Leonard (2006).
825: Finally, we obtain the following transformation equations
826: in the case of weighted moment calculations:
827: \begin{eqnarray}
828: \label{eq:Lbf}
829: \zeta^{(s)}&
830: \approx&
831: \frac{1}{1-\kappa}
832: \left[
833: \zeta^a -
834: \left(\frac{9}{4}+\frac{3{\upsilon^a_0}'}{4\xi^a\sigma^2}-
835: 2\frac{\trQ^a}{\xi^a}\Delta^0_L-\frac{{\xi^a}'}{\xi^a\sigma^2}\Delta^0_L
836: \right)
837: F
838: \right],\\
839: %%%
840: \delta^{(s)}
841: &\approx&
842: \frac{1}{1-\kappa}
843: \left[
844: \delta^a-\left(\frac{3}{4}+\frac{{\upsilon^a_0}'}{4\xi^a\sigma^2}\right)G
845: \right]
846: \end{eqnarray}
847: %% @@ edited by KU (2008/02/13)
848: %%%We show the details of these calculations in Appendix \ref{sec:TRNS}.
849: We show the details of these calculations in Full Appendix B.3.
850:
851:
852:
853: \section{Isotropic and Anisotropic PSF Corrections for HOLICs Measurements}
854:
855: In this section we present a detailed prescription for
856: the PSF anisotropy and circularization correction in HOLICs-based
857: flexion measurements by extending the KSB formalism.
858: We closely follow the treatment and the notation
859: given in \S 4.6.1 of Bartelmann \& Schneider (2001).
860:
861:
862: \subsection{General Description for PSF}
863:
864:
865: The observed surface brightness distribution $I^{\rm obs}(\btheta)$
866: can be expressed as the true surface brightness $I(\btheta)$
867: convolved with an effective PSF $P(\btheta)$,
868: \begin{eqnarray}
869: \label{eq:PSF}
870: I^{\rm obs}(\btheta)
871: = \int\! d^2\vartheta\,
872: I(\bvtheta)P(\btheta-\bvtheta).
873: \end{eqnarray}
874: Following the KSB formalism,
875: we assume that $P$ is nearly isotropic, so that the
876: anisotropic part of $P$ is small. We then define the
877: isotropic part of $P$ as the azimuthal
878: average over $P$, and decompose $P$
879: into an isotropic part, $P^{iso}$,
880: and an anisotropic part, $q$,
881: as
882: \begin{eqnarray}
883: \label{eq:PSFP}
884: P(\bvtheta)
885: = \int\! d^2\phi\, q(\bphi) P^{iso}(\bvtheta - \bphi),
886: \end{eqnarray}
887: where both $P^{iso}$ and $q$ are normalized to unity.
888: We then define $I^{iso}(\btheta)$ as the surface brightness distribution
889: smeared by the isotropic part $P^{iso}$,
890: \begin{eqnarray}
891: \label{eq:PSFiso}
892: I^{iso}(\btheta)
893: = \int\! d^2\vartheta\, I(\btheta)P^{iso}(\btheta-\bvtheta).
894: \end{eqnarray}
895: The observed surface brightness $I^{obs}(\btheta)$ is obtained by
896: convolving $I^{iso}(\btheta)$ with the anisotropy kernel $q(\btheta)$
897: as
898: \begin{eqnarray}
899: \label{eq:PSFaniso}
900: I^{obs}(\btheta) = \int\! d^2\vartheta\,
901: q(\btheta - \bvtheta)I^{iso}(\bvtheta).
902: \end{eqnarray}
903:
904: In the original KSB method for quadrupole weak lensing,
905: only the spin-2 PSF anisotropy described by
906: the quadrupole moment of the anisotropy kernel $q$
907: has to be taken into account.
908: However,
909: in order to correct HOLICs with spin-1 and spin-3 properties
910: for the anisotropic PSF effects,
911: we need to take into account the corresponding
912: dipole and octopole moments of the anisotropy kernel $q$
913: having spin-1 and spin-3 properties.
914: %as well as the quadrupole anisotropy.
915: %%%%%
916: We expand the integral of arbitrary function $f(\btheta)$
917: and $I^{obs}(\btheta)$ to obtain
918: \begin{eqnarray}
919: \label{eq:psfexp}
920: &&\int\! d^2\theta\,
921: f(\btheta) I^{obs}(\btheta) =
922: \int\! d^2\phi\, I^{iso}(\bphi)
923: \int\! d^2\theta\, f(\btheta +\bphi) q(\btheta)
924: \approx \int\! d^2\phi\,
925: I^{iso}(\bphi) f(\bphi)\nonumber\\
926: %%%
927: &&\hspace{1cm}
928: + q_{k}\int\! d^2\phi\,
929: I^{iso}(\bphi)
930: \frac{\partial f}{\partial\phi_k}
931: + \frac{1}{2}q_{kl}
932: \int\! d^2\phi\,
933: I^{iso}(\bphi) \frac{\partial^2f}{\partial\phi_k\partial\phi_l}\nonumber\\
934: &&\hspace{1cm}
935: +
936: \frac{1}{6} q_{klm}
937: \int\! d^2\phi\,
938: I^{iso}(\bphi)\frac{\partial^3f}{\partial\phi_k\partial\phi_l\partial\phi_m}
939: + .........,
940: \end{eqnarray}
941: where
942: \begin{eqnarray}
943: \label{eq:qijk}
944: &q_{i} = \int\! d^2\phi\, q(\bphi)\phi_i,\\
945: &q_{ij} = \int\! d^2\phi\, q(\bphi)\phi_i\phi_j,\\
946: &q_{ijk} = \int\! d^2\phi\, q(\bphi)\phi_i\phi_j\phi_k
947: \end{eqnarray}
948: are the dipole, quadrupole, and octopole moments of the
949: PSF anisotropy kernel $q$, respectively.
950:
951:
952:
953:
954: \subsection{Centroid Shift due to PSF Anisotropy}
955:
956: The PSF anisotropy can cause a centroid shift between the
957: images defined in terms of
958: $I^{iso}(\btheta)$ and $I^{obs}(\btheta)$.
959: This centroid shift is essentially due to the
960: spin-1 PSF anisotropy.
961:
962:
963: %Center of images we observed and images smeared by isotropic PSF are
964: %obtained as
965: The observed weighted-center of image can be expressed
966: in the complex form as
967: \begin{equation}
968: \label{eq:Cb_01}
969: \bar \theta^{obs} = \frac{\int\! d^2\theta\, \theta I^{obs}(\theta)
970: W(|\theta|^2/\sigma^2)}
971: {\int\! d^2\theta\, I^{obs}(\theta)W(|\theta|^2/\sigma^2)},
972: \end{equation}
973: where $\theta=\theta_1+i\theta_2$ is the complex angular position.
974: %in the image plane.
975: Similarly,
976: the weighted-center of a hypothetical image defined in terms of
977: $I^{iso}$ is
978: \begin{equation}
979: \label{eq:Cb_02}
980: \bar \theta^{iso}=\frac{\int\! d^2\theta\,\theta I^{iso}(\theta)
981: W(|\theta|^2/\sigma^2)}{\int\! d^2\theta\, I^{iso}(\theta)
982: W(|\theta|^2/\sigma^2)}.
983: \end{equation}
984: Then, the offset between the two centers $\bar{\theta}^{obs}$ and
985: $\bar{\theta}^{iso}$ is given as
986: \begin{eqnarray}
987: \Delta_P \equiv\bar \theta^{obs} - \bar\theta^{iso}
988: = \dt^{iso}-\dt^{obs},
989: \end{eqnarray}
990: where $\dt^{iso} \equiv \theta-\bar{\theta}^{iso}$ and
991: $\dt^{obs} \equiv \theta-\bar{\theta}^{obs}$ are the complex displacements
992: to an arbitrary point $\theta$.
993: %%%%
994: %The PSF-induced centroid shift is obtained as
995: Expanding $\Delta_P$ to the first order of $q$ yields
996: \begin{eqnarray}
997: \Delta_P
998: = \Delta\theta^{iso} - \Delta\theta^{obs}
999: \approx D_q + \frac{\frac{M}{M'\sigma^2}
1000: + \frac{2\trQ''}{M\sigma^4}
1001: + \frac{\xi'''}{2\sigma^6 M}}{1+\frac{\trQ'}{M\sigma^2}}\zeta_q
1002: \equiv D_q + \frac{P^0_D}{P^\Delta_D}\zeta_q,
1003: \end{eqnarray}
1004: where
1005: quantities with $\prime\prime$ and $\prime\prime\prime$
1006: refer to those calculated with
1007: $W''(x)$ and $W^{\prime\prime\prime}(x)$
1008: as the weight function, respectively,
1009: %%%%
1010: $(D_q,\zeta_q)$ are
1011: spin-1 complex moments of the anisotropy kernel defined by
1012: \begin{eqnarray}
1013: D_q &=& \int d^2\phi q(\phi) \phi,\\
1014: \zeta_q &=& \int d^2\phi q(\phi) \phi\phi\phi^*,\\
1015: \phi&=&\phi_1+i\phi_2,
1016: \end{eqnarray}
1017: and $(P^0_D,P^\Delta_D)$ are
1018: certain combinations of
1019: shape moments defined as the coefficients
1020: %% @@ edited by KU (2008/02/13)
1021: %%in equation (\ref{eq:Q_obs}) of Appendix \ref{sec:CSP},
1022: in equation (C11) of Full Appendix C.1.2,
1023: associated with the complex dipole moment $D$ with spin-1 properties.
1024: %%%%
1025: We note that $\Delta_P$, the PSF-induced centroid shift, cannot be
1026: constrained from observations; however, as we shall see in the next
1027: subsection, one can constrain the spin-1 octopole component in
1028: $\Delta_P$, namely $\zeta_q \propto \Delta_P-D_q$,
1029: in the first order approximation.
1030: %%%% @@ edited by KU (2008/02/13)
1031: %%More detailed calculations are presented in Appendix \ref{sec:CSP}.
1032: More detailed calculations are presented in Full Appendix C.1.2.
1033:
1034:
1035: \subsection{Anisotropic PSF Correction for HOLICs}
1036:
1037: We use equation (\ref{eq:psfexp}) to relate
1038: observable HOLICs (with subscript ``obs'')
1039: to those defined in terms of $I^{iso}$ (with subscript ``iso'').
1040: To the first order of $q$, we have the following relations
1041: for the spin-1 and spin-3 quantities:
1042: \begin{eqnarray}
1043: \label{eq:HOLICisoz}
1044: %%%
1045: %%%
1046: D^{obs}
1047: &\approx&
1048: D^{iso} + \Delta_P
1049: \approx
1050: D^{iso} + D_q + \frac{P^0_D}{P^\Delta_D}\zeta_q,\\
1051: %%%
1052: \zeta^{obs}
1053: &\approx&
1054: \zeta^{iso} + \frac{1}{\xi}\left( 2\trQ
1055: + \frac{\xi'}{\sigma^2} \right)(D_q-\Delta_P)
1056: + \frac{1}{\xi}\left(M+\frac{5\trQ'}{\sigma^2}
1057: + \frac{7\xi''}{2\sigma^4}+\frac{\upsilon'''_0}{2\sigma^6}\right)
1058: \zeta_q\nonumber\\
1059: &=& \zeta^{iso}
1060: + \left( P^0_\zeta - \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta
1061: \right)\zeta_q,\\
1062: \label{eq:HOLICisod}
1063: %%%%
1064: \delta^{obs}
1065: &\approx&
1066: \delta^{iso}
1067: +\frac{1}{\xi}\left(M+\frac{3\trQ'}{\sigma^2}
1068: +\frac{3\xi''}{2\sigma^4}+\frac{\upsilon'''_0}{6\sigma^6}\right)
1069: \delta_q=\delta^{iso}+P^0_{\delta}\delta_q,
1070: \end{eqnarray}
1071: where we have deffined the weighted first moment $D$ by
1072: \begin{equation}
1073: D=\frac{\int\!d^2\theta\, W(|\Delta\theta|^2/\sigma^2)\Delta\theta}{M},
1074: \end{equation}
1075: and the spin-3 PSF anisotropy $\delta_q$ by
1076: \begin{equation}
1077: \delta_q = \int\! d^2\theta\, q(\phi)\phi\phi\phi.
1078: \end{equation}
1079: %%%
1080: In general, the PSF varies spatially over the field.
1081: If the spatial variation of PSF is sufficiently smooth,
1082: then one can measure $\zeta_q$ and $\delta_q$ for a set of stars.
1083: Since $\zeta^{iso}$ and $\delta^{iso}$ vanish for stars,
1084: the spin-1 and spin-3 PSF anisotropies can be obtained as
1085: \begin{eqnarray}
1086: \label{eq:qstar}
1087: \zeta_q&=&\frac{
1088: (\zeta^{obs})_{*}
1089: }
1090: {\left( P^0_\zeta - \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta
1091: \right)_{*}
1092: },\\
1093: %%%%
1094: \delta_q &=&
1095: \frac{
1096: (\delta^{obs})_{*}
1097: }
1098: { (P^0_{\delta})_{*}},
1099: \end{eqnarray}
1100: where quantities with asterisk denote those measured for stellar objects.
1101: Note that unlike the higher-order PSF anisotropies $\zeta_q$
1102: and $\delta_q$, $\Delta_p$ cannot be determined from observations,
1103: so that $D^{obs}$ cannot be corrected for the anisotropic PSF effect.
1104: %% @@ edited by KU (2008/02/13)
1105: %%%We show detailed calculations in Appendix \ref{sec:anisoPSF}.
1106: We show detailed calculations in Full Appendix C.1.3.
1107:
1108:
1109: \subsection{Isotropic PSF Correction for HOLICs}
1110:
1111: This subsection provides the method of correcting for the
1112: isotropic PSF effect on the flexion measurement.
1113: Firstly, from Liouvelle's theorem we have
1114: $I(\btheta)=I^{(s)}(\bbeta(\btheta))$ with $I^{(s)}(\bbeta)$
1115: being the surface brightness distribution of the unlensed source.
1116: We then consider
1117: \begin{eqnarray}
1118: I^{iso}(\btheta) &=& \int\!d^2\phi\,
1119: I^{(s)}(\cA \bphi)P^{iso}(\btheta-\bphi)\nonumber\\
1120: &=& \frac{1}{{\rm det}\cA}\int\!d^2\vartheta\, I^{(s)}(\bvtheta)
1121: P^{iso}(\btheta-\cA^{-1}\bvtheta) \equiv \hat{I}(\cA\btheta),
1122: \end{eqnarray}
1123: where we have defined the brightness distribution $\hat{I}(\bbeta)$
1124: convolved with the hypothetical PSF $\hat{P}(\bbeta)$,
1125: \begin{eqnarray}
1126: \hat{I}(\bbeta) &=& \int\!d^2\phi\, I^{(s)}(\bphi)\hat{P}(\bbeta-\bphi),\\
1127: \hat{P}(\bbeta) &=& \frac{1}{{\rm det}\cA} P^{iso}(\cA^{-1}\bbeta).
1128: \end{eqnarray}
1129: The $\hat{P}$ can be regarded as an effective PSF relating
1130: $\hat{I}$ to
1131: $I^{(s)}$; the anisotropic part of $\hat{P}$ is caused by lensing.
1132: The octopole moment defined with $\hat I(\beta)$ is written as
1133: \begin{eqnarray}
1134: \hat Q_{ijk} =\int\! d^2\beta\,
1135: \Delta\beta_i
1136: \Delta\beta_j
1137: \Delta\beta_k
1138: \hat I(\bbeta)
1139: W\paren{\frac{|\Delta\bbeta|^2}{\hat \sigma^2}}.
1140: \end{eqnarray}
1141: Then, $\hat \zeta$ and $\hat \delta$ of HOLICs are defined
1142: in terms of $\hat{Q}_{ijk}$ and $\hat{Q}_{ijkl}$
1143: as
1144: \begin{eqnarray}
1145: \hat \zeta
1146: &=&\frac{\hat Q_{111}+\hat Q_{122}+i(\hat Q_{112}+\hat Q_{222})}
1147: {\hat Q_{1111}+2\hat Q_{1122}+\hat Q_{2222}},\\
1148: \hat \delta &=&\frac{\hat Q_{111}-3\hat Q_{122}+i(3\hat Q_{112}-\hat
1149: Q_{222})}
1150: {\hat Q_{1111}+2\hat Q_{1122}+\hat Q_{2222}}.
1151: \end{eqnarray}
1152: %%%%%%
1153: Substituting the expression for $\db$
1154: into the above equations and using equation (\ref{eq:Lbf}),
1155: we obtain
1156: \begin{eqnarray}
1157: \label{hatiso}
1158: \hat \zeta
1159: &\approx&
1160: \zeta^{iso}- \left(\frac{9}{4}
1161: +\frac{3{\upsilon_0^{iso}}'}{4\xi^{iso}}
1162: -2\frac{\trQ^{iso}}{\xi^{iso}}\Delta^0_L
1163: -\frac{{\xi^{iso}}'}{\xi^{iso}\sigma^2}\Delta^0_L\right)F
1164: =\zeta^{iso} -
1165: \left(C^0_{\zeta}+C^\Delta_\zeta\Delta^0_L\right) F,\\
1166: %%%%
1167: \hat \delta &\approx&
1168: \delta^{iso}-\left(\frac{3}{4}
1169: +\frac{{\upsilon_0^{iso}}'}{4\xi^{iso}}\right)G
1170: = \delta^{iso} - C^0_\delta G
1171: \end{eqnarray}
1172: to the first order.
1173: Here $C^0$s are certain combinations of
1174: shape moments defined in terms of $I^{iso}$
1175: %% @@ edited by KU (2008/02/13)
1176: %%(see Appendix \ref{sec:isoPSF}).
1177: (see Full Appendix C.2).
1178: %%%%
1179: In practice, however,
1180: one can replace $I^{iso}$ with $I^{obs}$ for calculating $C^0$s
1181: of first order in flexion.
1182:
1183: Next, we
1184: decompose
1185: $\hat P(\bbeta)$ into an isotropic part, $\hat P^{iso}$,
1186: and an anisotropic part, $\hat q$, as
1187: \begin{eqnarray}
1188: \hat P(\bbeta) =
1189: \int\! d^2\phi\, \hat q(\bbeta - \bphi)\hat P^{iso}(\bphi).
1190: \end{eqnarray}
1191: With $\hat P^{iso}$ we define the surface brightness distribution
1192: $\hat I^0$ smeared by the isotropic part $\hat P^{iso}$
1193: as
1194: \begin{eqnarray}
1195: \hat I^0(\bbeta) = \int\! d^2\phi\,
1196: I^{(s)}(\bphi)\hat P^{iso}(\bbeta - \bphi).
1197: \end{eqnarray}
1198: Then, the relationship between $\hat I(\bbeta)$ and $\hat I^0(\bbeta)$
1199: is the same as that between $I^{obs}(\btheta)$
1200: and $I^{iso}(\btheta)$
1201: except that $\hat P$ instead of $P$.
1202: Therefore, we obtain the following relations to the first order:
1203: \begin{eqnarray}
1204: \label{hat0}
1205: \hat \zeta
1206: &\approx&\hat \zeta^0 + \hat P^0_\zeta\zeta_{\hat q},\\
1207: \hat \delta &\approx&\hat \delta^0 + \hat P^0_\delta\delta_{\hat q},
1208: \end{eqnarray}
1209: where $\zeta_{\hat q}$ and $\delta_{\hat q}$ are defined as
1210: \begin{eqnarray}
1211: \zeta_{\hat q} &=& \int\!d^2\theta\, \hat q(\phi)\phi\phi\phi^*,\\
1212: \delta_{\hat q} &=& \int\!d^2\theta\, \hat q(\phi)\phi\phi\phi,
1213: \end{eqnarray}
1214: $\hat P^0_\zeta$ and $\hat P^0_\delta$
1215: are spin-0 coefficients defined by certain combinations of shape moments
1216: %% @@ edited by KU (2008/02/13)
1217: %%(see Appendix \ref{sec:isoPSF}),
1218: (see Full Appendix C.2),
1219: and are calculated using $I^0$ instead
1220: of $I^{iso}$ to the first order approximation.
1221: %%%%
1222: For stellar objects, $\hat\zeta_*^0=\hat\delta_*^0=0$,
1223: so that higher-order PSF anisotropies of $\hat q$ are obtained
1224: using equations (\ref{hatiso}) and
1225: (\ref{hat0}) as
1226: \begin{eqnarray}
1227: \zeta_{\hat q}&=&
1228: -\frac{\left(C^0_{\zeta}+C^\Delta_\zeta\Delta^0_L\right)_{*}}
1229: {(1-\kappa)\left( \hat P^0_\zeta
1230: - \frac{\hat P^0_D}{\hat P^\Delta_D}\hat P^\Delta_\zeta \right)_{*}}F,\\
1231: %%%%%
1232: \delta_{\hat q} &=&
1233: -\frac{ (C^0_{\delta})_* }{(1-\kappa) (\hat P^0_{\delta})_* }G.
1234: \end{eqnarray}
1235: Further, to the first order approximation,
1236: we have
1237: \begin{eqnarray}
1238: \left(
1239: \hat P^0_\zeta - \frac{\hat P^0_D}{\hat P^\Delta_D}\hat P^\Delta_\zeta
1240: \right)
1241: &\approx&
1242: (1-\kappa)^{-3}
1243: \left(
1244: P^0_\zeta - \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta
1245: \right),
1246: \\
1247: \hat P^0_\delta&\approx&\frac{P^0_\delta}{(1-\kappa)^3}
1248: \end{eqnarray}
1249: Finally,
1250: we can relate
1251: unlensed HOLICs in terms of $\hat I^0(\bbeta)$ to
1252: observed HOLICs corrected for the PSF anisotropy
1253: by
1254: \begin{eqnarray}
1255: \label{iso0}
1256: \hat \zeta^0 &\approx&
1257: \frac{1}{1-\kappa}
1258: \left[
1259: \zeta^{iso}- \left(C^0_{\zeta}+C^\Delta_\zeta\Delta^0_L\right) F
1260: +
1261: \frac{\left(C^0_{\zeta}+C^\Delta_\zeta\Delta^0_L\right)_{*}}
1262: {\left(P^0_\zeta
1263: -
1264: \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta\right)_{*}}
1265: \left(P^0_\zeta - \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta\right) F
1266: \right],
1267: \\
1268: %%%%%
1269: \hat \delta^0 &\approx&
1270: \frac{1}{1-\kappa}
1271: \left[
1272: \delta^{iso}- C^0_\delta G + \frac{ (C^0_{\delta})_* }
1273: { (P^0_{\delta})_* }P^0_\delta G
1274: \right].
1275: \end{eqnarray}
1276: Assuming $\langle \hat\zeta^0 \rangle = \langle \hat\delta^0\rangle =0 $
1277: for unlensed sources,
1278: we obtain the desired expressions for flexion as
1279: \begin{eqnarray}
1280: \label{eq:HtoF}
1281: F &\approx&
1282: \left\langle
1283: \frac{\zeta^{iso}}{
1284: \left(C^0_{\zeta} + C^\Delta_\zeta\Delta^0_L\right)
1285: -
1286: \frac{\left(C^0_{\zeta}+C^\Delta_\zeta\Delta^0_L\right)_{*}}
1287: {\left(P^0_\zeta -
1288: \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta\right)_{*}}
1289: \left(P^0_\zeta - \frac{P^0_D}{P^\Delta_D}P^\Delta_\zeta\right)}
1290: \right\rangle,\\
1291: %%%%
1292: G &\approx& \left\langle
1293: \frac{\delta^{iso}}{C^0_\delta
1294: - \frac{ (C^0_{\delta})_* }{ (P^0_{\delta})_* }P^0_\delta}
1295: \right\rangle.
1296: \end{eqnarray}
1297: A detailed derivation of the above equations is provided in
1298: %% @@ edited by KU (2008/02/13)
1299: %%Appendix \ref{sec:isoPSF}.
1300: Full Appendix C.2.
1301:
1302: \section{Simulations and Observations}
1303:
1304:
1305: \subsection{Simulated PSF Anisotropies and Corrections}
1306: \label{subsec:sim}
1307:
1308:
1309: We use simulations to test and assess the limitations of
1310: our PSF correction scheme for the flexion measurement.
1311: To do this, we assume particular models for describing the
1312: isotropic/anisotropic PSF and the surface brightness distribution
1313: for a source. In the present simulations
1314: observational noise and lensing effects are not taken into account.
1315:
1316: First, we assume for
1317: the stellar surface brightness distribution
1318: a two-dimensional Dirac delta function,
1319: $I_*(\btheta)=\delta_D^2(\btheta)$,
1320: while for the galaxy surface brightness distribution
1321: a truncated Gaussian as
1322: \begin{equation}
1323: \label{eq:sim_gal}
1324: I_{\rm gal}(\btheta) =
1325: \exp\left(-\frac{|\btheta|^2}{2R_{\rm gal}^2}\right)
1326: -
1327: \exp\left(-\frac{R_{\rm max}^2}{2R_{\rm gal}^2}\right)
1328: \ \ \ \ {\rm for} \ |\btheta|\le R_{\rm max},
1329: \end{equation}
1330: where $R_{\rm gal}$ and $R_{\rm max}$
1331: are the Gaussian dispersion and the truncation radius
1332: of $I_{\rm gal}$, respectively.
1333: %%%
1334: In the following we set $R_{\rm max}=3R_{\rm gal}$.
1335: %%%%
1336: Next, we assume the isotropic part of PSF, $P^{iso}$, also follows
1337: a truncated Gaussian of the form:
1338: \begin{equation}
1339: \label{eq:sim_iso}
1340: P^{iso}(\btheta) = \frac{1}{2\pi\sigma_{iso}^2}\exp\left(
1341: -\frac{|\btheta|^2}{2\sigma_{iso}^2}
1342: \right)
1343: -
1344: \frac{1}{2\pi\sigma_{iso}^2}\exp\left(
1345: -\frac{\theta_{\rm max}^2}{2\sigma_{iso}^2}
1346: \right) \ \ \ \ {\rm for} \ |\btheta|\le \theta_{\rm max},
1347: \end{equation}
1348: where $\sigma_{iso}$ and $\theta_{\rm max}$ are the Gaussian dispersion
1349: and the truncation radius of $P^{iso}$, respectively.
1350: In the following we set $\theta_{\rm max}=3\sigma_{iso}$.
1351: Finally, we adopt the PSF anisotropy kernel $q(\btheta)$ of the
1352: following form:
1353: \begin{equation}
1354: \label{eq:sim_aniso}
1355: q(\btheta) = A_{\rm aniso}\frac{\theta_1}{|\btheta|^2}
1356: \ \ \ \ {\rm for} \ |\btheta| \le \theta_{\rm aniso},
1357: \end{equation}
1358: where $A_{\rm aniso}$ is the normalization factor that controls the
1359: strength of PSF anisotropy, and $\theta_{\rm aniso}$ is the truncation
1360: radius. We have chosen the direction of anisotropy along the $x$-axis
1361: ($\theta_1$-axis). Thus, the anisotropy kernel $q(\btheta)$ has
1362: two free parameters, $(A_{\rm aniso},\theta_{\rm aniso})$.
1363: We vary the parameters $(A_{\rm aniso},\theta_{\rm aniso})$ to
1364: test our anisotropic PSF correction scheme
1365: as a function of degree of PSF anisotropy. We take $\theta_{\rm aniso}$
1366: in the range of $\theta_{\rm aniso}\in (0,\sigma_{iso})$.
1367:
1368:
1369: Having set up the models, we then produce pixelized images for
1370: model stars and galaxies using the surface brightness distributions
1371: $I_{*}(\btheta)$ and $I_{\rm gal}(\btheta)$, which are then
1372: convolved with the model PSF to yield
1373: $I^{obs}_*(\btheta)$
1374: and $I^{obs}_{\rm gal}(\btheta)$.
1375: No observational noise or
1376: intrinsic/gravitational flexion has been added in the present
1377: simulations. Here we consider the following set of Gaussian source
1378: radii, $R_{\rm gal}=1\sigma_{iso}, 2\sigma_{iso}, 3\sigma_{iso}$.
1379: For each set of $(A_{\rm aniso},\theta_{\rm aniso})$, we measure
1380: various PSF moments, such as $D_q-\Delta_P\propto \zeta_q$,
1381: from mock stellar images using the Gaussian weight
1382: function of dispersion $r_g=\sigma_{iso}$. On the other hand,
1383: we measure various shape moments for mock galaxy images of
1384: $I^{obs}_{\rm gal}(\btheta)$
1385: using the Gaussian weight of
1386: $r_g=\sqrt{R_{\rm gal}^2+\sigma_{iso}^2}$ (i.e., Gaussian dispersion of
1387: the PSF-convolved image).
1388: Then,
1389: we correct observed HOLICs $(\zeta^{obs},\delta^{obs})$
1390: for the PSF
1391: anisotropy to obtain $(\zeta^{iso}.\delta^{iso})$, which should
1392: vanish
1393: for a perfect PSF correction.
1394:
1395:
1396: We show in Figures \ref{fig:psfsim_zeta} and \ref{fig:psfsim_zratio}
1397: results of anisotropic PSF correction for the first HOLICs $\zeta$ of
1398: spin 1.
1399: Figure \ref{fig:psfsim_zeta} shows the dimensionless
1400: $r_g|\zeta|$ before (solid) and after (dashed)
1401: the anisotropic PSF correction
1402: as a function of degree of PSF anisotropy:
1403: $A_{\rm aniso}$ (left panel) and $\theta_{\rm aniso}$ (right panel).
1404: %%%%
1405: Firstly, it clearly shows that the spin-1 PSF anisotropy induced in
1406: mock galaxy images is considerably reduced after applying our PSF
1407: correction method outlined in \S \ref{sec:ksb+}.
1408: Secondly, we see a clear trend that the smaller the galaxy size,
1409: the more severe the anisotropic PSF effect; or that $r_g|\zeta|$
1410: increases with decreasing galaxy size, $R_{\rm gal}$.
1411: %%%%
1412: In particular, when the source size is comparable to the size of PSF
1413: ($R_{\rm gal}=\sigma_{iso}$), the effect is larger
1414: about one order of magnitude than that of large sources with
1415: $R_{\rm gal}=3\sigma_{iso}$.
1416: %%%%
1417: Figure \ref{fig:psfsim_zratio} shows
1418: the ratio
1419: $|\zeta^{iso}|/|\zeta^{obs}|$
1420: of residual to observed PSF anisotropy in mock galaxy images.
1421: We see that, overall, the fractional correction factor
1422: $|\zeta^{iso}|/|\zeta^{obs}|$ is larger for larger galaxy images.
1423: %%%%
1424: Similar trends are also found for the second
1425: HOLICs $\delta$ of spin 3, as shown in
1426: Figures \ref{fig:psfsim_delta} and \ref{fig:psfsim_dratio}.
1427:
1428:
1429: \subsection{Flexion Analysis of Subaru A1689 Data}
1430:
1431: We apply our flexion analysis method
1432: based on the HOLICs moment approach to
1433: Subaru imaging observations of the cluster A1689.
1434: %%%%
1435: A1689 is
1436: %one of the best studied cluster located at
1437: a rich cluster of galaxies at a
1438: moderately low redshift of $z=0.183$,
1439: having a large Einstein radius of $\approx 45$ arcsec
1440: ($z_s\sim 1$; Broadhurst et al. 2005b).
1441: A1689 is one of the best studied lensing clusters
1442: (e.g.,
1443: Tyson \& Fisher 1995;
1444: King, Clowe, \& Schneider et al. 2002;
1445: Bardeau et al. 2005;
1446: Broadhurst et al. 2005a;
1447: Broadhurst et al. 2005b;
1448: Halkola et al. 2006;
1449: Medezinski et al. 2007;
1450: Leonard et al. 2007;
1451: Limousin et al. 2007;
1452: Umetsu, Broadhurst, Takada 2007;
1453: Umetsu \& Broadhurst 2007),
1454: and therefore serves as an ideal target
1455: for testing our flexion analysis pipeline.
1456: %%%%%
1457: %%%%%
1458: Deep HST/ACS imaging of the central region of A1689
1459: %with the superb resolution and
1460: %sensitivity
1461: has revealed
1462: %has provided high-quality photometry of the central region
1463: %of A1689, revealing
1464: $\sim 100$ multiply lensed images of $\sim 30$ background
1465: galaxies (Broadhurst et al. 2005b),
1466: which allowed a detailed reconstruction of the mass
1467: distribution in the cluster core
1468: ($10h^{-1} {\rm kpc} \simlt r \simlt 200 h^{-1}{\rm kpc}$).
1469: %revealed a more than hundred of multiply
1470: %The superb resolution {\it HST/ACS}
1471: %A model-independent method has been developed\cite{BTU05} for
1472: Broadhurst et al. (2005a)
1473: developed a method for
1474: reconstructing the cluster mass profile
1475: by combining
1476: weak-lensing
1477: tangential shear and magnification bias
1478: measurements,
1479: %azimuthally-averaged
1480: %weak shear and magnification measurements,
1481: and derived a model-independent projected mass profile of
1482: the cluster out to its virial radius ($r\simlt 2 h^{-1}$ Mpc)
1483: based on wide-field Subaru/Suprime-Cam data.
1484: %%%
1485: The combination of weak shear and magnification data breaks the
1486: mass sheet degeneracy (Broadhurst, Taylor, \& Peacock 1995)
1487: inherent in all reconstruction methods based
1488: solely on the shape-distortion information
1489: (Schneider \& Seitz 1995).
1490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1491: Broadhurst et al. (2005a) found that
1492: the combined ACS and Subaru profile of the cluster
1493: is well fitted by an
1494: NFW profile (Navarro, Frenk, \& Frenk 1997)
1495: with
1496: a virial mass of $M_{\rm vir}=1.35\times 10^{15}h^{-1}M_{\odot}$
1497: and
1498: a concentration of $c_{\rm vir} \sim 13.7$,
1499: which is significantly larger than theoretically expected
1500: ($c_{\rm vir}\simeq 5$) for the standard LCDM model
1501: (Bullock et al. 2001; Neto et al. 2007).
1502: %%%%%
1503: Based on these Subaru data,
1504: Umetsu \& Broadhurst (2007) used
1505: a maximum likelihood method
1506: to reconstruct the two-dimensional mass map of A1689
1507: from combined shear and magnification data,
1508: and found the azimuthally-averaged mass profile from the
1509: full two-dimensional reconstruction is in good agreement with the
1510: earlier results from the one-dimensional analysis by
1511: Broadhurst et al. (2005a),
1512: supporting the assumption of quasi-circular symmetry in the
1513: projected mass distribution of A1689.
1514: Recently, Leonard et al. (2007) performed a weak lensing analysis
1515: of A1689 based on the ACS data
1516: by incorporating measurements of flexion as well as
1517: weak shear and strong lensing, and their flexion reconstruction
1518: has revealed mass substructures associated with
1519: small clumps of galaxies, while no flexion signal has been detected
1520: at the cluster center, showing an under density around the location
1521: of the cD galaxy.
1522:
1523:
1524: For our flexion analysis of A1689,
1525: we used Suprime-Cam $i'$-imaging data
1526: (Broadhurst et al. 2005a; Umetsu \& Broadhurst 2007),
1527: covering a field of $\sim 30'\times 25'$ with
1528: $0\farcs 202 \,{\rm pixel}^{-1}$ sampling.
1529: The seeing FWHM in the co-added $i'$
1530: image is $0\farcs 88$, and the limiting
1531: magnitude is $i'=25.9$ for a $3\sigma$ detection within a $2''$ aperture
1532: (see Broadhurst et al. 2005a).
1533: Since the flexion signal is weaker at larger angular scales
1534: (see OUF for detailed discussions), in the present flexion analysis
1535: we discarded outer boundaries from the analysis and
1536: only used the central $3000\times 3000$ pixel region,
1537: corresponding an angular scale of $\approx 10'$ on a side,
1538: or a physical scale of $1.3h^{-1}$ Mpc at the cluster redshift of
1539: $z=0.183$.
1540:
1541:
1542: We used our weak lensing analysis pipeline based on IMCAT (Kaiser et
1543: al. 1995) extended to include our HOLICs moment method.
1544: %%%%%
1545: We selected a stellar sample of $N_*=73$ objects for measuring
1546: the anisotropic/isotropic PSF effects.
1547: %%%%%
1548: On the basis of
1549: the simulation results, we excluded from our background galaxy sample
1550: those small objects
1551: whose half-light radius ($r_h$) and Gaussian detection radius ($r_g$)
1552: are smaller than or comparable to
1553: the PSF; here we selected galaxies with
1554: $0\farcs 6 < r_h < 2''$ and $r_g > 0\farcs 38$,
1555: whereas the median values of stellar $r_h$ and $r_g$ are
1556: $\langle r_{h*}\rangle \approx 0\farcs 48$ and
1557: $\langle r_{g*}\rangle \approx 0\farcs 34$ using $N_*=73$ stars.
1558: %$3 < r_h 10$ pixels and
1559: %$r_g > 1.9$ pixels
1560: These lower cutoffs in the galaxy size are essential
1561: for reliable flexion measurements,
1562: because
1563: the smaller the object,
1564: the noisier its shape measurement due to pixelization noise;
1565: and the shape of an image whose intrinsic size is smaller than
1566: or comparable to the PSF size can be highly distorted and smeared,
1567: as we have seen in \S \ref{subsec:sim}.
1568: Faint objects will also yield noisy shape measurements, in particular
1569: for the case of higher order shape moments. We thus selected bright
1570: galaxies with $20 < i' < 25$ in the AB magnitude system.
1571: Further we excluded from our background sample
1572: those objects whose
1573: flexion estimates are significantly larger than the model prediction
1574: ($|F|\sim 0.1$ arcsec$^{-1}$)
1575: using the best-fitting NFW profile
1576: derived from the joint Subaru and ACS analysis
1577: (Broadhurst et al. 2005a);
1578: %%%%%%
1579: with this model,
1580: we set an upper cutoff in flexion of $|F|<0.4$
1581: arcsec$^{-1}$, and an upper cutoff in the first HOLICs of
1582: $|\zeta^{iso}|<0.03$ arcsec$^{-1}$ (see equation [\ref{eq:HtoF}]).
1583: We note that the measurements of spin-3 HOLICs $\delta$ were found to
1584: be quite noisy (see OUF for detailed discussions), so that we discarded
1585: the second flexion measurements from the present study.
1586: Finally, these selection criteria yielded a sample of $791$ galaxies
1587: usable for our flexion analysis, corresponding to a mean surface number
1588: density of $\bar{n}_g= 7.75$ arcmin$^{-2}$.
1589: Using the background galaxy sample above we found
1590: a mean value of
1591: $\langle F\rangle = 0.000223$ arcsec$^{-1}$ and a dispersion of
1592: $\sigma_F= 0.11245$ arcsec$^{-1}$.
1593: %%%%
1594: In Figure \ref{fig:zetaq_field} we show the spatial distribution of
1595: spin-1 PSF anisotropy as measured from stellar images before (left panel)
1596: and after (right panel) the anisotropic PSF correction.
1597: Figure \ref{fig:zetaq} compares the distribution of
1598: two components of complex spin-1 PSF anisotropy before (left panel) and
1599: after(right panel) the anisotropic PSF correction.
1600: As clearly shown in Figures \ref{fig:zetaq_field} and \ref{fig:zetaq},
1601: the higher-order PSF anisotropy in observational data is indeed
1602: significant, so that one needs to take into account the higher-order PSF
1603: anisotropy correction in practical flexion measurements.
1604:
1605:
1606:
1607:
1608: We use our first flexion measurements obtained with our moment-based
1609: analysis method to reconstruct the projected mass distribution of
1610: A1689.
1611: To do this we utilize
1612: the Fourier-space relation between the first flexion $F$
1613: and the lensing convergence $\kappa$
1614: (\S 2.4 of OUF; Bacon et al. 2006)
1615: with the weak lensing approximation.
1616: The field size for the mass reconstruction is
1617: $9'\times 9'$, sampled with a grid of $256\times 256$ pixels,
1618: over which the unconstrained $k=0$ mode is set to zero.
1619: %%%
1620: Figures \ref{fig:A1689_Emode} and \ref{fig:A1689_Bmode} show
1621: the $E$-mode convergence $\kappa=\kappa_E$ due to
1622: lensing and the $B$-mode convergence $\kappa_B$ which is expected to
1623: vanish in the weak lensing limit and can thus be used to monitor the
1624: reconstruction error in the $E$-mode $\kappa$ map.
1625: %%%
1626: A central $8'\times 8'$ region is
1627: displayed in Figures \ref{fig:A1689_Emode} and \ref{fig:A1689_Bmode}.
1628: The reconstructed $\kappa$ maps were smoothed with a Gaussian filter
1629: of ${\rm FWHM}=0\farcm 33$.
1630: %%%%%
1631: Table \ref{tab:stats} lists basic statistics of the reconstructed
1632: $E$- and $B$-mode $\kappa$ fields measured in the central $8'\times 8'$
1633: field.
1634: The rms dispersion in the Gaussian smoothed $B$-mode $\kappa$ map is
1635: obtained as
1636: $\sigma_B\approx 0.51$. The maximum and minimum values
1637: in the $B$-mode convergence field are $1.52$ and $-1.61$,
1638: corresponding to $2.9 \sigma$ and $-3.1\sigma$ fluctuations
1639: (see Table \ref{tab:stats}).
1640: %$\sigma_B\approx 0.45$
1641: %over the central $8'\times 8'$ region.
1642: %%%%
1643: Figure \ref{fig:A1689_Emode} reveals two significant mass concentrations
1644: in the $E$-mode $\kappa$ map
1645: associated with clumps of bright galaxies.
1646: The first peak has a peak value of $\kappa_E=2.66$, and
1647: is detected at
1648: %$5.9\sigma$ significance.
1649: $5.2\sigma$ significance.
1650: This first peak is associated with
1651: the central concentration of bright cluster galaxies including
1652: the cD galaxy, as shown in Figure \ref{fig:A1689_image}.
1653: This central mass concentration associated with the brightest cluster
1654: galaxies was not detected in the earlier ACS flexion analysis by Leonard et
1655: al. (2007).
1656: The ACS/Subaru best-fitting NFW model predicts $\kappa(z_s=1)\approx
1657: 2.5$ at $\theta \sim 0\farcm 1$.
1658: %%%%%
1659: The second peak
1660: %($5\sigma$ significance),
1661: ($4.4\sigma$ significance),
1662: on the other hand,
1663: is located $\approx 0\farcm 9$ to the northeast direction,
1664: and is associated with a
1665: local clump of bright galaxies (see Figure \ref{fig:A1689_image}),
1666: having a peak value of $\kappa_E=2.23$.
1667: This second mass peak has been detected in the earlier lensing studies
1668: based on the high resolution HST/ACS data
1669: (e.g, Broadhurst et al. 2005b; Leonard et al. 2007),
1670: and its likely bimodality in the central region
1671: has been discussed previous studies
1672: (Miralda-Escude \& Babul 1995;
1673: Halkola, Seitz, \& Pannella 2006;
1674: Limousin et al 2007;
1675: Saha, Williams, \& Ferreras 2007).
1676: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1677: However, as compared to the Subaru weak lensing analysis by
1678: Umetsu \& Broadhurst (2007),
1679: the flexion-based mass reconstruction cannot recover
1680: the global cluster structure on larger angular scales
1681: ($\simgt$ a few arcmin), as demonstrated by OUF.
1682:
1683:
1684:
1685:
1686: \section{Discussion and Conclusions}
1687:
1688: In the present paper, we have developed a method for
1689: weak lensing flexion analysis by fully extending the KSB method
1690: to include the measurement of HOLICs (OUF).
1691: %%%%
1692: In particular, we take into
1693: account explicitly the weight function in calculations of noisy
1694: shape moments and the effects of spin-1 and spin-3 PSF anisotropies,
1695: as well as isotropic PSF smearing, in the limit of
1696: weak lensing and
1697: small PSF anisotropy ($q$).
1698: %%%%
1699: The higher order weak lensing effect induces a centroid shift in the
1700: observed image of the background (Goldberg \& Bacon 2005; OUF; Goldberg
1701: \& Leonard 2007). In weighted moment calculations, this will yield
1702: in the flexion measurement
1703: additional correction terms (relevant to $W'(x), W''(x),
1704: W'''(x)$) that must be taken into
1705: account by properly expanding the weight function $W(x)$.
1706: It is found that neglecting these additional terms originated from the
1707: Taylor expansion of $W(x)$ yields the same result as obtained by
1708: Goldberg \& Leonard (2007; see Appendix therein).
1709: We extended the KSB formalism to include
1710: the higher-order
1711: isotropic and anisotropic PSF effects relevant to spin-1 and spin-3
1712: HOLICs by following the
1713: prescription given by KSB and Bartelmann \& Schneider (2001), which
1714: provides direct relations between
1715: the observable HOLICs and underlying flexion in the weak lensing limit.
1716: %%%%%
1717:
1718: We have implemented in our analysis pipeline
1719: our flexion analysis algorithm based on the HOLICs
1720: moment approach, and tested the reliability and limitation of our
1721: PSF correction scheme using numerical simulations.
1722: %%%%
1723: Our simulation results show that (i)
1724: after applying our PSF correction method
1725: the PSF-induced anisotropies
1726: in HOLICs of mock galaxy images can be considerably reduced
1727: by a factor of $10$--$100$,
1728: depending on the strength of PSF anisotropy,
1729: (ii) those small galaxies whose angular size is smaller than or
1730: comparable to the size of PSF suffer from severe anisotropic PSF
1731: effects,
1732: %the smaller the galaxy image, the stronger the anisotropic PSF
1733: %effect,
1734: and that (iii) there is an overall trend that
1735: the fractional correction factor is larger for larger galaxy images.
1736: Therefore, our simulation results
1737: support the reliability of our PSF-correction scheme and
1738: its practical implementation.
1739:
1740:
1741: Based on the simulation results, we have applied our flexion analysis
1742: pipeline to
1743: ground-based $i'$ imaging data of the rich cluster A1689 ($z=0.183$)
1744: taken with Subaru/Suprime-Cam.
1745: Our flexion analysis of Subaru A1689 data revealed
1746: a non-negligible, significant effect of higher-order PSF anisotropy
1747: induced in stellar images
1748: (Figures \ref{fig:zetaq_field} and \ref{fig:zetaq}).
1749: %%%%
1750: It is therefore important in practical
1751: flexion measurements to quantify and correct for the higher-order
1752: anisotropic PSF effects.
1753: %Despite the apparently small number density of background galaxies,
1754: %$\bar{n}_g=7.75$ arcmin$^{-2}$, usable for weak lensing flexion
1755: %analysis,
1756: %Thanks to high-quality imaging with Subaru/Suprime-Cam,
1757:
1758: Our mass reconstruction from
1759: the first-flexion measurements
1760: shows two significant
1761: %($>5\sigma$)
1762: ($>4\sigma$)
1763: mass structures associated
1764: with concentrations of bright galaxies in the central cluster region:
1765: the first peak ($5.2\sigma$)
1766: associated with the central concentration of bright
1767: galaxies including the cD galaxy,
1768: and the second peak ($4.4\sigma$) associated with a clump of bright
1769: galaxies located $\sim 1'$ northeast of the cluster center.
1770: %%%%%
1771: This significant detection of the second peak
1772: %, associated
1773: %with a local concentration of bright galaxies,
1774: confirms
1775: earlier ACS results from the strong lensing analysis
1776: (Broadhurst et al. 2005b; Halkola et al. 2006; Leonard et al. 2007)
1777: and the combined strong lensing, weak shear,
1778: and flexion analysis by Leonard et al. (2007).
1779: %%%%
1780: The central mass peak, however,
1781: was not recovered in the earlier flexion analysis
1782: by Leonard et al. (2007) based on HST/ACS data.
1783: Leonard et al. (2007) attributed this to
1784: their relatively large reconstruction error at the cluster center,
1785: although they have a very large number density of
1786: background galaxies,
1787: $\bar{n}_g\approx 75$ arcmin$^{-2}$.
1788: %%%%
1789: On the other hand,
1790: owing to our conservative selection criteria for the background sample,
1791: the mean number density of background galaxies used for the present analysis
1792: is $\bar{n}_g=7.75$ arcmin$^{-2}$, which is almost one order of
1793: magnitude smaller than that of the ACS data,
1794: and is about $20\%-30\%$
1795: of a typical number density of
1796: magnitude/size-selected background galaxies usable for the
1797: quadrupole shape measurements
1798: in ground-based Subaru observations
1799: ($\bar n_g\sim 30-40 {\rm arcmin}^{-2}$).
1800: %%%%
1801: However, we found that it is rather
1802: important to remove small/faint galaxy images and noisy outliers in
1803: flexion measurements since they are likely to be affected by
1804: the residual PSF anisotropy and/or observational noise in the shape
1805: measurement (\S \ref{subsec:sim}).
1806: Besides, the smaller the object, the larger the amplitude of
1807: intrinsic flexion contributions. Recall that
1808: flexion and HOLICs have a dimension of length inverse:
1809: The response to flexion is size-dependent, and
1810: the amplitude of intrinsic flexion is inversely proportional to
1811: the object size. Indeed, we find that inclusion of smaller objects
1812: results in a noisy reconstruction.
1813: %%%%%
1814: Similar values of the background number density, $\bar n_g\simlt 10 {\rm
1815: arcmin}^{-2}$, have been used in
1816: recent quadrupole weak lensing analyses based on Subaru observations
1817: (e.g., Broadhurst et al. 2005a; Umetsu \& Broadhurst 2007;
1818: weak lensing cluster mass measurements of Okabe \& Umetsu 2008).
1819: In their studies only objects redder than the cluster sequence
1820: are selected in color-magnitude space
1821: for their weak lensing analysis,
1822: because such a red population is expected to comprise only background
1823: galaxies ($\bar z_s\sim 0.9$; see, e.g., Medezinski et al. 2007),
1824: made redder by relatively large $k$-corrections and with
1825: negligible contamination by cluster galaxies (Broadhurst et al. 2005a;
1826: Medezinski et al. 2007).
1827: %%% with the smaller # of .. one needs to account for..
1828: However, the smaller number of objects implies a coarser angular resolution
1829: in the map-making for achieving a proper signal-to-noise ratio
1830: (e.g., per-pixel ${\rm S/N}\simgt 1$). With $\bar n_g\sim 10 {\rm
1831: arcmin}^{-2}$ for cluster quadrupole weak lensing,
1832: typical angular resolutions are about $1-2$ arcmin
1833: (e.g., Gaussian FWHM, or boxcar width).
1834: On the other hand, flexion measures essentially
1835: the gradient of the tidal gravitational shear field
1836: (i.e., $F,G\propto \phi(r)/r^3$), and hence is relatively sensitive
1837: to small-scale structures.
1838: Therefore, our successful reconstruction of the mass substructures
1839: with a small
1840: background density, $\bar n_g\sim 8 {\rm arcmin}^{-2}$, could be
1841: attributed to the superior sensitivity of flexion to small scale
1842: structures (see OUF for detailed discussions)
1843: and the here-adopted selection criteria for a background
1844: galaxy sample for weak lensing flexion analysis.
1845:
1846:
1847: %Due to
1848: %Thanks to the superior sensitivity of
1849: %flexion to small-scale structure,
1850: %we can still recover the mass substructures even with a small background
1851: %density of $\bar{n}_g\simlt 10$ galaxies arcmin$^{-2}$.
1852: %%%%%
1853: %The significant detection of the second peak, associated
1854: %with a local concentration of bright galaxies,
1855: %confirms
1856: %earlier ACS results from the strong lensing analysis
1857: %(Broadhurst et al. (2005b); Halkola et al. 2006; Leonard et
1858: %al. 2007)
1859: %and the combined strong lensing, weak shear,
1860: %and flexion analysis by Leonard et al. (2007).
1861: %%%
1862: Finally, we emphasize that our HOLICs formalism here is different from
1863: the earlier work by Goldberg \& Leonard (2007) in that
1864: (1) additional correction terms for the centroid shift, relevant
1865: to the derivatives of the weight function,
1866: have been included
1867: and
1868: (2) the spin-1 and spin-3 PSF anisotropies, as well as the isotropic PSF
1869: smearing, have been taken into account under the assumption of small PSF
1870: anisotropy ($q[\btheta]$), as done in the KSB formalism.
1871: %%%%
1872: Our flexion-based mass reconstruction of A1689 demonstrates the power of
1873: the generalized flexion analysis techniques
1874: for quantitative and accurate measurements of the weak gravitational
1875: lensing effects.
1876:
1877:
1878: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1879: %%%
1880: %%% Acknowledgments
1881: %%%
1882: %%%
1883: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1884:
1885: \acknowledgments
1886:
1887: We thank Masahiro Takada for valuable discussions.
1888: We thank the anonymous referee for invaluable comments and suggestions.
1889: The work is partially supported by the COE program at Tohoku University.
1890: This work in part supported by
1891: the National Science Council of Taiwan
1892: under the grant NSC95-2112-M-001-074-MY2.
1893:
1894: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1895: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1896: %%%
1897: %%% Appendix
1898: %%%
1899: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1900: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1901:
1902: \appendix
1903:
1904:
1905: \section{HOLICs Formalism in the Complex Form}
1906: \label{sec:basis}
1907:
1908: \subsection{Complex Displacement}
1909:
1910: In this Appendix, we adopt the following notations
1911: to represent complex quantities with
1912: the degree $N$ and the spin number $M$:
1913: \begin{eqnarray}
1914: X^N_M &\equiv&
1915: (\dt_1 + i\dt_2)^{\frac{N+M}{2}}(\dt_1 +
1916: i\dt_2)^{*\frac{N-M}{2}}\nonumber\\
1917: &=& (\dt_1 + i\dt_2)^{\frac{N+M}{2}}(\dt_1 - i\dt_2)^{\frac{N-M}{2}},\\
1918: Y^N_M &\equiv&
1919: (\db_1 + i\db_2)^{\frac{N+M}{2}}(\db_1 + i\db_2)^{*\frac{N-M}{2}}\nonumber\\
1920: &=& (\db_1 + i\db_2)^{\frac{N+M}{2}}(\db_1 - i\db_2)^{\frac{N-M}{2}}.
1921: \end{eqnarray}
1922: Here a quantity with the spin number of $-M$
1923: is equivalent to the complex conjugate
1924: of the corresponding spin-$M$ quantity.
1925: Unless otherwise noted,
1926: we shall use $X^N_M$ to represent complex quantities in
1927: the image plane, and $Y^N_M$ those in the source plane.
1928:
1929:
1930:
1931: The product of $X^N_M$ and $X^L_K$ is expressed as
1932: \begin{eqnarray}
1933: X^N_MX^L_K
1934: &=& X^{N+L}_{M+K},\nonumber\\
1935: X^N_MX^{L*}_K
1936: &=& X^{N+L}_{M-K} \ \ \ (M>K),\nonumber\\
1937: &=& (X^{N*}_MX^L_K)^* = \paren{X^{N+L}_{K-M}}^* \ \ \
1938: (M<K).
1939: \end{eqnarray}
1940: For arbitrary complex numbers $W$ and $Z$
1941: the following identities hold:
1942: \begin{eqnarray}
1943: 2\Real{WX^{N*}_M}X^L_K
1944: &=&
1945: \paren{WX^{N*}_M + W^*X^N_M}X^L_K
1946: = WX^{N+L}_{K-M} + W^*X^{N+L}_{M+K},\\
1947: 4\Real{WX^{N*}_M}\Real{ZX^{L*}_K}&=&
1948: \paren{WX^{N*}_M
1949: + W^*X^N_M}\paren{ZX^{L*}_K + Z^*X^L_K},\nonumber\\
1950: &=&
1951: WZ\paren{X^{N+L}_{K+M}}^*
1952: + WZ^*X^{N+L}_{K-M} + W^*ZX^{N+L}_{-K+M} + W^*Z^*X^{N+L}_{K+M},\nonumber\\
1953: &=&2\Real{WZ\paren{X^{N+L}_{K+M}}^*} + 2\Real{WZ^*X^{N+L}_{K-M}}.
1954: \end{eqnarray}
1955:
1956:
1957: \subsection{HOLICs Family}
1958:
1959: We summarize in the complex form
1960: a family of complex shape moments including HOLICs
1961: relevant to the weak lensing flexion analysis.
1962:
1963: \begin{eqnarray}
1964: M&=\int d^2\theta I(\theta)W(X^2_0/\sigma^2)& {\rm(spin-0)}\nonumber\\
1965: D&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^1_1}{M}
1966: &{\rm(spin-1)},\nonumber\\
1967: \trQ&=\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^2_0
1968: &{\rm(spin-0)},\nonumber\\
1969: \chi&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^2_2}{\trQ}
1970: &{\rm(spin-2)},\nonumber\\
1971: \zeta&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^3_1}{\xi}&{\rm(spin-1)},\nonumber\\
1972: \delta&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^3_3}{\xi}&{\rm(spin-3)},\nonumber\\
1973: \xi&={\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^4_0}&{\rm(spin-0)},\nonumber\\
1974: \eta&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^4_2}{\xi}&{\rm(spin-2)},\nonumber\\
1975: \lambda&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^4_4}{\xi}&{\rm(spin-4)},\nonumber\\
1976: \iota_I&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^5_1}{\xi}&{\rm(spin-1)},\nonumber\\
1977: \iotaIII&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^5_3}{\xi}&{\rm(spin-3)},\nonumber\\
1978: \iota_V&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^5_5}{\xi}&{\rm(spin-5)},\nonumber\\
1979: \upsilon_0&=\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^6_0&{\rm(spin-0)},\nonumber\\
1980: \upsilonII&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^6_2}{\xi}&{\rm(spin-2)},\nonumber\\
1981: \upsilonIV&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^6_4}{\xi}&{\rm(spin-4)},\nonumber\\
1982: \upsilonVI&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^6_6}{\xi}&{\rm(spin-6)},\nonumber\\
1983: \tau_I&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^7_1}{\xi}&{\rm(spin-1)},\nonumber\\
1984: \tauIII&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^7_3}{\xi}&{\rm(spin-3)},\nonumber\\
1985: \tau_V&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^7_5}{\xi}&{\rm(spin-5)},\nonumber\\
1986: \tauVII&=\frac{\int d^2\theta I(\theta)W(X^2_0/\sigma^2) X^7_7}{\xi}&{\rm(spin-7)}.
1987: \end{eqnarray}
1988:
1989: \subsection{Differential Operators}
1990:
1991: %In this section, we differentiate $X^N_M$ and weight function
1992: %$W\paren{\frac{X_0^2}{\sigma^2}}$.
1993:
1994: Let us first define the complex gradient operator $\partial$ as
1995: \begin{equation}
1996: \partial = \left(\frac{\partial}{\partial \theta_1} + i\frac{\partial}
1997: {\partial \theta_2}\right).
1998: \end{equation}
1999:
2000: Operating $\partial$ on complex $X^N_M$ yields:
2001: \begin{eqnarray}
2002: \label{eq:GENdif}
2003: &\hspace{-2cm}\partial X^1_1 &= \left(\frac{\partial}{\partial \theta_1} + i\frac{\partial}{\partial \theta_2}\right)(\theta_1 + i\theta_2)=(1-1)=0,\nonumber\\
2004: &\hspace{-2cm}\partial^* X^1_1
2005: &= \left(\frac{\partial}{\partial \theta_1} - i\frac{\partial}{\partial
2006: \theta_2}\right)
2007: (\theta_1 +i\theta_2)=(1+1)=2,\nonumber\\
2008: &\hspace{-2cm}\partial X^N_M
2009: &= \partial\left( \left(X^1_1\right)^\frac{N+M}{2}\left(X^{1*}_1
2010: \right)^\frac{N-M}{2} \right)
2011: = (N-M) \left(X^1_1\right)^{\frac{N+M}{2}}
2012: \left(X^{1*}_1\right)^\frac{N-M-2}{2} = (N-M)X^{N-1}_{M+1},\\
2013: &\hspace{-2cm}\partial^* X^N_M
2014: &= \partial^*\left( \left(X^1_1\right)^\frac{N+M}{2}
2015: \left(X^{1*}_1\right)^\frac{N-M}{2} \right)
2016: = (N+M) \left(X^1_1\right)^\frac{N+M-2}{2}\left(X^{1*}_1
2017: \right)^{\frac{N-M}{2}} = (N+M)X^{N-1}_{M-1}.
2018: \end{eqnarray}
2019:
2020: Similarly, by operating $\partial$ on the weight function
2021: $W(X_0^2/\sigma^2)$, one finds the following:
2022: \begin{eqnarray}
2023: \label{eq:GENdif2}
2024: \partial W\left(\frac{X^2_0}{\sigma^2}\right) &=& \frac{2}{\sigma^2}X^1_1W'\left(\frac{X^2_0}{\sigma^2}\right),\nonumber\\
2025: \partial^* W\left(\frac{X^2_0}{\sigma^2}\right) &=& \frac{2}{\sigma^2}X^{1*}_1W'\left(\frac{X^2_0}{\sigma^2}\right),
2026: \end{eqnarray}
2027: where $F$ and $G$ are the first and the second flexion defined as
2028: $F=\cF/(1-\kappa)$ and $G=\cG/(1-\kappa)$, respectively.
2029: Then,
2030: the complex displacement in the source plane, $\Delta\beta$,
2031: is expressed in terms of the lensing convergence, shear and flexion as
2032: \begin{equation}
2033: \db = \db_1 +i\db_2 \equiv Y^1_1 \approx (1-\kappa)
2034: \left[
2035: X^1_1 - gX^{1*}_1 - \frac{1}{4}
2036: \paren{2FX^2_0 + F^*X^2_2 + GX^{2*}_2}
2037: \right].
2038: \end{equation}
2039: The integration measures in the source and image planes are related
2040: in the following way
2041: \begin{eqnarray}
2042: d^2\beta &=&(1\tm\kappa)^2\Biggl(1 \tm 2\Real{FX^{1*}_1} \tm |g|^2 \tp \frac{1}{4}|F|^2X^2_0 \tm \frac{1}{4}|G|^2X^2_0\nonumber\\
2043: &\tp& \frac{1}{2}\Real{F^2X^{2*}_2} \tm \Real{g^*FX^1_1} \tm \Real{g^*GX^{1*}_1} \tm \frac{1}{2}\Real{FG^*X^2_2}\Biggr)d^2\theta\nonumber\\
2044: &\approx&(1-\kappa)^2\paren{1- 2\Real{FX^{1*}_1}}d^2\theta,
2045: \end{eqnarray}
2046: to the first order of reduced flexion (e.g. OUF).
2047:
2048:
2049:
2050:
2051: \clearpage
2052:
2053: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2054: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2055: %%%
2056: %%% References
2057: %%%
2058: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2059: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2060:
2061:
2062: \begin{thebibliography}{99}
2063:
2064:
2065: \bibitem{Bacon00}
2066: %% cosmic shear statistics
2067: Bacon, D.~J., Refregier, A.~R., Ellis, R.~S. 2000, MNRAS, 318,
2068: 625
2069:
2070: \bibitem{Bacon06}
2071: Bacon, D. J.,
2072: Goldberg, D. M.,
2073: Rowe, B. T. P.,
2074: \& Taylor, A. N. 2006, \mnras, 365, 414
2075:
2076: \bibitem{BARD1689}
2077: Bardeau, S. et al. 2005 A\&A, 434, 433
2078:
2079: \bibitem{review}
2080: Bartelmann, M., \& Schneider, P. 2001, Phys.Rep., 340, 291
2081: 652, 937.
2082:
2083: \bibitem{BTP}
2084: Broadhurst, T., Taylor, A.~N., \&
2085: Peacock, J.~A. 1995, \apj, 438, 49
2086:
2087:
2088: \bibitem{B05b}
2089: Broadhurst, T. et al. 2005, \apj, 621, 53 (Broadhurst et al. 2005b)
2090:
2091: \bibitem{B05a}
2092: Broadhurst, T.,
2093: Takada, M.,
2094: Umetsu, K.,
2095: Kong, X.,
2096: Arimoto, N.,
2097: Chiba, M., \& Futamase, T. 2005, 619,
2098: 143L (Broadhurst et al. 2005a)
2099:
2100:
2101: \bibitem{Bullock}
2102: Bullock, J.~S. et al. 2001, MNRAS, 321, 559
2103:
2104:
2105: \bibitem{Erben00}
2106: Erben, T. et al. 2000, A\&A, 355, 23
2107:
2108:
2109: \bibitem{GN02}
2110: Goldberg, D.~M. \& Natarajan, P. 2002,
2111: \apj, 564, 65
2112:
2113:
2114: \bibitem{Flexion}
2115: Goldberg, D. M., \& Bacon, D. J. 2005, ApJ, 619, 741
2116:
2117: \bibitem{Flexion2}
2118: Goldberg, D. M., \& Leonard, A. 2006, ApJ, 660, 1003
2119:
2120: \bibitem{halkola07}
2121: Halkola, A.
2122: Seitz, S.,
2123: \& Pannella, M. 2006, \mnras, 372, 1425
2124:
2125: \bibitem{Hamana03}
2126: % cosmic shear
2127: Hamana, T. et al. 2003, ApJ, 597, 98
2128:
2129:
2130: \bibitem{STEP1}
2131: Heymans, C. et al. 2006, MNRAS, 368, 1323
2132:
2133: \bibitem{Sextupole}
2134: Irwin, J., \& Shmakova, M. 2006, ApJ, 645, 17
2135:
2136:
2137: \bibitem{KS93}
2138: Kaiser, N., \& Squires, G. 1993, ApJ, 404, 441
2139:
2140: \bibitem{KSB}
2141: Kaiser, N., Squires, G., Broadhurst, T. 1995, ApJ, 449, 460
2142:
2143: \bibitem{KA1689}
2144: King, L. J., Clowe, D. I., Schneider, P. 2002 A\&A, 383, 118
2145:
2146: \bibitem{GA1689}
2147: Leonard, A., Goldberg, D.~M., Haaga, J.~L.,
2148: Massey, R. 2007, ApJ, 666, 51L
2149:
2150: \bibitem{elinor}
2151: Medezinski, E., Broadhurst, T., Umetsu, K.,
2152: Coe, D., Benitez, N., Ford, H., Rephaeli, Y.,
2153: Arimoto, N., \& Kong, X. 2007, \apj, 663, 717
2154:
2155:
2156: \bibitem{SW-A1689}
2157: Limousin, M. et al. 2007, ApJ, 668, 643
2158:
2159:
2160: \bibitem{STEP2}
2161: Massey, R. et al. 2007, MNRAS, 376, 13
2162:
2163: \bibitem{NFW}
2164: Navarro, J.~F., Frenk, C.~S., White, S.~D.~M., 1997, ApJ, 490, 493
2165:
2166: \bibitem{Neto07}
2167: Neto, A.~F. et al. 2007, \mnras, 381, 1450
2168:
2169:
2170: \bibitem{Okabe}
2171: Okabe, N. \& Umetsu, K. 2008, \pasj, in press
2172: (arXiv:astro-ph/0702649)
2173:
2174:
2175: %\bibitem{HOLICs}
2176: % Okura, Y., Umetsu, K., \& Futamase, T., 2007, ApJ, 660, 995
2177:
2178: \bibitem{HOLICs}
2179: Okura, Y., Umetsu, K., \& Futamase, T. 2007, ApJ, 660, 995
2180:
2181: \bibitem{shapelet}
2182: Refregier, A. 2003, MNRAS, 338, 35
2183:
2184: \bibitem{saha07}
2185: Saha, P.,
2186: Williams, L.~L.~R.,
2187: Ferreras, I. 2007, \apj, 663, 29
2188:
2189: \bibitem{ss95}
2190: Schneider, P. \& Seitz, C. 1995, \aap 294, 411
2191:
2192:
2193: \bibitem{Schneider07}
2194: Schneider, P. \& Er, X. 2007, submitted to A\&A (astro-ph/0709.1003)
2195:
2196: \bibitem{TA1689}
2197: Tyson, J. A., \& Fisher, P., 1995 ApJ 446 L55
2198:
2199:
2200: \bibitem{UTF1999}
2201: Umetsu, K, Tada, M., \& Futamase, T. 1999,
2202: Prog.~Theor.~Phys.~Suppl., 133, 53
2203:
2204: \bibitem{UF2000}
2205: Umetsu, K, \& Futamase, T. 2000, ApJ, 539, L5
2206:
2207: \bibitem{UA1689}
2208: Umetsu, K., Takada, M., Broadhurst, T. 2007,
2209: Mod.~Phys.~Lett.~A, 22, 2099 (arXiv:astro-ph/0702096)
2210:
2211: \bibitem{Umetsu07}
2212: Umetsu, K. \& Broadhurst, T. 2007, submitted to ApJ
2213: (arXiv:astro-ph/0712.3441)
2214:
2215: \bibitem{vanWaerbeke}
2216: Van Waerbeke et al. 2001, A\&A, 374, 757
2217:
2218: \bibitem{Wittman}
2219: Wittman, D. et al. 2001, ApJ, 557, 89L
2220:
2221:
2222: \end{thebibliography}
2223:
2224: \clearpage
2225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2226: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2227: %%%
2228: %%% Tables
2229: %%%
2230: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2231: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2232:
2233: %%% Table 1
2234:
2235: \begin{deluxetable}{cllllll}
2236: \tabletypesize{\scriptsize}
2237: \tablecaption{
2238: \label{tab:stats}
2239: Basic statistics of the reconstructed $E$ and $B$ mode convergence fields}
2240: \tablewidth{0pt}
2241: \tablehead{
2242: \colhead{field} &
2243: \colhead{mean ($\bar\kappa$)} &
2244: \colhead{$\sigma$} &
2245: \colhead{skewness \tablenotemark{a}}&
2246: \colhead{kurtosis \tablenotemark{b}} &
2247: \colhead{minimum}&
2248: \colhead{maximum}
2249: }
2250: \startdata
2251: $B$ & 0.058 & 0.512 & $-0.334 \pm 0.2$ & $0.043 \pm 0.4$ & -1.61 & 1.52\\
2252: $E$ &-0.058 & 0.701 & $-0.239 \pm 0.2$ & $0.125 \pm 0.4$ & -2.09 & 2.66
2253: %$B$ & 0.058 & 0.512 & -0.334 & 0.043 & -1.61 & 1.52\\
2254: %$E$ &-0.058 & 0.701 & -0.239 & 0.125 & -2.09 & 2.66
2255: \enddata
2256: \tablecomments{The moments are calculated from
2257: the convergence within the central $8'\times 8'$ region.}
2258: \tablenotetext{a}{Skewness defined as
2259: $\langle(\kappa-\bar\kappa)^3\rangle/\sigma^3$.}
2260: \tablenotetext{b}{Kurtosis defined as
2261: $\langle(\kappa-\bar\kappa)^4 \rangle/\sigma^4-3$.}
2262: \end{deluxetable}
2263:
2264: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2265: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2266: %%%
2267: %%% Figures
2268: %%%
2269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2270: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2271:
2272:
2273: %% Use the figure environment and \plotone or \plottwo to include
2274: %% figures and captions in your electronic submission.
2275: %% To embed the sample graphics in
2276: %% the file, uncomment the \plotone, \plottwo, and
2277: %% \includegraphics commands
2278: %%
2279: %% If you need a layout that cannot be achieved with \plotone or
2280: %% \plottwo, you can invoke the graphicx package directly with the
2281: %% \includegraphics command or use \plotfiddle. For more information,
2282: %% please see the tutorial on "Using Electronic Art with AASTeX" in the
2283: %% documentation section at the AASTeX Web site,
2284: %% http://www.journals.uchicago.edu/AAS/AASTeX.
2285: %%
2286: %% The examples below also include sample markup for submission of
2287: %% supplemental electronic materials. As always, be sure to check
2288: %% the instructions to authors for the journal you are submitting to
2289: %% for specific submissions guidelines as they vary from
2290: %% journal to journal.
2291:
2292:
2293: %% This example uses \plotone to include an EPS file scaled to
2294: %% 80% of its natural size with \epsscale. Its caption
2295: %% has been written to indicate that additional figure parts will be
2296: %% available in the electronic journal.
2297:
2298:
2299: \begin{figure*}
2300: \epsscale{1.0}
2301: \plotone{f1.ps}
2302: \caption{
2303: \label{fig:psfsim_zeta}
2304: Test of anisotropic PSF correction for the first HOLICs of spin-1,
2305: $\zeta$, based on numerical simulations.
2306: No observational noise and lensing signal is included.
2307: %%%
2308: The values of first HOLICs multiplied with the object detection radius,
2309: $r_g|\zeta|$, are shown as a function of
2310: parameters $A_{\rm aniso}$ ({\it left}: $\theta_{\rm aniso}=0.92\sigma_{iso}$)
2311: and $\theta_{\rm aniso}$ ({\it right}: $A_{\rm aniso}=0.1$)
2312: for a model PSF anisotropy.
2313: %%%%
2314: Solid lines indicate the observed values of first HOLICs ($\zeta^{obs}$)
2315: for Gaussian source images smeared with the model PSF,
2316: and dashed lines indicate the residual values ($\zeta^{iso}$)
2317: after correcting for the spin-1 PSF anisotropy.
2318: No observational noise or lensing signal has been added.
2319: %%%
2320: The PSF consists of an isotropic part $P^{iso}$
2321: described by a truncated Gaussian
2322: with dispersion $\sigma_{iso}$ and an isotropic part
2323: $q(\theta)=A_{aniso}\theta_1/|\btheta|^2$
2324: truncated at $\theta=\theta_{aniso}$.
2325: %%%
2326: Filled circles, open triangles, and crosses
2327: represent the measurements for a Gaussian source of
2328: dispersion $R_{\rm gal}=1\sigma_{iso}, 2\sigma_{iso}, 3\sigma_{iso}$,
2329: respectively.
2330: The $\zeta_{obs}$ is measured with a Gaussian weight function
2331: of dispersion $r_g=\sqrt{R_{\rm gal}^2+\sigma_{iso}^2}$
2332: from the surface brightness distribution smeared with the model PSF.
2333: }
2334: \end{figure*}
2335:
2336:
2337: \begin{figure*}
2338: \epsscale{1.0}
2339: %%%
2340: \plotone{f2.ps}
2341: \caption{
2342: \label{fig:psfsim_zratio}
2343: Ratio of residual to observed spin-1 PSF anisotropy,
2344: $|\zeta^{iso}|/|\zeta^{obs}|$,
2345: as a function of
2346: model parameters $A_{\rm aniso}$ ({\it left}) and $\theta_{\rm aniso}$ ({\it
2347: right}) for the PSF anisotropy kernel.
2348: The PSF consists of an isotropic part $P^{iso}$
2349: described by a truncated Gaussian
2350: with dispersion $\sigma_{iso}$ and an isotropic part
2351: $q(\theta)=A_{aniso}\theta_1/|\btheta|^2$
2352: truncated at $\theta=\theta_{aniso}$.
2353: %%%
2354: Filled circles, open triangles, and crosses
2355: represent the measurements for a Gaussian source of
2356: dispersion $R_{\rm gal}=1\sigma_{iso}, 2\sigma_{iso}, 3\sigma_{iso}$,
2357: respectively.
2358: }
2359: \end{figure*}
2360:
2361:
2362: \begin{figure*}
2363: \epsscale{1.0}
2364: %%%
2365: \plotone{f3.ps}
2366: \caption{
2367: \label{fig:psfsim_delta}
2368: Same as Figure 1 but for the second HOLICs, $\delta$, of spin-3.
2369: }
2370: \end{figure*}
2371:
2372:
2373: \begin{figure*}
2374: \epsscale{1.0}
2375: %%%
2376: \plotone{f4.ps}
2377: \caption{
2378: \label{fig:psfsim_dratio}
2379: Same as Figure 2 but for the second HOLICs, $\delta$, of spin-3.
2380: }
2381: \end{figure*}
2382:
2383:
2384:
2385: \begin{figure*}
2386: \epsscale{1.0}
2387: %%%
2388: \plotone{f5.ps}
2389: \caption{
2390: \label{fig:zetaq_field}
2391: The spin-1 PSF anisotropy field $\zeta^{obs}_*(x,y)$
2392: before ({\it left}) and after ({\it right}) the PSF correction
2393: %measured from stellar shape moments
2394: over the Subaru $i'$-band image of A1689.
2395: The spin-1 PSF anisotropy was measured from stellar shape moments
2396: following the HOLICs formalism outlined in \S \ref{sec:ksb+}.
2397: The orientation of the vectors shows the direction of the spin-1
2398: anisotropy, and the length is proportional to the magnitude of
2399: anisotropy.
2400: A vector of $0.005 {\rm pixel}^{-1}$ is displayed in the inset panel.
2401: %%as displayed in the inset panel.
2402: }
2403: \end{figure*}
2404:
2405:
2406: \begin{figure*}
2407: \epsscale{1.0}
2408: %%%
2409: \plotone{f6.ps}
2410: \caption{
2411: \label{fig:zetaq}
2412: Comparison of spin-1 PSF anisotropy components before ({\it left})
2413: and after ({\it right})
2414: the PSF correction.
2415: }
2416: \end{figure*}
2417:
2418:
2419: \begin{figure*}
2420: \epsscale{1.2}
2421: \plotone{f7.ps}
2422: \caption{
2423: \label{fig:A1689_Emode}
2424: The dimensionless surface mass density $\kappa$
2425: of the galaxy cluster A1689 ($z=0.183$)
2426: in the central $4'\times 4'$ region
2427: reconstructed using the first flexion
2428: observed with Subaru telescope/Suprime-Cam.
2429: %The contours are spaced in units of $1\sigma$
2430: %reconstruction error
2431: %($\approx 0.45$)
2432: The lowest contour and the contour interval are
2433: at a $1\sigma$ level of
2434: the reconstruction error ($\approx 0.51$)
2435: estimated from the rms of the $B$-mode reconstruction.
2436: %
2437: The black, solid circle in the lower-right corner indicates the
2438: Gaussian FWHM ($=0\farcm 33$) used for the mass reconstruction.
2439: }
2440: \end{figure*}
2441:
2442:
2443:
2444: \begin{figure*}
2445: \epsscale{1.2}
2446: \plotone{f8.ps}
2447: \caption{
2448: \label{fig:A1689_Bmode}
2449: Same as Figure \ref{fig:A1689_Emode}
2450: but for the
2451: B-mode reconstruction
2452: from the first flexion measured from Subaru data.
2453: The color scale is the same as in Figure
2454: \ref{fig:A1689_Emode}.
2455: %The lowest contour level is
2456: %set to the mean $B$-mode convergence ($8'\times 8'$ region),
2457: %$\bar{\kappa}_B \approx 0.48$,
2458: The lowest contour and the contour interval are at a
2459: $1\sigma$ level of the B-mode reconstruction
2460: ($\sigma_B\approx 0.51$)
2461: over the $8'\times 8'$ region.
2462: }
2463: \end{figure*}
2464:
2465:
2466: \begin{figure*}
2467: \epsscale{1.0}
2468: \plotone{f9.ps}
2469: \caption{
2470: \label{fig:A1689_image}
2471: %% @@ edited by KU (2008/02/14)
2472: % Subaru $i'$-band image of the central $4'\times 4'$ cluster region.
2473: False-color image of the central $4'\times 4'$ cluster region
2474: composed of the Subaru/Suprime-Cam $V$ and $i'$ images.
2475: %%
2476: Overlayed are contours of the lensing $\kappa$-field reconstructed from
2477: the first flexion measurements using
2478: % Subaru/Suprime-Cam data.
2479: the $i'$-band data.
2480: The contours are spaced in units of $1\sigma (\approx 0.51)$
2481: reconstruction error estimated from the rms of the $B$-mode
2482: reconstruction.
2483: North is to the top, and East to the left.
2484: }
2485: \end{figure*}
2486:
2487:
2488: \end{document}
2489:
2490: %%
2491: %% End of file `sample.tex'.
2492: