0805.2155/ms.tex
1: \documentclass[prd,twocolumn,showpacs,nofootinbib]{revtex4}
2: \usepackage{times}
3: \usepackage{natbib}
4: \usepackage{epsfig}
5: 
6: \newcommand{\apjl}{Astrophys. J. Lett.}
7: \newcommand{\apjs}{Astrophys. J. Suppl. Ser.}
8: \newcommand{\aap}{Astron. Astrophys.}
9: \newcommand{\mnras}{Mon. Not. R. Astron. Soc.}
10: \newcommand{\araa}{Annu. Rev. Astron. Astrophys.}
11: \newcommand{\physrep}{Phys. Rep.}
12: 
13: \newcommand{\mk}{$\bullet$}
14: \newcommand{\etc}{\noindent \mk}
15: \newcommand{\ang}{\hat\nabla}
16: 
17: \newcommand{\bdv}[1]{{\bf #1}}
18: 
19: \newcommand{\up}[1]{{\rm #1}}
20: \newcommand{\kms}{{\rm km\, s}^{-1}}
21: \newcommand{\mpc}{{\rm Mpc}}
22: \newcommand{\hmpc}{{h^{-1}\mpc}}
23: \newcommand{\msun}{M_{\odot}}
24: \newcommand{\hmsun}{{h^{-1}\msun}}
25: \newcommand{\Rvir}{R_\up{vir}}
26: \newcommand{\OM}{\Omega_m}
27: \newcommand{\OB}{\Omega_b}
28: \newcommand{\rms}{\sigma_8}
29: 
30: \newcommand{\vL}{\bdv{l}}
31: \newcommand{\Vang}{\bdv{\hat n}}
32: \newcommand{\Vangm}{\bdv{\hat m}}
33: \newcommand{\lenT}{\tilde T}
34: \newcommand{\spix}{\sigma_\up{pix}}
35: \newcommand{\Npix}{N_\up{pix}}
36: \newcommand{\fsky}{f_\up{sky}}
37: \newcommand{\Tobs}{\tilde T^\up{obs}}
38: \newcommand{\Tobss}{\tilde T^{\up{obs}*}}
39: \newcommand{\That}{\hat T}
40: \newcommand{\lenC}{\tilde C}
41: \newcommand{\Cobs}{\tilde C^\up{obs}}
42: \newcommand{\sbeam}{\sigma_b}
43: \newcommand{\tfwhm}{\theta_\up{FWHM}}
44: \newcommand{\khat}{\hat\kappa}
45: \newcommand{\Cphi}{C^{\phi\phi}}
46: \newcommand{\Ckap}{C^{\kappa\kappa}}
47: \newcommand{\Cdd}{C^{dd}}
48: \newcommand{\Nkap}{N^{\kappa\kappa}}
49: \newcommand{\tE}{\theta_\up{E}}
50: \newcommand{\tEm}{\theta_\up{E}^m}
51: \newcommand{\tEms}{\theta_\up{E}^{m2}}
52: \newcommand{\tEst}{\theta_\up{E}^\star}
53: \newcommand{\tEhat}{\hat\tE}
54: \newcommand{\tEQE}{\hat\tE^\up{QE}}
55: \newcommand{\tEML}{\hat\tE^\up{ML}}
56: \newcommand{\tEopt}{\hat\tE^\up{opt}}
57: \newcommand{\beeq}{\vspace{10pt}\begin{equation}} 
58: \newcommand{\eneq}{\vspace{10pt}\end{equation}}
59: \newcommand{\bear}{\vspace{10pt}\begin{eqnarray}}
60: \newcommand{\enar}{\end{eqnarray}{\\\\ \noindent}}
61: \newcommand{\Fisher}{\mathcal{F}}
62: \newcommand{\Lik}{\mathcal{L}}
63: \newcommand{\lcut}{l_\up{cut}}
64: \newcommand{\muK}{\mu\up{K}}
65: \newcommand{\Minit}{M_\up{init}}
66: \newcommand{\bhat}{\bdv{\hat s}}
67: \newcommand{\dsmall}{\bdv{\hat\Delta}}
68: 
69: 
70: \begin{document}
71: 
72: \title{Improved estimation of cluster mass profiles 
73: from the cosmic microwave background}
74: 
75: \author{Jaiyul Yoo$^1$}
76: \altaffiliation{Electronic address: jyoo@cfa.harvard.edu} 
77: \author{Matias Zaldarriaga$^{1,2}$}
78: \affiliation{$^1$Harvard-Smithsonian Center for Astrophysics, Harvard 
79: University, 60 Garden Street, Cambridge, MA 02138}
80: \affiliation{$^2$Jefferson Physical Laboratory, Harvard University, 
81: 17 Oxford Street, Cambridge, MA 02138}
82: 
83: \begin{abstract}
84: We develop a new method for reconstructing cluster mass profiles and 
85: large-scale structure from the cosmic microwave background (CMB). By analyzing
86: the likelihood of CMB lensing, we analytically prove that standard quadratic
87: estimators for CMB lensing are unbiased and achieve the optimal condition
88: only in the limit of no lensing; they become progressively biased and
89: sub-optimal, when the lensing effect is large, especially for clusters that
90: can be found by ongoing Sunyaev-Zel'dovich surveys. Adopting an alternative
91: approach to the CMB likelihood, we construct a new maximum likelihood 
92: estimator that utilizes delensed CMB temperature fields based on an assumed
93: model. We analytically show that this estimator asymptotically approaches
94: the optimal condition as our assumed model is refined, and we numerically
95: show that as we iteratively apply it to CMB maps
96: our estimator quickly converges to the true model with a factor of ten
97: less number of clusters than standard quadratic estimators need.
98: For realistic CMB experiments, we demonstrate the 
99: applicability of the maximum likelihood estimator with tests against
100: numerical simulations in the presence of CMB secondary contaminants. With 
101: significant improvement on the signal-to-noise ratio, our new maximum 
102: likelihood estimator can be used to measure the cluster-mass 
103: cross-correlation functions at different redshifts, probing the evolution
104: of dark energy. 
105: \end{abstract}
106: 
107: \pacs{98.62.Sb, 98.70.Vc, 98.80.Es}
108: 
109: \maketitle
110: 
111: \section{Introduction}
112: \label{sec:intro}
113: As the most distant observable sources, the cosmic microwave background (CMB)
114: anisotropies provide a unique channel to probe the universe after the 
115: cosmological recombination epoch. In particular, weak gravitational lensing of 
116: the CMB can be used to map the matter distribution in the universe at higher 
117: redshift than weak lensing of faint background galaxies can ever achieve. 
118: Recent work \citep{HIPAET04,SMZADO07,HIHOET08,CASLET08}
119: has focused on measuring the lensing signature in the CMB
120: by large-scale structure between the last scattering surface and the present
121: universe, but relatively little attention has been paid to weak lensing 
122: of the CMB by clusters of galaxies.
123: 
124: The abundance of massive clusters is exponentially sensitive to the growth of 
125: the underlying matter distribution, and hence it has been recognized as a
126: powerful probe of the evolution of dark energy (e.g., \citep{DETF06}).
127: However, the constraining power as a cosmological probe
128: can be only realized, if the cluster masses are accurately measured.
129: To achieve this goal, many cluster surveys are designed to detect massive
130: clusters and measure their mass using the Sunyaev-Zel'dovich (SZ)
131: effect, and some of the planned
132: surveys are already operational using the South Pole Telescope (SPT),
133: and the Atacama Cosmology Telescope (ACT).
134: Weak lensing of the CMB can be applied to the same clusters found in the
135: SZ surveys without additional observations,
136: providing independent measurements of their mass. 
137: Furthermore, the CMB provides the highest redshift source plane with precision 
138: measurements of its distance, which can be combined with galaxy weak lensing 
139: measurements of the same lensing clusters to obtain angular diameter 
140: distance ratio estimates that are independent of the mass distribution,
141: substantially increasing the leverage to constrain cosmological
142: parameters \citep{HUHOVA07}.
143: 
144: Gravitational lensing by clusters imprints a unique signature in the CMB 
145: anisotropies. On arcminute scales, the primordial CMB anisotropies decay
146: exponentially due to the photon diffusion from the baryon-photon fluid around
147: the recombination epoch \citep{SILK68},
148: and to a good approximation the CMB can be considered as a pure temperature
149: gradient 
150: on small scales. Based on this approximation, \citet{SEZA95} showed that 
151: clusters create dipole-like wiggles in the CMB temperature by remapping the
152: otherwise smooth gradient field,
153: and this unique feature can be used to isolate the lensing effect
154: by clusters and to reconstruct the deflection angle, once the temperature
155: gradient is separately measured on large scales.
156: \citet*{VAAMWH04} and \citet{HOKO04} used $N$-body simulations 
157: to model realistic lensing clusters, and they found that 
158: the mass reconstruction for individual clusters is compromised,
159: since it is hard to measure the large-scale temperature gradient accurately
160: and secondary anisotropies in the CMB can partially 
161: mimic the lensing signature. 
162: 
163: However, it has been realized that one can apply the same technique developed 
164: for reconstructing large-scale structure to clusters of galaxies, measuring
165: the statistical properties of a sample of clusters.
166: Unlike galaxy weak lensing, CMB anisotropies have no characteristic shape, 
167: even statistically, from which the deviation is a measure of the lensing 
168: effect. Gravitational lensing, however, gives rise to a deviation 
169: of the two-point correlation function of the CMB temperature anisotropies
170: from statistical isotropy. 
171: The standard technique is to construct a lensing estimator that is quadratic in 
172: observed temperature anisotropies, measuring the  
173: correlation between different Fourier modes, which is directly proportional
174: to the lensing effect \citep{HU01b}.
175: 
176: This method is easy to implement in analyzing real data compared to the full
177: likelihood analysis \citep{HISE03a} and no separate measurement is required
178: to obtain the large-scale temperature gradient. However,
179: \citet{MABAMEET05} showed that standard quadratic estimators need a 
180: modification to be an unbiased estimator in a region around massive clusters.
181: \citet*{HUDEVA07} quantitatively demonstrated that
182: standard quadratic estimators based on the linear approximation ignore
183:  higher-order terms in the lensing effect that coherently contribute to
184: the lensing reconstruction, and hence the reconstruction is biased low when 
185: the lensing effect is large. Furthermore, they proposed modified quadratic 
186: estimators that remove the higher-order terms in violation of the linear
187: approximation by low-pass filtering observed temperature fields, and they
188: showed that the modified quadratic estimators recover cluster mass profiles
189: with no significant bias.
190: However, the cutoff scale of the low-pass filter is somewhat arbitrary and
191: it depends on the lensing effect, which we want to measure with the estimators.
192: 
193: Here we develop a new maximum likelihood estimator for reconstructing cluster
194: mass profiles and large-scale structure by analyzing the likelihood of CMB
195: lensing. Our approach is similar in making full use of the likelihood 
196: information to one advocated by \citet{HISE03a}. While they derive an 
197: analytic expression for a maximum likelihood estimator, it is impractical to
198: apply to a realistic problem, because the solution is too general and 
199: computationally expensive. However, our maximum likelihood estimator is 
200: different from theirs and it is easy
201: to use in practice, because we adopt an alternative approach to setting up
202: the likelihood: it takes a similar form of standard quadratic estimators and it 
203: approaches the optimal condition as it is iteratively applied to CMB maps.
204: Furthermore, we show that our maximum likelihood estimator can reconstruct
205: cluster mass profiles with a factor of ten less number of clusters than 
206: standard or modified quadratic estimators need.
207: 
208: The rest of the paper is organized as follows.
209: We first derive a quadratic
210: estimator, accounting for the telescope beam effect
211: in Sec.~\ref{sec:formalism}. This consideration
212: makes a difference compared to the usual practice in the literature,
213: where quadratic estimators are 
214: often applied to beam deconvolved CMB maps.
215: In Sec.~\ref{sec:mle} we analytically show that the
216: quadratic estimators are unbiased and optimal only when the lensing effect 
217: vanishes, and why the modified quadratic estimators outperform the 
218: standard quadratic 
219: estimators when the lensing effect is large. Based on this observation, we 
220: construct a delensed temperature field and derive a maximum likelihood 
221: estimator using the delensed temperature field. 
222: We demonstrate its applicability to realistic CMB experiments using numerical
223: simulations in Sec.~\ref{sec:num}. 
224: We discuss the impact
225: of the telescope beam and instrumental noise in the delensing process
226: and we conclude in Sec.~\ref{sec:con}.
227: 
228: In this paper we will only
229: consider lensing estimators based on CMB temperature
230: anisotropies, since the planned surveys are not yet sensitive to CMB
231: polarization anisotropies on arcminute scales.
232: However, it is straightforward to extend our formalism
233: to lensing estimators based on CMB polarization anisotropies. 
234: Throughout the paper
235: we assume a flat $\Lambda$CDM universe with the matter density 
236: parameter $\OM h^2=0.127$, the baryon density parameter $\OB h^2=0.0222$, the 
237: Hubble constant $h=0.73$, the spectral index $n_s=0.95$, the optical depth to
238: the last scattering surface $\tau=0.09$, and the primordial
239: curvature perturbation amplitude $A_s=2.5\times10^{-9}$ (corresponding to the
240: matter power spectrum normalization $\rms=0.75$), consistent with the recent
241: cosmological parameter estimation (e.g., \citep{TESTET06,SPBEET07,KODUET08})
242: 
243: \section{Formalism}
244: \label{sec:formalism}
245: Here we describe our notations for weak lensing of the
246: CMB and derive a quadratic estimator for CMB lensing reconstruction.
247: 
248: \subsection{Weak Lensing of the CMB}
249: Gravitational lensing deflects light rays as they propagate through
250: fluctuating gravitational fields, and the deflection vector 
251: $\bdv{d}(\Vang)$ at the angular position $\Vang$ on the sky is related to
252: the line-of-sight projection of the gravitational potential $\psi$ as
253: $\bdv{d}(\Vang)=\bdv{\ang}\phi(\Vang)$, where the projected potential is
254: \beeq
255: \phi(\Vang)=-2\int_0^{D_\star}\!\! 
256: dD~{D_\star-D\over D D_\star}~\psi(D\Vang,D),
257: \eneq
258: $\ang$ is the derivative with respect to $\Vang$, 
259: and $D_\star$ is the comoving angular diameter distance to the last scattering
260: surface. Here we have assumed a flat universe and $c\equiv1$. 
261: The projected potential is further
262: related to the convergence $\kappa$ as $\ang^2\phi(\Vang)=-2\kappa(\Vang)$.
263: 
264: 
265: Since gravitational lensing conserves the surface brightness of diffuse 
266: backgrounds, the lensed temperature field $\lenT(\Vang)$ 
267: of the CMB is simply the intrinsic (unlensed)
268: temperature field $T(\Vang)$ remapped by
269: the deflection vector,
270: \beeq
271: \lenT(\Vang)=T\left[\Vang+\ang\phi(\Vang)\right].
272: \label{eq:lensing}
273: \eneq
274: We will use notation with (or without) tilde to represent lensed (or unlensed)
275: quantities.
276: Note that we mainly work in the Rayleigh-Jeans tail and express the surface 
277: brightness in terms of temperature.
278: 
279: In a sufficiently small patch of the sky, it significantly simplifies the
280: manipulations to work in Fourier space
281: \citep[see][for all-sky formalism]{HU00,OKHU03,CHLE05}. In Fourier space
282: the lensed temperature is 
283: \bear
284: \label{eq:lensexp}
285: \lenT_\vL&=&\int d^2\Vang~\lenT(\Vang)~e^{-i\vL\cdot\Vang} \\ 
286: &=&T_\vL-\int{d^2\vL'\over(2\pi)^2}\left[(\vL-\vL')\cdot\vL'\right]T_{\vL'}
287: \phi_{\vL-\vL'}+\cdots, \nonumber
288: \enar
289: where we Taylor expanded $\lenT_\vL$ to the first order in $\phi_\vL$.
290: We kept the same notation for Fourier components, while the functional
291: dependence is indicated as a subscript (e.g., $T(\Vang)$ and $T_\vL$ are
292: Fourier counterparts).
293: The rms deflection angle $\langle\bdv{d}\cdot\bdv{d}\rangle^{1/2}$
294: is a few arcminutes and the deflection power peaks at a few degree scale,
295: comparable to the angular sizes of clusters.
296: However, the large-scale deflection field is coherent over the
297: scales of the temperature fluctuations, resulting in an unobservable overall
298: shift of the temperature field \citep{MAZA06}, and the linear approximation
299: remains valid. In Sec.~\ref{sec:mle} we discuss the limitation of this 
300: approximation when the lensing effect is large in a region around massive
301: clusters.
302: 
303: Since the intrinsic CMB is Gaussian and isotropic,
304: the statistical properties
305: of the temperature field can be completely described by the power spectrum
306: $C_l$,
307: \beeq
308: \langle T_{\vL_1}T^*_{\vL_2}\rangle=(2\pi)^2~\delta(\vL_1-\vL_2)~C_{l_1},
309: \eneq
310: where the asterisk represents complex conjugation and $\delta$ is the
311: Dirac delta function.
312: Analogously, we define the projected potential power spectrum $\Cphi_l$. 
313: Thus the deflection and the convergence
314: power spectra are $\Cdd_l=l^2\Cphi_l$ and $\Ckap_l=l^4\Cphi_l/4$, respectively.
315: Note that $\Cphi_l$ can always be defined in this way, though it may be an
316: incomplete description of the statistical properties of the projected potential
317: when $\phi_\vL$ is non-Gaussian. 
318: Finally, the power spectrum of the lensed temperature field is 
319: \beeq
320: \tilde C_l=\left[1-l^2R\right]C_l+\int{d^2\vL'\over(2\pi)^2}
321: \left[(\vL-\vL')\cdot\vL'\right]^2C_{l-l'}\Cphi_{l'},
322: \eneq
323: where $R\equiv(1/4\pi)\int d\ln l~l^4\Cphi_l$ 
324: is the half of the rms deflection angle \citep{HU00,LECH06}.
325: 
326: In practice, the observed temperature field has two additional contributions:
327: detector noise independent of the signal, and telescope beam
328: convolving the signals from different patches of the sky.
329: We assume that the detector noise is white, so that the noise
330: power spectrum is constant,
331: \beeq
332: C_l^N\equiv\Delta_T^2=\spix^2{4\pi \fsky\over \Npix},
333: \eneq
334: where $\spix$ is the rms error in each pixel of the detector 
335: in units of $\mu$K, 
336: $\fsky$ is the fraction of the survey area on the sky, and $\Npix$ is the
337: total number of detector pixels \citep{KNOX95}.
338: Convolution is simply a multiplication in Fourier space, and the
339: beam factor for a simple Gaussian beam we consider is
340: $B_l=\exp\left[-{1\over2}l^2\sbeam^2\right]$.
341: The beam width $\sbeam$ is related to the full-width half-maximum (FWHM) as
342: $\sbeam=\tfwhm/\sqrt{8\ln2}$.
343: The observed temperature field and its power spectrum are then
344: \bear
345: \label{eq:tobs}
346: \Tobs_\vL&=&\lenT_\vL~e^{-{1\over2}l^2\sbeam^2}+T^N_\vL,\\
347: \Cobs_l&=&\tilde C_l~e^{-l^2\sbeam^2}+C^N_l.
348: \enar
349: In reality, one needs to consider other contributions to $\Tobs$, such as
350: residual foregrounds, point radio sources, and 
351: CMB secondary anisotropies. We will
352: only consider secondary contributions in Sec.~\ref{ssec:sz}.
353: 
354: \subsection{Quadratic Estimator}
355: \label{ssec:qe}
356: Here we consider a convergence estimator $\khat(\Vang)$ that is quadratic
357: in the observed temperature field, accounting for  telescope beam and
358: detector noise.\footnote{We will use quantities with hat to represent
359: estimators of the quantities without hat, e.g., a convergence estimator
360: is denoted as $\khat$ and a true convergence field is denoted as $\kappa$.
361: However, this notational convention should not be confused with that
362: used for temperature fields: $T$, $\lenT$, $\Tobs$, and $\That$ represent
363: the intrinsic (unlensed), the lensed [Eq.~(\ref{eq:lensing})], the
364: observed [Eq.~(\ref{eq:tobs})], and the delensed [Eq.~(\ref{eq:delensT})]
365: temperature fields, respectively.}
366: We require that the estimator be unbiased when averaged
367: over an ensemble of CMB maps, $\langle\khat(\Vang)\rangle=\kappa(\Vang)$.
368: With these conditions, the estimator takes the general form in Fourier space
369: \beeq
370: \khat_\bdv{L}={N_L\over2}\int{d^2\vL_1\over(2\pi)^2}~F(\vL_1,\vL_2)~
371: \Tobs_{\vL_1}~\Tobs_{\vL_2},
372: \label{eq:khat}
373: \eneq
374: where $\vL_2=\bdv{L-l}_1$ and $N_L$ is a normalization coefficient, which
375: only depends on $L=|\bdv{L}|$.
376: The functional form of $F(\vL_1,\vL_2)$ can be obtained by minimizing the
377: variance of $\khat_\bdv{L}$ and imposing the normalization condition 
378: \beeq
379: F(\vL_1,\vL_2)={\left[\bdv{L}\cdot\vL_1C_{l_1}+
380: \bdv{L}\cdot\vL_2C_{l_2}\right]\over2~\Cobs_{l_1}\Cobs_{l_2}}~
381: e^{-{1\over2}l_1^2\sbeam^2}~e^{-{1\over2}l_2^2\sbeam^2},
382: \label{eq:fctform}
383: \eneq
384: and the normalization coefficient is
385: \beeq
386: {1\over N_L}={1\over L^2}\int{d^2\vL_1\over(2\pi)^2}
387: {\left[\bdv{L}\cdot\vL_1C_{l_1}+\bdv{L}\cdot\vL_2C_{l_2}\right]^2
388: \over2~\Cobs_{l_1}\Cobs_{l_2}}~e^{-l_1^2\sbeam^2}~e^{-l_2^2\sbeam^2}.
389: \eneq
390: Finally, the variance of the estimator is
391: \beeq
392: \langle\khat_\bdv{L}\khat^*_\bdv{L'}\rangle=(2\pi)^2~\delta(\bdv{L-L'})
393: (\Ckap_L+\Nkap_L),
394: \eneq
395: where $\Nkap_L=L^2N_L/4$ is the noise power spectrum of $\khat_\bdv{L}$.
396: One can think of 
397: $\Ckap_L/N_L^{\kappa\kappa}$ as a signal-to-noise ratio, and the reconstruction
398: becomes difficult at the angular scale $L$, where
399: $\Ckap_L\simeq N_L^{\kappa\kappa}$.
400: Given experimental specifications, the noise power spectrum 
401: $\Nkap_L$, as a function of the intrinsic CMB power spectrum $C_L$,
402: becomes smallest, when there exists substantial power in
403: $C_L$ at the scale of interest,
404: with its shape deviating from the scale-invariance ($L^2C_L=$constant)
405: \citep{ZAZA06}.
406: 
407: Our estimator recovers the general form of the standard quadratic estimators
408: as $\sbeam\rightarrow0$, and
409: $N_L$ corresponds to the noise power spectrum of a deflection
410: estimator $\hat\bdv{d}_\bdv{L}=2\bdv{L}~\khat_\bdv{L}/L^2$ 
411: used in the literature
412: \citep{HU01b}.
413: 
414: The estimator can be decomposed as two Wiener-filtered temperature functions
415: in real space, which essentially correlates the gradient of the lensed
416: temperature field with the unlensed temperature field to isolate the lensing
417: effect,
418: \bear
419: \label{eq:G}
420: \bdv{G}(\Vang)&=&\int{d^2\vL\over(2\pi)^2}~i\vL~\Tobs_\vL{C_l\over\Cobs_l}~
421: e^{-{1\over2}l^2\sbeam^2+i\vL\cdot\Vang} \\
422: \label{eq:W}
423: W(\Vang)&=&\int{d^2\vL\over(2\pi)^2}~\Tobs_\vL{1\over\Cobs_l}~
424: e^{-{1\over2}l^2\sbeam^2+i\vL\cdot\Vang},
425: \enar
426: and the convergence estimator can be expressed in terms of 
427: $\bdv{G}(\Vang)$ and $W(\Vang)$ as
428: \beeq
429: \khat_\bdv{L}=-{N_L\over2}i\bdv{L}\cdot\int d^2\Vang~\bdv{G}(\Vang)W(\Vang)
430: ~e^{-i\bdv{L}\cdot\Vang}.
431: \label{eq:GW}
432: \eneq
433: This approach of using the two Wiener-filtered functions is more convenient 
434: for computing $\khat_\bdv{L}$ by using Fast Fourier Transform (FFT) routines 
435: than by directly computing Eq.~(\ref{eq:khat}). Furthermore, it is more 
436: physically intuitive than the general derivation, though the latter has clear
437: advantage in its transparency and understanding
438: the uniqueness of the functional form $F(\vL_1,\vL_2)$.
439: A modified quadratic estimator can be constructed by removing the signals
440: in Eq.~(\ref{eq:G}) at $l\geq\lcut$, while Eq.~(\ref{eq:W}) remains unchanged.
441: 
442: \begin{figure}
443: \centerline{\psfig{file=f1.eps, width=3.0in}}
444: \caption{Convolution filter $H(\theta)$ as a function of separation 
445: $\theta=|\Vang|$ for CMB experiments
446: with $\spix=5\muK$ and~$10\muK$ in Sec.~\ref{sec:num}.
447: The insets show details of $H(\theta)$ at the center ($left$)
448: and its tail ($right$).}
449: \label{fig:filter}
450: \end{figure}
451: 
452: To better understand how quadratic estimators operate, we Fourier transform and
453: rearrange Eq.~(\ref{eq:GW}) as
454: \bear
455: \label{eq:div}
456: {1\over2}\ang\cdot\left[\bdv{G}(\Vang)W(\Vang)\right]&=&
457: \int{d^2\bdv{L}\over(2\pi)^2}{-\khat_\bdv{L}\over N_L}~
458: e^{i\bdv{L}\cdot\Vang} \\
459: &=&\int d^2\Vangm~H(\Vangm-\Vang)\khat(\Vangm). \nonumber
460: \enar
461: The divergence of the two Wiener-Filtered functions is a convolution of the
462: convergence estimate $\khat(\Vang)$ and the filter
463: \beeq
464: H(\Vang)=\int{d^2\bdv{L}\over(2\pi)^2}~{-1\over N_L}~e^{i\bdv{L}\cdot\Vang}.
465: \label{eq:filterH}
466: \eneq
467: Figure~\ref{fig:filter} plots the filter $H(\theta)$ as a function of
468: separation $\theta=|\Vang|$ for experiments with $\spix=5\muK$ and~$10\muK$,
469: to which we apply quadratic estimators in Sec.~\ref{sec:num}. The filter
470: peaks at the center and its width is $\simeq3'$, roughly set by the scale 
471: that the intrinsic CMB and detector noise power spectra become comparable.
472: While the filter is highly oscillating at its tail, it is negligible
473: at $\theta\geq10'$ due to the large weight near the center.
474: A factor of two change in $\spix$ has little impact on the width of the 
475: filter, because the crossing scale is already at the CMB damping tail.
476: 
477: \section{Maximum Likelihood Estimator}
478: \label{sec:mle}
479: In this section, we analyze the likelihood of CMB lensing by singular
480: isothermal clusters. We first derive a quadratic estimator for singular 
481: isothermal clusters and compare the estimator to the optimal estimator
482: from the likelihood. With the simple singular isothermal model, 
483: our analysis will
484: be carried out analytically, showing that
485: (1)~the standard quadratic estimators are
486: unbiased and optimal in the limit of no lensing, (2)~they progressively become
487: biased and sub-optimal when the lensing effect increases, 
488: and (3)~why the modified quadratic
489: estimators perform better than the standard quadratic estimators. Finally, we
490: develop a unbiased maximum likelihood estimator to reconstruct 
491: cluster mass profiles as well as large-scale structure.
492: We demonstrate its applicability to CMB experiments with tests
493: against numerical simulations using more realistic cluster models
494: in Sec.~\ref{sec:num}.
495: 
496: \subsection{Quadratic Estimator for a Singular Isothermal Cluster}
497: A singular isothermal cluster has a density profile $\rho(r)\propto r^{-2}$
498: and its enclosed mass increases with $r$, which requires truncation at some
499: radius to be a viable model for real clusters. However, this
500: model has advantage in its simplicity: its 
501: properties are described by one parameter, Einstein radius
502: 
503: \beeq
504: \tE=4\pi\sigma^2~{D_\star-D_L\over D_\star},
505: \eneq
506: where $\sigma$ is one-dimensional velocity dispersion of a cluster
507: and $D_L$ is
508: the comoving angular diameter distance to the lensing cluster. CMB
509: lensing has a well-defined single plane of the source redshift and
510: the comoving angular diameter distance to the last scattering surface 
511: $D_\star=14.12$~Gpc is now measured with less than 1\% uncertainty 
512: \citep{KODUET08}.
513: The convergence is $\kappa(\Vang)=\tE/2\theta$ and the deflection 
514: vector is $\bdv{d}(\Vang)=-\tE\Vang$ given the angular separation 
515: $\theta=|\Vang|$ from the origin
516: in a cluster centric coordinate. When a virial radius $\Rvir$
517: is defined as the radius inside which the mean density is 200 times the 
518: cosmic mean matter density,
519: a singular isothermal cluster of mass $M=10^{14}\hmsun$ 
520: within the virial
521: radius at $z_L=1$ has an Einstein radius $\tE=8.\!\!''0$ and a velocity
522: dispersion $\sigma=2.0\times10^{-3}(=610~\kms)$, 
523: and they scale as $\tE\propto M^{2/3}$ and $\sigma\propto M^{1/3}$.
524: 
525: A quadratic estimator $\tEQE$
526: for singular isothermal clusters can be readily derived
527: using the method described in Sec.~\ref{ssec:qe}, but here we take an idealized
528: approach for the purpose of comparison, where we assume $\spix=\sbeam=0$.
529: Under the condition
530: that the estimator is unbiased $\langle\tEQE\rangle=\tE$ and it has the
531: minimum variance, the quadratic estimator is
532: \bear
533: \label{eq:teqe}
534: \tEQE&=&{1\over\Fisher}\int{d^2\vL_1\over(2\pi)^2}\int{d^2\vL_2\over(2\pi)^2}
535: \\
536: &\times&{\lenT_{\vL_1}\lenT_{\vL_2}\over \lenC_{l_1}\lenC_{l_2}}
537: {\pi(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)\over|\vL_1+\vL_2|^3},
538: \nonumber
539: \enar
540: with the normalization coefficient 
541: \bear
542: \label{eq:fisher}
543: \Fisher&=&\int{d^2\vL_1\over(2\pi)^2}\int{d^2\vL_2\over(2\pi)^2}\\
544: &\times&{2\pi^2\over \lenC_{l_1}\lenC_{l_2}} 
545: \left[{(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)
546: \over|\vL_1+\vL_2|^3}\right]^2. \nonumber
547: \enar
548: The variance of the estimator is 
549: $\langle(\tEQE-\tE)(\tEQE-\tE)\rangle=1/\Fisher$.
550: Here we Taylor expanded $\lenT_\vL$ and kept terms only to the first order
551: in $\tE$ in deriving $\tEQE$.
552: 
553: \subsection{Relation to the Optimal Estimator}
554: \label{ssec:optimal}
555: The likelihood function $P(\lenT|\tEm)$ simply represents the probability that
556: a singular isothermal model 
557: with $\tEm$ can have the lensed temperature field $\lenT(\Vang)$. Since the
558: intrinsic CMB follows a Gaussian distribution and gravitational lensing only
559: remaps the intrinsic CMB, the distribution of $\lenT(\Vang)$ is also
560: Gaussian and its statistical properties are fully described by the covariance
561: matrix of $\lenT(\Vang)$
562: \beeq
563: \lenC(\Vang,\Vang')=\langle\lenT(\Vang)\lenT(\Vang')\rangle
564: =\int{d^2\vL\over(2\pi)^2}~\lenC_l ~e^{i\vL\cdot(\Vang-\Vang')}.
565: \eneq
566: For convenience, we take a negative logarithm of $P(\lenT|\tEm)$ and call it
567: likelihood,
568: \bear
569: \label{eq:likeli}
570: \Lik(\lenT|\tEm)&\equiv&-\ln P(\lenT|\tEm) \\
571: &=&{1\over2}\lenT(\Vang)~\lenC^{-1}(\Vang,\Vang'|\tEm)~\lenT(\Vang')+
572: {1\over2}\ln\det\lenC(\tEm), \nonumber
573: \enar
574: where the summation over $\Vang$ and $\Vang'$ is implicitly assumed and 
575: hereafter we will suppress the angular dependence for simplicity.
576: In general, the likelihood is a functional with its argument of a scalar field,
577: such as $\kappa(\Vang)$ or $\phi(\Vang)$. However, in our case
578: it reduces to a function 
579: with its argument of a scalar $\tEm$, substantially simplifying
580: the manipulation.
581: 
582: We take a derivative of $\Lik$ with respect to $\tEm$,
583: \bear
584: \label{eq:first}
585: {\partial\Lik\over\partial\tEm}&=&-{1\over2}\lenT~\lenC^{-1}~
586: {\partial\lenC\over\partial\tEm}~\lenC^{-1}~\lenT \\
587: &=&-\int\!\!\!{d^2\vL_1\over(2\pi)^2}\int\!\!\!{d^2\vL_2\over(2\pi)^2}
588: {\lenT_{\vL_1}\lenT_{\vL_2}\over\lenC_{l_1}\lenC_{l_2}}
589: {\pi(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)\over|\vL_1+\vL_2|^3}, \nonumber
590: \enar
591: where we computed the derivative to the first order in $\tEm$.
592: Since gravitational lensing only redistributes the intrinsic CMB,
593: the last term (log determinant) in Eq.~(\ref{eq:likeli}) is 
594: independent of $\tEm$ and
595: hence the derivative with respect to $\tEm$ vanishes in Eq.~(\ref{eq:first}).
596: However, in the presence of non-white instrumental noise, 
597: and/or other secondary
598: contaminants, the derivative acquires a nonzero value but it is in general 
599: negligible compared to the quadratic term in Eq.~(\ref{eq:first}). We will
600: neglect this effect in the remainder of this paper.
601: In the presence of significant contaminants
602: from secondaries, the assumption that the likelihood function is Gaussian
603: becomes invalid before the log determinant term becomes non-negligible.
604: 
605: With the derivative of $\Lik$, we can compute the Fisher information matrix
606: \beeq
607: \label{eq:fisher2}
608: \Fisher=\left\langle{\partial^2\Lik\over\partial\tEms}\right\rangle
609: =\left\langle{\partial\Lik\over\partial\tEm}{\partial\Lik\over\partial\tEm}
610: \right\rangle 
611: \eneq
612: where for the second equality we used the normalization condition of the
613: likelihood function
614: $1=\int d\lenT~ P(\lenT|\tEm)=\int d\lenT~ e^{-\Lik}$. Within the Gaussian
615: approximation, $\Fisher$ can be evaluated at any value of $\tEm$.
616: Note that $\Fisher$ is identical to the normalization coefficient in 
617: Eq.~(\ref{eq:fisher}). 
618: 
619: In statistical parameter estimation, there exists a powerful theorem, known
620: as the Cram\'er-Rao inequality that error bars in a parameter estimation
621: have a definite lower bound  $\sigma(\tEm)\geq\Fisher^{-1/2}$
622: set by the Fisher matrix.
623: Moreover, this theorem provides a necessary and sufficient condition for an
624: estimator to saturate the Cram\'er-Rao inequality, i.e., to be an optimal 
625: estimator $\tEopt$ \citep{BABIC05},
626: \beeq
627: {\partial\Lik\over\partial\tEm}=\Fisher~(\tEm-\tEopt).
628: \label{eq:optimality}
629: \eneq
630: Now it is apparent that only in the limit of no lensing 
631: (the true Einstein radius $\tE=\tEm=0$) does the
632: quadratic estimator $\tEQE$ become an optimal estimator $\tEopt$ with the
633: smallest variance attainable from the data. Conversely, $\tEQE$ becomes
634: progressively biased and sub-optimal as the lensing effect increases.
635: This can be also understood by the validity of the linear approximation:
636: since the quadratic estimator is constructed to be unbiased and to minimize
637: the variance when $\lenT_\vL$ is expanded to the linear order in $\phi_\vL$, 
638: it is natural to expect that this condition breaks down when higher-order terms
639: in $\phi_\vL$ become dominant over the linear order term. The modified 
640: quadratic estimator, on the other hand, removes the angular modes of the
641: signals at $l\geq\lcut$ by explicitly setting the integrand zero
642: in Eq.~(\ref{eq:teqe}), where the linear approximation breaks down, and this 
643: process helps suppress the contributions from the higher-order terms 
644: in $\phi_\vL$
645: because the higher-order terms are related to multiple integrals over the modes
646: that are suppressed most. Precisely for this reason could the modified
647: quadratic estimators be more robust than the standard quadratic estimators
648: even when the lensing effect is large.
649: 
650: However, the modified quadratic estimator requires a rather arbitrary choice 
651: of the cutoff scale $\lcut$, which depends on the lensing effect,
652: though it may be possible to calibrate against simulations \citep{HUDEVA07}. 
653: Furthermore, the removal of the lensing signals at $l\geq\lcut$ inevitably
654: results in lower signal-to-noise ratio, making the reconstruction noisier.
655: We discuss this issue with numerical simulations in Sec.~\ref{ssec:com}.
656: 
657: 
658: \subsection{Maximum Likelihood Estimator}
659: Given the Gaussian probability distribution of the CMB, 
660: the likelihood retains all the information of the observed data.
661: Even when there exists no optimal estimator, one can always find an estimator,
662: if not analytically, that maximizes the likelihood:
663: the maximum likelihood estimator $\tEML$ is the solution of
664: \beeq
665: {\partial\Lik\over\partial\tEm}\Bigg|_{\tEm=\tEML}=0.
666: \label{eq:mle}
667: \eneq
668: However, this equation is highly non-linear in general and requires 
669: approximations to be solved even numerically. 
670: Equations~(\ref{eq:optimality}) and~(\ref{eq:mle}) show that an optimal 
671: estimator is always the maximum likelihood estimator. However, note that
672: while the converse is not true in general, the 
673: maximum likelihood estimator asymptotically 
674: approaches to the optimal condition. 
675: 
676: Having understood that the quadratic estimator becomes an optimal (and maximum
677: likelihood) estimator in the limit of no lensing in Sec.~\ref{ssec:optimal},
678: we present an alternative approach to modeling the likelihood and derive a
679: new maximum likelihood estimator for singular isothermal clusters. We then 
680: generalize this approach to clusters with arbitrary mass distributions.
681: 
682: Consider a model with $\tEm$ and its deflection field 
683: $\bdv{d}^m(\Vang)=-\tEm\Vang$.
684: We construct a delensed
685: temperature field $\That(\Vang)$ by delensing the observed
686: $\lenT(\Vang)$ with $\bdv{d}^m(\Vang)$, and $\That(\Vang)$ is related to
687: the intrinsic temperature field $T(\Vang)$ as
688: \bear
689: \label{eq:delensT}
690: \That(\Vang)&\equiv&\lenT(\Vang-\bdv{d}^m) \\
691: &=&T(\Vang-\bdv{d}^m+\bdv{d})=T\left[(1+\Delta)\Vang\right], \nonumber
692: \enar
693: with $\Delta=\tEm-\tE$. Now we can write the likelihood in terms of the 
694: delensed temperature field $\That(\Vang)$
695: \beeq
696: \Lik(\That|\tEm)={1\over2}\That(\tEm)~C^{-1}~\That(\tEm)+{1\over2}\ln\det C,
697: \eneq
698: where we emphasized the dependence of $\That(\Vang)$ on $\tEm$,
699: and $C$ is the covariance matrix of $T(\Vang)$.
700: Taking a derivative of $\Lik$ with respect to $\tEm$ gives
701: \bear
702: {\partial\Lik\over\partial\tEm}&=&{1\over2}\left[{\partial\That\over\partial
703: \tEm}C^{-1}~\That+\That~ C^{-1}{\partial\That\over\partial\tEm}\right] \\
704: &=&-\int\!\!\!{d^2\vL_1\over(2\pi)^2}\int\!\!\!{d^2\vL_2\over(2\pi)^2}
705: {T_{\vL_1}T_{\vL_2}\over C_{l_1}C_{l_2}}
706: {\pi(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)\over|\vL_1+\vL_2|^3}. 
707: \nonumber
708: \enar
709: The second equality is obtained by evaluating the derivative at $\Delta=0$.
710: Assuming that our initial
711: model with $\tEst$ is a good approximation to the true model with $\tE$
712: ($\Delta_\star=\tEst-\tE\simeq0$), the 
713: likelihood can be expanded around $\Delta_\star$
714: \bear
715: \Lik&=&\Lik_\star+\left({\partial\Lik\over\partial\tEm}\right)_\star
716: (\Delta-\Delta_\star) \\
717: &+&{1\over2}\left({\partial^2\Lik\over\partial\tEms}
718: \right)_\star(\Delta-\Delta_\star)^2+{\mathcal O}(\Delta^3),\nonumber
719: \enar
720: and we can use the standard Newton-Raphson method to solve Eq.~(\ref{eq:mle})
721: and obtain a maximum likelihood estimator $\tEML$,
722: \bear
723: \label{eq:iterD}
724: \Delta(\tEML)-\Delta_\star&=&\tEML-\tEst \\
725: &=&-\left({\partial\Lik\over\partial\tEm}\right)_\star\bigg/
726: \left({\partial^2\Lik\over\partial^2\tEms}\right)_\star.\nonumber
727: \enar
728: It is important to note that the validity of our solution for $\tEML$
729: is independent of the linear approximation, but
730: the convergence of $\tEML$ depends on the goodness of $\tEst$ 
731: to $\tE$. Eq.~(\ref{eq:iterD}) still involves computationally
732: intensive evaluations of the second derivative, or the curvature matrix.
733: We further simplify $\tEML$ by replacing the curvature matrix with its 
734: ensemble average, Fisher matrix
735: \bear
736: \hat\Fisher&=&\int{d^2\vL_1\over(2\pi)^2}\int{d^2\vL_2\over(2\pi)^2} \\
737: &\times&{2\pi^2\over C_{l_1}C_{l_2}}
738: \left[{(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)
739: \over|\vL_1+\vL_2|^3}\right]^2, \nonumber
740: \enar
741: and by evaluating the derivatives at $\Delta_\star=0$.
742: Finally, our new maximum likelihood estimator is
743: \bear
744: \label{eq:iter}
745: \tEML&=&\tEst+{1\over\hat\Fisher}\int{d^2\vL_1\over(2\pi)^2}
746: \int{d^2\vL_2\over(2\pi)^2} \\
747: &\times&
748: {\That_{\vL_1}\That_{\vL_2}\over C_{l_1}C_{l_2}}
749: {\pi(\vL_1C_{l_1}+\vL_2C_{l_2})\cdot(\vL_1+\vL_2)\over|\vL_1+\vL_2|^3}. \nonumber
750: \enar
751: This equation is readily recognizable as the standard quadratic estimator
752: in Eq.~(\ref{eq:teqe}), except $\lenC_l$ and $\lenT_\vL$ replaced with $C_l$ and
753: $\That_\vL$. The resemblance should not be surprising, and in hindsight one
754: could have expected this outcome
755: given the result in Sec.~\ref{ssec:optimal}: the quadratic
756: estimator becomes optimal when the lensing effect is vanishingly small; 
757: as we delens $\lenT(\Vang)$ well enough that $\That(\Vang)$ is close to 
758: $T(\Vang)$, the residual lensing effect in $\That(\Vang)$ is substantially 
759: reduced and therefore the maximum likelihood estimator takes
760: the form of the quadratic estimator, returning diminishing change of the
761: second term in Eq.~(\ref{eq:iter}), i.e., 
762: $\tEML\simeq\tEst\simeq\tE$.
763: 
764: We want to emphasize that this new estimator in the form of quadratic 
765: estimators is derived by iteratively solving for the maximum likelihood
766: in Eq.~(\ref{eq:mle}) and updating the initial model $\tEst$ as in the
767: standard Newton-Raphson method,
768: i.e., it is a maximum likelihood estimator and is
769: independent of the linear approximation, to which the validity of the standard
770: quadratic estimator is limited. One may be concerned about replacing the
771: curvature matrix with the Fisher matrix in Eq.~(\ref{eq:iter}) and obtaining
772: a solution quadratic in $\That_\vL$ instead of a solution rational in 
773: $\That_\vL$ (quadratic in $\That_\vL$ both in numerator and in denominator).
774: However, both procedures guarantee that the correct solution of 
775: Eq.~(\ref{eq:mle}) is iteratively found reaching the same peak of the 
776: likelihood, while the error estimation of parameters is approximated by using
777: the Fisher matrix, rather than the full curvature matrix. In Sec.~\ref{sec:num}
778: we demonstrate that this is a good approximation and the initial
779: model converges quickly to the true model.
780: Given the nomenclature of the existing quadratic estimators, now let us call
781: our new maximum likelihood estimator an improved quadratic 
782: estimator.\footnote{However, note 
783: that since our new estimator takes the result
784: of the previous iteration as an initial model, another iteration makes use 
785: of $\That(\Vang)$ that is constructed by using the initial model and this
786: initial model is also a function of $\That(\Vang)$ in the previous iteration,
787: which makes the estimator a rational function of temperature, instead of a
788: quadratic function. Therefore,  it is technically incorrect to call it 
789: a quadratic estimator.}
790: 
791: In practice we can use the standard quadratic
792: estimators to obtain an initial model and 
793: then proceed with our improved quadratic
794: estimator to refine the solution, even when the lensing effect is large. 
795: In general, the reconstruction of cluster mass profiles
796: is too noisy to provide a good initial model.
797: However, we can adopt an initial model for clusters from other 
798: observations (e.g., galaxy weak lensing and X-ray measurement) or theoretical 
799: expectations (e.g., Navarro-Frenk-White (NFW) profiles \citep{NFW97}). As 
800: opposed to the modified quadratic estimators, there is no arbitrary choice of
801: $\lcut$ in our method. 
802: 
803: The toy model developed here can be readily generalized and our
804: improved quadratic estimator can be used to reconstruct mass profiles of
805: realistic clusters and large-scale structure. However, in the presence of
806: the telescope beam and detector noise, the delensing process becomes 
807: non-optimal because it does not commute with the beam smoothing. In the
808: absence of detector noise, one can deconvolve the beam factor, delens the
809: temperature field, and convolve the beam again, which can solve the problem
810: of non-commutativity. 
811: 
812: However, in the presence of detector noise, the beam
813: deconvolved noise can produce unwanted power on all scales when it is
814: delensed due to the non-white power below the beam scale. One can in principle
815: filter out or remove these small scales before delensing to mitigate the 
816: problem \citep{HUDEVA07}, 
817: which however introduces additional ad hoc scale to the problem.
818: The impact of telescope beam and detector noise is small in practice
819: for surveys like SPT ($\Delta_T\simeq6\muK$-arcmin) and ACT 
820: ($\Delta_T\simeq10\muK$-arcmin)
821: as we numerically demonstrate in 
822: Sec.~\ref{sec:num}. We explicitly show in Appendix~\ref{app:delens}
823: that the delensing process suppresses the beam effect by a factor of the
824: average magnification by clusters,
825: since it corresponds to a mapping from the image plane to the source plane.
826: Non-white instrumental noise and boundary effect of
827: detectors may affect the delensing process. However, compared to the survey
828: area, the lensing signals are limited to a relatively small region 
829: around clusters where none of those effect is expected to be significant.
830: 
831: 
832: \section{Reconstructing Cluster Mass Profiles}
833: \label{sec:num}
834: Here we use numerical simulations of the CMB and cluster lensing potential to 
835: demonstrate the applicability of our improved quadratic estimator to 
836: CMB experiments. First, we adopt a more realistic model for massive 
837: clusters and investigate the dependence of our improved
838: quadratic estimator on assumed initial models in Sec.~\ref{ssec:iQE}. Then we 
839: reconstruct cluster mass profiles using the standard, modified, and
840: improved quadratic estimators, and we compare their performance 
841: in Sec.~\ref{ssec:com}.
842: Finally, we discuss the effects of contaminants and investigate the robustness
843: of our improved quadratic estimators in the presence of the Sunyaev-Zel'dovich
844: (SZ) effects. 
845: 
846: 
847: \begin{figure}
848: \centerline{\psfig{file=f2.eps, width=3.0in}}
849: \caption{(color online)
850: Reconstructed convergence fields of a $30'\times30'$ region around
851: a cluster at $z_L=1$ from an ideal experiment with $\Delta_T=0$. 
852: Cluster mass is set $M=5\times10^{14}\hmsun$.
853: Improved quadratic estimators are applied once with initial mass models of 
854: $\Minit=5\times10^{14}\hmsun$ ({\it left})
855: and $\Minit=1\times10^{14}\hmsun$ ({\it right}) to a single patch of sky. 
856: The bottom
857: panels show the residual after the true cluster convergence field is 
858: subtracted from the top panels.}
859: \label{fig:residual}
860: \end{figure}
861: 
862: 
863: 
864: \subsection{Improved Quadratic Estimator}
865: \label{ssec:iQE}
866: A singular isothermal model used in Sec.~\ref{sec:mle} is useful in developing
867: an analytic solution of the likelihood approach. However, it is rather an
868: academic model than a realistic model for massive clusters.
869: Recent numerical simulations show that there exist a universal mass profile
870: for dark matter halos, NFW profiles \citep{NFW97}
871: \beeq
872: \rho(r)={\rho_s\over r/r_s(1+r/r_s)^2}.
873: \eneq
874: The scale radius $r_s$ is described by the concentration parameter 
875: $c=\Rvir/r_s$ and the normalization coefficient $\rho_s$ 
876: is related to the mass of clusters $M=4\pi r_s^3\rho_s[\ln(1+c)-c/(1+c)]$. 
877: We now use NFW profiles to model massive clusters.
878: 
879: The convergence field $\kappa(\Vang)$
880: of NFW profiles can be obtained by the ratio of 
881: the projected mass density $\Sigma(r)$ to the critical surface density 
882: $\Sigma_\up{crit}$ of the lensing cluster at $z_L$,
883: \beeq
884: \kappa\left(\theta={r\over D_L}\right)={\Sigma(r)\over\Sigma_\up{crit}}=
885: {2~r_s\rho_s\over\Sigma_\up{crit}}
886: P\left({r\over r_s}\right)(1+z_L)^2,
887: \label{eq:kappa}
888: \eneq
889: where the functional form $P(x)$ of the projected density is 
890: \citep{BARTE96,WRBR00}
891: \bear
892: P(x)&=&{1\over x^2-1}\left[1-{2\over\sqrt{1-x^2}}~\up{tanh}^{-1}
893: \sqrt{1-x\over1+x}\right],~~~(x<1) \nonumber \\
894: &=&{1\over3},~~~(x=1)\\
895: &=&{1\over x^2-1}\left[1-{2\over\sqrt{x^2-1}}~\up{tan}^{-1}\sqrt{x-1\over x+1}
896: \right],~~~(x>1)\nonumber
897: \enar
898: and the critical surface density 
899: $\Sigma_\up{crit}^{-1}=4\pi G D_L(D_\star-D_L)/D_\star(1+z_L)$ is only a 
900: function
901: of $z_L$ given the precise measurement of $D_\star$. Note that the convergence
902: field $\kappa$ of NFW profiles depend only on the angular separation
903: $\theta=|\Vang|$ due to spherical symmetry.
904: The redshift dependence in Eq.~(\ref{eq:kappa})
905: arises due to our use of comoving coordinates, reflecting higher densities 
906: of the universe at $z_L>0$. For reference, $D_L=850\hmpc$ and $2400\hmpc$, and
907: $\Sigma_\up{crit}=2.8\times10^3hM_\odot\up{pc}^{-2}$ 
908: and $1.8\times10^3hM_\odot\up{pc}^{-2}$ for
909: $z_L=0.3$ and~1, respectively. For clusters of $M=5\times10^{14}\hmsun$ and 
910: $1\times10^{14}\hmsun$, $\Rvir=2.1\hmpc$ and $1.2\hmpc$ appear subtended by
911: $3.\!'0$ and $4.\!'9$ on the sky at $z_L=1$ and 0.3.
912: 
913: We use {\scriptsize CMBFAST} \citep{SEZA96} to 
914: generate CMB temperature maps of $200'\times200'$ ($1000\times1000$ pixels)
915: and set the pixel scale $0.\!'2$ smaller than detector beam sizes. Given
916: a cluster mass $M$ and redshift $z_L$, we first compute the convergence field
917: $\kappa(\Vang)$ using Eq.~(\ref{eq:kappa}). The lensing potential $\phi(\Vang)$ 
918: and its deflection vector $\bdv{d}(\Vang)$ of the cluster are then computed
919: in Fourier space, where their relations to $\kappa(\Vang)$ become a simple
920: multiplication.
921: The lensed temperature field $\lenT(\Vang)$ is computed by displacing
922: the intrinsic temperature field $T(\Vang)$ with $\bdv{d}(\Vang)$ according to
923: Eq.~(\ref{eq:lensing}). Finally, we smooth $\lenT(\Vang)$ with a telescope
924: beam and add detector noises to obtain $\Tobs(\Vang)$.
925: Standard quadratic estimators can be used to reconstruct a convergence
926: field $\khat(\Vang)$ by using Eqs.~(\ref{eq:G}), (\ref{eq:W}), and 
927: (\ref{eq:GW}) with $\Tobs(\Vang)$, and so can modified quadratic estimators
928: with a choice of $\lcut$, beyond which the integrand in Eq.~(\ref{eq:G})
929: is set zero.
930: 
931: Similarly, our new estimation process begins with finding a solution
932: $\bhat$ to the delensing equation $\bhat=\Vang+\ang\phi^m(\Vang)$ given the
933: lensing potential $\phi^m(\Vang)$ of an assumed initial model. We then 
934: construct a delensed temperature field $\That(\bhat)=\Tobs(\Vang)$ and
935: use the same equations with $\Tobs(\Vang)$ replaced by $\That(\bhat)$
936: to reconstruct $\khat_\bdv{L}$. Imposing a consistency condition between
937: the assumed model and the estimation result can provide a criterion for the
938: iteration convergence of our improved quadratic estimators.
939: 
940: ACT and SPT will find $\sim2\times10^4$
941: massive clusters mainly by the spectral distortion of 
942: the CMB arising from the inverse Compton scattering of hot electrons in 
943: clusters, so called the SZ effect \citep{SUZE70,SUZE72},
944: with roughly redshift-independent threshold mass 
945: $M\geq2\times10^{14}\hmsun$. To test our improved quadratic estimators,
946: we consider a typical cluster of $M=5\times10^{14}\hmsun$ and $c=3$.
947: Figure~\ref{fig:residual} shows the reconstructed $\khat(\Vang)$ of a 
948: massive cluster at $z_L=1$ in an ideal experiment with $\Delta_T=0$. 
949: Here we simply adopt a NFW profile with fixed concentration
950: $c=3$ for our initial model 
951: and allow mass $\Minit$ of the model to vary. Even with fixed concentration,
952: $r_s$ changes as a function of $\Minit$, and hence our assumption allows
953: for changes in the shape as well as the scaling of initial mass models. 
954: However, note that while we use this parametrized model of clusters,
955: our reconstruction is general and non-parametric, such that we recover
956: 2-D structure of
957: $\kappa(\Vang)$ at each pixel rather than obtain model parameters $M$ and $c$
958: (see \citep{DODEL04,LEKI06} for reconstructing a parametrized cluster model). 
959: We assume that the cluster center is known from other observations with 
960: uncertainty less than our pixel scale $0.\!'2$.
961: The upper panels show the reconstructed $\khat(\Vang)$ from our
962: improved quadratic estimator using an initial model of
963: $\Minit=5\times10^{14}\hmsun$ ({\it left}) and $1\times10^{14}\hmsun$ 
964: ({\it right}), and the bottom panels
965: show the residual after the true $\kappa(\Vang)$ is subtracted from the top
966: panels.
967: 
968: \begin{figure}
969: \centerline{\psfig{file=f3.eps, width=3.0in}}
970: \caption{Dependence of reconstructed mass profiles on an initial mass model
971: $\Minit$. Thick and thin solid lines represent the true cluster mass profile 
972: and the mean of reconstructed mass profiles from 500 clusters. 
973: The mass profiles are obtained by averaging reconstructed convergence
974: over the annulus of each cluster. 
975: The uncertainties in the mean profile are shown as shaded regions.
976: Dashed lines show an assumed
977: initial mass model and the cluster virial radius is shown as vertical
978: dotted lines. In Panels~($c$) and~($d$), the initial mass models are 
979: taken as the mean mass profile from the previous
980: iteration. The reconstruction quickly converges to the true mass profile 
981: in two iterations even with an incorrect 
982: choice of $\Minit=1\times10^{14}\hmsun$, 
983: exhibiting no detectable bias in an ideal experiment.}
984: \label{fig:iQE}
985: \end{figure}
986: 
987: With the perfect initial model in the left panels, the delensed temperature
988: field $\That(\Vang)$ is identical to the intrinsic $T(\Vang)$, and our improved
989: quadratic estimator returns $no$ change on average to the initial model
990: ({\it bottom}). However, there exist random noises in $\khat(\Vang)$ over the
991: map, arising from the fluctuations of the intrinsic temperature gradient,
992: though they are evidently small and discernible from the massive cluster
993: ({\it top}). In the right panels, $\lenT(\Vang)$ is delensed with the imperfect
994: initial model, so that $\That(\Vang)$ is not identical to $T(\Vang)$ but 
995: the lensing effect is significantly reduced.
996: In this regime, quadratic estimators become asymptotically
997: optimal and reconstruct $\kappa(\Vang)$ unbiased. The top panel exhibits
998: small anisotropy and some residual remains in the bottom panel. In a single
999: patchy of the sky, the CMB anisotropy has a gradient direction and 
1000: gravitational lensing of the CMB makes no difference orthogonal to the 
1001: gradient direction, in which reconstruction is completely degenerate,
1002: resulting in the asymmetry in $\khat(\Vang)$.
1003: However, since the CMB has no preferred direction, this obstacle can be 
1004: overcome by stacking clusters in different patches of the sky. In practice,
1005: this stacking process provides the average $\kappa(\Vang)$ of the clusters,
1006: or the cluster-mass
1007: cross-correlation function \citep{HUDEVA07}. Hereafter we assume that identical
1008: clusters are stacked for simplicity.
1009: 
1010: We now quantify the ability to reconstruct $\kappa(\Vang)$ with varying
1011: accuracy of assumed models. Figure~\ref{fig:iQE} plots the reconstructed
1012: cluster mass profiles from 500 clusters ({\it thin solid}). The mass profiles
1013: are obtained by averaging reconstructed $\khat(\Vang)$ over the annulus of 
1014: each cluster, and the uncertainties in the mean mass profile
1015: are shown as shaded 
1016: regions. Figure~\ref{fig:iQE}$a$ shows that our improved quadratic
1017: estimator is unbiased when our assumed model is perfect; it 
1018: recovers the true model ({\it thick solid}) with no bias. If an 
1019: assumed initial model is significantly different from the true model
1020: in Fig.~\ref{fig:iQE}$b$,
1021: the improved quadratic estimator suffers from the same problem
1022: that the standard quadratic estimators have, and the reconstruction is again
1023: biased low when the residual lensing effect is large. However, the 
1024: reconstructed $\khat(\Vang)$ is inconsistent with our assumed model
1025: ({\it dashed}), implying that it has not converged to the correct solution.
1026: In Fig.~\ref{fig:iQE}$c$ we take the reconstructed $\khat(\Vang)$ as a new
1027: initial model and apply our improved quadratic estimator to the same clusters.
1028: The reconstructed $\khat(\Vang)$ is now close to the true $\kappa(\Vang)$,
1029: but still inconsistent with the assumed model. We iterate once more in 
1030: Fig.~\ref{fig:iQE}$d$ and the reconstructed $\khat(\Vang)$ is identical to
1031: the true $\kappa(\Vang)$. One more iteration results in no further change
1032: and the estimate is consistent with the assumed and also the true models,
1033: indicating the convergence of our estimates.
1034: 
1035: Even with the imperfect initial model, the
1036: reconstruction quickly converges to the true $\kappa(\Vang)$ and 
1037: no significant bias develops even beyond $\Rvir$ ({\it dotted}).
1038: When the reconstructed $\khat(\Vang)$ is inconsistent with the assumed model,
1039: one can in principle adopt a different initial model for a faster convergence
1040: before applying the estimator iteratively. Note that the asymmetry seen
1041: in Fig.~\ref{fig:residual} disappears and the reconstructed $\khat(\Vang)$
1042: restores symmetry,
1043: once many clusters are stacked. Furthermore, the uncertainties in the mean
1044: profile decrease as our assumed model converges to the true model, because
1045: it solely results from the intrinsic fluctuations of the CMB in the case of
1046: perfect delensing.
1047: 
1048: \begin{figure}
1049: \centerline{\psfig{file=f4.eps, width=3.0in}}
1050: \caption{Mass profile reconstruction for low mass clusters 
1051: of $M=1~\times10^{14}\hmsun$ at $z_L=0.3$ from standard
1052: (sQE) and improved (iQE) quadratic estimators (in the same format as in 
1053: Fig.~\ref{fig:iQE}). 10,000 ({\it left}) and 1000 ({\it right}) clusters 
1054: are used to obtain the mean profile, and the shaded region shows the 
1055: uncertainties in the mean profile. Both estimators recover the true mass
1056: profiles within $\Rvir$ in the low mass regime.
1057: Approximately ten times more clusters are needed for sQE to achieve
1058: the same accuracy than for iQE. However, for comparison we plot the mean 
1059: profile from 1000 clusters as the dot-dashed line in the left panel.}
1060: \label{fig:lowM}
1061: \end{figure}
1062: 
1063: \begin{figure*}
1064: \centerline{\psfig{file=f5.eps, width=5.5in}}
1065: \caption{Comparison of reconstructed mass profiles from standard (sQE), 
1066: modified (mQE), and improved (iQE) quadratic estimators 
1067: in realistic experiments with $\spix=5\muK$. 
1068: The reconstruction is more difficult in the presence of detector noise and
1069: telescope beam. For the mean of reconstructed mass profiles,
1070: 10,000 clusters of $M=5\times10^{14}\hmsun$ at $z_L=1$ are stacked when sQE or 
1071: mQE is used, while iQE is iteratively applied to only 1000 clusters. 
1072: The shaded regions show the uncertainties in the mean profile.
1073: The dot-dashed line (panel~$d$) shows the shape distortion in $\khat(\Vang)$
1074: when mQE is applied after beam-deconvolution, and the line is displaced to 
1075: avoid confusion (see the text). With $\tfwhm=1'$ (panel~$f$),
1076: iQE can recover the mean mass profile 
1077: with small bias below the beam scale. For comparison, we plot
1078: the reconstructed mass profile ({\it dot-dashed}) using mQE in Panel~($f$).}
1079: \label{fig:com}
1080: \end{figure*}
1081: 
1082: 
1083: \subsection{Performance Comparison}
1084: \label{ssec:com}
1085: Before we assess the performance of the three lensing estimators 
1086: in realistic experiments,
1087: we first compare our improved quadratic estimator to the standard quadratic
1088: estimator, when the lensing effect is small. Figure~\ref{fig:lowM} plots
1089: the reconstructed cluster mass profiles in the same format as 
1090: Fig.~\ref{fig:iQE}. For clusters of $M=1\times10^{14}\hmsun$ at $z_L=0.3$ 
1091: ($\kappa\ll1$),
1092: the improved quadratic estimator recovers the true mass profile with no 
1093: detectable bias after two iterations. With signals smaller by
1094: a factor of five than in Fig.~\ref{fig:iQE}, 1000 clusters are
1095:  stacked to obtain the mean mass profile, while 10,000 clusters are
1096: required for the standard quadratic estimator. As we quantify the
1097: difference in the signal-to-noise ratio below, 
1098: the standard quadratic estimator needs approximately 
1099: ten times as many clusters as the improved quadratic estimator needs to
1100: achieve the same accuracy, 
1101: but we show the mean profile ({\it dot-dashed}) obtained by applying the
1102: standard quadratic estimator to 1000 clusters for comparison. Once enough
1103: clusters are stacked, the standard quadratic estimator
1104: works well within $\Rvir$, though it shows some hint of deviation
1105: at the core. Thus, the standard quadratic
1106: estimator may be safely used to reconstruct mass profiles of clusters with
1107: $M<1\times10^{14}\hmsun$ at $z_L=0.3$. However, given the source of the CMB
1108: at $z_\star=1090$, the lensing effect becomes larger
1109: as $z_L$ increases, until $\Sigma_\up{crit}$ reaches the minimum at 
1110: $z_L\simeq2.5$, where $D_L$ becomes a half of $D_\star$. Therefore, the standard
1111: quadratic estimator cannot be used to reconstruct unbiased mass profiles of
1112: clusters that are either at $z_L\geq0.3$ or massive 
1113: $M\geq1\times10^{14}\hmsun$. Since
1114: ACT and SPT will find clusters of $M\geq2\times10^{14}\hmsun$ 
1115: at higher redshift, modified
1116: or improved quadratic estimators are preferred to the standard quadratic 
1117: estimator.
1118: 
1119: Now we consider realistic experiments with $\spix=5\muK$ and compare
1120: the performance of the lensing estimators in Fig.~\ref{fig:com}. 
1121: Since the reconstruction becomes noisier in the
1122: presence of detector noise and telescope beam,
1123: 10,000 clusters are stacked for the mean mass profiles when the standard
1124: or modified quadratic estimator is used, while the improved quadratic estimator
1125: is iteratively applied to only 1000 clusters. For clusters
1126: of $M=5\times10^{14}\hmsun$ at $z_L=1$, Fig.~\ref{fig:com}$a$ shows that 
1127: the standard 
1128: quadratic estimators become substantially biased  in a region around
1129: massive clusters, consistent with the previous results 
1130: \citep{MABAMEET05,HUDEVA07}. Quadratic terms in $\phi_\vL$ ignored in
1131: the linear approximation coherently contribute to $\khat_\vL$, and hence the
1132: reconstructed $\khat(\Vang)$ is biased 
1133: low where the linear approximation is violated \citep{HUDEVA07}.
1134: 
1135: Next we consider a modified quadratic estimator in Fig.~\ref{fig:com}$b$ 
1136: and adopt $\lcut=1500$. The modified quadratic
1137: estimator recovers the true mass profile within $\Rvir$ but with small
1138: deviation beyond $\Rvir$. The modified quadratic estimators operate
1139: in the same way of the standard quadratic estimators, except signals are
1140: removed on small scales ($l\geq\lcut$), 
1141: where the linear approximation is violated.
1142: However, the choice of $\lcut$ is rather arbitrary and should be calibrated
1143: against simulations: lower $\lcut$ is needed for more massive clusters.
1144: Note that the modified quadratic estimator with 
1145: $\lcut\rightarrow\infty$ 
1146: exactly reduces to the standard quadratic estimator
1147: (in practice $\lcut\gtrsim10^4$ can achieve this limit because of the Silk 
1148: damping). In other words, a modified quadratic estimator 
1149: with $\lcut\simeq10^4$ fails to reconstruct
1150: the mass profile (born out by Fig.~\ref{fig:com}$a$).
1151: Moreover, we had to adopt $\lcut=1500$ to reconstruct the mass profile
1152: in Fig.~\ref{fig:com}$b$ and~\ref{fig:com}$d$, a more aggressive choice
1153: than $\lcut=2000$ proposed in \citep{HUDEVA07}, with which
1154: we cannot recover the mass profile.  This reflects the
1155: sensitivity of the modified quadratic estimator to $\lcut$ as a function of
1156: cluster mass.
1157: Larger number of clusters are also required to reconstruct the true mean
1158: mass profile due to the reduction in the signal-to-noise ratio.
1159: 
1160: Figure~\ref{fig:com}$c$ shows the reconstruction by our improved quadratic
1161: estimator with $\Minit=1\times10^{14}\hmsun$. The improved quadratic 
1162: estimator recovers
1163: the true mass profile with no significant bias in the presence of detector
1164: noise. After a few iterations, the estimates quickly converge to the true
1165: model and the scatter around the mean is greatly reduced compared to 
1166: Fig.~\ref{fig:com}$b$. Note that we iteratively applied the improved quadratic
1167: estimator to the same 1000 clusters.
1168: 
1169: In Fig.~\ref{fig:com}$d$ and \ref{fig:com}$e$, we consider the effect of
1170: telescope beam with $\tfwhm=0.\!'5$. Both estimators
1171: in Fig.~\ref{fig:com}$d$ and \ref{fig:com}$e$ recover the true mass profile
1172: unbiased in the presence of detector beam, while there exist some deviations
1173: in both cases.
1174: However, note that we explicitly account for the beam 
1175: effect using the formulas developed in Sec.~\ref{ssec:qe}, rather than 
1176: deconvolve the beam before applying the lensing estimators. The latter
1177: approach often used in the literature suffers from deconvolved detector noise
1178: exponentiating on small scales. This problem requires a low-pass filtering
1179: of reconstructed $\khat(\Vang)$, additionally
1180: removing the signals below the beam scale,
1181: which results in a distortion of its shape of $\khat(\Vang)$,
1182: making it hard to compare directly to theoretical predictions. 
1183: However, in reality 
1184: beam convolution suppresses detector noises (of course  lensing signals
1185: as well), and it simply makes the reconstruction noisy below the beam scale.
1186: The dot-dashed line in Fig.~\ref{fig:com}$d$
1187: contrasts the reconstruction when we explicitly remove 
1188: $\khat(\Vang)$ at $l\geq1/\sbeam$, where $\khat(\Vang)$ is
1189: obtained by applying the modified quadratic estimator with
1190: beam-deconvolved data (the line is displaced to avoid confusion with other
1191: lines). Significant shape distortion in $\khat(\Vang)$ 
1192: complicates the interpretation.
1193: 
1194: For a larger beam size comparable to the scale radius of the clusters
1195: ($\tfwhm\simeq1'$), the reconstruction becomes more challenging: modified
1196: quadratic estimators cannot recover the cluster mass profile without 
1197: significant shape distortion ({\it dot-dashed}).
1198: The improved quadratic estimator in Fig.~\ref{fig:com}$f$
1199: recovers the true mass profile beyond $\Rvir$, while it develops small
1200: bias below the beam scale. 
1201: 
1202: Figure~\ref{fig:err} plots the fractional difference between the lensing 
1203: estimates and the true cluster mass profile in Fig.~\ref{fig:com}, 
1204: comparing their uncertainty in the mean profile. The difference ({\it lines})
1205: is computed from the mean mass profiles by stacking 10,000 clusters for 
1206: both estimators, while the statistical uncertainty
1207: ({\it gray bands}) in the difference is scaled for 500 clusters for 
1208: comparison.
1209: The left panel shows that both estimators recover the cluster mass profile
1210: at the 5\% level or better in the absence of telescope beam, 
1211: while the modified quadratic estimator may need fine-tuning of $\lcut$ to
1212: achieve better accuracy. However, the difference in their measurement 
1213: uncertainty is in stark contrast: the improved quadratic estimator has a
1214: significantly higher signal-to-noise ratio than the modified quadratic
1215: estimator. While the reconstruction becomes harder especially beyond $\Rvir$
1216: in the presence of telescope beam shown in the right panel,
1217: this trend of signal-to-noise ratio difference persists.
1218: Note that due to the beam smoothing effect the uncertainty in the estimates 
1219: at $\theta\leq\tfwhm$ is reduced while it is highly correlated among adjacent
1220: bins.
1221: 
1222: So far 
1223: we have numerically demonstrated the performance of the lensing estimators
1224: in Figs.~\ref{fig:com} and~\ref{fig:err}:
1225: standard quadratic estimators are significantly 
1226: biased; modified and improved quadratic estimators recover the cluster 
1227: mass profile with no bias, while they show substantial difference
1228: in the number of clusters that is required to obtain the mean mass profile.
1229: To quantify this difference, we evaluate 
1230: $\Delta\chi^2$ of each lensing estimator
1231: \beeq
1232: \Delta\chi^2=\sum_{\theta,\theta'}\kappa(\theta)~C_{\khat}^{-1}(\theta,\theta')
1233: ~\kappa(\theta'),
1234: \eneq
1235: where the covariance matrix of $\khat(\theta)$ is 
1236: \beeq
1237: C_{\khat}(\theta,\theta')=\left\langle\left[\khat(\theta)-\kappa(\theta)\right]
1238: \left[\khat(\theta')-\kappa(\theta')\right]\right\rangle.
1239: \eneq
1240: Since $\khat(\Vang)$ is computed from the two Wiener-filtered functions of
1241: the CMB temperature anisotropies, the covariance matrix is non-diagonal.
1242: The finite width of the convolution filter $H(\Vang)$ in Eq.~(\ref{eq:filterH})
1243: also reflects that the lensing estimators are a non-local function of the CMB
1244: temperature anisotropies, and hence non-zero $C_{\khat}$ 
1245: when $\theta\neq\theta'$.
1246: 
1247: \begin{figure}
1248: \centerline{\psfig{file=f6.eps, width=3.0in}}
1249: \caption{Fractional difference between the lensing estimates and the true
1250: cluster mass profile in Fig.~\ref{fig:com}. The difference ({\it lines})
1251: is computed from the mean mass profiles obtained by stacking 10,000 clusters
1252: for both estimators,
1253: while the statistical uncertainty ({\it gray bands}) in the difference is 
1254: scaled for 500 clusters. The vertical dotted lines show
1255: the cluster virial radius.}
1256: \label{fig:err}
1257: \end{figure}
1258: 
1259: In the absence of telescope beam in Figs.~\ref{fig:com}$b$,~\ref{fig:com}$c$,
1260: and~\ref{fig:err}$a$, the ratio of $\Delta\chi^2$ for the modified quadratic
1261: estimator relative to the improved quadratic estimator is 8.1:
1262: a factor of eight larger number
1263: of clusters is required for the modified quadratic estimator to achieve the
1264: same level of accuracy than that for the improved quadratic estimator.
1265: In the presence of telescope beam in Figs.~\ref{fig:com}$d$,~\ref{fig:com}$e$,
1266: and~\ref{fig:err}$b$, 
1267: beam smoothing substantially degrades the ability to recover the
1268: true cluster mass profile for both estimators, and its effect is
1269: relatively larger for the modified quadratic estimator, increasing
1270: the ratio to 10.4. 
1271: 
1272: \subsection{Sunyaev-Zel'dovich Effects}
1273: \label{ssec:sz}
1274: On small scales ($l>2000$), the primordial CMB temperature anisotropies decay
1275: exponentially due to the Silk damping \citep{SILK68} and the dominant source
1276: of secondary anisotropies is the thermal Sunyaev-Zel'dovich (tSZ) effect,
1277: arising from scattering off hot electrons in massive clusters. However, the
1278: tSZ effect imprints a unique frequency dependence in the CMB temperature
1279: anisotropies, which in principle can be used to remove the tSZ signals.
1280: The same Compton scattering process also gives rise to a Doppler effect in the 
1281: CMB temperature anisotropies due to the bulk motion of electron gas, or
1282: the kinetic Sunyaev-Zel'dovich (kSZ) effect (see \citep{BIRKI99,CAHORE02}
1283: for recent reviews). These kSZ signals, albeit smaller than tSZ signals,
1284: are spectrally indistinguishable from the intrinsic CMB temperature 
1285: anisotropies, introducing an artifact in the lensing reconstruction. Here
1286: we assume that the tSZ signals can be cleaned perfectly, and 
1287: we investigate how the kSZ signals deteriorate the lensing reconstruction.
1288: 
1289: For simplicity, we assume that the gas density traces the dark matter 
1290: distribution in a massive cluster, with the same NFW profile. Given the 
1291: line-of-sight velocity $v_\up{los}$ of the cluster, the kSZ effect results
1292: in temperature anisotropies
1293: \beeq
1294: \Delta T(\theta)=-v_\up{los}~\tau(\theta)~T_\up{CMB}\equiv-\Delta T_\up{kSZ}
1295: ~{\Sigma(\theta)\over\Sigma(0)},
1296: \label{eq:kszT}
1297: \eneq
1298: where $\tau(\theta)$ is the Thompson scattering
1299: optical depth, proportional to the projected density $\Sigma(r=D_L~\theta)$. 
1300: We parametrized the product of 
1301: $v_\up{los}$ and $\tau(0)$ as $\Delta T_\up{kSZ}$. 
1302: Note that since the intrinsic CMB and kSZ induced anisotropies dilute in the
1303: same way as the universe expands, there is no $(1+z_L)$ factor in 
1304: Eq.~(\ref{eq:kszT}) and $T_\up{CMB}=2.725$K is the CMB temperature today.
1305: 
1306: For a typical cluster with
1307: electron number density $\sim0.01~\up{cm}^{-3}$ and core radius $\sim100$~kpc,
1308: the Thompson scattering optical depth is $\tau(0)=2\times10^{-3}$ at the core. 
1309: The rms velocity
1310: dispersion in linear theory is $\sigma_v=1.3\times10^{-3}(=390~\kms)$ 
1311: at $z_L=1$,
1312: and this results in the rms temperature fluctuation $\Delta T_\up{kSZ}=3.7\muK$
1313: at the core. We randomly draw $\Delta T(0)$ from a Gaussian distribution with
1314: zero mean and dispersion $\sigma=\Delta T_\up{kSZ}$, 
1315: then we add $\Delta T(\Vang)$ to $\lenT(\Vang)$ for observations of 
1316: each cluster. 
1317: 
1318: \begin{figure}
1319: \centerline{\psfig{file=f7.eps, width=3.0in}}
1320: \caption{(color online)
1321: Cluster lensing and kinetic Sunyaev-Zel'dovich (kSZ) effects on 
1322: the CMB. For comparison, we plot $6'\times6'$ regions of CMB temperature maps
1323: around a cluster of $M=5\times10^{14}\hmsun$ ($\theta_\up{vir}=3.\!\!'0$) 
1324: at $z_L=1$.
1325: {\it Upper panels}: lensed temperature map ({\it left})
1326: and its difference from the intrinsic temperature map ({\it right}).
1327: {\it Bottom panels}: assuming that the cluster is 
1328: moving toward an observer, the kSZ effect is set
1329: $\Delta T_\up{kSZ}=3$ ({\it left}) and $15\muK$ ({\it right}) at the center.
1330: The color scales in each panel represent the same temperature except in the 
1331: upper right panel, where the color represents the difference ranging from
1332: $-5\muK$ to $5\muK$.}
1333: \label{fig:kszshow}
1334: \end{figure}
1335: 
1336: First, we compare the cluster lensing and kSZ effects on the CMB temperature
1337: field. Figure~\ref{fig:kszshow} plots a 6'$\times$6' regions of CMB maps
1338: around a cluster of $M=5\times10^{14}\hmsun$ ($\theta_\up{vir}=3.\!\!'0$) 
1339: at $z_L=1$.
1340: The top panels show the lensed temperature field ({\it left}) and the
1341: difference from the intrinsic temperature field ({\it right}). Gravitational
1342: lensing imprints dipole-like wiggles in the CMB map on top of the smooth
1343: large-scale gradient field. Perpendicular to the gradient direction there
1344: exists no temperature change and hence lensing reconstruction is degenerate
1345: along the direction. The bottom panels show the kSZ effect with
1346: $\Delta T_\up{kSZ}=3\muK$ ({\it left}) and $15\muK$ ({\it right}). We assume
1347: that the cluster is moving toward the observer. With the small optical depth
1348: in the left panel, the kSZ effect is relatively small compared to the lensing
1349: effect. Larger optical depth in the right panel substantially enhances the
1350: kSZ effect, dominating over the lensing effect at the center. However, since
1351: the lensing effect is much less concentrated than the kSZ effect as the 
1352: dipole-like wiggles peak at a few scale radii ({\it top right}),
1353: the reconstruction is still possible.
1354: 
1355: Figure~\ref{fig:ksz} shows the impact of the kSZ effect on reconstructing
1356: mass profiles. For clusters of $M=5\times10^{14}\hmsun$ at $z_L=1$ in 
1357: an experiment
1358: with $\tfwhm=1'$ and $\spix=5\muK$, we iteratively
1359: use improved quadratic estimators with $\Minit=1\times10^{14}\hmsun$.
1360: The mean and the uncertainties are computed
1361: from 1000 clusters. Figure~\ref{fig:ksz}$a$ shows that the kSZ effect with
1362: $\Delta T_\up{kSZ}=3\muK$ has relatively little impact on the reconstruction:
1363: the kSZ effect becomes negligible beyond $r_s$ because the density profile
1364: declines $r^{-3}$ (the gas density in reality would be steeper and
1365: more confined to the center than we assumed here). The lensing effect, on the
1366: other hand, is sensitive to the deflection field and remains strong beyond
1367: $r_s$, declining less rapidly than the kSZ effect \citep{SEZA96}.
1368: In Fig.~\ref{fig:ksz}$b$,
1369: we consider a larger kSZ effect with $\Delta T_\up{kSZ}=15\muK$, expected
1370: either from higher electron number density or from higher matter fluctuation
1371: normalization $\rms\propto\sigma_v$. With the temperature anisotropies 
1372: comparable to the lensing effect, the reconstruction becomes difficult and
1373: it starts to develop bias around $\Rvir$ as $\Delta T_\up{kSZ}$ increases.
1374: Note that the bias at the center is largely due to the telescope beam effect.
1375: 
1376: \begin{figure}
1377: \centerline{\psfig{file=f8.eps, width=3.0in}}
1378: \caption{Impact of kinetic Sunyaev-Zel'dovich (kSZ) effects on the mass profile
1379: reconstruction. Assuming that the gas distribution traces
1380: the dark matter distribution in clusters, the kSZ effect is computed by 
1381: assigning a Gaussian random velocity to each cluster
1382: with rms temperature change $\Delta T_\up{kSZ}=3$ ({\it left}) and 
1383: $15\muK$ ({\it right}) at the center, respectively.}
1384: \label{fig:ksz}
1385: \end{figure}
1386: 
1387: In the presence of contaminants such as residual foreground or tSZ effect,
1388: radio point sources, and large kSZ effect, the lensing estimators based on
1389: temperature anisotropies need to be complemented by using lensing estimators
1390: based on combination of temperature and E- and B-mode polarization
1391: \citep{OKHU03}, since there exists no significant source of contamination that
1392: mimics the intrinsic CMB polarization. Furthermore, the unique relation between
1393: the E- and B-mode polarization signals \citep{SEZA97,KAKOST97} can be used to
1394: provide a robust consistency check. However, measurements of the lensed
1395: polarization fields would require an experiment with higher angular resolution
1396: and sensitive detectors than experiments that are currently available.
1397: 
1398: \section{Discussion}
1399: \label{sec:con}
1400: Weak gravitational lensing of the CMB gives rise to a deviation of the 
1401: two-point correlation function of the CMB temperature anisotropies from 
1402: otherwise statistically isotropic function. Quadratic estimators 
1403: \citep{HU01b} have been
1404: widely used to reconstruct cluster mass profiles and large-scale structure
1405: by measuring the induced anisotropies in the two-point correlation function.
1406: We have shown that standard quadratic estimators become optimal in the
1407: limit of no lensing, saturating the Cram\'er-Rao bound, while they become
1408: progressively biased and sub-optimal as the lensing effect increases. 
1409: Especially for clusters that can be found by the ongoing SZ surveys like 
1410: ACT and SPT, the standard quadratic estimators start to be biased at 
1411: $z_L\simeq0.3$, and at higher redshift, where the lensing effect is larger,
1412: other estimators should be used to reconstruct cluster mass profiles.
1413: 
1414: It is recently proposed \citep{HUDEVA07}
1415: that this obstacle in the standard quadratic estimators
1416: can be overcome by explicitly removing the signals in the CMB temperature
1417: gradient field at $l\geq\lcut$, where the lensing effect is large in violation
1418: of the linear approximation. However, although these
1419: modified quadratic estimators
1420: recover cluster mass profiles with no significant bias, the choice of $\lcut$
1421: is somewhat arbitrary and it depends on the lensing effect, which requires
1422: prior calibrations against numerical simulations before one can apply the
1423: modified quadratic estimators to CMB maps.
1424: 
1425: We have developed a new maximum likelihood estimator for reconstructing 
1426: cluster mass profiles and large-scale structure. We first construct a CMB
1427: temperature field by delensing the observed temperature field based on an
1428: assumed mass model. We have proved that the delensed temperature field is 
1429: close to the unlensed temperature field with telescope beam smoothed and
1430: detector noise added, if the assumed mass model is a good approximation to 
1431: the true mass model. The delensed temperature field can then be used to set up
1432: the likelihood of the CMB, and our new estimator that maximizes this 
1433: likelihood takes a similar form of the standard quadratic estimators,
1434: because it approaches to an optimal estimator as the assumed model becomes
1435: the true model. Our maximum likelihood estimator can be iteratively applied
1436: as we
1437: update the assumed mass model, until it converges (to the true model) and
1438: the estimate is consistent with the assumed model. Our maximum likelihood
1439: estimator, named as an improved quadratic estimator, is easy to implement
1440: in practice and it has no free parameter.
1441: 
1442: Our improved quadratic estimators quickly converge to the true mass model
1443: after a few iterations, even when an assumed initial model is 
1444: significantly different from the true model. 
1445: When the estimate is inconsistent with the
1446: assumed model, one can adopt another initial model for iterations for faster
1447: convergence of the improved quadratic estimators. The telescope beam and 
1448: detector noise renders the reconstruction harder, but we have demonstrated
1449: that the improved quadratic estimators recover cluster mass profiles with a beam
1450: size comparable to the cluster scale radius.
1451: Furthermore, our new estimator significantly improves the signal-to-noise
1452: ratio over the standard or modified quadratic estimators by a factor of 
1453: ten in number of clusters, because when
1454: an assumed model is close to the true mass model, the only source of noise
1455: for our estimator is the intrinsic fluctuations of the CMB temperature 
1456: gradient.
1457: 
1458: We have tested the robustness of the improved quadratic estimators in the
1459: presence of the kSZ effect. The kSZ distortion $\Delta T_\up{kSZ}\leq15\muK$
1460: at the center results in relatively small bias in the reconstructed cluster
1461: mass profiles. However, since the optical depth is a function of 
1462: electron number density in the clusters, it is related to the true mass
1463: profile.
1464: Therefore, we could take a more aggressive approach to
1465: modeling kSZ signals from an assumed initial mass model and subtract the
1466: kSZ contributions before applying  improved quadratic estimators. 
1467: Furthermore, this template for kSZ signals can also be iteratively refined 
1468: as we update our assumed mass model.
1469: 
1470: Since the reconstruction is non-parametric, it is not limited to spherical
1471: clusters, while stacking many clusters ensures that irregular shapes of
1472: individual clusters become irrelevant. 
1473: Similar arguments can be applied to projection effects: 
1474: each cluster can be located at a line-of-sight with overdense or underdense
1475: regions, but projection effects become negligible once many lines-of-sight are 
1476: combined.
1477: Given a sample of clusters from SZ surveys, the average mass profile
1478: of stacked clusters would provide a cluster-mass cross-correlation function,
1479: which can be used to measure the growth rate of structure, probing the 
1480: evolution of dark energy, instead of individual cluster mass profiles.
1481: 
1482: However, in reality it would be harder to reconstruct
1483: cluster mass profiles than considered here, because there exist other
1484: contaminants such as point radio sources and residual foreground and/or tSZ
1485: effect, and other complications such as non-isolated clusters
1486: and internal bulk motion of gas in clusters. However, additional information
1487: from polarization measurements may be used to overcome some of the 
1488: difficulties, given the unique relation between the E- and B-mode polarization
1489: signals and relatively negligible primary and secondary contaminants.
1490: Finally we mention that our improved quadratic estimators can be applied 
1491: to reconstruct large-scale structure, while in this regime standard quadratic 
1492: estimators can be used without significant bias.
1493: 
1494: 
1495: \acknowledgments 
1496: We thank Oliver Zahn for useful discussion. J.~Y. thanks C.~K.~Chan for 
1497: technical help on FFTw routines. J.~Y. is supported by the Harvard College 
1498: Observatory under the Donald~H. Menzel fund.
1499: M.~Z. is supported by the David and Lucile Packard, the Alfred~P. Sloan,
1500: and the John~D. and Catherine~T. MacArthur Foundations.
1501: This work was further supported by NSF grant AST~05-06556 and NASA ATP
1502: grant NNG~05GJ40G.
1503: 
1504: 
1505: \appendix
1506: \section{Delensed Temperature Field}
1507: \label{app:delens}
1508: Here we derive a relation 
1509: $\That_\vL\simeq T_\vL ~e^{-{1\over2}l^2\sbeam^2}+T^N_\vL$ in the 
1510: presence of telescope beam and detector noise. Given the lensing potential
1511: $\phi^m(\Vang)$ of an assumed mass model, the lensing equation relates an image
1512: position $\Vang$ to a source position $\bhat^m=\Vang+\ang\phi^m(\Vang)$.
1513: Here we keep the superscript $m$ to indicate the relation to the assumed
1514: model. The true source position is then
1515: $\bhat=\Vang+\ang\phi(\Vang)$, where $\phi(\Vang)$ is the true lensing 
1516: potential. Now we construct a delensed temperature field
1517: \bear
1518: \That(\bhat^m)&=&\Tobs(\Vang) \\
1519: &=&\int d^2\Vangm ~B(\Vangm-\Vang)~\lenT(\Vangm)+T^N(\Vang), \nonumber
1520: \enar
1521: where $B(\Vangm)$ is the telescope beam function.
1522: Since the lensing equation is not analytically invertible in general, we
1523: keep both $\bhat^m$ and $\Vang$, but note that they are not independent
1524: variables. In Fourier space, the delensed temperature field is
1525: \beeq
1526: \That_\vL=\int d^2\bhat^m ~\That(\bhat^m)~e^{-i\vL\cdot\bhat^m}\equiv
1527: \That^S_\vL+\That^N_\vL,
1528: \eneq
1529: with a contribution from the CMB
1530: \bear
1531: \That^S_\vL&=&\int d^2\bhat^m \int d^2\Vangm~ B(\Vangm-\Vang)~\lenT(\Vangm)
1532: ~e^{-i\vL\cdot\bhat^m} \nonumber \\
1533: &=&\int d^2\vL_1~ B_{\vL_1}\lenT_{\vL_1}\int{d^2\bhat^m\over(2\pi)^2}~
1534: e^{i\vL_1\cdot\Vang-i\vL\cdot\bhat^m},
1535: \label{app:gen}
1536: \enar
1537: and a contribution from the detector noise
1538: \bear
1539: \That^N_\vL&=&\int d^2\bhat^m ~T^N(\Vang)~e^{-i\vL\cdot\bhat^m}\\
1540: &=&\int d^2\vL_1~ T^N_{\vL_1}\int{d^2\bhat^m\over(2\pi)^2}~
1541: e^{i\vL_1\cdot\Vang-i\vL\cdot\bhat^m}. \nonumber
1542: \enar
1543: 
1544: \begin{figure}
1545: \centerline{\psfig{file=f9.eps, width=3.0in}}
1546: \caption{Effects of telescope beam and detector noise on the delensing 
1547: process. The top panel compares $\That_\vL$ ({\it thin}) with 
1548: $T_\vL ~e^{-{1\over2}l^2\sbeam^2}+T^N_\vL$ ({\it thick}) in terms of their 
1549: power spectrum, and the bottom panel shows the fractional deviations.
1550: The vertical dotted line represents the beam scale $l=1/\sbeam$.
1551: CMB experiments
1552: with $\tfwhm=1'$ and $\spix=5\muK$ are considered for clusters of 
1553: $M=5\times10^{14}\hmsun$ at $z_L=1$. The noise only case is largely obscured by 
1554: the solid line.}
1555: \label{fig:app}
1556: \end{figure}
1557: 
1558: 
1559: The lensed temperature is $\lenT(\Vang)=T(\bhat)$ and its Fourier mode is
1560: \beeq
1561: \lenT_\vL=\int d^2\vL_1~T_{\vL_1}\int{d^2\Vang\over(2\pi)^2}~
1562: e^{-i\vL\cdot\Vang+i\vL_1\cdot\bhat}.
1563: \eneq
1564: With the linear approximation, one can expand the exponential term to the
1565: first order in $\phi_\vL$ and this equation reduces to Eq.~(\ref{eq:lensexp}).
1566: However, we keep the equation as general as possible to be valid, even when
1567: the lensing effect is large. Substituting $\lenT_{\vL_1}$ in Eq.~(\ref{app:gen})
1568: and changing the integration variable $\Vang$ to $\bhat^m$ gives
1569: \bear
1570: \That^S_\vL&=&\int d^2\vL_1\int d^2\vL_2~B_{\vL_1}T_{\vL_2}
1571: \int{d^2\Vang\over(2\pi)^2}\int{d^2\Vang_2\over(2\pi)^2}
1572: \left|{d^2\bhat^m\over d\Vang^2}\right|\nonumber \\
1573: &\times&e^{i\vL_1\cdot(\Vang-\Vang_2)}e^{i(\vL_2\cdot\Vang_2-\vL\cdot\Vang)}~
1574: e^{i\left[\vL_2\cdot\ang\phi(\Vang_2)-\vL\cdot\ang\phi^m(\Vang)\right]}.
1575: \enar
1576: Given the lensing potential $\phi(\Vang)$ (analogously for $\phi^m(\Vang)$),
1577: the Jacobian is related to the distortion matrix 
1578: \beeq
1579: \left|{d^2\bhat\over d\Vang^2}\right|=\left|\bdv{M}^{-1}\right|=
1580: \left|\bdv{I}+\ang\ang\phi\right|=
1581: \left|\left[1-\kappa(\Vang)\right]^2-\gamma^2(\Vang)\right|,
1582: \eneq
1583: and its inverse is the lensing magnification.
1584: 
1585: For a Gaussian beam $B_\vL=\exp\left[-{1\over2}l^2\sbeam^2\right]$, we can
1586: integrate over the beam factor
1587: \bear
1588: \That^S_\vL&=&\int d^2\vL_2~T_{\vL_2}\int{d^2\Vang\over(2\pi)^2}
1589: \int{d^2\Vang_2\over(2\pi)^2}\left|{d^2\bhat^m\over d\Vang^2}\right| \\
1590: &\times&
1591: {2\pi\over\sbeam^2}e^{-{|\Vang-\Vang_2|^2\over2\sbeam^2}}
1592: e^{i(\vL_2\cdot\Vang_2-\vL\cdot\Vang)}~
1593: e^{i\left[\vL_2\cdot\ang\phi(\Vang_2)-\vL\cdot\ang\phi^m(\Vang)\right]}. \nonumber
1594: \enar
1595: Now we parametrize $\Vang_2$ by a dimensionless displacement vector
1596: $\dsmall$ centered at $\Vang$ (i.e., $\Vang_2=\Vang+\sbeam\dsmall$).
1597: The Gaussian beam factor guarantees that the integrand is non-vanishing only
1598: when $\Delta=|\dsmall|$ is small. In order to get more intuition, we expand
1599: $\phi(\Vang_2)\simeq\phi(\Vang)+\ang\phi(\Vang)\cdot\sbeam\dsmall$
1600: to the linear order in $\Delta$, and integrating over $\dsmall$ gives
1601: \bear
1602: \That^S_\vL&=&\int d^2\vL_2~T_{\vL_2}\int{d^2\Vang\over(2\pi)^2}
1603: \left|{d^2\bhat^m\over d\Vang^2}\right|\\
1604: &\times&e^{i(\vL_2\cdot\bhat-\vL\cdot\bhat^m)}~
1605: e^{-{1\over2}\sbeam^2|\bdv{M}^{-1}\cdot\vL_2|^2}. \nonumber
1606: \enar
1607: This is the final expression for the delensed temperature field. The first
1608: exponential
1609: term of the integrand controls the delensing process: when the assumed model
1610: is close to the true model after a few iterations 
1611: ($\phi^m(\Vang)\simeq\phi(\Vang)$, $\bhat^m\simeq\bhat$), the integral becomes
1612: a Dirac delta function and $\That^S_\vL=T_\vL$, when the beam smoothing is
1613: negligible. The distortion matrix is close to the identity matrix beyond
1614: $\Rvir$ and $\That^S_\vL\simeq T_\vL ~e^{-{1\over2}l^2\sbeam^2}$. Around massive
1615: clusters, the distortion matrix deviates from the identity matrix and its
1616: determinant becomes smaller than one, making the exponential factor unity.
1617: This reflects that the beam size is reduced by a mapping from the image
1618: plane to the source plane, and practically 
1619: $\That^S_\vL\simeq T_\vL~ e^{-{1\over2}l^2\tilde\sbeam^2}$ with
1620: $\tilde\sbeam<\sbeam$.
1621: 
1622: For a white detector noise, the delensed detector noise is simply the
1623: redistributed white noise. However, since the delensing process alters the unit
1624: area on the sky,
1625: it becomes non-white but its deviation is confined to relatively small
1626: region; the noise power spectrum is
1627: \bear
1628: \langle\That^N_\vL\That^{N*}_{\vL'}\rangle&=&
1629: \int d^2\vL_1\int d^2\vL_2~\langle T^N_{\vL_1}T^{N*}_{\vL_2}\rangle \nonumber \\
1630: &\times&\int{d^2\bhat^m_1\over(2\pi)^2}\int{d^2\bhat^m_2\over(2\pi)^2}~
1631: e^{i\vL_1\cdot\Vang_1-i\vL\cdot\bhat^m_1}~e^{-i\vL_2\cdot\Vang_2+i\vL'\cdot\bhat^m_2}
1632: \nonumber\\
1633: &=&C^N\int{d^2\bhat^m_1}\int{d^2\bhat^m_2}~\delta(\Vang_1-\Vang_2)~
1634: e^{-i\vL\cdot\bhat^m_1+i\vL'\cdot\bhat^m_2} \nonumber \\
1635: &=&C^N\int{d^2\bhat^m_1}\left|{d^2\bhat^m_1\over d\Vang_1^2}\right|
1636: e^{-i(\vL-\vL')\cdot\bhat^m_1}.
1637: \enar
1638: It is the Jacobian of the distortion matrix that 
1639: makes white noise non-white in a region around massive clusters. Outside
1640: $\Rvir$, where the Jacobian is near unity, the integral becomes a Dirac
1641: delta function and the noise is again white.
1642: 
1643: Figure~\ref{fig:app} compares our delensing ($\That_\vL$: {\it thin})
1644: and perfect delensing ($T_\vL ~e^{-{1\over2}l^2\sbeam^2}+T^N_\vL$: {\it thick})
1645: processes in terms of their power spectrum. In the absence of detector noise
1646: ({\it dashed}), $\That^S_\vL$ starts to deviate from 
1647: $T_\vL~ e^{-{1\over2}l^2\sbeam^2}$ around the beam scale $l\simeq1/\sbeam$
1648: ({\it vertical dotted}), declining less rapidly. On scales below the beam scale,
1649: our approximation ($\Delta\ll1$) breaks down and $\bdv{M}^{-1}(\Vang)$ differs
1650: from the identity matrix, leading to the excess power. 
1651: However, at this scale, signals are dominated by the detector noise
1652: ({\it solid}). Since detector noises are unaffected by the beam distortion
1653: when delensed, the deviation of $\That^N_\vL$ from $T^N_\vL$
1654: is relatively mild and it is solely due to the (inverse) magnification effect
1655: of the mapping from the image plane to the source plane. The noise only
1656: case ({\it dotted}) is largely obscured by the solid line. In summary,
1657: telescope beam and detector noise has little impact on our delensing
1658: process at scales larger than the beam scale, where most of the information
1659: is contained.
1660: 
1661: \vfill
1662: 
1663: \begin{thebibliography}{37}
1664: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1665: \expandafter\ifx\csname bibnamefont\endcsname\relax
1666:   \def\bibnamefont#1{#1}\fi
1667: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1668:   \def\bibfnamefont#1{#1}\fi
1669: \expandafter\ifx\csname citenamefont\endcsname\relax
1670:   \def\citenamefont#1{#1}\fi
1671: \expandafter\ifx\csname url\endcsname\relax
1672:   \def\url#1{\texttt{#1}}\fi
1673: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1674: \providecommand{\bibinfo}[2]{#2}
1675: \providecommand{\eprint}[2][]{\url{#2}}
1676: 
1677: \bibitem[{\citenamefont{{Hirata} et~al.}(2004)\citenamefont{{Hirata},
1678:   {Padmanabhan}, {Seljak}, {Schlegel}, and {Brinkmann}}}]{HIPAET04}
1679: \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{{Hirata}}},
1680:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{{Padmanabhan}}},
1681:   \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}},
1682:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{{Schlegel}}},
1683:   \bibnamefont{and}
1684:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{{Brinkmann}}},
1685:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{70}},
1686:   \bibinfo{pages}{103501} (\bibinfo{year}{2004}),
1687:   \eprint{arXiv:astro-ph/0406004}.
1688: 
1689: \bibitem[{\citenamefont{{Smith} et~al.}(2007)\citenamefont{{Smith}, {Zahn}, and
1690:   {Dor{\'e}}}}]{SMZADO07}
1691: \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{{Smith}}},
1692:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{{Zahn}}}, \bibnamefont{and}
1693:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{{Dor{\'e}}}},
1694:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{76}},
1695:   \bibinfo{pages}{043510} (\bibinfo{year}{2007}), \eprint{arXiv:0705.3980}.
1696: 
1697: \bibitem[{\citenamefont{{Hirata} et~al.}(2008)\citenamefont{{Hirata}, {Ho},
1698:   {Padmanabhan}, {Seljak}, and {Bahcall}}}]{HIHOET08}
1699: \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{{Hirata}}},
1700:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{{Ho}}},
1701:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{{Padmanabhan}}},
1702:   \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}}, \bibnamefont{and}
1703:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{{Bahcall}}},
1704:   \bibinfo{journal}{ArXiv e-prints} \textbf{\bibinfo{volume}{801}}
1705:   (\bibinfo{year}{2008}), \eprint{0801.0644}.
1706: 
1707: \bibitem[{\citenamefont{{Calabrese} et~al.}(2008)\citenamefont{{Calabrese},
1708:   {Slosar}, {Melchiorri}, {Smoot}, and {Zahn}}}]{CASLET08}
1709: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{{Calabrese}}},
1710:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Slosar}}},
1711:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Melchiorri}}},
1712:   \bibinfo{author}{\bibfnamefont{G.~F.} \bibnamefont{{Smoot}}},
1713:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{{Zahn}}},
1714:   \bibinfo{journal}{ArXiv e-prints} \textbf{\bibinfo{volume}{803}}
1715:   (\bibinfo{year}{2008}), \eprint{0803.2309}.
1716: 
1717: \bibitem[{\citenamefont{{Albrecht} et~al.}(2006)}]{DETF06}
1718: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Albrecht}}}
1719:   \bibnamefont{et~al.}, \bibinfo{journal}{ArXiv Astrophysics e-prints}
1720:   (\bibinfo{year}{2006}), \eprint{astro-ph/0609591}.
1721: 
1722: \bibitem[{\citenamefont{{Hu} et~al.}(2007{\natexlab{a}})\citenamefont{{Hu},
1723:   {Holz}, and {Vale}}}]{HUHOVA07}
1724: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Hu}}},
1725:   \bibinfo{author}{\bibfnamefont{D.~E.} \bibnamefont{{Holz}}},
1726:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{{Vale}}},
1727:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{76}},
1728:   \bibinfo{pages}{127301} (\bibinfo{year}{2007}{\natexlab{a}}),
1729:   \eprint{arXiv:0708.4391}.
1730: 
1731: \bibitem[{\citenamefont{{Silk}}(1968)}]{SILK68}
1732: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{{Silk}}},
1733:   \bibinfo{journal}{\apjl} \textbf{\bibinfo{volume}{151}},
1734:   \bibinfo{pages}{L19+} (\bibinfo{year}{1968}).
1735: 
1736: \bibitem[{\citenamefont{{Seljak} and {Zaldarriaga}}(2000)}]{SEZA95}
1737: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}} \bibnamefont{and}
1738:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Zaldarriaga}}},
1739:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{538}}, \bibinfo{pages}{57}
1740:   (\bibinfo{year}{2000}), \eprint{arXiv:astro-ph/9907254}.
1741: 
1742: \bibitem[{\citenamefont{{Vale} et~al.}(2004)\citenamefont{{Vale}, {Amblard},
1743:   and {White}}}]{VAAMWH04}
1744: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{{Vale}}},
1745:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Amblard}}},
1746:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{White}}},
1747:   \bibinfo{journal}{New Astronomy} \textbf{\bibinfo{volume}{10}},
1748:   \bibinfo{pages}{1} (\bibinfo{year}{2004}), \eprint{arXiv:astro-ph/0402004}.
1749: 
1750: \bibitem[{\citenamefont{{Holder} and {Kosowsky}}(2004)}]{HOKO04}
1751: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{{Holder}}} \bibnamefont{and}
1752:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Kosowsky}}},
1753:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{616}}, \bibinfo{pages}{8}
1754:   (\bibinfo{year}{2004}), \eprint{arXiv:astro-ph/0401519}.
1755: 
1756: \bibitem[{\citenamefont{{Hu}}(2001)}]{HU01b}
1757: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Hu}}},
1758:   \bibinfo{journal}{\apjl} \textbf{\bibinfo{volume}{557}}, \bibinfo{pages}{L79}
1759:   (\bibinfo{year}{2001}), \eprint{arXiv:astro-ph/0105424}.
1760: 
1761: \bibitem[{\citenamefont{{Hirata} and {Seljak}}(2003)}]{HISE03a}
1762: \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{{Hirata}}} \bibnamefont{and}
1763:   \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}},
1764:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{67}},
1765:   \bibinfo{pages}{043001} (\bibinfo{year}{2003}),
1766:   \eprint{arXiv:astro-ph/0209489}.
1767: 
1768: \bibitem[{\citenamefont{{Maturi} et~al.}(2005)\citenamefont{{Maturi},
1769:   {Bartelmann}, {Meneghetti}, and {Moscardini}}}]{MABAMEET05}
1770: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Maturi}}},
1771:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Bartelmann}}},
1772:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Meneghetti}}},
1773:   \bibnamefont{and}
1774:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{{Moscardini}}},
1775:   \bibinfo{journal}{\aap} \textbf{\bibinfo{volume}{436}}, \bibinfo{pages}{37}
1776:   (\bibinfo{year}{2005}), \eprint{arXiv:astro-ph/0408064}.
1777: 
1778: \bibitem[{\citenamefont{{Hu} et~al.}(2007{\natexlab{b}})\citenamefont{{Hu},
1779:   {DeDeo}, and {Vale}}}]{HUDEVA07}
1780: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Hu}}},
1781:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{{DeDeo}}}, \bibnamefont{and}
1782:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{{Vale}}},
1783:   \bibinfo{journal}{New Journal of Physics} \textbf{\bibinfo{volume}{9}},
1784:   \bibinfo{pages}{441} (\bibinfo{year}{2007}{\natexlab{b}}),
1785:   \eprint{arXiv:astro-ph/0701276}.
1786: 
1787: \bibitem[{\citenamefont{{Tegmark} and et~al.}(2006)}]{TESTET06}
1788: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Tegmark}}} \bibnamefont{and}
1789:   \bibinfo{author}{\bibnamefont{et~al.}}, \bibinfo{journal}{\prd}
1790:   \textbf{\bibinfo{volume}{74}}, \bibinfo{pages}{123507}
1791:   (\bibinfo{year}{2006}), \eprint{arXiv:astro-ph/0608632}.
1792: 
1793: \bibitem[{\citenamefont{{Spergel} and et~al.}(2007)}]{SPBEET07}
1794: \bibinfo{author}{\bibfnamefont{D.~N.} \bibnamefont{{Spergel}}}
1795:   \bibnamefont{and} \bibinfo{author}{\bibnamefont{et~al.}},
1796:   \bibinfo{journal}{\apjs} \textbf{\bibinfo{volume}{170}}, \bibinfo{pages}{377}
1797:   (\bibinfo{year}{2007}), \eprint{arXiv:astro-ph/0603449}.
1798: 
1799: \bibitem[{\citenamefont{{Komatsu} and et~al.}(2008)}]{KODUET08}
1800: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{{Komatsu}}} \bibnamefont{and}
1801:   \bibinfo{author}{\bibnamefont{et~al.}}, \bibinfo{journal}{ArXiv e-prints}
1802:   \textbf{\bibinfo{volume}{803}} (\bibinfo{year}{2008}), \eprint{0803.0547}.
1803: 
1804: \bibitem[{\citenamefont{{Hu}}(2000)}]{HU00}
1805: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Hu}}}, \bibinfo{journal}{\prd}
1806:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{043007}
1807:   (\bibinfo{year}{2000}), \eprint{arXiv:astro-ph/0001303}.
1808: 
1809: \bibitem[{\citenamefont{{Okamoto} and {Hu}}(2003)}]{OKHU03}
1810: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{{Okamoto}}} \bibnamefont{and}
1811:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{Hu}}},
1812:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{67}},
1813:   \bibinfo{pages}{083002} (\bibinfo{year}{2003}),
1814:   \eprint{arXiv:astro-ph/0301031}.
1815: 
1816: \bibitem[{\citenamefont{{Challinor} and {Lewis}}(2005)}]{CHLE05}
1817: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Challinor}}} \bibnamefont{and}
1818:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Lewis}}},
1819:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{71}},
1820:   \bibinfo{pages}{103010} (\bibinfo{year}{2005}),
1821:   \eprint{arXiv:astro-ph/0502425}.
1822: 
1823: \bibitem[{\citenamefont{{Mandel} and {Zaldarriaga}}(2006)}]{MAZA06}
1824: \bibinfo{author}{\bibfnamefont{K.~S.} \bibnamefont{{Mandel}}} \bibnamefont{and}
1825:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Zaldarriaga}}},
1826:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{647}}, \bibinfo{pages}{719}
1827:   (\bibinfo{year}{2006}), \eprint{arXiv:astro-ph/0512218}.
1828: 
1829: \bibitem[{\citenamefont{{Lewis} and {Challinor}}(2006)}]{LECH06}
1830: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Lewis}}} \bibnamefont{and}
1831:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Challinor}}},
1832:   \bibinfo{journal}{\physrep} \textbf{\bibinfo{volume}{429}},
1833:   \bibinfo{pages}{1} (\bibinfo{year}{2006}), \eprint{arXiv:astro-ph/0601594}.
1834: 
1835: \bibitem[{\citenamefont{{Knox}}(1995)}]{KNOX95}
1836: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{{Knox}}},
1837:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{52}}, \bibinfo{pages}{4307}
1838:   (\bibinfo{year}{1995}), \eprint{arXiv:astro-ph/9504054}.
1839: 
1840: \bibitem[{\citenamefont{{Zahn} and {Zaldarriaga}}(2006)}]{ZAZA06}
1841: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{{Zahn}}} \bibnamefont{and}
1842:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Zaldarriaga}}},
1843:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{653}}, \bibinfo{pages}{922}
1844:   (\bibinfo{year}{2006}), \eprint{arXiv:astro-ph/0511547}.
1845: 
1846: \bibitem[{\citenamefont{{Babich}}(2005)}]{BABIC05}
1847: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{{Babich}}},
1848:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{72}},
1849:   \bibinfo{pages}{043003} (\bibinfo{year}{2005}),
1850:   \eprint{arXiv:astro-ph/0503375}.
1851: 
1852: \bibitem[{\citenamefont{{Navarro} et~al.}(1997)\citenamefont{{Navarro},
1853:   {Frenk}, and {White}}}]{NFW97}
1854: \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{{Navarro}}},
1855:   \bibinfo{author}{\bibfnamefont{C.~S.} \bibnamefont{{Frenk}}},
1856:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~D.~M.}
1857:   \bibnamefont{{White}}}, \bibinfo{journal}{\apj}
1858:   \textbf{\bibinfo{volume}{490}}, \bibinfo{pages}{493} (\bibinfo{year}{1997}),
1859:   \eprint{astro-ph/9611107}.
1860: 
1861: \bibitem[{\citenamefont{{Bartelmann}}(1996)}]{BARTE96}
1862: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Bartelmann}}},
1863:   \bibinfo{journal}{\aap} \textbf{\bibinfo{volume}{313}}, \bibinfo{pages}{697}
1864:   (\bibinfo{year}{1996}), \eprint{arXiv:astro-ph/9602053}.
1865: 
1866: \bibitem[{\citenamefont{{Wright} and {Brainerd}}(2000)}]{WRBR00}
1867: \bibinfo{author}{\bibfnamefont{C.~O.} \bibnamefont{{Wright}}} \bibnamefont{and}
1868:   \bibinfo{author}{\bibfnamefont{T.~G.} \bibnamefont{{Brainerd}}},
1869:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{534}}, \bibinfo{pages}{34}
1870:   (\bibinfo{year}{2000}).
1871: 
1872: \bibitem[{\citenamefont{{Seljak} and {Zaldarriaga}}(1996)}]{SEZA96}
1873: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}} \bibnamefont{and}
1874:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Zaldarriaga}}},
1875:   \bibinfo{journal}{\apj} \textbf{\bibinfo{volume}{469}}, \bibinfo{pages}{437}
1876:   (\bibinfo{year}{1996}), \eprint{astro-ph/9603033}.
1877: 
1878: \bibitem[{\citenamefont{{Sunyaev} and {Zeldovich}}(1970)}]{SUZE70}
1879: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{{Sunyaev}}}
1880:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{Y.~B.}
1881:   \bibnamefont{{Zeldovich}}}, \bibinfo{journal}{Comments on Astrophysics and
1882:   Space Physics} \textbf{\bibinfo{volume}{2}}, \bibinfo{pages}{66}
1883:   (\bibinfo{year}{1970}).
1884: 
1885: \bibitem[{\citenamefont{{Sunyaev} and {Zeldovich}}(1972)}]{SUZE72}
1886: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{{Sunyaev}}}
1887:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{Y.~B.}
1888:   \bibnamefont{{Zeldovich}}}, \bibinfo{journal}{Comments on Astrophysics and
1889:   Space Physics} \textbf{\bibinfo{volume}{4}}, \bibinfo{pages}{173}
1890:   (\bibinfo{year}{1972}).
1891: 
1892: \bibitem[{\citenamefont{{Dodelson}}(2004)}]{DODEL04}
1893: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{{Dodelson}}},
1894:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{70}},
1895:   \bibinfo{pages}{023009} (\bibinfo{year}{2004}),
1896:   \eprint{arXiv:astro-ph/0402314}.
1897: 
1898: \bibitem[{\citenamefont{{Lewis} and {King}}(2006)}]{LEKI06}
1899: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Lewis}}} \bibnamefont{and}
1900:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{{King}}},
1901:   \bibinfo{journal}{\prd} \textbf{\bibinfo{volume}{73}},
1902:   \bibinfo{pages}{063006} (\bibinfo{year}{2006}),
1903:   \eprint{arXiv:astro-ph/0512104}.
1904: 
1905: \bibitem[{\citenamefont{{Birkinshaw}}(1999)}]{BIRKI99}
1906: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Birkinshaw}}},
1907:   \bibinfo{journal}{\physrep} \textbf{\bibinfo{volume}{310}},
1908:   \bibinfo{pages}{97} (\bibinfo{year}{1999}), \eprint{arXiv:astro-ph/9808050}.
1909: 
1910: \bibitem[{\citenamefont{{Carlstrom} et~al.}(2002)\citenamefont{{Carlstrom},
1911:   {Holder}, and {Reese}}}]{CAHORE02}
1912: \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{{Carlstrom}}},
1913:   \bibinfo{author}{\bibfnamefont{G.~P.} \bibnamefont{{Holder}}},
1914:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.~D.}
1915:   \bibnamefont{{Reese}}}, \bibinfo{journal}{\araa}
1916:   \textbf{\bibinfo{volume}{40}}, \bibinfo{pages}{643} (\bibinfo{year}{2002}),
1917:   \eprint{arXiv:astro-ph/0208192}.
1918: 
1919: \bibitem[{\citenamefont{{Seljak} and {Zaldarriaga}}(1997)}]{SEZA97}
1920: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{{Seljak}}} \bibnamefont{and}
1921:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Zaldarriaga}}},
1922:   \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{78}}, \bibinfo{pages}{2054}
1923:   (\bibinfo{year}{1997}), \eprint{arXiv:astro-ph/9609169}.
1924: 
1925: \bibitem[{\citenamefont{{Kamionkowski}
1926:   et~al.}(1997)\citenamefont{{Kamionkowski}, {Kosowsky}, and
1927:   {Stebbins}}}]{KAKOST97}
1928: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{{Kamionkowski}}},
1929:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Kosowsky}}},
1930:   \bibnamefont{and}
1931:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{{Stebbins}}},
1932:   \bibinfo{journal}{\prl} \textbf{\bibinfo{volume}{78}}, \bibinfo{pages}{2058}
1933:   (\bibinfo{year}{1997}), \eprint{arXiv:astro-ph/9609132}.
1934: 
1935: \end{thebibliography}
1936: 
1937: 
1938: \end{document}
1939: 
1940: