1: \documentclass{emulateapj}
2: %\documentclass[manuscript]{aastex}
3: %\documentclass[12pt,preprint]{aastex}
4:
5:
6: \usepackage{graphicx}
7: \usepackage{epsf}
8: %\usepackage{apjfonts}
9: \bibliographystyle{apj}
10:
11:
12: %%%
13: \newcommand{\nM}{\frac{dn}{dM}}
14: \newcommand{\Mobs}{M_{\rm obs}}
15: \newcommand{\nMobs}{\frac{dn}{d\rm\Mobs}}
16: \newcommand{\Npix}{N_{\rm pix}}
17: \newcommand{\Nz}{N_{\rm z}}
18: \newcommand{\NM}{N_{\rm M}}
19: \newcommand{\Var}{\mbox{Var}}
20: \newcommand{\tc}{c'}
21: \newcommand{\bm}[1]{\mathbf{#1}}
22: \newcommand{\bOmega}{\bm{\Omega}}
23: \newcommand{\Mbias}{M_{\rm bias}}
24: \newcommand{\sigmalnM}{\sigma_{\rm \ln M}}
25: \newcommand{\avg}[1]{\langle#1\rangle}
26: \newcommand{\mbar}{{\bar m}}
27: \newcommand{\C}{\bold C}
28: \newcommand{\D}{\bold D}
29: \newcommand{\bS}{\bold S}
30: \newcommand{\himpc}{h^{-1}{\rm Mpc}}
31: \newcommand{\hiMsun}{h^{-1}{M_\odot}}
32: \newcommand{\cvir}{c_{\rm vir}}
33: \newcommand{\OmegaDE}{\Omega_{\rm DE}}
34: %%%
35:
36:
37: \begin{document}
38:
39:
40: \journalinfo{The Astrophysical Journal, 688:729--741, 2008 December 1 }
41: \submitted{Received 2008 March 11; accepted 2008 July 13}
42: \shortauthors{WU, ROZO, \& WECHSLER}
43: \shorttitle{EFFECTS OF HALO ASSEMBLY BIAS ON SELF-CALIBRATION}
44: \title{The Effects of Halo Assembly Bias on Self-Calibration in Galaxy Cluster Surveys}
45:
46: \author{Hao-Yi Wu \altaffilmark{1}, Eduardo Rozo \altaffilmark{2},
47: Risa H. Wechsler \altaffilmark{1}}
48:
49: \altaffiltext{1}{Kavli Institute for Particle Astrophysics and
50: Cosmology, Physics Department, Stanford Linear Accelerator Center,
51: Stanford University, Stanford, CA 94305; hywu@stanford.edu,
52: rwechsler@stanford.edu}
53:
54: \altaffiltext{2}{The Center for Cosmology and Astro-Particle Physics,
55: The Ohio State University, Columbus, OH 43210;
56: erozo@mps.ohio-state.edu}
57:
58:
59: \begin{abstract}
60:
61: Self-calibration techniques for analyzing galaxy cluster counts
62: utilize the abundance and the clustering amplitude of dark matter
63: halos. These properties simultaneously constrain cosmological
64: parameters and the cluster observable--mass relation. It was recently
65: discovered that the clustering amplitude of halos depends not only on
66: the halo mass, but also on various secondary variables, such as the
67: halo formation time and the concentration; these dependences are
68: collectively termed ``assembly bias.'' Applying modified Fisher
69: matrix formalism, we explore whether these secondary variables have a
70: significant impact on the study of dark energy properties using the
71: self-calibration technique in current (SDSS) and the near future (DES,
72: SPT, and LSST) cluster surveys. The impact of the secondary dependence
73: is determined by (1) the scatter in the observable--mass relation and
74: (2) the correlation between observable and secondary variables. We
75: find that for optical surveys, the secondary dependence does not
76: significantly influence an SDSS-like survey; however, it may affect a
77: DES-like survey (given the high scatter currently expected from
78: optical clusters) and an LSST-like survey (even for low scatter values
79: and low correlations). For an SZ survey such as SPT, the impact of
80: secondary dependence is insignificant if the scatter is 20\% or lower
81: but can be enhanced by the potential high scatter values introduced by
82: a highly -correlated background. Accurate modeling of the assembly
83: bias is necessary for cluster self-calibration in the era of precision
84: cosmology.
85:
86: \end{abstract}
87:
88:
89: \keywords{cosmology: theory --- cosmological parameters ---
90: large-scale structure of universe --- galaxies: clusters: general ---
91: galaxies: halos --- methods: statistical}
92:
93:
94: \section{Introduction}
95:
96: The observed accelerating expansion of the Universe, which is often
97: interpreted as evidence of dark energy, is one of the most surprising
98: results of modern cosmology. In the $\Lambda$CDM paradigm, dark
99: energy governs the late time expansion of the Universe, halting the
100: growth of structures. Consequently, the evolution of the number of
101: massive galaxy clusters provides one of the most powerful probes of
102: dark energy
103: \citep[e.g.][]{Wang98,Haiman01,Holder01,Levine02,Hu03,Majumdar03,Rozo07a,Gladders07,Mantz07}.
104:
105:
106: Several planned and ongoing surveys will identify massive clusters
107: over substantial volumes using a variety of techniques, including
108: optical galaxy counts \cite[e.g.][]{SDSS, DES, LSST}, the
109: Sunyaev-Zel'dovich effect \cite[e.g.][]{SPT, ACT}, and X-ray emissions
110: \cite[e.g.][]{Ebeling07MACS,400d}. These cluster surveys will
111: complement a variety of future dark energy measurements using tools
112: such as Type Ia supernovae, weak lensing, and baryon acoustic
113: oscillations. Since each of these methods is subject to different
114: systematics, combining them thus provides cross checks necessary to
115: avoid biased inferences on the properties of dark energy
116: \citep{Albrecht06}.
117:
118:
119: While the abundance of clusters as a function of mass is well
120: understood from a theoretical standpoint, measuring this abundance
121: relies on observable mass tracers. This reliance is the single most
122: significant obstacle confronting the use of clusters as cosmological
123: probes. In particular, the statistical observable--mass relation
124: needs to be understood to high accuracy in order to avoid systematic
125: errors in the inference of cosmological parameters. Alternatively,
126: additional observable quantities that depend on halo mass allow one to
127: simultaneously constrain cosmology and the aforementioned
128: observable--mass relation. One such observable quantity is the
129: clustering amplitude of clusters, which depends sensitively on mass
130: and can be determined through a counts-in-cells analysis. This
131: general method is often referred to as ``self-calibration''
132: \citep{Majumdar04,LimaHu04,LimaHu05,LimaHu07}.
133:
134:
135: In this work, we explore a possible systematic that arises in the
136: self-calibration analysis, namely, the dependence of the clustering
137: amplitude of halos on secondary variables. The clustering amplitude
138: of halos is characterized by the halo bias, and recent studies have
139: shown that halo bias depends not only on halo mass but also on
140: additional halo properties, such as concentration, formation time,
141: spin, substructure fraction, etc. \citep[e.g.][]{Gao05, Harker06,
142: Wechsler06, GaoWhite07, Wetzel07, Bett07, Jing07}. These dependences
143: are often interpreted as arising from the different assembly histories
144: of halos of the same mass, and we refer to these dependences
145: collectively as ``assembly bias'' \citep[e.g.][]{Croton07}. If
146: cluster selection is biased with respect to any of these variables,
147: the observed clustering amplitude of clusters will deviate from the
148: mean clustering amplitude of clusters with the same mass distribution.
149: This deviation will lead to a biased inference of the observable--mass
150: relation, and therefore to biased estimates for the cosmological
151: parameters of interest.
152:
153:
154: We herein take the secondary parameter to be the halo concentration,
155: which has been shown to play a role in halo bias for massive clusters
156: by Wechsler et al.\@ (2006; see also \citealt{Wetzel07};
157: \citealt{Jing07}). We then incorporate the concentration dependence
158: of halo bias into the standard self-calibration formalism developed in
159: \cite{LimaHu04, LimaHu05}. With modified Fisher matrix formalism, we
160: investigate the impact of this additional dependence on cosmological
161: parameter estimates from self-calibration. We specifically calculate
162: the expected effects for four example galaxy cluster surveys, which
163: represent the Sloan Digital Sky Survey (SDSS; assuming clusters
164: selected from the photometric data), the Dark Energy Survey (DES), the
165: South Pole Telescope (SPT), and the Large Synoptic Survey Telescope
166: (LSST). We also explore various assumptions about the correlation
167: between cluster observable and concentration. In detail, the
168: significance of this systematic effect depends on the strength of this
169: correlation as well as on the observable--mass scatter. We find that
170: the resulting bias in the inferred cosmological parameters is
171: insignificant for the current SDSS photometric surveys, but it can be
172: significant for upcoming photometric surveys such as DES and LSST. On
173: the other hand, for SZ this systematic is less likely to be
174: significant if the scatter is small and mainly intrinsic, but may
175: still be significant if the correlated background dominates the
176: scatter.
177:
178:
179: This paper is organized as follows. In \S\ref{sec:bias} we discuss
180: why assembly bias may lead to biased cosmological parameter estimates
181: in cluster counting experiments. In \S\ref{sec:self-calibration} we
182: review the standard self-calibration formalism, and then proceed in
183: \S\ref{sec:assemblybias} to include assembly bias into this formalism.
184: Our statistical methodology for estimating the systematic errors due
185: to assembly bias is described in \S\ref{sec:fisher}. Details of our
186: implementation can be found in
187: \S\ref{sec:implementation}. Section \ref{sec:results} presents our results
188: and discussion. We summarize in \S\ref{sec:conclusions}.
189:
190: %-----------------------------------------------
191: %-----------------------------------------------
192: %-----------------------------------------------
193: %-----------------------------------------------
194: %-----------------------------------------------
195: %-----------------------------------------------
196:
197:
198: \section{Halo Bias and Dark Energy: Why Assembly Bias Matters}\label{sec:bias}
199:
200: Halo bias characterizes the clustering amplitude of dark matter halos,
201: and it is typically defined as the ratio between the density contrast
202: of halos and that of the dark matter. In the hierarchical structure
203: formation predicted by CDM, halo bias is a strong function of mass,
204: increasing for more massive halos. This dependence on mass is now
205: well calibrated from numerical simulations and can be approximated
206: analytically with the excursion-set theory
207: \citep[e.g.][]{MoWhite96,Sheth01, SeljakWarren04, Zentner07}. Halo
208: bias depends sensitively on dark energy in a way that is complementary
209: to the dependence of the mass function on dark energy; thus, including
210: the halo bias information in cluster counting experiments improves the
211: dark energy constraints from using mean halo abundances alone.
212:
213:
214: Much work on halo bias has made the simplifying assumption that halo
215: bias depends only on halo mass. However, recent studies based on
216: \emph{N}-body simulations have found evidence that secondary variables such
217: as halo formation time and concentration do impact halo bias
218: \citep[e.g.][]{Gao05,Harker06,Wechsler06, GaoWhite07,
219: Wetzel07,Bett07,Jing07}. In this work, we focus on the impact of halo
220: concentration on halo bias, principally because among all secondary
221: parameters, this dependence is the strongest at cluster scales and is
222: the best understood statistically. The halo concentration describes
223: the halo density profile and is defined as $c = R_{\rm vir}/r_s$, where
224: $r_s$ is the radius where the density profile has a log slope of $-2$.
225: The halo concentration has been shown to correlate tightly with the
226: halo formation epoch by e.g.\@ \cite{Wechsler02}.
227:
228:
229: We specifically use the fitting formula given by
230: \citet[][eq.\@ 6]{Wechsler06}:
231: %
232: \begin{equation}
233: b^{\rm ab}(M,c) = b_{\rm avg}(M)\times b_c(c|M/M_*)
234: \label{eq:W06}
235: \end{equation}
236: %
237: where $b_{\rm avg}(M)$ is the mean halo bias at fixed mass,
238: $b_c(c|M/M_*)$ characterizes the concentration dependence of halo
239: bias, and $M_*$ is the characteristic mass of gravitational collapse
240: [quantitatively defined as $\sigma(M_*)=1.686$, where $\sigma(M)$ is
241: the r.m.s density fluctuation inside a sphere that encloses mass $M$].
242: The superscript ``ab'' refers to ``assembly bias,'' which we use as a
243: generic term for the dependence of halo bias on secondary variables,
244: based on the conjecture that these dependences arise through the
245: different formation histories of halos of the same mass. We assume
246: this formula holds for all clusters included in our fiducial surveys,
247: although part of these clusters are outside the range where this
248: formula has been calibrated with simulations. In addition, we note
249: that \cite{Wechsler06} calibrated this formula with $M_{\rm vir}$, while
250: the mass function and the halo bias we use are not always
251: well-calibrated with the same mass definition. We ignore the
252: systematic errors that may be caused by these uncertainties.
253:
254:
255: \begin{figure*}[t!]
256: %\plotone{assemblybias.eps}
257: \plotone{f1.eps}
258: \caption{
259: {\it Left:} Dependence of halo bias on concentration at
260: $z=0$ assumed in this work, based on the fitting formula of
261: $b^{\rm ab}(M,c)$ in \protect\cite{Wechsler06}. We assume a WMAP3
262: cosmology and log-normally distributed concentrations at a given halo
263: mass. Halos are binned by concentration into four quartiles, and the
264: halo bias of each quartile systematically deviates from the average
265: halo bias (\emph{solid curve}). Above $10^{13.5} \hiMsun$, low concentration
266: halos (\emph{red and orange dotted curves}) are more clustered than high
267: concentration ones (\emph{green and blue dashed curves}) of the same mass.
268: The bottom panel shows the residual compared with the average halo
269: bias. {\it Right panel:} Degeneracy between high dark energy density
270: and assembly bias. The solid curve shows the cumulative bias
271: (eq.\@~[\ref{eq:cbias}], with the selection function nonzero above a
272: threshold $M_{\rm th}$) for the fiducial WMAP3 cosmology. The dashed
273: curve shows the effect of assembly bias with the assumption of
274: perfectly anti-correlated cluster observable and concentration (see
275: \S\ref{sec:bias} and \ref{sec:assemblybias} for details). This
276: correlation can mimic the effect of high dark energy density (here
277: assumed to be $\OmegaDE = 0.9$), shown as the dotted curve.
278: }
279: \label{fig:assemblybias}
280: \end{figure*}
281:
282:
283: The left panel of Figure~\ref{fig:assemblybias} illustrates how
284: concentration impacts halo bias in the fitting formula of
285: \citet[][]{Wechsler06}. As can be seen, for $M\gtrsim10^{13.5}
286: \hiMsun$, low concentration halos are more clustered than high
287: concentration ones of the same mass. This difference is potentially
288: significant: if the cluster observable is correlated with
289: concentration, one might measure cluster bias that differs from the
290: mean halo bias for random halos of the same mass.
291:
292:
293: The right panel of Figure~\ref{fig:assemblybias} shows how the effect
294: of assembly bias can resemble that of a high dark energy density, with
295: an extreme assumption of perfectly anti-correlated observable and
296: concentration. Cumulative bias, which is relevant for halo samples
297: above a certain observable threshold (see eq.~[\ref{eq:cbias}]), is
298: plotted here. As can be seen, if we tend to observe low concentration
299: halos, the effect of assembly bias (\emph{dashed curve}) makes the
300: observed halo bias higher than the mean halo bias (averaged over
301: random halos samples of the same mass) for the same fiducial cosmology
302: (\emph{solid curve}). This effect mimics a high dark energy density
303: $\OmegaDE = 0.9$ (\emph{dotted curve}), since high $\OmegaDE$ will make
304: structures rarer and more clustered. Thus, a wrong inference of
305: $\OmegaDE$ is possible if assembly bias is ignored in this case. In
306: the following sections, we provide detailed formalism and analyses of
307: such systematics under the framework of the self-calibration of
308: observable--mass relation.
309:
310:
311: %-----------------------------------------------
312: %-----------------------------------------------
313: %-----------------------------------------------
314: %-----------------------------------------------
315: %-----------------------------------------------
316: %-----------------------------------------------
317:
318:
319: \section{Formalism}
320:
321: \subsection{Counts-in-Cells Analysis and Basic Self-Calibration: A Review}\label{sec:self-calibration}
322:
323: In a pixelated galaxy cluster survey, halo bias is related to the
324: sample variance of cluster counts within the small sub-volumes of the
325: survey \citep{HuKravtsov03}. Including the sample variance in a
326: counts-in-cells analysis allows one to ``self-calibrate'' the
327: observable--mass distribution, which is one of the main uncertainties
328: in modeling the surveys. This approach can thereby improve the dark
329: energy constraints relative to ``counts only'' experiments
330: \citep{Majumdar04,LimaHu04,LimaHu05,LimaHu07}. In this section, we
331: review the basic self-calibration, closely following the formalism
332: developed by \cite{LimaHu04}.
333:
334:
335: Given a large-volume survey, consider a redshift slice which is
336: sufficiently thin to make evolution ignorable. We then divide the
337: area of this slice into equal-area cells and count the clusters in
338: each cell.\footnote{ Here we suppress all redshift dependence in our
339: notation for simplicity. In practice, we consider the redshift
340: dependence of the mass function, the halo bias, the observable--mass
341: distribution, and the comoving survey volume. For readers of
342: \cite{LimaHu04}, note that our notation is slightly different. Since
343: we consider a single redshift slice, our subscript $i$ indicates the
344: cell label of the same redshift, while in \cite{LimaHu04}, their
345: subscript $i$ indicates a cell of redshift $z_i$. } The number of
346: clusters in cell $i$, denoted by $N_i$, is affected by the Poisson
347: shot noise, which is modeled as $N_i \sim {\rm Poisson}(m_i)$. This
348: Poisson mean $m_i$ varies from cell to cell due to the large-scale
349: clustering of matter and halos, and this fluctuation can be modeled as
350: a normal distribution $m_i \sim {\rm N}(\mbar, S)$, where $\mbar$ is
351: the mean halo abundance and $S$ is the sample variance.
352:
353:
354: In a given mass range, the mean number counts of clusters in cell $i$
355: depend on $\bar m$, the bias integrated over the mass range $\bar b$,
356: and the mass overdensity $\delta_i$ within this cell with respect to
357: the background:
358: %
359: \begin{equation}
360: m_i = \mbar (1+ \bar b\ \delta_i)\ .
361: \end{equation}
362: %
363: The sample variance then has the form
364: \begin{eqnarray}
365: S&=& \avg{(m_i-\mbar)^2} \nonumber\\
366: &=& \mbar^2 \bar b^2 \sigma_{\rm V}^2 \ ,
367: \end{eqnarray}
368: where
369: %
370: \begin{equation}
371: \sigma_{\rm V}^2 = \frac{1}{V^2}\int\frac{d^3\vec k}{(2 \pi)^3}W(\vec k)W^*(\vec k) P(k) \ .
372: \end{equation}
373: %
374: Here $P(k)$ is the matter power spectrum and $W(\vec k)$ is the
375: \emph{k}-space window function of a cell of volume $V$, normalized such that
376: $V = \int d^3\vec x W(\vec x)$. Applying a counts-in-cells analysis,
377: $N_i$ of each cell can be measured, and $\mbar$ and $S$ can be
378: obtained from a likelihood analysis. With additional knowledge of the
379: matter power spectrum, $\bar b$ can be obtained.
380:
381:
382: Note that this sample variance should be more rigorously defined as
383: the sample {\it covariance}
384: %
385: \begin{eqnarray}
386: S_{ij}&=& \avg{(m_i-\mbar)(m_j-\mbar)} \\
387: &=& \mbar^2 \bar b^2 \sigma_{ij}^2 \ , \nonumber
388: \end{eqnarray}
389: %
390: with
391: %
392: \begin{equation}
393: \sigma_{ij}^2 = \frac{1}{V_i V_j}\int\frac{d^3\vec k}{(2 \pi)^3}W_i(\vec k)W^*_j(\vec k) P(k) \ .
394: \end{equation}
395: %
396: In practice, our cell size is much larger than the correlation length
397: of clusters; thus, the correlations between different cells are
398: negligible. The off-diagonal elements are therefore much smaller then
399: the diagonal ones, and the matrix $S_{ij}$ reduces to a diagonal
400: matrix $S_{ij} = \delta_{ij} S$, whose dimension equals $n_c$, the
401: number of cells in the redshift slice.
402:
403:
404: We next relate these measurable quantities to theoretical models. Let
405: $\Mobs$ denote the observed mass proxy (the observable) of galaxy
406: clusters. Given a differential mass function $dn/dM$ and an
407: observable--mass distribution $P(\Mobs|M)$, the differential observed
408: cluster abundance is given as
409: %
410: \begin{equation}
411: \nMobs = \int dM\ \nM P(\Mobs|M) \ .
412: \end{equation}
413: %
414: In terms of the binning function $\phi(\Mobs)$---which is defined to
415: be equal to unity if $\Mobs$ falls in the bin corresponding to the
416: observable range, and zero otherwise---and the cell volume $V$, the
417: mean observed cluster abundance reads
418: %
419: \begin{equation}
420: \bar m = V \int d\Mobs\ \nMobs \phi(\Mobs) \ ,
421: \end{equation}
422: %
423: which can be further simplified as
424: %
425: \begin{equation}
426: \bar m = V \int dM\ \nM \avg{\phi|M}
427: \label{eq:mean}
428: \end{equation}
429: %
430: if we define the selection function to be
431: %
432: \begin{equation}
433: \avg{\phi|M} = \int d\Mobs\ P(\Mobs|M) \phi(\Mobs) \ .
434: \end{equation}
435: %
436: Given the halo bias $b(M)$, the bias integrated over the observable
437: bin similarly reads
438: %
439: \begin{equation}
440: \bar b = \frac{V}{\bar m}\int dM\ \nM b(M) \avg{\phi|M} \ .
441: \label{eq:cbias}
442: \end{equation}
443: %
444: From equations~\ref{eq:mean} and~\ref{eq:cbias} we can see that if
445: both $\bar m$ and $\bar b$ are measured in the survey, the selection
446: function $\avg{\phi|M}$ can be self-calibrated.
447:
448:
449: In large-volume surveys, we often have several redshift bins and need
450: to consider how $\mbar$ and $S$ vary with redshift: $\mbar(z)$ and
451: $S(z)$. The sample variance is then generalized to the matrix $\bold
452: S = {\rm diag} (\bold S_{ij}(z_1),\bold S_{ij}(z_2),...)$, where each
453: $\bold S_{ij}(z_k)$ has the dimension $n_c\times n_c$. Similarly,
454: $\mbar$ is generalized as $\bold \mbar = (\bold\mbar(z_1),
455: \bold\mbar(z_2),...)$, with each $\bold\mbar(z_k) $ being a $n_c$
456: component vector. For future reference, we further define $\bold M =
457: {\rm diag}(\bold\mbar)$ and $\C = \bold M +\bold S$; $\C$ is the
458: covariance matrix in the limit of large cluster numbers in a cell
459: ($m_i\gg 1$; see \citealt{LimaHu04}).
460:
461:
462: Constraints on dark energy parameters are extracted from the
463: likelihood function that involves the counts-in-cells data, the
464: theoretical mean abundance, and the theoretical sample variance. For
465: theoretical forecasts, the Fisher matrix---the expectation value of
466: the second derivative of the minus log-likelihood function---is often
467: applied. For a combination of the Poisson shot noise and the Gaussian
468: sample variance, the Fisher matrix reads \citep{LimaHu04}
469: %
470: \begin{equation}
471: F_{\alpha \beta}=\bold{\bar m}^{T}_{,\alpha} \bold C^{-1} \bold{\bar m}_{,\beta} + \frac{1}{2} {\rm Tr}[\bold C^{-1} \bold S_{,\alpha}\bold C^{-1} \bold S_{,\beta}]
472: \ ,
473: \label{eq:fisher}
474: \end{equation}
475: %
476: where the comma and subscript ${\alpha}$ indicates the partial
477: derivative with respect to model parameter $\theta_\alpha$. The
478: Fisher matrix approach essentially approximates the likelihood
479: function as a Gaussian distribution near its maximum likelihood point,
480: and the curvature at this point is related to the constraints on the
481: model parameters. The covariance matrix for model parameters is
482: approximated by the inverse of the Fisher matrix. This basic picture
483: will play a key role in \S\ref{sec:fisher}, where we modify the Fisher
484: matrix formalism for assessing the systematic errors.
485:
486: %-----------------------------------------------
487: %-----------------------------------------------
488: %-----------------------------------------------
489: %-----------------------------------------------
490: %-----------------------------------------------
491: %-----------------------------------------------
492:
493:
494: \subsection{Incorporating Assembly Bias into Self-Calibration}\label{sec:assemblybias}
495:
496: We now incorporate assembly bias into the self-calibration formalism.
497: The formalism we outline below is relevant for any secondary parameter
498: which both affects the halo bias and correlates with the cluster mass
499: proxy. We specifically consider the secondary parameter to be the halo
500: concentration $c$ and refer to this dependence throughout as
501: ``assembly bias.'' Note that although the halo concentration and
502: assembly history are generally expected to be tightly correlated
503: \citep{NFW97, Wechsler02}, they may not have exactly the same effect
504: on halo bias \citep[see e.g.][]{GaoWhite07}.
505:
506:
507: Let $b^{\rm ab}(M,c)$ be the halo assembly bias, which now depends on both
508: mass and concentration, and let $f(c|M)$ be the distribution of
509: concentrations for halos of mass $M$. In this case, the
510: observable--mass distribution $P(\Mobs|M)$ needs to be generalized to
511: an observable--mass--concentration distribution $P(\Mobs|M,c)$. With
512: the secondary parameter $c$, the mean abundance $\bar m$ takes the
513: form
514: %
515: \begin{equation}
516: \mbar = V \int dM\ \nM \int dc\ f(c|M) \avg{\phi|M,c} \ ,
517: \end{equation}
518: %
519: where
520: %
521: \begin{equation}
522: \avg{\phi|M,c} = \int d\Mobs\ P(\Mobs|M,c) \phi(\Mobs) \ .
523: \end{equation}
524: %
525: This mean abundance remains the same as equation~\ref{eq:mean} since
526: the concentration dependence only affects the halo bias but not the
527: mass function. We thus require
528: %
529: \begin{equation}
530: \int dc\ f(c|M) \avg{\phi|M,c} = \avg{\phi|M} \ .
531: \label{eq:marginalization}
532: \end{equation}
533: %
534: On the other hand, the bias integrated over the observable range is
535: affected, and the analog of Equation \ref{eq:cbias} is
536: %
537: \begin{equation}
538: \bar b^{\rm ab} = \frac{V}{\mbar} \int dM\ \nM \int dc\ b^{\rm ab}(M,c) f(c|M) \avg{\phi|M,c} \ .
539: \end{equation}
540: %
541: The corresponding sample variance in this case reads
542: \begin{equation}
543: S_{ij}^{\rm ab} = \mbar^2 (\bar b^{\rm ab})^2 \sigma_{ij}^2 \ ,
544: \end{equation}
545: %
546: and we analogously define $\C^{\rm ab} = {\bold M} + {\bold S}^{\rm ab}$.
547: Replacing the corresponding matrices in Equation~\ref{eq:fisher}, we
548: obtain the Fisher matrix incorporating assembly bias.
549:
550:
551: The difference between $P(\Mobs|M,c)$ and $P(\Mobs|M)$ depends on how
552: $\Mobs$ correlates with $c$. We leave these details to
553: \S\ref{sec:correlation} and simply state here that our parametrization
554: depends on the cross-correlation coefficient $r$ relating $\Mobs$ and
555: $c$ at fixed halo mass. When $r=0$, assembly bias has no impact on
556: self-calibration; when $r=\pm 1$, the impact of assembly bias is
557: maximized. Figure~\ref{fig:EllipseAB} demonstrates the formalism
558: described above (with an SPT survey assumption and a WMAP3 cosmology,
559: see \S\ref{sec:implementation}) and shows how the correlation between
560: $\Mobs$ and $c$ changes the constraints on dark energy parameters,
561: assuming that we have thorough knowledge of assembly bias and that
562: $r=\pm1$ (\emph{dotted and dashed curves}). As can be seen, correlation
563: between $\Mobs$ and $c$ actually improves the dark energy constraints
564: if $r$ is known a priori. This improvement is presumably due the
565: dependence of bias on $M_*$, which is also sensitive to dark energy,
566: although we also note that the assumption of self-similarity in
567: $M/M_*$ needs to be assessed in the dark energy-dominated regime. In
568: addition, with the knowledge of $r$, the scatter in $\Mobs$ actually
569: contains the information of halo concentration, which may also improve
570: cosmological constraints. These extreme cases are mainly for
571: demonstration, since we are unlikely to have sufficient astrophysical
572: knowledge to specify both the assembly bias and this correlation.
573: However, if individual concentrations can be measured for the most
574: massive clusters (where the impact of assembly bias is most severe),
575: they could provide observational evidence of assembly bias and
576: increase the efficacy of self-calibration.
577:
578:
579: In the following sections, we explore the question: if one were to
580: perform the self-calibration analysis ignoring the effects of assembly
581: bias (effectively, assuming $r=0$), how would the estimated
582: cosmological parameters be biased? As we shall see, the answer
583: sensitively depends on $r$ and on the scatter in the observable--mass
584: distribution. We next include $r$ as a free parameter in the Fisher
585: matrix analysis and consider the effect of marginalization over $r$.
586: However, a caveat for applying the Fisher matrix here is that since
587: $r$ is bound to the range $[-1,1]$, the likelihood function for $r$
588: may not be well-approximated as Gaussian if $r$ is close to $\pm 1$.
589: Because the Fisher matrix is based on this Gaussian approximation, it
590: may not apply to the case when $r$ approaches $\pm1$. On the other
591: hand, our fiducial choices of this parameter, which are in the range
592: $|r| \leq 0.5$, may circumvent this problem.
593:
594:
595: \begin{figure}[t!]
596: %\plotone{selfcalibration_ab.eps}
597: \plotone{f2.eps}
598: \caption{
599: Improvement of dark energy constraints assuming a thorough modeling of
600: assembly bias and knowledge of the cross-correlation relating $\Mobs$
601: ---the cluster's mass estimate based on a cluster observable---and
602: $c$, the halo's concentration parameter. All error ellipses include
603: the $68\%$ confidence regions in the $\OmegaDE$--$w$ plane. The solid
604: ellipse shows the fiducial model of zero observable--concentration
605: correlation ($r=0$), in which case assembly bias has no effect. The
606: dotted/dashed ellipse corresponds to an observable which is perfectly
607: correlated/anti-correlated with concentration ($r=1$/$-1$). If
608: assembly bias is correctly modeled, the sensitivity of assembly bias
609: to $M_*$ slightly improves dark energy constraints.
610: }
611: \label{fig:EllipseAB}
612: \end{figure}
613:
614:
615: %-----------------------------------------------
616: %-----------------------------------------------
617: %-----------------------------------------------
618: %-----------------------------------------------
619: %-----------------------------------------------
620: %-----------------------------------------------
621:
622:
623: \subsection{Biased Parameter Estimation from Ignored Systematics: A Modified Fisher Matrix Formalism}\label{sec:fisher}
624:
625: In \S\ref{sec:bias}, we described how ignoring the impact of assembly
626: bias can potentially lead to biased cosmological parameter estimates.
627: In this section, we modify Fisher matrix formalism to quantitatively
628: assess the significance of this systematic. We focus on how the
629: parameter estimates are biased due to a wrong model assumption, and
630: how significant this systematic error is when compared with
631: statistical uncertainties. This formalism is motivated by the
632: standard Fisher matrix formalism as presented in \cite{Tegmark97KL}.
633:
634:
635: We generally consider two models, denoted by model $A$ and model $B$,
636: each of which describes a data set $\vec x$ based on a parameter
637: $\theta$. Here $\theta$ can be generalized to a vector denoting a set
638: of parameters ($\theta_i$ values). We assume that the observed data set
639: $\vec x$ is well described by model $B$ but is mistakenly analyzed
640: according to model $A$. If $\theta_t$ denotes the true parameter in
641: model $B$ that corresponds to the observed data set $\vec x$, we are
642: interested in how the estimated parameter $\hat \theta$ recovered
643: based on model $A$ differs from $\theta_t$. Our quantitative analysis
644: can be summarized as follows:
645: %
646: \begin{enumerate}
647: %
648: \item Our starting point is the likelihood function $L_A(\vec
649: x|\theta)$ for model $A$. The data set $\vec x$ is assumed to be
650: drawn from the probability distribution $P_B(\vec x|\theta_t)$ for
651: model $B$; in order to relate $\theta$ to $\theta_t$, we take average
652: over $\vec x$ to compute $\avg{\ln L_A(\theta)|\theta_t}$.
653: %
654: \item We take the point $\hat\theta$ which maximizes $\avg{\ln
655: L_A(\theta)|\theta_t}$ as our estimator for the recovered cosmology.
656: This step defines the function $\hat\theta(\theta_t)$, the recovered
657: model parameter varying with the input parameter $\theta_t$. We are
658: particularly interested in $\delta\theta =
659: \hat\theta(\theta_t)-\theta_t$, which is the systematic error in
660: parameter inference due to assuming an incorrect model.\footnote{An
661: alternative approach is to first use $L_A(\vec x|\theta)$ to compute
662: the maximum likelihood estimator $\hat\theta(\vec x)$. Since
663: $\hat\theta$ is now a function of the data $\vec x$, one could use
664: $P_B(\vec x|\theta_t)$ to compute the expectation value
665: $\avg{\hat\theta|\theta_t}$. However, this approach is not
666: analytically tractable. }
667: %
668: \item In order to assess the significance of the systematic error
669: $\delta\theta$, we compare it against the statistical uncertainty in
670: $\theta$. We calculate the modified Fisher matrix $ \bold{\tilde
671: F}_{ij}(\theta_t) =\avg{\partial^2(-\ln L_A)/\partial\theta_i\partial
672: \theta_j| \theta_t}$ and obtain the corresponding error bar
673: $\sigma_{\theta_i}^2= {(\bold{\tilde F}^{-1})_{ii}}$. The systematic
674: error is significant if $\delta\theta_{i} \gtrsim \sigma_{\theta_i}$.
675: %
676: \end{enumerate}
677: %
678: A detailed derivation when both $P_A(\vec x|\theta)$ and $P_B(\vec
679: x|\theta)$ are Gaussian can be found in the Appendix.
680:
681:
682: In this study, model $A$ represents the standard self-calibration
683: analysis that ignores assembly bias, while model $B$ is
684: self-calibration analysis that includes the impact of assembly bias.
685: The data set $\vec x$ is the number counts in each of the cells under
686: consideration. The systematic errors of the recovered parameters are
687: given by
688: \begin{equation}
689: \delta\theta_j =\sum_i (\bold{F}^{-1})_{ij} {\rm Tr}\{\frac{1}{2}{\C^{-1}\C_{,i}\C^{-1}(\C^{\rm ab}-\C)}\} \ ,
690: \end{equation}
691: where $\C^{\rm ab}$ is the covariance matrices with assembly bias, and
692: $\C$ and ${\bold F}$ are the same as those in
693: equation~\ref{eq:fisher}. The modified Fisher matrix reads
694: %
695: \begin{equation}
696: {\tilde F}_{ij}={\bold{\bar m}}^{T}_{,i} \bold C^{-1} {\bold{\bar m}}_{,j} + \frac{1}{2} {\rm Tr}[\bold C^{-1} \bold S_{,i}\bold C^{-1} \bold S_{,j}\C^{-1}\C^{\rm ab}] \ ,
697: \end{equation}
698: %
699: in which the modification comes from the change of covariance matrix
700: due to assembly bias (see the Appendix).
701: %
702: We note that similar formalisms arising from different approaches can
703: be found in e.g.\@ \citet{Knox98}, \citet{Huterer01},
704: \citet{HutererLinder07}, and \citet{Amara07}.
705:
706:
707:
708:
709: \begin{figure*}[t!]
710: %\plotone{deviations.eps}
711: \plotone{f3.eps}
712: \caption{
713: Systematic errors due to ignoring existent assembly bias. Here we
714: assume two sets of scatter and correlation values and perform the
715: analysis discussed in \S\ref{sec:fisher}, with an SPT survey
716: assumption and a WMAP3 cosmology (see \S\ref{sec:implementation}).
717: The open circles and dashed ellipses show the true parameter values
718: and the $68\%$ confidence regions with assembly bias correctly
719: included. The solid circles and the solid ellipses show the estimated
720: values and $68\%$ confidence regions if assembly bias is completely
721: ignored. The left panel shows that for a moderate assumption of
722: $\sigma_{\rm\ln M }=0.1$ and $r=-0.5$, the systematic errors are
723: $0.22$ and $0.23\sigma$ for $\OmegaDE$ and $w$, respectively; in
724: this case the effects of assembly bias are ignorable. On the other
725: hand, the right panel shows that for an extreme assumption of
726: $\sigma_{\rm\ln M }=0.25$ and $r=-1$, the systematic errors are
727: $1.14$ and $1.2\sigma$ for $\OmegaDE$ and $w$, respectively; in
728: this case the effects of assembly bias are significant.
729: }
730: \label{fig:estimator}
731: \end{figure*}
732:
733:
734: Figure \ref{fig:estimator} illustrates the results of our formalism as
735: applied to the self-calibration analysis for an SPT-like survey in the
736: specified WMAP3 cosmology (see \S\ref{sec:implementation} for details
737: of implementation and assumptions). In each panel, the open circles
738: indicate the assumed true values, while the filled circles show the
739: recovered parameters from a self-calibration analysis that ignores
740: assembly bias. The ellipses include the $68\%$ confidence regions in
741: the $\OmegaDE$--$w$ plane; the dashed ellipses correspond to
742: correctly-modeled assembly bias (assuming that we know the correlation
743: coefficient $r$ a priori; $r$ will be mathematically defined in
744: \S\ref{sec:correlation}), while the solid ellipses correspond to the
745: ignored assembly bias. Note that the shape of the confidence regions
746: can also be changed by this systematic. The left panel shows the
747: assumption of a small $\Mobs$--$M$ scatter and low $\Mobs$--$c$
748: correlation ($\sigma_{\rm\ln M }=0.1$ and $r=-0.5$), and the systematic
749: errors are $0.22$ and $0.23\sigma$ for $\OmegaDE$ and $w$,
750: respectively; the deviations of the parameter estimates are much less
751: than the statistical uncertainties. The right panel shows the
752: assumption of a larger scatter and perfectly anti-correlated $\Mobs$
753: and $c$ ($\sigma_{\rm\ln M }=0.25$ and $r=-1$), and the resulting
754: systematic errors are $1.14$ and $1.2\sigma$ for $\OmegaDE$ and
755: $w$, respectively; these deviations are significant and cannot be
756: ignored. We thus expect the impact of assembly bias will be stronger
757: if the observable--mass relation has a large scatter and if $\Mobs$ is
758: strongly correlated with $c$. The exact dependence of systematic
759: error on these two quantities will be fully explored in \S
760: \ref{sec:results}.
761:
762:
763: %--------------------------------------------------------------------------------------------------
764: %--------------------------------------------------------------------------------------------------
765: %--------------------------------------------------------------------------------------------------
766: %--------------------------------------------------------------------------------------------------
767: %--------------------------------------------------------------------------------------------------
768: %--------------------------------------------------------------------------------------------------
769: %--------------------------------------------------------------------------------------------------
770: %--------------------------------------------------------------------------------------------------
771:
772:
773: \section{Implementation}\label{sec:implementation}
774:
775: \subsection{Parameterizing the Observable-Concentration Correlation}\label{sec:correlation}
776:
777: In the absence of assembly bias, we follow \citet{LimaHu05} to
778: parameterize the observable--mass relation $P(\ln\Mobs|M)$. Given halo
779: mass $M$, the corresponding log observables $\ln\Mobs$ are modeled as
780: a Gaussian distribution with mean $\ln M + \ln\Mbias$---where $\Mbias$
781: specifies the offset between the estimate mass and the true mass---and
782: variance $\sigmalnM^2$. This parameterization serves as the standard
783: case as we generalize $P(\ln\Mobs|M)$ to $P(\ln\Mobs|M,c)$ for
784: analyzing the effect of assembly bias.
785:
786: A priori, we do not know exactly how the estimated mass of a cluster
787: $\Mobs$ will depend on the cluster's concentration $c$, that is, the
788: correct parameterization for $P(\ln\Mobs|M,c)$. In detail, this
789: relation may depend on both physical and observational
790: effects. However, we would like to demand a simple wish-list of
791: properties of our parameterization:
792: %
793: \begin{enumerate}
794: \item When marginalized over concentration, $P(\ln\Mobs|M,c)$ should
795: reduce to the Gaussian distribution $P(\ln\Mobs|M)$ of the fiducial
796: case (as required by eq.~[\ref{eq:marginalization}]), independent
797: of any new parameters introduced (i.e.\@ we should keep the total
798: $\ln\Mobs$--$ \ln M$ scatter fixed).
799: \item In order to study how self-calibration is affected as the
800: dependence of $\Mobs$ on $c$ is ``turned on,'' the parameterization
801: should have a tunable parameter. When this tunable parameter is set to
802: zero, our analysis should reduce to the standard case.
803: \end{enumerate}
804: %
805: In the interest of simplicity, we take $P(\ln\Mobs|M,c)$ to be
806: Gaussian in $\ln\Mobs$, and assume that the halo concentration
807: slightly shifts $\ln\Mobs$ relative to $\ln M$, so that the mean and
808: the variance of $\ln\Mobs$ are given by
809: %
810: \begin{eqnarray}
811: \avg{\ln \Mobs|M,\tc} &=& \ln M + \ln M_{\rm bias} + r\sigmalnM\tc \\
812: \Var(\ln \Mobs|M,c) &=& \sigmalnM^2(1-r^2) \ .
813: \end{eqnarray}
814: %
815: In the above expressions, $r$ is the correlation coefficient between
816: $\ln\Mobs$ and $\tc$ at fixed $\ln M$, $\sigmalnM$ is the scatter in
817: $\ln\Mobs$ at fixed $M$, and $\tc$ is defined via
818: %
819: \begin{equation}
820: \tc = \frac{\ln c-\avg{\ln c|M}}{\sqrt{\Var(\ln c|M)}} \ .
821: \end{equation}
822: %
823: Note that when $r=0$, all of the observed scatter in $\ln\Mobs$ at
824: fixed $\ln M$ is intrinsic, and our model reduces to the standard
825: case. Conversely, for $r=1$, the scatter in $\ln\Mobs$ at fixed $\ln
826: M$ is entirely due to the scatter in halo concentration at fixed mass.
827: As a consistency check, we find that if we marginalize $P(\ln
828: \Mobs|M,c)$ over concentration (assuming a log-normal distribution for
829: $c$ at fixed mass, see e.g.\@ \citealt{Jing00}, \citealt{Bullock01},
830: and \citealt{Neto07}), the resulting distribution $P(\ln\Mobs|M)$ is
831: exactly that of the standard case; that is, our parameterization
832: preserves the total scatter in $\ln\Mobs$ at a given $\ln M$.
833:
834:
835: %--------------------------------------------------------------------------------------------------
836: %--------------------------------------------------------------------------------------------------
837: %--------------------------------------------------------------------------------------------------
838: %--------------------------------------------------------------------------------------------------
839: %--------------------------------------------------------------------------------------------------
840: %--------------------------------------------------------------------------------------------------
841:
842:
843: \subsection{Survey Assumptions, Cosmological Models, and Nuisance Parameters}
844:
845: % \begin{turnpage}
846: \begin{deluxetable}{ccccccc}
847: \tabletypesize{\scriptsize}
848: % \rotate
849: \tablecaption{Survey Assumptions\label{tab:surveys}}
850: \tablewidth{0pt}
851: \tablehead{
852: \colhead{Survey} &
853: \colhead{$M_{\rm th}$} &
854: \colhead{Bin Size} &
855: \colhead{$N_{\rm bins}$} &
856: \colhead{Area} &
857: \colhead{$z_{\rm max}$}\\
858: \colhead{} &
859: \colhead{$(\hiMsun)$} &
860: \colhead{$(\Delta {\rm log}_{10}\Mobs)$} &
861: \colhead{} &
862: \colhead{$(\rm{deg^2})$}
863: }
864: \startdata
865: SDSS (optical) & $10^{13.5}$ & 0.5 & 3 & 7500 & 0.3 \\
866: DES (optical) & $10^{13.5}$ & 0.5 & 3 & 5000 & 1 \\
867: SPT (SZ) & $10^{14.2}$ & 1 & 1 & 4000 & 2 \\
868: LSST (optical) & $10^{13.5}$ & 0.5 & 3 & 20000 & 2 \\
869: \enddata
870: %%%%%%%
871: %\tablenotetext{a}{}
872: \tablecomments{All surveys use cells of area $10\deg^2$ and $\Delta z = 0.1$}
873: \end{deluxetable}
874: % \end{turnpage}
875:
876:
877: With the Fisher matrix analysis, we statistically forecast the
878: systematic effects for four galaxy cluster surveys: the Sloan Digital
879: Sky Survey (SDSS, \citealt{SDSS}; assuming the volume using
880: photometric data), the Dark Energy Survey
881: (DES\footnote{See http://www.darkenergysurvey.org/}), the South Pole
882: Telescope (SPT\footnote{See http://pole.uchicago.edu/}), and the Large
883: Synoptic Survey Telescope (LSST\footnote{See http://www.lsst.org/}). The
884: survey areas are assumed to be $7500\ \deg^2$ for SDSS, $5000\ \deg^2$
885: for DES, $4000\ \deg^2$ for SPT, and $20000\ \deg^2$ for LSST, with
886: survey depths of $z_{\rm max}=0.3, 1.0, 2.0$ and $2.0$ respectively. The
887: cells used for the counts-in-cells analysis are assumed to have an
888: area $10\ \deg^2$ and redshift interval $\Delta z =0.1$. We assume
889: clusters with $\Mobs \geq10^{14.2}\ \hiMsun$ are observed by SPT, and
890: perform no mass binning. For SDSS, DES, and LSST, the observational
891: threshold is assumed to be $\Mobs\geq 10^{13.5}\ \hiMsun$, and the
892: counts in each of these surveys are binned in three observable bins.
893: The survey parameters for all four surveys are detailed in
894: Table~\ref{tab:surveys}.
895:
896: While the mass threshold of SZ observations has little redshift
897: dependence \citep[e.g.][]{Carlstrom02}, the mass threshold of optical
898: surveys has more uncertainties. Clusters with mass $10^{13.5}\
899: \hiMsun$ have been shown to be detectable, with high purity and
900: completeness, with more than 10 bright red galaxies ($\sim 0.4 L_*$)
901: in the SDSS photometric survey out to $z \sim 0.3$ \citep{Koester07,
902: Johnston07}. We note that our choice of the minimum mass for the
903: optical surveys assumes that such clusters can still be detected with
904: high purity and completeness out to the maximum redshift $z_{\rm max}$.
905: This assumption may be reasonable out to $z=1$, where clusters have
906: been shown to have a robust red sequence, but the efficacy of this
907: method will eventually break down at higher redshifts. In any case,
908: it will need to be tested in detail with both realistic simulations
909: and the data itself. We note that for LSST, one may wish to detect
910: clusters using peaks in the lensing shear instead of from assumptions
911: about the galaxy distribution \citep[e.g.][]{Kaiser95,Hennawi05}, in
912: which case self-calibration could serve as a consistency check for the
913: predictions for the observed shear signal made directly from
914: simulations. In \S\ref{sec:results}, we consider one example case for
915: LSST, which has similar assumptions to the lower $z_{\rm max}$ optical
916: surveys, for reference.
917:
918:
919: % \begin{turnpage}
920: \begin{deluxetable*}{ccccccc}
921: \tabletypesize{\scriptsize}
922: % \rotate
923: \tablecaption{Fiducial Cosmologies\label{tab:cosmology}}
924: \tablewidth{0pt}
925: \tablehead{
926: \colhead{Cosmology} &
927: \colhead{$\OmegaDE$} &
928: \colhead{$w$} &
929: \colhead{$\delta_\zeta(k=0.05{\rm Mpc^{-1}})$} &
930: \colhead{$n$} &
931: \colhead{$\Omega_bh^2$} &
932: \colhead{$\Omega_mh^2$}
933: }
934: \startdata
935: WMAP1 & 0.73 & -1 & $5.07\times10^{-5}$ & 1 & 0.024 & 0.14\\
936: WMAP3 & 0.76 & -1 & $4.53\times10^{-5}$ & 0.958 & 0.0223 & 0.128\\
937: \enddata
938: %\tablenotetext{}{}
939: %%%%%%%
940: %\tablenotetext{a}{}
941: \tablecomments{All of our forecasts assume Plank-like priors: $\sigma(\ln \zeta)=\sigma(\ln \Omega_mh^2) = \sigma(\ln \Omega_bh^2) = \sigma(n) =0.01$, except for $\OmegaDE$ and $w$.}
942: \end{deluxetable*}
943: % \end{turnpage}
944:
945:
946: In this work, we consider two sets of cosmological parameters, namely
947: the best fit cosmologies to WMAP1 \citep{WMAP1} and WMAP3
948: \citep{WMAP3}, whose parameter values are listed in
949: Table~\ref{tab:cosmology}. Both of them are flat $\Lambda$CDM
950: cosmologies but differ mainly in the relative contribution of dark
951: energy to the global energy density, in the normalization of
952: fluctuations ($\delta_\zeta$ or $\sigma_8$), and in the spectral index
953: ($n$). The impact of these differences on our analysis will be
954: presented in \S\ref{sec:results}. In our statistical forecast, we do
955: not put any priors on dark energy parameters, but we assume
956: Planck-like priors on the rest of the cosmological parameters (see
957: Table~\ref{tab:cosmology}). Finally, in our forecast models we use
958: the halo mass function by \citet{Jenkins01}, the bias function by
959: \citet{Sheth01}, and the assembly bias $b^{\rm ab}(M,c)$ found by
960: \citet[][this assumption was shown in Fig.\@ 1]{Wechsler06}.
961:
962:
963: With regard to the observable--mass relation, our model involves three
964: nuisance parameters: the bias in the estimated mass ($\ln M_{\rm bias}$),
965: the scatter of $\ln\Mobs$ given $\ln M$ ($\sigma_{\rm \ln M}$), and the
966: cross-correlation coefficient between $\ln\Mobs$ and the normalized
967: halo concentration $\tc$ ($r$). Throughout, we take $\ln M_{bias}=0$
968: as our fiducial model. Our choice for the fiducial values for the
969: scatter and the cross-correlation coefficient in each of the surveys
970: requires further discussion.
971:
972:
973: Let us first focus on the scatter. For an SPT-like survey, the
974: observational mass proxy is the SZ decrement of the cosmic microwave
975: background due to the hot, ionized gas permeating the inter-cluster
976: medium. At present, this scatter has only been predicted from
977: numerical simulations but has not been determined from observations.
978: \cite{WhiteM02} argued that the three main sources of scatter are the
979: evolution of the $M$--$T$ relation, asphericity in the matter
980: distribution, and line-of-sight projection. \cite{Motl05} and
981: \cite{Nagai06} showed that the scatter is 10\%-15\%, and the scaling
982: relation is insensitive to the detailed physical processes involved in
983: galaxy formation, with a good agreement with self-similar models.
984: However, \cite{Shaw07} showed that at least 20\% intrinsic scatter
985: exists due to the internal properties of galaxy clusters. They also
986: demonstrated that this scatter could be reduced by choosing different
987: aperture radius for defining $M$ and $Y$, or by removing cluster samples
988: with many substructures. Moreover, it may be possible to reduce the
989: scatter even further using cluster structural properties. For
990: example, \cite{Afshordi07} proposed a ``fundamental plane'' among the
991: cluster mass, the total SZ flux, and the SZ half-light radius
992: $R_{\rm SZ,2}$; in simulations, this relation reduced the scatter in mass
993: estimates to $\sim$ 14\%. Further, \cite{Haugboelle07} found that by
994: constructing an empirical model for the SZ profile, which includes a
995: scaling parameter $r_0$, they could reduce the scatter down to 4\%.
996: In this work, we take the largest of these range of values, namely
997: 20\%, as our fiducial scatter for SPT. If SPT is insensitive to halo
998: assembly bias for this largest possible scatter, then it will also be
999: insensitive for smaller values of scatter.
1000:
1001:
1002: In optical surveys, the usual observational mass proxy is the optical
1003: richness, namely the galaxy number in a galaxy cluster. Other choices
1004: are also possible, including the total optical luminosity or
1005: combinations of parameters \citep[e.g.][]{Rykoff07,Becker07,Reyes08}.
1006: Determining a reasonable choice for the scatter for a DES-like survey
1007: is somewhat less straightforward, as predictions from simulations are
1008: less robust and the scatter can highly depend on both the richness
1009: measure and the cluster finder. \cite{Gladders07} applied a
1010: self-calibration analysis to a catalog from the Red-Sequence Cluster
1011: Surveys (RCS), finding the fractional scatter $f_{\rm sc}$ to be $0.69 \pm
1012: 0.20$ and $0.71^{+0.19}_{-0.17}$ (based on different priors). Using
1013: the velocity information of galaxies in maxBCG clusters,
1014: \cite{Becker07} estimated that the optical richness had a mass
1015: dependent scatter which varied from about $0.75$ for massive clusters
1016: to $1.2$ for group scale objects. Cross-correlation with the X-ray
1017: data on these same clusters suggests a considerably smaller scatter of
1018: about $0.5$ (E.\@ Rozo et al.\@ in preparation). Here, we choose $0.5$ as
1019: our fiducial scatter value for two reasons: First, as we shall see,
1020: even with this amount of scatter, halo assembly bias has a significant
1021: impact on self-calibration studies; this small scatter thus provides a
1022: baseline value for the impact of assembly bias. Second, we note that
1023: current optical richness estimates have all used fairly crude measures
1024: of richness. We think it is highly probable that in the near future
1025: we will start seeing richness measures that are considerably more
1026: strongly correlated with mass than those used at present. Thus, we
1027: have opted to select a scatter value that is closer to the lowest
1028: scatter estimated in current samples.
1029:
1030:
1031: We now turn to the correlation relating observable and concentration,
1032: $r$. Currently, there are no observational constraints on $r$ for
1033: either optical or SZ mass proxies. We also find no quoted values for
1034: the correlation between SZ and halo concentration in the literature,
1035: although we note that \citet[][]{Reid06} and \citet[][]{Shaw07}
1036: investigated the impact of halo concentration on the scatter in
1037: $Y_{\rm SZ}$ and found that a considerable fraction of the scatter in SZ
1038: is due to variations in halo concentration at fixed mass. In this
1039: work, we choose the fiducial value to be $r=0.4$, which is the value
1040: observed in simulations of clusters using the hydrodynamical ART code
1041: (D.\@ Rudd 2008, private communication).
1042:
1043:
1044: The $\Mobs$--$c$ correlation $r$ for optical clusters is somewhat
1045: better understood. We are not aware of current observational
1046: constraints on the correlation between optical cluster richness and
1047: halo concentrations, although with a sufficiently large sample, this
1048: value could in principle be measured from lensing data.
1049: \citet[][]{Zentner05} and \citet[][]{Wechsler06} have shown that the
1050: amount of substructures in a cluster-size halo is negatively
1051: correlated with halo concentration. On the other hand, selection
1052: effects could modify this correlation. For instance, high
1053: concentration halos might, on average, be assigned higher richness
1054: than low concentration halos of the same mass due to the larger galaxy
1055: density near the cluster core. In this work, we choose $r=-0.5$ as
1056: our fiducial value, which is roughly consistent with the numerical
1057: results of \citet[][]{Zentner05} and \citet[][]{Wechsler06}. We note
1058: that the results presented here assume that all of these parameters
1059: are constant with both redshift and mass.
1060:
1061:
1062: %--------------------------------------------------------------------------------------------------
1063: %--------------------------------------------------------------------------------------------------
1064: %--------------------------------------------------------------------------------------------------
1065: %--------------------------------------------------------------------------------------------------
1066: %--------------------------------------------------------------------------------------------------
1067: %--------------------------------------------------------------------------------------------------
1068:
1069:
1070: \begin{figure*}[t!]
1071: %\plotone{Contours.eps}
1072: \plotone{f4.eps}
1073: %\centerline{\epsfxsize=4.7truein\epsffile{f4.eps}}
1074: \caption{
1075: Systematic errors for $\OmegaDE$ (\emph left) and $w$ (\emph right)
1076: estimators, as a function of scatter in the observable given mass
1077: ($\sigmalnM$) and the correlation between observable and halo
1078: concentration ($r$). The ratios of the systematic error and the
1079: statistical uncertainty ($|\delta\theta|/\sigma_{\rm\theta}$) are shown for
1080: three of our main survey assumptions: SDSS, DES, and SPT, from top to
1081: bottom. High scatter value and strong correlation/anti-correlation
1082: correspond to high deviation of estimators. We also mark the fiducial
1083: values of $\sigmalnM$ and $r$ in each panel according to our current
1084: knowledge from observations and numerical simulations. We note that
1085: these plots are also applicable to other surveys (e.g. X-ray) for
1086: which the survey volume and mass threshold are the same as those
1087: assumed here. See \S\ref{sec:results} for discussion.
1088: }
1089: \label{fig:Contours}
1090: \end{figure*}
1091:
1092:
1093: \section{Results and Discussion}\label{sec:results}
1094:
1095: We now present the effect of assembly bias for a set of specific
1096: assumptions about galaxy cluster surveys. We focus on the systematic
1097: errors in the two dark energy parameters $\OmegaDE$ and $w$ when
1098: compared with the statistical errors expected from each survey,
1099: $|\delta \OmegaDE|/{\sigma_{\Omega_{\rm DE}}}$ and $|\delta
1100: w|/\sigma_w$. Figure~\ref{fig:Contours} shows how these two ratios
1101: vary with the scatter in $\ln\Mobs$--$\ln M$ ($\sigma_{\rm \ln M}$) and
1102: the cross-correlations coefficient of $\ln\Mobs$--$\tc$ ($r$) for our
1103: fiducial SDSS, DES, and SPT surveys. All plots assume a WMAP3
1104: cosmology. As can be seen, a high degree of correlation and/or large
1105: scatter can result in significantly biased cosmological estimates for
1106: both DES and SPT, while for the current SDSS the statistical
1107: uncertainty is sufficiently large that halo assembly bias is
1108: insignificant.
1109:
1110:
1111: How halo assembly bias differently affects DES and SPT is worth
1112: discussing. For a fixed scatter and correlation coefficient, the
1113: cosmological constraints from DES are considerably less biased than
1114: those of SPT. The reason for this difference is two-fold. First, DES
1115: clusters probe a lower mass scale than SPT clusters do. This
1116: difference is important because the effect of concentration on halo
1117: bias is important for high mass halos, but non-existent for halos of
1118: mass near $M_*\sim 10^{13} \hiMsun$ \citep{Wechsler06}. Consequently,
1119: the cosmological constraints coming from low mass clusters (groups)
1120: should be unbiased.\footnote{Although the low mass clusters (groups)
1121: are less affected by assembly bias, they are subjected to more
1122: statistical errors. For the most constraining power, the choice of
1123: mass threshold should be made based on the scatter, the completeness,
1124: and the purity.} The second important difference between SPT and DES
1125: is that in our fiducial surveys we have assumed binned counts for DES
1126: clusters but only thresholded counts for SPT clusters. Consequently,
1127: all the cosmological information provided by the shape of the halo
1128: mass function (which is unaffected by assembly bias) does not
1129: contribute to the SPT constraints. Thus, SPT constraints are
1130: considerably more sensitive to the effects of assembly bias than the
1131: DES constraints given the same scatter and correlation coefficient.
1132:
1133:
1134: That is not, however, the end of the story. In order to fairly
1135: compare SPT to DES, one also needs to consider the regions of
1136: parameter space relevant to each of these surveys. We noted earlier
1137: that numerical simulations predict that the intrinsic scatter in the
1138: SZ signal is approximately $20\%$ or even less
1139: \citep[e.g.][]{Motl05,Nagai06,Shaw07,Haugboelle07}. As can be seen
1140: from the bottom panels of Figure~\ref{fig:Contours}, these scatter
1141: values do not result in significant biasing of the recovered
1142: cosmological parameters for any value of $r$. Although the expected
1143: intrinsic scatter is small, the projection effect may raise or even
1144: dominate the total scatter \citep[see e.g.][]{WhiteM02,Hallman07}.
1145: \citet{Holder07} found that the SZ background can generate errors
1146: larger than $20\%$ in recovered flux if $\sigma_8$ is near 0.7. If
1147: the extra scatter is due to the randomly aligned structures along the
1148: line of sight, then we do not expect assembly bias to have a significant
1149: impact. On the other hand, if the projection effect is dominated by
1150: nearby structures, the extra scatter due to projection will be
1151: strongly correlated with the environment, resulting in higher
1152: correlation between concentration and observable. In fact, if the
1153: scatter due to projection is as high as \citet{Holder07} predicted and
1154: is also dominated by nearby correlated structure, the effect of
1155: assembly bias may be strengthened.
1156:
1157:
1158: Photometric surveys like DES, in contrast, are very likely to be
1159: sensitive to the impact of assembly bias. In this case, we know that
1160: the optical richness--mass relation has a scatter $\gtrsim 50\%$
1161: (e.g. \citealt{Gladders07, Becker07}; E.\@Rozo et al.\@ in preparation).
1162: As can be seen in the middle panels of Figure~\ref{fig:Contours}, even
1163: moderate correlations between $\Mobs$ and $c$, e.g.\@ $|r|\gtrsim
1164: 0.5$, can result in significant biasing of the recovered cosmological
1165: parameters. It is likely, therefore, that cosmological analysis of
1166: the DES optical cluster sample will need to include halo assembly bias
1167: in order to avoid systematic errors in dark energy inference, unless
1168: the analysis can be done with an observable that is more tightly
1169: correlated with mass. On the other hand, DES will have additional
1170: mass measurements, including its weak lensing and SZ signals from SPT.
1171: If these measurements are included in the analysis, the effect of
1172: assembly bias may be diminished. Observational inference of assembly
1173: bias may even be possible with mass profile measurements.
1174:
1175:
1176: We note that in these figures, the only differences are between the
1177: survey volumes and the threshold in $M_{\rm obs}$, without assuming any
1178: other information specific to the optical or SZ surveys. Therefore,
1179: these results are applicable to other surveys (e.g.\@ X-ray surveys)
1180: with the same survey conditions.
1181:
1182:
1183: In Figure~\ref{fig:surveys} we (1) explore the systematic effects due
1184: to assembly bias under different cosmological parameters (WMAP1 and
1185: WMAP3) and (2) extend the calculation to include an assumption for an
1186: LSST-like optical cluster survey. (For SDSS, DES, LSST: $r=-0.5$ and
1187: $\sigma_{\rm\ln M}=0.5$. For SPT: $r=0.4$ and $\sigma_{\rm\ln M}=0.2$.)
1188: First, the most relevant difference between WMAP1 and WMAP3 is that
1189: WMAP3 has higher $\OmegaDE$ and lower normalization $\delta_\zeta$ or
1190: $\sigma_8$ values, as listed in Table~\ref{tab:cosmology}. As a
1191: result, the WMAP3 cosmology has fewer clusters, and the sample
1192: variance of the clusters is smaller. These differences increase the
1193: statistical errors of the surveys \citep[see also][]{LimaHu07}, thus
1194: making the impact of assembly bias less significant in the WMAP3
1195: cosmology than in the WMAP1 cosmology. Overall, our main conclusions
1196: remain unchanged. Second, for LSST, we find the systematic due to
1197: assembly bias is very significant for our fiducial values; this
1198: systematic is likely to be significant for even small values of
1199: $\sigmalnM$ and $r$.
1200:
1201:
1202: We especially note that the systematic of assembly bias impacts
1203: $\OmegaDE$ and $w$ differently. From the survey point of view,
1204: increasing $z_{\rm max}$ above $1$ largely improves the constraints on
1205: $w$, but barely improves the constraints on $\OmegaDE$. For $w$, the
1206: systematic error due to assembly bias increases monotonically with
1207: $z_{\rm max}$, while for $\OmegaDE$, this systematic error somewhat
1208: cancels and then changes its sign as $z_{\rm max}$ increases. This
1209: difference is due to the fact that $\OmegaDE$ and $w$ affect the
1210: observed large-scale structure differently in different regimes. Dark
1211: energy affects the observed large-scale structure through two
1212: mechanisms: the growth function and the comoving volume. High
1213: $\OmegaDE$ and high $w$ both result in stronger suppression of
1214: structure. On the other hand, the volume dependence works
1215: differently: high $\OmegaDE$ and low $w$ correspond to larger volumes.
1216: The effects work in opposite directions for $\OmegaDE$; higher
1217: $\OmegaDE$ leads to less structure but more volume. Before the onset
1218: of dark energy domination, the comoving volume effect dominates; after
1219: dark energy take-over, the growth function effect dominates. Thus
1220: near the onset of dark energy domination, the observed structure is
1221: insensitive to $\OmegaDE$, leading to no extra information from this
1222: regime. The effects of assembly bias on $\OmegaDE$ before and after
1223: the dark energy domination have opposite signs and thus cancel each
1224: other. On the other hand, for $w$, both effects work in the same
1225: direction; thus, including more survey volume will always increase the
1226: amount of information on $w$, and the systematic effects do not
1227: cancel. That is why the systematic effect of assembly bias on $w$
1228: increases monotonically with $z_{\rm max}$.
1229:
1230:
1231: \begin{figure}[t!]
1232: %\plotone{surveys.eps}
1233: \plotone{f5.eps}
1234: \caption{
1235: Impact of assembly bias for two different cosmologies and four
1236: survey conditions. The ratio $|\delta\theta|/\sigma_\theta$ for
1237: $\OmegaDE$ and $w$ are plotted as circles and squares respectively.
1238: Solid and open symbols are for WMAP1 and WMAP3 cosmologies
1239: respectively. DES and LSST are clearly sensitive to assembly bias,
1240: while SPT is marginally sensitive to it, with the effect being
1241: stronger for WMAP1 than WMAP3. A current SDSS-like survey is not
1242: sensitive to assembly bias. (Fiducial values assumed for other
1243: parameters include: $r=-0.5$ and $\sigma_{\rm \ln M}=0.5$ for SDSS, DES,
1244: and LSST, and $r=0.4$ and $\sigma_{\rm \ln M}=0.2$ for SPT)
1245: }
1246: \label{fig:surveys}
1247: \end{figure}
1248:
1249:
1250: Another interesting question is how the constraints on cosmological
1251: parameters are degraded if we include $r$ as an additional nuisance
1252: parameter that needs to be marginalized over. However, as we
1253: mentioned earlier, the fact that the likelihood function is
1254: non-Gaussian in $r$ if $r$ is close to $\pm 1$ implies that the Fisher
1255: matrix estimates may not apply. Therefore, the following constraints
1256: with marginalization over $r$ are only to be taken as rough
1257: indicators.
1258:
1259:
1260: We use moderate values for $r$ ($0.4$ for SZ and $-0.5$ for optical)
1261: to compare the cosmological constraints assuming (1) fixed $r$ values,
1262: and (2) $r$ to be a free parameter in the Fisher matrix.
1263: Table~\ref{tab:constraints} contains our results for three of the
1264: survey assumptions. As can be seen, while the error bars for DES are
1265: only slightly affected by marginalization over $r$, those for SPT
1266: increase by a factor of 2 to 3. The reason is again related to
1267: the mass binning; since our fiducial SPT survey does not include mass
1268: binning, there is no information about the shape of the halo mass
1269: function, which, if present, can improve the constraints on the
1270: scatter in the observable--mass relation. In the absence of this shape
1271: information, the constraints on the scatter is modest, which means
1272: that marginalization of $\OmegaDE$ and $w$ over the acceptable region
1273: of the parameter space will reach areas with very large scatter.
1274: Since those areas are highly sensitive to the effects of halo assembly
1275: bias, the marginalized errors will be significantly larger. In the
1276: last row of Table~\ref{tab:constraints} (SPT5), we assume five narrow
1277: observable bins for SPT with bin size $\Delta {\rm log}_{0}\Mobs =
1278: 0.2$. In this case, the dark energy constraints are barely degraded
1279: after marginalizing over $r$. Thus, mass binning is a crucial
1280: component of the data analysis for both DES and SPT to maximize their
1281: potential as cosmological probes. Note that, in all cases, $r$ itself
1282: cannot be well-constrained like other nuisance parameters. Since the
1283: dependence on $r$ only affects the sample variance but not the
1284: abundance, the information for constraining $r$ is insufficient.
1285:
1286:
1287: % \begin{turnpage}
1288: \begin{deluxetable*}{ccccccccccc}
1289: \tabletypesize{\scriptsize}
1290: % \rotate
1291: \tablecaption{Self-Calibration Constraints. \label{tab:constraints}}
1292: \tablewidth{0pt}
1293: \tablehead{
1294: &
1295: \multicolumn{4}{c}{Self-Calibration with Fixed $r$} & & \multicolumn{5}{c}{Self-Calibration Marginalized over $r$} \\
1296: \cline{2-5} \cline{7-11}\\
1297: \colhead{Survey} &
1298: \colhead{$\OmegaDE$} &
1299: \colhead{$w$} &
1300: \colhead{$\ln\Mbias$} &
1301: \colhead{$\sigmalnM^2$} &
1302: &
1303: \colhead{$\OmegaDE$} &
1304: \colhead{w} &
1305: \colhead{$\ln\Mbias$} &
1306: \colhead{$\sigmalnM^2$} &
1307: \colhead{$r$}
1308: }
1309: \startdata
1310: SDSS & 0.066 & 0.240 & 0.411 & 0.086 & & 0.074 & 0.251 & 0.460 & 0.108 & 0.294 \\
1311: DES & 0.006 & 0.045 & 0.051 & 0.022 & & 0.006 & 0.047 & 0.053 & 0.025 & 0.125 \\
1312: SPT & 0.010 & 0.076 & 0.104 & 0.028 & & 0.025 & 0.177 & 0.355 & 0.149 & 1.300 \\
1313: SPT5 & 0.009 & 0.061 & 0.079 & 0.017 & & 0.010 & 0.062 & 0.087 & 0.027 & 0.357 \\ \enddata
1314: %%%%%%%
1315: %\tablenotetext{}{}
1316: \tablecomments{
1317: Cosmological constraints with fixed cross-correlation coefficient $r$,
1318: and with marginalized $r$. We assume a WMAP3 cosmology, and the
1319: nuisance parameters are the same as those in Fig.\@~\ref{fig:surveys}.
1320: After marginalization over $r$, the constraints from binned cluster
1321: samples (SDSS, DES, and SPT5) are barely degraded, while the
1322: constraints from thresholded samples (SPT) are degrade by a factor of
1323: 2 to 3. This result demonstrates the importance of mass binning. In
1324: all cases, $r$ cannot be well-constrained like other nuisance
1325: parameters since it only affects the sample variance but not the
1326: abundance. We emphasize, however, that the second part of this table
1327: are to be interpreted as rough indicators, since the likelihood
1328: function may not be Gaussian in $r$.
1329: }
1330: \end{deluxetable*}
1331: % \end{turnpage}
1332:
1333:
1334: %--------------------------------------------------------------------------------------------------
1335: %--------------------------------------------------------------------------------------------------
1336: %--------------------------------------------------------------------------------------------------
1337: %--------------------------------------------------------------------------------------------------
1338: %--------------------------------------------------------------------------------------------------
1339: %--------------------------------------------------------------------------------------------------
1340:
1341:
1342: \section{Summary}
1343: \label{sec:conclusions}
1344:
1345: Self-calibration analysis in galaxy cluster surveys relies on the
1346: dependence of the halo bias on mass to simultaneously constrain
1347: cosmology and the cluster observable--mass distribution. Recent work
1348: has shown that halo bias is sensitive not only to halo mass, but also
1349: to secondary parameters related to the assembly history. Here we
1350: consider the effect of halo concentration on the bias as a specific
1351: case of the secondary parameters (generally termed assembly bias), and
1352: show how it might affect self-calibration analyses. In particular, if
1353: halo selection depends on halo concentration, the observed clustering
1354: amplitude of the corresponding cluster sample will deviate from that
1355: of a random selection of clusters with the same mass distribution.
1356: This deviation in the observed clustering amplitude can result in
1357: biased inferences of cosmological parameters, depending on (1) the
1358: amount of scatter between halo mass and the observational mass proxy,
1359: and (2) the correlation between the mass proxy and halo concentration.
1360: For current surveys like SDSS, the statistical uncertainty is still
1361: sufficiently large that the systematic error due to assembly bias is
1362: negligible. On the other hand, for an SPT-like survey, the expected
1363: small amount of intrinsic scatter between the SZ decrement and halo
1364: mass suggests that the impact of assembly bias on parameter estimation
1365: is negligible; however, if the projection effect results in higher
1366: scatter in high density regions, assembly bias may have significant
1367: impact. For a DES-like survey, where the mass proxy is likely to have
1368: considerably larger scatter, we estimate that assembly bias can
1369: displace the recovered dark energy parameters from their true values
1370: by about $1\sigma$. For an LSST-like survey, this systematic error
1371: can exceed $2\sigma$ in $w$. In the last two cases, halo assembly
1372: bias may need to be explicitly included in the cosmological analysis
1373: to avoid biasing of the recovered dark energy parameters. We
1374: emphasize, however, that our analysis has assumed the specific
1375: dependence of halo bias on halo concentration found by
1376: \citet{Wechsler06}. If this dependence is shown to be smaller at high
1377: masses, if the correlation relating the observable mass proxy and halo
1378: concentration can be shown to be small, or if observables that are
1379: more tightly correlated with mass can be found, the effect will be
1380: mitigated. We have shown that binning in mass is crucial for both
1381: optical and SZ surveys, as marginalization over this correlation
1382: coefficient can increase the expected errors of dark energy parameters
1383: by a factor of a few if we only use thresholded counts.
1384:
1385:
1386: \acknowledgments
1387: We thank Marcos Lima, Michael Busha, David Rapetti, Doug Rudd, Gil
1388: Holder, David Weinberg, and Andrew Zentner for useful discussions. We
1389: are grateful to the anonymous referee for helpful comments. The IDL
1390: contour plots are modified from the routine provided at
1391: http://www.davidpace.com. We also thank OSU CCAPP and KIPAC for hospitality
1392: and support during our visits. HW and RHW were supported in part by
1393: the U.S. Department of Energy under contract number DE-AC02-76SF00515.
1394: ER was supported by the Center for Cosmology and Astro-Particle
1395: Physics (CCAPP) at the Ohio State University.
1396:
1397:
1398: %--------------------------------------------------------------------------------------------------
1399: %--------------------------------------------------------------------------------------------------
1400: %--------------------------------------------------------------------------------------------------
1401: %--------------------------------------------------------------------------------------------------
1402: %--------------------------------------------------------------------------------------------------
1403: %--------------------------------------------------------------------------------------------------
1404:
1405:
1406: \appendix
1407: \section{Biased Parameter Estimation from Incorrect Models}
1408: \label{app:fisher-implementation}
1409:
1410: In this section, we explicitly implement the modified Fisher matrix
1411: formalism developed in \S\ref{sec:fisher} for the case in which both
1412: $P_A(\vec x|\theta)$ and $P_B(\vec x|\theta)$ are Gaussian. Let
1413: $\vec\mu(\theta)$ and $\C(\theta)$ be the mean and covariance matrix
1414: defining $P_A(\vec x|\theta)$ in model $A$, which is related to the
1415: likelihood function; let $\vec\mu^B(\theta)$ and $\C^B(\theta)$ be the
1416: corresponding quantities in model $B$, which represent the observed
1417: data. Note that $\vec\mu(\theta)$ and $\C(\theta)$ contain the model
1418: parameter $\theta$ that we are trying to fit, while $\vec\mu^B$ and
1419: $\C^B$ contain the true parameter value $\theta_t$. The
1420: log-likelihood function of model $A$ reads (up to a constant)
1421: %
1422: \begin{equation}
1423: 2\mathcal{L} = -2\ln L(\vec x|\theta)=\ln\det \C+(\vec x - \vec\mu)^T\C^{-1} (\vec x - \vec\mu) \ .
1424: \end{equation}
1425: %
1426: Taking the derivative with respect to $\theta$ and averaging over
1427: $\vec x$, the maximum likelihood estimator $\hat\theta$ can be found
1428: by solving
1429: %
1430: \begin{equation}
1431: \avg{2\mathcal L_{,i}} = {\rm Tr}[\C^{-1}\C_{,i}(1-\C^{-1}\avg{\D})+\C^{-1}\avg{\D_{,i}}]|_{\theta = \hat\theta} =0 \ ,
1432: \end{equation}
1433: %
1434: where $\avg{\D} = \C^B + (\vec\mu^B-\vec\mu)(\vec\mu^B-\vec\mu)^T$ and
1435: $\avg{\D_{,i}} = -2\vec\mu_{,i}(\vec\mu^B-\vec\mu)^T$. We then set
1436: $\hat\theta = \theta_t + \delta\theta$, linearize this equation with
1437: respect to $\delta\theta$, and solve for $\delta\theta$.
1438:
1439:
1440: To proceed further, we focus on two simple examples of interest. The
1441: first example is the effect of assembly bias; model $A$ corresponds
1442: the standard self-calibration, while model $B$ corresponds to
1443: self-calibration with assembly bias. In this case, model $B$ changes
1444: the sample variance but not the mean; thus
1445: $\vec\mu(\theta)=\vec\mu^B(\theta)$ for all $\theta$ values, but
1446: $\C(\theta)\neq \C^B(\theta)$. After linearizing with respect to
1447: $\delta\theta$, the linear equations for $\delta\theta$ read
1448: %
1449: \begin{equation}
1450: {\rm Tr}\{ \C^{-1}\C_{,i}(1-\C^{-1}\C^B + \sum_j\C^{-1}\C_{,j}\C^{-1}\C^B\delta\theta_j
1451: )\}+ 2 \sum_j\vec\mu_{,j}^T\C^{-1}\vec\mu_{,i}\delta\theta_j = 0 \ .
1452: \end{equation}
1453: %
1454: After solving the linear equations, we obtain the parameter deviation $\delta\theta$
1455: %
1456: \begin{equation}
1457: \delta\theta_j =\sum_i (\bold{F}^{-1})_{ij} {\rm Tr}\{\frac{1}{2}{\C^{-1}\C_{,i}\C^{-1}(\C^{\rm B}-\C)}\} \ ,
1458: \end{equation}
1459: %
1460: where
1461: %
1462: \begin{equation}
1463: {F}_{ij} = \vec\mu_{,i}^T\C^{-1}\vec\mu_{,j} + \frac{1}{2}{\rm Tr}\{\C^{-1}\C_{,i}\C^{-1}\C_{,j}\}
1464: \end{equation}
1465: %
1466: is the Fisher matrix of the Gaussian likelihood function. Note that
1467: the bias in the recovered parameters is proportional to the difference
1468: between models $A$ and $B$.
1469:
1470:
1471: Note that in analyzing the data generated by model $B$ using model $A$
1472: changes not only the recovered parameters but also their error bars.
1473: By performing a similar calculation, the modified Fisher matrix with
1474: systematics now reads
1475: %
1476: \begin{equation}
1477: {\tilde F}_{ij}=\vec\mu^{T}_{,i} \bold C^{-1} \vec\mu_{,j} +
1478: \frac{1}{2} {\rm Tr}[\bold C^{-1} \bold C_{,i}\bold C^{-1} \bold C_{,j}\C^{-1}\C^{\rm B}] \ .
1479: \end{equation}
1480: %
1481: The error bar for all parameters estimated in model $A$ using the data
1482: generated by model $B$ can be recovered by inverting ${\bf \tilde F}$.
1483: However, in the case of counts-in-cells, the likelihood function is
1484: not perfectly Gaussian; it is convolution of Poisson and Gaussian
1485: \citep[see e.g.][]{LimaHu04,HuCohn06}. The modified Fisher matrix
1486: thus reads
1487: %
1488: \begin{equation}
1489: {\tilde F}_{ij}=\vec\mu^{T}_{,i} \bold C^{-1} \vec\mu_{,j}
1490: + \frac{1}{2} {\rm Tr}[\bold C^{-1} \bold S_{,i}\bold C^{-1} \bold S_{,j}\C^{-1}\C^{\rm B}] \ .
1491: \end{equation}
1492: %
1493:
1494:
1495: As a second example, we consider the case in which model $B$ changes
1496: the mean but not the variance of the data. One example is the effect
1497: of modified gravity on the weak lensing shear cross power spectrum
1498: \citep[e.g.][]{HutererLinder07}. Here model $A$ is the General
1499: Relativity prediction, while model $B$ is the modified gravitational
1500: prediction. In this case, $\vec\mu(\theta) \neq \vec\mu^B(\theta)$
1501: while $\C(\theta) = \C^B(\theta)$. The linear equation for
1502: $\delta\theta_j$ reads
1503: %
1504: \begin{eqnarray}
1505: {\rm Tr}\{
1506: {\sum_j\C^{-1}\C_{,i}\C^{-1}\C_{,j}\delta\theta_j}
1507: \}
1508: -2 (\vec\mu^{\rm B}-\vec\mu)^T\C^{-1}\vec\mu_{,i} +2 \sum_j \vec\mu_{,j}^T\C^{-1}\vec\mu_{,i}\delta\theta_j
1509: = 0 \ ,
1510: \end{eqnarray}
1511: %
1512: which is equivalent to
1513: %
1514: \begin{equation}
1515: \delta\theta_j=\sum_i ({\bf F}^{-1})_{ij} \{ (\vec\mu^{\rm B}-\vec\mu)^T\C^{-1}\vec\mu_{,i} \} \ .
1516: \end{equation}
1517: %
1518: Our formalism thus provides a different and generalizable route of
1519: obtaining the systematic error.
1520:
1521: %\bibliography{/Users/hao-yiwu/Desktop/master_refs}
1522:
1523: \bibliography{ms}
1524:
1525: \end{document}
1526: