1: \documentclass[prd,twocolumn,showpacs,nofootinbib]{revtex4}
2: \usepackage{times}
3: \usepackage{natbib}
4: \usepackage{epsfig}
5:
6: \newcommand{\OM}{\Omega_m}
7: \newcommand{\OB}{\Omega_b}
8: \newcommand{\OD}{\Omega_\Lambda}
9: \newcommand{\ODE}{\Omega_\up{DE}}
10: \newcommand{\rms}{\sigma_8}
11: \newcommand{\up}[1]{{\rm #1}}
12: \newcommand{\pc}{\ {\rm pc}}
13: \newcommand{\kpc}{\ {\rm kpc}}
14: \newcommand{\mpc}{{\rm Mpc}}
15: \newcommand{\hpc}{{h^{-1}\pc}}
16: \newcommand{\hkpc}{{h^{-1}\kpc}}
17: \newcommand{\hmpc}{{h^{-1}\mpc}}
18: \newcommand{\hpci}{{h\pc^{-1}}}
19: \newcommand{\hkpci}{{h\kpc^{-1}}}
20: \newcommand{\hmpci}{{h\mpc^{-1}}}
21: \newcommand{\mpci}{{\mpc^{-1}}}
22:
23: \newcommand{\apjl}{Astrophys. J. Lett.}
24: \newcommand{\aj}{Astron. J.}
25: \newcommand{\aap}{Astron. Astrophys.}
26: \newcommand{\mnras}{Mon. Not. R. Astron. Soc.}
27: \newcommand{\apjs}{Astrophys. J. Suppl. Ser.}
28: \newcommand{\apss}{Astrophys. Space Sci.}
29: \newcommand{\physrep}{Phys. Rep.}
30:
31: \newcommand{\mk}{$\bullet$}
32: \newcommand{\etc}{\noindent \mk}
33: \newcommand{\ang}{\hat\nabla}
34:
35: \newcommand{\bdv}[1]{{\bf #1}}
36: \newcommand{\beeq}{\vspace{10pt}\begin{equation}}
37: \newcommand{\eneq}{\vspace{10pt}\end{equation}}
38: \newcommand{\bear}{\vspace{10pt}\begin{eqnarray}}
39: \newcommand{\enar}{\end{eqnarray}{\\\\ \noindent}}
40:
41: \newcommand{\Vang}{\bdv{\hat n}}
42: \newcommand{\Sang}{\bdv{\hat s}}
43: \newcommand{\zobs}{z_\up{obs}}
44: \newcommand{\ztobs}{\tilde z}
45: \newcommand{\tobs}{\tilde\theta}
46: \newcommand{\pobs}{\tilde\phi}
47: \newcommand{\dobs}{\delta_\up{obs}}
48: \newcommand{\nobs}{n_\up{obs}}
49: \newcommand{\ntobs}{\tilde n}
50: \newcommand{\fobs}{f_\up{obs}}
51: \newcommand{\Ntr}{n_\up{t}}
52: \newcommand{\dMB}{\delta_\up{mb}}
53: \newcommand{\devo}{\delta_\up{evo}}
54: \newcommand{\zdist}{\delta_\up{z}}
55: \newcommand{\rtd}{r_3}
56:
57: \newcommand{\xim}{\xi_m}
58: \newcommand{\Pm}{P_m}
59: \newcommand{\xiMB}{\xi_\up{mb}}
60: \newcommand{\xiobs}{\xi_\up{obs}}
61: \newcommand{\xievo}{\xi_\up{evo}}
62: \newcommand{\xint}{\xi_\up{int}}
63: \newcommand{\xidMB}{\xi_\up{\delta~\up{mb}}}
64: \newcommand{\xizz}{\xi_\up{zz}}
65: \newcommand{\xidz}{\xi_\up{\delta\up{z}}}
66: \newcommand{\xizd}{\xi_\up{\up{z}\delta}}
67: \newcommand{\clcor}[1]{C_l^{#1}}
68:
69: \begin{document}
70:
71: \title{Complete treatment of galaxy two-point statistics: \\
72: gravitational lensing effects and redshift-space distortions}
73:
74: \author{Jaiyul Yoo}
75: \altaffiliation{Electronic address: jyoo@cfa.harvard.edu}
76: \affiliation{Harvard-Smithsonian Center for Astrophysics, Harvard University,
77: 60 Garden Street, Cambridge, MA 02138}
78:
79:
80: \date{\today}
81:
82: \begin{abstract}
83: We present a coherent theoretical framework for computing gravitational
84: lensing effects and redshift-space distortions in an inhomogeneous universe
85: and investigate their impacts on galaxy two-point statistics.
86: Adopting the linearized Friedmann-Lema\^\i tre-Robertson-Walker metric,
87: we derive the gravitational lensing and the
88: generalized Sachs-Wolfe effects that
89: include the weak lensing distortion, magnification, and time delay effects,
90: and the redshift-space distortion, Sachs-Wolfe, and integrated Sachs-Wolfe
91: effects, respectively.
92: Based on this framework, we first compute their effects on observed source
93: fluctuations, separating them
94: as two physically distinct origins: the volume effect that involves
95: the change of volume and is always present in galaxy two-point statistics,
96: and the source effect that depends on the intrinsic properties of source
97: populations. Then
98: we identify several terms that are ignored in the
99: standard method, and we compute the observed
100: galaxy two-point statistics, an ensemble average of all the combinations
101: of the intrinsic source fluctuations and the additional contributions from
102: the gravitational lensing and the generalized Sachs-Wolfe effects.
103: This unified treatment of galaxy two-point statistics clarifies the relation
104: of the gravitational lensing and the generalized Sachs-Wolfe effects
105: to the metric perturbations and the underlying matter fluctuations.
106: For near future dark energy surveys, we compute
107: additional contributions to the observed galaxy two-point statistics
108: and analyze their impact on the anisotropic structure.
109: Thorough theoretical modeling of galaxy two-point statistics would be
110: not only necessary to analyze precision measurements from upcoming dark energy
111: surveys, but also provide further
112: discriminatory power in understanding the underlying physical mechanisms.
113: \end{abstract}
114:
115: \pacs{98.80.-k,98.65.-r,98.80.Jk,98.62.Py}
116:
117: \maketitle
118:
119: \section{Introduction}
120: \label{sec:intro}
121: The standard inflationary models with a single inflaton potential predict
122: a nearly perfect Gaussian spectrum of primordial fluctuations
123: \citep{BASTTU83,STARO82,HAWKI82,GUPI82,MUFEBR92}. Two-point
124: statistics, correlation function in real space and power spectrum in Fourier
125: space, constitutes a complete description of Gaussian random fields, and it
126: has been widely used to understand the physics of the early universe from
127: measurements of the cosmic microwave background and large-scale structure.
128: The recent discovery \citep{RIFIET98,PEADET99}
129: of the late time acceleration of the universe has spurred
130: extensive investigations of a mysterious
131: energy component with negative pressure, dubbed dark energy.
132: Observationally,
133: upcoming dark energy surveys will measure galaxy two-point statistics with
134: unprecedented precision from millions of galaxies, constraining the expansion
135: history and the spatial curvature of the universe. Consequently, accurate
136: theoretical modeling of galaxy two-point statistics would be
137: crucial to take full advantage of the promise that these
138: future surveys will deliver.
139:
140: In achieving this goal, complications arise notably from the nonlinear
141: evolution of matter and scale-dependence of galaxy bias.
142: In this paper we limit ourselves to the linear bias model \citep{KAISE84}
143: and study the linear theory predictions and its corrections, considering that
144: recent attention has been paid to measuring galaxy two-point statistics
145: in the linear regime (e.g., \citep{EIBLET05,TEEIST06,PENIET07,PASCET07}).
146: However, measurement precision is often highest on nonlinear scales, and
147: proper modeling of galaxy bias on nonlinear scales can substantially increase
148: the leverage to constrain the underlying physics (see, e.g.,
149: \citep{JIMOBO98,SELJA00,MAFR00,PESM00,SCSHET01,BEWE02,COSH02}).
150:
151: Further complication arises from the distortion of
152: redshift-space structure by peculiar velocities, which results in anisotropy
153: from otherwise isotropic two-point statistics \citep{KAISE87,HAMIL92}.
154: The standard practice is to analyze the angle-averaged correlation function
155: or power spectrum, or to construct a linear combination of their multipole
156: components, suppressing the angular dependence of two-point statistics.
157: However, analyzing the full anisotropic structure, though observationally
158: challenging, can utilize additional information that is lost to some degree
159: in the standard practice \citep{MATSU04,SEEI07,OKMAET08,GACAHU08}.
160:
161: Gravitational lensing, often assumed to be negligible in galaxy two-point
162: statistics, deflects the
163: propagation of light rays, displacing the position of observed
164: galaxies, and it alters the unit area on the sky and magnifies
165: the observed flux, changing the observed number density of galaxies.
166: The former effect on two-point statistics is to convolve it with the power
167: spectrum of the lensing potential, smoothing out the features in galaxy
168: two-point statistics \citep{SELJA96}.
169: The latter effect, known as the magnification bias \citep{NARAY89},
170: is often used to measure the galaxy-matter cross-correlation function
171: from two source populations separated by large line-of-sight distance
172: \citep{SCMEET05,BLPOET06}. Recent work \citep{MATSU00,VADOET07,HUGALO07}
173: showed that these effects on galaxy two-point statistics are
174: non-negligible at the level of accuracy adequate for upcoming dark energy
175: surveys.
176:
177: However, it is unclear whether this list of additional contributions
178: on galaxy two-point statistics is exhaustive, and what are the contribution
179: terms that are ignored
180: in the standard method but need to be considered if higher accuracy is
181: dictated by observations.
182: Here we present a coherent theoretical framework for computing
183: gravitational lensing effects and redshift-space distortions, and investigate
184: their impacts on galaxy two-point statistics in an inhomogeneous universe.
185: Our treatment generalizes the early work \citep{MATSU00}
186: and complements the recent work \citep{VADOET07,HUGALO07},
187: providing a unified description of galaxy two-point statistics.
188: However, we emphasize that these effects naturally arise from metric
189: perturbations in our approach, comprising a complete and exhaustive set of
190: additional (linear order) contributions to galaxy two-point statistic.
191:
192: The rest of this paper is organized as follows.
193: In Sec.~\ref{sec:formalism}, we describe our notation for the
194: Friedmann-Lema{\^\i}tre-Robertson-Walker (FLRW) metric
195: and derive the gravitational lensing and the generalized Sachs-Wolfe effects.
196: In Sec.~\ref{sec:density}, we study their impacts on source galaxy
197: fluctuations and discuss their correspondence to the standard redshift-space
198: distortion and gravitational lensing effect. In Sec.~\ref{ssec:two},
199: we derive the observed galaxy two-point statistics
200: in real space and in Fourier space, and we compare the effects
201: of each contribution term on the observed galaxy two-point statistics
202: in Sec.~\ref{ssec:com}. We conclude in Sec.~\ref{sec:discuss} with a
203: discussion of the further improvement of our approach.
204:
205: \section{Formalism}
206: \label{sec:formalism}
207: Here we describe our notation for a background metric in an inhomogeneous
208: universe and derive
209: governing equations for non-relativistic matter in Sec.~\ref{ssec:metric}.
210: Combining these with photon geodesic equations, we derive the generalized
211: Sachs-Wolfe (Sec.~\ref{ssec:sachs}) and the gravitational lensing
212: (Sec.~\ref{ssec:lensing}) effects, developing a coherent
213: framework for describing how matter fluctuations affect observable quantities.
214:
215: \subsection{Metric and Perturbations}
216: \label{ssec:metric}
217: We assume that the homogeneous and isotropic background of the universe is
218: described by the Friedmann-Lema\^\i tre-Robertson-Walker (FLRW)
219: metric and its inhomogeneous part is represented by the perturbations for the
220: cosmological fluids and the spacetime geometry:
221: \beeq
222: ds^2=-a^2(\eta)\left[1+2\psi\right]d\eta^2+a^2(\eta)\left[1+2\phi\right]
223: g_{\alpha\beta}^{(3)}dx^\alpha dx^\beta,
224: \label{eq:metric}
225: \eneq
226: with the metric tensor for a three-space of constant spatial curvature
227: $K=-H^2_0(1-\Omega_0)$,
228: \beeq
229: g_{\alpha\beta}^{(3)}dx^\alpha dx^\beta=d\chi^2+r^2(\chi)d\Omega^2,
230: \eneq
231: where $a(\eta)$ is the scale factor for the expansion of the background
232: as a function of the conformal time $\eta$, and the comoving angular diameter
233: distance is
234: $r(\chi)=K^{-1/2}\sin(\sqrt{K}\chi)$ for a closed universe $K>0$ and
235: $(-K)^{-1/2}\sinh(\sqrt{-K}\chi)$ for an open universe
236: $K<0$, where $\chi$ is the comoving line-of-sight distance.
237: The flat limit can be obtained as $K\rightarrow0$.
238: We will denote the covariant derivative of a three-tensor
239: with respect to $g^{(3)}_{\alpha\beta}$ as a vertical bar and the covariant
240: derivative in the spacetime metric as a semicolon in the following.
241: Here Latin indices represent 4D space-time components, and
242: Greek indices run from~1 to~3, representing the spatial part of the metric.
243: Throughout the paper, we set the speed of light $c\equiv1$
244:
245: We express the perturbations in the conformal Newtonian gauge, where $\psi$
246: and $\phi$ correspond to the intuitive physical quantities, i.e., Newtonian
247: potential and Newtonian
248: curvature. This choice of gauge condition leaves no residual
249: degree of freedom up to the first-order in
250: perturbations. Here we only consider scalar perturbations, as primordial
251: vector perturbations decay quickly in a universe with ordinary components
252: and the current upper limit on tensor perturbations is order of magnitude
253: smaller than the amplitude of scalar perturbations
254: (e.g., \citep{TESTET06,SPBEET07,KODUET08})
255:
256: Given the stress energy tensor $T^{ab}$ of cosmological components,
257: the evolution of the matter and metric perturbations is governed by the
258: Einstein equations $G_{ab}=8\pi GT_{ab}$, and the Bianchi identities
259: $T^{ab}{_{;b}}=0$ guarantee the conservation of energy and momentum
260: (e.g., \citep{BARDE80,KOSA84,MUFEBR92,HWNO02,HU04}).
261: Current cosmological observations favor a universe dominated by
262: dark energy, but with non-relativistic matter as the major source of metric
263: perturbations. In this universe, the scalar Einstein equations are
264: \bear
265: (k^2-3K)~\phi&=&{3H_0^2\over2}~\OM~\left[{\delta\over a}+3H~{v\over k}\right],
266: \\
267: \psi&=&-~\phi,
268: \enar
269: where $\delta$ is the density perturbation in non-relativistic matter.
270: The Hubble parameter is $H=\dot a/a$, where the overdot
271: is the derivative with respect to time, $dt=a~d\eta$. The matter density and
272: the Hubble parameters at the present day $a_0$ are denoted as $\OM$ and
273: $H_0$, respectively.
274: The Newtonian curvature is identical to the Newtonian potential with the
275: opposite sign ($\psi=-~\phi$)
276: in the matter-dominated era, where there is vanishing
277: anisotropic stress.
278: The conservation of energy momentum provides the continuity and Euler
279: equations,
280: \bear
281: \dot\delta+{k\over a}~v&=&-~3~\dot\phi, \\
282: \dot v+Hv&=&{k\over a}~\psi,
283: \enar
284: where $v$ is the velocity of non-relativistic matter in units of $c$.
285: In the conformal Newtonian gauge, the relativistic equations on sub-horizon
286: scales correspond to the usual Newtonian equations,
287: \bear
288: \label{eq:poisson}
289: \nabla^2\psi&=&{3H_0^2\over2}~\OM~{\delta\over a},\\
290: \label{eq:peculiar}
291: \bdv{v}&=&-~{2\over3}~{a^2Hf\over\OM H_0^2}~\nabla\psi,
292: \enar
293: where $f=d\ln D/d\ln a$ and $D$ is a growth factor of the matter density
294: perturbation. The evolution of the density perturbation is related to the
295: Newtonian potential by
296: \beeq
297: \ddot \delta+2H\dot\delta=\nabla^2\psi,
298: \eneq
299: and the growth factor $D$ is a growing solution of this differential equation,
300: normalized to a unity at $a_0$.
301: Full relativistic consideration results in additional multiple terms in the
302: right-hand side of the equation (e.g., \citep{MUFEBR92}),
303: but they are suppressed at least by the
304: ratio of a characteristic scale $1/k$ to the Hubble distance $1/H$.
305: Note that we have interchangeably expressed equations in Fourier space and
306: configuration space, which is valid to the linear order in perturbations
307: and significantly simplifies the manipulations.
308:
309: \subsection{Geodesic Equations and Sachs-Wolfe Effects}
310: \label{ssec:sachs}
311: The propagation of light rays is described by a photon geodesic
312: $x^a(\lambda)$ with an affine parameter $\lambda$, and a null vector
313: $k^a=dx^a/d\lambda$ tangent to $x^a$ is determined by the null equation
314: ($ds^2=k^ak_a=0$) and the geodesic equations ($k^a{_{;b}}k^b=0$).
315: In a perturbed FLRW universe, the null vector can be expressed as
316: \beeq
317: k^0={\nu\over a}~(1+\delta\nu),~~~
318: k^\alpha=-~{\nu\over a}~(e^\alpha+\delta e^\alpha),
319: \label{eq:null}
320: \eneq
321: where $\nu$ and $e^\alpha$ are the photon frequency and its (time-reversed)
322: propagation direction from the observer, and the dimensionless quantities
323: $\delta\nu$ and
324: $\delta e^\alpha$ represent their perturbations.
325: In a homogeneous expanding universe,
326: the null vector follows the usual relations $\nu\propto1/a$,
327: $e^\alpha e_\alpha=1$, and $de^\alpha/d\eta=e^\beta e^\alpha{_{|\beta}}$, and
328: indeed Eq.~(\ref{eq:null}) may be derived from the null and
329: the geodesic equations.
330: For a comoving observer whose rest frame has vanishing energy flux,
331: the four velocity is $u^a=(1/a,0)$ and the observed frequency $\nu_\up{obs}$
332: of a photon source is related to the frequency $\nu_\up{e}$ at the emission
333: by a redshift parameter,
334: \beeq
335: 1+z={(k^a~u_a)_\up{e}\over(k^a~u_a)_\up{obs}}={\nu_\up{e}\over\nu_\up{obs}}
336: ={1\over a_\up{e}},
337: \label{eq:redshift}
338: \eneq
339: where we assumed $a_\up{obs}=a_0=1$.
340:
341: In an inhomogeneous universe, the observed redshift $\zobs$ deviates from
342: the true redshift $z$. Perturbations in the null equation is
343: \beeq
344: e^\alpha~\delta e_\alpha=\delta\nu+\psi-\phi,
345: \eneq
346: and perturbations in the geodesic equations for the temporal
347: and spatial components are
348: \bear
349: {d\over dy}(\delta\nu+\psi)&=&\psi_{,\alpha}~e^\alpha-{d\phi\over d\eta},
350: \label{eq:zerogeo}\\
351: {d\over dy}(\delta e^\alpha+2\phi ~e^\alpha)&=&\delta e^\beta~ e^\alpha{_{|\beta}}
352: -\delta\nu~{de^\alpha\over d\eta}+\psi^{|\alpha}-\phi^{|\alpha},~~~~~
353: \label{eq:ageo}
354: \enar
355: where we used the zeroth order null geodesic
356: $d/dy\equiv\partial_\eta-e^\alpha\partial_\alpha=(a/\nu)(d/d\lambda)$
357: and kept the terms to the first order in perturbations.
358:
359: The four velocity of a comoving observer is now
360: $u^a=((1-\psi)/a,~v^\alpha/a)$ and the observed redshift is
361: \beeq
362: 1+\zobs={(k^a~u_a)_\up{e}\over(k^a~u_a)_\up{obs}}
363: =(1+z)\left[1+(\delta\nu+\psi+v_\alpha~ e^\alpha)^\up{e}_\up{o}\right].
364: \eneq
365: This can be further simplified by using Eq.~(\ref{eq:zerogeo}) as
366: \bear
367: \label{eq:gSW}
368: 1+\zobs&=&(1+z)\times\bigg[1+V(z)-V(0)\\
369: &-&\psi(z)+\psi(0)+\int_0^y dy~{\partial\over\partial\eta}
370: (\phi-\psi)\bigg],\nonumber
371: \enar
372: where $V=v_\alpha ~e^\alpha$ is the line-of-sight velocity
373: \citep{SAWO67,HWNO99,MATSU00}. The additional terms in the square
374: bracket alter the simple redshift-distance relation in
375: Eq.~(\ref{eq:redshift}), giving rise to the standard redshift-space distortion
376: by peculiar velocities, the Sachs-Wolfe effect by gravitational redshift, and
377: the integrated Sachs-Wolfe effect by the time evolution of gravitational
378: potential across which photons propagate.
379: Hereafter
380: we will collectively refer to these effects as the generalized Sachs-Wolfe
381: effect.
382:
383: \subsection{Gravitational Lensing}
384: \label{ssec:lensing}
385: In a homogeneous universe, the gravitational lensing effects vanish
386: and light rays propagate with the direction unchanged.
387: For a photon source at $\bdv{\hat z}$-axis in an inhomogeneous universe,
388: the propagation direction from the observer is
389: $\Vang=e^\alpha=(0,0,1)$ and the null vector is
390: $k^{x,y}=-(\nu/a)\delta e^{x,y}$. The null vector is further
391: related to the photon position $r(\chi)\Vang=(x,y)$ on the
392: sky at any time by
393: \beeq
394: k^{x,y}={d\over d\lambda}(r\Vang)={\nu\over a}{d\over dy}(r\Vang),
395: \eneq
396: where we replace the derivative with respect to the affine parameter by using
397: the zeroth order null geodesic, consistent to the first order in perturbations.
398: The spatial component of the geodesic equation (Eq.[\ref{eq:ageo}]) is then
399: \beeq
400: {d\over dy}(\delta e^{x,y})=-{d^2\over dy^2}(r\Vang)=
401: \psi^{|\alpha}-\phi^{|\alpha}=2~\psi^{|\alpha}.
402: \label{eq:path}
403: \eneq
404:
405: Since gravitational lensing conserves the surface brightness, the observed
406: surface brightness $I_\up{obs}(\Vang)$ on the sky is simply
407: the intrinsic surface
408: brightness at the source position $\Sang$: $I_\up{obs}(\Vang)=I(\Sang)$,
409: and the source
410: position $\Sang$ can be obtained by integrating Eq.~(\ref{eq:path})
411: along the photon geodesic
412: \beeq
413: \Sang=\Vang+\ang~\Psi(\Vang),
414: \label{eq:lens}
415: \eneq
416: with the projected lensing potential
417: \bear
418: \Psi(\Vang)&=&-~2\int_0^{y_s}dy'\int_0^{y'}dy ~{\psi(y)\over
419: r(\chi_s)~r(\chi)} \nonumber \\
420: &=&-~2\int_0^{y_s}dy~\psi(y)~{r(\chi_s-\chi)\over r(\chi_s)~r(\chi)},
421: \enar
422: where $\ang$ is the derivative with respect to $\Vang$, and
423: $\chi_s=\int_0^{z_s}dz/H(z)$ is the comoving line-of-sight distance to the
424: source redshift $z_s$.
425: The integration along the unperturbed photon geodesic $dy$ is often called
426: the Born approximation. Following the literature, we take the geodesic
427: as the photon radial direction $d\chi$, but note that
428: $d/d\chi=\partial_\eta-\partial_\chi$.
429:
430: The convergence $\kappa(\Vang)$ is defined as
431: $\ang^2\Psi(\Vang)=-2\kappa(\Vang)$ and it is further related
432: to density fluctuations along the geodesic by Poisson's equation
433: (Eq.[\ref{eq:poisson}])
434: \bear
435: \kappa(\Vang)&=&\int_0^{\chi_s}d\chi~(\nabla^2-\nabla^2_\chi)~
436: \psi[r(\chi)\Vang,\chi]~
437: {r(\chi_s-\chi)~r(\chi)\over r(\chi_s)}~~~ \nonumber \\
438: &=&{3H_0^2\over2}\OM\int_0^{\chi_s}\!\!d\chi~
439: {\delta[r(\chi)\Vang,\chi]\over a(\chi)}
440: ~{r(\chi_s-\chi)~r(\chi)\over r(\chi_s)}.
441: \label{eq:conv}
442: \enar
443: The contribution from the radial derivatives $\nabla^2_\chi$
444: is proportional to the potential
445: difference between the source and observer, and this boundary term is
446: negligible compared to the first term \citep{JASEWH00,HISE03a}.
447: Numerical ray tracing experiments through
448: $N$-body simulations show that the weak lensing approximation to the first
449: order in perturbations is accurate even in nonlinear regime when nonlinear
450: matter power spectrum is used in place of linear matter power spectrum
451: \citep{JASEWH00}. Also note that all the prior results for a single source
452: redshift can be readily generalized to a source population with a redshift
453: distribution $W(\chi_s)$ by integrating the results over $\chi_s$ with
454: $W(\chi_s)$ in the integrand.
455:
456: While conservation of surface brightness guarantees that photons are neither
457: destroyed nor created, gravitational deflection
458: distorts the cross-section of a bundle of light rays, magnifying (or
459: de-magnifying) observed fluxes. Gravitational lensing
460: magnification $\mu(\Vang)$
461: is related to the Jacobian of a mapping from the image plane to the source
462: plane by
463: \bear
464: \mu(\Vang)^{-1}&=&\left|{d^2\Sang\over d^2\Vang}\right|
465: =\left|\bdv{I}+\ang\ang~\Psi(\Vang)\right| \\
466: &=& \left|\left[1-\kappa(\Vang)\right]^2-\gamma^2(\Vang)\right|, \nonumber
467: \enar
468: where $\bdv{I}$ is a unit $2\times2$ matrix and
469: $\gamma(\Vang)$ is the tangential shear. In the weak lensing regime,
470: $\mu(\Vang)=1+2~\kappa(\Vang)$.
471:
472: Gravitational lensing also modifies the propagation time of light rays in two
473: ways, compared to the light travel time in the absence of the gravitational
474: lensing effects: it distorts the photon geodesic, increasing the path length
475: that photons travel, and the gravitational potential retards the light
476: travel time. The former is referred to as the geometric time delay
477: \citep{BLNA86}
478: \beeq
479: \tau_\up{geo}(\Vang)={1\over2}~{r(\chi_l)~r(\chi_s)\over r(\chi_s-\chi_l)}~
480: \ang\Psi(\Vang)\cdot\ang\Psi(\Vang),
481: \eneq
482: and the latter is the potential or Shapiro time delay \citep{SHAPIRO64}
483: \beeq
484: \tau_\up{pot}(\Vang)={r(\chi_l)~r(\chi_s)\over r(\chi_s-\chi_l)}~
485: \Psi(\Vang).
486: \eneq
487: These effects can be derived by using the small angle approximation in
488: deflection and the relation $d\eta=(1-\psi+\phi)d\chi$ from the metric
489: in Eq.~(\ref{eq:metric}).
490: Note that the proper time delay can be obtained by multiplying the lens
491: redshift $1+z_l$ in the limit of a single lens case, and this derivation in
492: a cosmological context recovers the standard relation for time delay.
493:
494: \section{Source Fluctuations}
495: \label{sec:density}
496: Inhomogeneous matter fluctuations in the universe deflect the propagation
497: of light rays, giving rise to the gravitational lensing effects.
498: The generalized Sachs-Wolfe effect also
499: arises from the same matter fluctuations responsible for the gravitational
500: lensing effects. Having discussed the basic mechanism of the
501: gravitational lensing and the generalized Sachs-Wolfe effects
502: that complicate the simple interpretation of observable quantities,
503: we now investigate their impact on an observed overdensity field
504: $\dobs(\Vang,z)$ of source galaxies. Contributions to $\dobs(\Vang,z)$
505: come from matter fluctuations in addition
506: to the intrinsic overdensity $\delta(\Vang,z)$ of source galaxies.
507: Noting that the contributions can be linearized and added
508: to the first order in perturbations, we separate these contributions
509: as two physically distinct parts: one that involves the
510: change of volume, and one that involves the intrinsic properties
511: of source galaxies.
512: The impact on galaxy two-point statistics will be discussed in
513: the following section.
514:
515: \subsection{Volume Effect}
516: \label{ssec:vol}
517: Consider a unit comoving volume $dV=r^2(\chi)d\Omega dz/H(z)$
518: and a unit flux interval $df$, and let
519: $n(\Vang,z,f)$ be the comoving number density of source galaxies.
520: The generalized Sachs-Wolfe effect alters the unit comoving volume $dV$.
521: Note, however, that it not only changes the unit redshift interval $dz$,
522: but also changes both the angular diameter distance $r(\chi)$ and the
523: Hubble parameter $H(z)$. By imposing the number conservation,
524: the observed number density of the source galaxies can be obtained by
525: \beeq
526: \label{eq:nSW}
527: \nobs(\zobs)=n(z)\left[1+\delta V\right]
528: {r^2(\chi)\over r^2(\chi_\up{obs})}
529: {H(\zobs)\over H(z)}{dz\over d\zobs},
530: \eneq
531: where $\delta V=(2\phi+\varepsilon)^e_o$ represents the distortion of volume
532: element, when it is transformed from the conformal Newtonian gauge to
533: the local Lorentz frame, where the velocity of non-relativistic matter
534: vanishes. We give a more rigorous derivation in Appendix~\ref{app:ngal}.
535: The solid angle $d\Omega$ remains unaffected by the generalized Sachs-Wolfe
536: effect.
537:
538: If the mean comoving number density evolves slowly compared
539: to the redshift change due to the generalized Sachs-Wolfe effect
540: $\bar n(z)=\bar n(\zobs)$,
541: contributions to $\dobs(\zobs)$ arise solely from the change in
542: volume element $dV$,
543: \beeq
544: \label{eq:dgSW}
545: \dobs(\zobs)=\delta(z)-2~{1+z\over H\chi}~\varepsilon
546: -(1+z)H~{d\over dz}\left({\varepsilon\over H}\right)
547: -\varepsilon+\delta V,
548: \eneq
549: where we rewrote Eq.~(\ref{eq:gSW}) as $1+\zobs=(1+z)(1+\varepsilon)$ and
550: the contribution $\varepsilon$
551: from the generalized Sachs-Wolfe effect is
552: \beeq
553: \varepsilon(z)= V(z)-V(0)-\psi(z)+\psi(0)-2
554: \int_0^\chi d\chi{\partial\psi\over\partial\eta}.
555: \label{eq:varep}
556: \eneq
557: In the Einstein-de~Sitter universe,
558: the Newtonian potential is constant and hence the
559: integrated Sachs-Wolfe effect vanishes. In general, as we show in the next
560: section, the peculiar velocity effect is dominant over the
561: Sachs-Wolfe and the integrated Sachs-Wolfe effects, and
562: $\varepsilon(z)\simeq V(z)-V(0)$.
563: Note that while our derivation so far is valid for nonflat universes,
564: in deriving Eq.~(\ref{eq:dgSW}) we assumed that the spatial curvature $K$
565: is close to zero. The second term in Eq.~(\ref{eq:dgSW}) has a multiplicative
566: factor $\sqrt{K}\chi/\tan(\sqrt{K}\chi)$ for a closed universe $K>0$ and
567: $\sqrt{-K}\chi/\tanh(\sqrt{-K}\chi)$ for an open universe $K<0$, which
568: becomes a unity as $K\rightarrow0$.
569:
570: With a proper line-of-sight distance $r_p=\chi(z)/(1+z)$ and a
571: normalized peculiar velocity
572: $u=V(z)/H(z)$, Eq.~(\ref{eq:dgSW}) can be rearranged as
573: \beeq
574: \dobs(\zobs)=\delta(z)-{2u\over r_p}-{du\over dr_p},
575: \label{eq:stdz}
576: \eneq
577: if we ignore the Sachs-Wolfe and the integrated Sachs-Wolfe effects in
578: Eq.~(\ref{eq:varep}).
579: This recovers the standard relation for redshift-space distortions
580: \citep{KAISE87,STWI95,HAMIL98}. Note that
581: the standard method ignores the contributions
582: in Eq.~(\ref{eq:dgSW}) from the Sachs-Wolfe and the integrated Sachs-Wolfe
583: effects. We discuss their impact in Sec.~\ref{sec:two}.
584:
585: Gravitational lensing magnification increases
586: the flux interval $df$ and the solid angle $d\Omega$ by a factor of
587: $\mu$, respectively. With the number conservation in $dV$ and $df$,
588: the observed number density is therefore
589: \beeq
590: \nobs(\fobs)=n(f)~{df\over d\fobs}~{d\Omega\over d\Omega_\up{obs}}=
591: {1\over\mu^2}~n(f).
592: \label{eq:mag}
593: \eneq
594: Similarly, if the mean comoving number density is the same over the
595: flux change due to lensing magnification (i.e., the source
596: luminosity function is flat), the observed overdensity is then
597: \beeq
598: \dobs(\Vang)=\delta(\Vang)-4~\kappa(\Vang),
599: \eneq
600: reflecting the change in volume and flux.
601:
602: Gravitational lensing displaces the source position on the sky
603: according to Eq.~(\ref{eq:lens}), and the observed number density is
604: $\nobs(\Vang)=n\left[\Vang+\ang\Psi(\Vang)\right]$. By Taylor expanding
605: $\nobs(\Vang)$ to the first order in $\Psi(\Vang)$,
606: the observed overdensity can be written as
607: \beeq
608: \dobs(\Vang)=\delta(\Vang)+\ang\Psi(\Vang)\cdot\ang\delta(\Vang).
609: \label{eq:gld}
610: \eneq
611: Note that the additional contribution is already in the second order in
612: perturbations and furthermore it vanishes on average, because the deflection
613: angle $\ang\Psi(\Vang)$ has no preferred direction. The first non-vanishing
614: effect from gravitational
615: lensing displacement comes in the second order in $\Psi(\Vang)$
616: \citep{VADOET07}, and we therefore ignore this effect.
617:
618: Finally the gravitational time delay decreases the arrival time of photons
619: in an overdense region, compared to that in the absence of
620: lensing. The net effect is therefore that we sample sources at farther
621: distance in the fixed time interval \citep{HUCO01}.
622: However, for discrete sources
623: the effect vanishes as long as the life time of the sources is longer
624: than the time delay.
625:
626: \subsection{Source Effect}
627: \label{ssec:src}
628: The generalized Sachs-Wolfe and the gravitational lensing effects modify a
629: unit volume and a unit flux interval, leading to the contributions to
630: $\dobs(\Vang,z,f)$. Furthermore, the changes in observed redshift and flux
631: can result in different mean number densities, if the redshift distribution of
632: the source galaxy population varies
633: in the redshift interval or the luminosity function is non-trivial
634: over the flux change. These additional contributions from the change
635: in mean number densities are related to the intrinsic properties of source
636: galaxies, and we collectively refer to these effects as the source effect.
637: However, note that while the source effect may be absent for some galaxy
638: populations, the volume effect is always present. Therefore, we keep together
639: the contributions from the volume effect in considering the source effect.
640:
641: We first consider the effect of gravitational lensing magnification.
642: Lensing magnification not
643: only increases $d\Omega$ and $df$ in Eq.~(\ref{eq:mag}), but also changes
644: the number count of source galaxies,
645: if the luminosity function is non-flat, i.e.,
646: $\bar\nobs(\fobs)\ne\bar n(f)$. Assuming $\bar n(f)df\propto f^{-s}df$
647: with a constant slope $s$ over a narrow flux range $df$, the observed
648: number density can be expressed as
649: \beeq
650: \bar\nobs(\fobs)={\bar n(\fobs/\mu)\over\mu^2}=\bar n(\fobs)~\mu^{s-2},
651: \eneq
652: and the observed overdensity is now
653: \bear
654: \label{eq:mgb}
655: \dobs(\Vang)&=&\delta(\Vang)+(2s-4)\kappa(\Vang) \\
656: &=&\delta(\Vang)+5(p-0.4)\kappa(\Vang), \nonumber
657: \enar
658: where we used the logarithmic slope $p=d\log\bar n(m)/dm=0.4(s-1)$
659: in a sample with limiting magnitude $m$. In the literature,
660: these contributions from both the volume and the source effects
661: are referred to as the magnification bias
662: \citep{NARAY89,BARTE95,JAIN02,JASCSH03,MOJA98}. Note that this bias can be
663: either positive or negative, depending on the slope $p$, and the volume
664: effect can be canceled by the source effect with $p=0.4$
665: (see \citep{SCMEET05} for the recent detection from the Sloan Digital
666: Sky Survey).
667:
668: The redshift distribution of source galaxies
669: also affects the mean number counts due
670: to the generalized Sachs-Wolfe effect. For a redshift distribution
671: $\bar n(z)dz\propto z^\alpha\exp\left[-(z/z_0)^\beta\right]dz$,
672: the observed overdensity can be obtained by substituting $\bar n(z)$
673: with $\bar n\left[\zobs-(1+\zobs)\varepsilon\right]$,
674: \bear
675: \label{eq:zevo}
676: \dobs(z)&=&\delta(z)-{1+z\over z}\left[\alpha-\beta
677: \left({z\over z_0}\right)^\beta\right]\varepsilon \\
678: &-&2~{1+z\over H\chi}~\varepsilon
679: -(1+z)H{d\over dz}
680: \left({\varepsilon\over H}\right)-\varepsilon+\delta V, \nonumber
681: \enar
682: where the second term in the right-hand side is
683: the additional contribution related to the evolution of source galaxies, and
684: the rest of the additional terms come from the volume effect in
685: Eq.~(\ref{eq:dgSW}).
686:
687:
688: \subsection{Summary}
689: \label{ssec:sum}
690: We have investigated the effects of inhomogeneous matter fluctuations on
691: observed overdensity fields. Here we summarize their contributions and
692: clarify the functional dependence. We then compare their impact on galaxy
693: two-point statistics in Sec.~\ref{sec:two}.
694:
695: For a sample of galaxies at redshift $z$ selected with a limiting flux $f$ and
696: narrow intervals of $dz$ and $df$, the observed overdensity
697: $\dobs(\Vang,z,f)$ is the sum of
698: the intrinsic overdensity field $\delta(\Vang,z,f)$
699: and the contributions from the gravitational
700: lensing and the generalized Sachs-Wolfe effects:
701: \beeq
702: \dobs(\Vang,z,f)=\delta+\dMB+\zdist+\devo.
703: \label{eq:summary}
704: \eneq
705: From Eq.~(\ref{eq:mgb}), the magnification bias is defined as
706: \beeq
707: \dMB(\Vang,z,f)=5\left[p(f)-0.4\right]\kappa(\Vang,z),
708: \eneq
709: with redshift $z$ being the source redshift of the convergence $\kappa(\Vang)$
710: in Eq.~(\ref{eq:conv}). Considering $\varepsilon(z)\simeq V(z)-V(0)$, we
711: call the volume effect in Eq.~(\ref{eq:dgSW}) as
712: the redshift-space distortion bias,
713: \bear
714: \label{eq:zfive}
715: \zdist(\Vang,z)&=&-2~{1+z\over H\chi}~\varepsilon
716: -(1+z)H{d\over dz}
717: \left({\varepsilon\over H}\right)-\varepsilon+\delta V \nonumber \\
718: &=&-2~{1+z\over H\chi}~\varepsilon +{1+z\over H}~\varepsilon~{dH\over dz}
719: \nonumber \\
720: &&-{1+z\over H}~{\partial\varepsilon\over\partial\chi}
721: -\varepsilon+\delta V.
722: \enar
723: Note that the generalized Sachs-Wolfe effect $\varepsilon(z)$ implicitly
724: depends on the direction $\Vang$ via the line-of-sight velocity
725: $V(z)=v_\alpha e^\alpha=\Vang\cdot\bdv{v}(\Vang,z)$, but it is independent
726: of the limiting flux $f$, provided that galaxies have no velocity bias
727: (i.e., galaxies and matter follow the same velocity field).
728: Finally, the evolution bias is defined from Eq.~(\ref{eq:zevo}) as
729: \beeq
730: \devo(\Vang,z,f)=-{1+z\over z}\left[\alpha-\beta
731: \left({z\over z_0}\right)^\beta\right]\varepsilon,
732: \label{eq:bevo}
733: \eneq
734: where the directional dependence comes from $\varepsilon$ and the evolution
735: coefficients $(\alpha,\beta,z_0)$ depend on the galaxy sample selected with
736: the limiting flux $f$. While the evolution bias arising from the difference
737: between $\bar n(z)$ and $\bar n(\zobs)$ was recognized
738: \citep{KAISE87,HAMIL98,MATSU04},
739: it has been ignored in the literature.
740: However, we show in Sec.~\ref{sec:two} that the evolution bias
741: can be significantly enhanced. Last, we want to emphasize that
742: equation~(\ref{eq:summary}) is gauge-invariant as is
743: written in the conformal Newtonian gauge.
744:
745: \section{Galaxy Two-Point Statistics}
746: \label{sec:two}
747: We have derived additional contributions of the gravitational lensing and
748: the generalized Sachs-Wolfe effects to the intrinsic density fluctuations in
749: Sec.~\ref{sec:density}, fully consistent up to the first order
750: in perturbations. Given two samples of galaxies with limiting fluxes
751: $f_1$ and $f_2$, the observed galaxy correlation function is then
752: $\xiobs(\Vang_1,z_1,\Vang_2,z_2)=\langle\dobs(\Vang_1,z_1)~\dobs(\Vang_2,z_2)
753: \rangle$ and the observed power spectrum is
754: $\langle\dobs(\bdv{k}_1,z_1)~\dobs^*(\bdv{k}_2,z_2)\rangle=(2\pi)^3
755: \delta^D(\bdv{k}_1-\bdv{k}_2)P_\up{obs}(k_1)$.
756: In Sec.~\ref{ssec:two},
757: we derive this ensemble average of all the combinations of each component in
758: $\dobs$ in Eq.~(\ref{eq:summary}), after we simplify the equation.
759: We then discuss their impact on the observed galaxy two-point statistics
760: by analyzing specific examples in Sec.~\ref{ssec:com}.
761:
762: \subsection{Correlation Function and Power Spectrum}
763: \label{ssec:two}
764: Here we compute the observed galaxy correlation function $\xiobs$ and
765: power spectrum $P_\up{obs}$. However, as some components in $\dobs$ are
766: smaller than other components, their combinations are even smaller
767: by an order-of-magnitude. We therefore
768: start by estimating the auto-correlation functions of each component
769: and simplify the equation before we compute all the
770: cross-correlation functions and power spectra.
771:
772: We first consider the correlation of the redshift-space distortion bias
773: $\xizz=\langle\zdist(\Vang_1,z_1)~\zdist(\Vang_2,z_2)
774: \rangle$. The redshift-space distortion bias $\zdist$ in Eq.~(\ref{eq:zfive})
775: has five components that depend either $\varepsilon(z)$ or its partial
776: derivative with respect to $z$ or $\chi$, and the contribution
777: $\varepsilon(z)$ from the generalized Sachs-Wolfe effect in
778: Eq.~(\ref{eq:varep}) has also three different components that depend on the
779: peculiar velocity, the Newtonian potential, and its time derivative.
780: The Newtonian potential and the peculiar velocity in
781: Eqs.~(\ref{eq:poisson}) and~(\ref{eq:peculiar}) take the simple form in
782: Fourier space
783: \bear
784: \psi_\bdv{k}&=&-{3H_0^2\over2}~{\OM\over a}~{\delta_\bdv{k}\over k^2},\\
785: \bdv{v}_\bdv{k}&=&iHfa~\delta_\bdv{k}~{\bdv{k}\over k^2}.
786: \enar
787: On a typical correlation scale $1/k$, they scale as
788: $H^2\delta_k/k^2$ and $H\delta_k/k$ with $f\simeq1$ at $z\gtrsim1$:
789: $\psi_k$ is smaller than $v_k$ by the ratio of the correlation scale $1/k$
790: to the Hubble distance $1/H$. Similarly, the integrated Sachs-Wolfe effect
791: is of the same order as the Newtonian potential and it vanishes in the limit
792: of zero cosmological constant, i.e., Einstein-de~Sitter universe,
793: because it is proportional to the time derivative of the ratio of the growth
794: factor to the expansion scale factor $D(z)/a$.
795: Therefore, we can safely ignore the Sachs-Wolfe and the integrated
796: Sachs-Wolfe effects and we assume $\varepsilon(z)\simeq V(z)$. Note that
797: given a particular realization of the observer's rest frame,
798: its peculiar velocity $V(0)$ is uncorrelated and the unobservable potential
799: $\psi(0)$ in Eq.~(\ref{eq:varep})
800: can be absorbed by a gauge transformation.
801:
802: With the assumption $\varepsilon(z)\simeq V(z)$, we further simplify
803: Eq.~(\ref{eq:zfive}) by comparing the five components in the redshift-space
804: distortion bias, and similar justification was made in \citep{MATSU00}.
805: Respectively, each component scales as
806: $\delta_k/k\chi$, $H\delta_k/k$,
807: $\partial\delta_k/k\partial\chi$, $H\delta_k/k$,
808: and $H\delta_k/k$, and hence
809: they are smaller than $\delta_k$ by the ratio of
810: correlation scale $1/k$ to the Hubble distance $1/H$ or the
811: line-of-sight distance $\chi$ (roughly of order $1/H$), except the third
812: component: the partial derivative with respect to $\chi$ cancels
813: the correlation scale $1/k$ and hence the amplitude of the third component
814: is of order $\delta_k$, larger than the other components in the redshift-space
815: distortion bias. Therefore, we only keep the dominant component in
816: the redshift-space distortion bias \citep{MATSU00},
817: \beeq
818: \zdist(\Vang,z)\simeq-{1+z\over H}~{\partial V\over\partial\chi},
819: \eneq
820: consistent with the standard relation for the redshift-space distortion,
821: justifying its nomenclature. However, note that all these ignored components
822: are proportional to $\varepsilon$. At low redshift, they contribute
823: to galaxy two-point statistics at the sub-percent level, while we show
824: in Sec.~\ref{ssec:com} that at higher
825: redshift their contribution is somewhat larger.
826:
827: Having substantially reduced the number of combinations for an ensemble
828: average, we are now well positioned to compute correlation functions and
829: their power spectra.
830: For two galaxy positions $\bdv{x}_1=\left[r(\chi_1)\Vang_1,\chi_1\right]$
831: and $\bdv{x}_2=\left[r(\chi_2)\Vang_2,\chi_2\right]$, the auto-correlation
832: of the redshift-space distortion bias is
833: \bear
834: \label{eq:pzz}
835: \xizz&=&\langle\zdist(\Vang_1,z_1)~\zdist(\Vang_2,z_2)\rangle\\
836: &=&f_1f_2\int{d^3\bdv{k}\over(2\pi)^3}~e^{i\bdv{k}\cdot(\bdv{x_1-x_2})}
837: \Pm(k;z_1,z_2)~{k_z^4\over k^4} \nonumber \\
838: &=&f_1f_2\int_0^\infty{dk\over k}~{k^3\over 2\pi^2}~\Pm(k;z_1,z_2) \nonumber \\
839: &\times&
840: \left[{1\over5}j_0(k\rtd)P_0(\gamma)-{4\over7}j_2(k\rtd)P_2(\gamma)+
841: {8\over35}j_4(k\rtd)P_4(\gamma)\right], \nonumber
842: \enar
843: where the 3D comoving separation is
844: $\rtd=\left[r(\bar\chi)^2\Delta\theta^2+(\chi_2-\chi_1)^2\right]^{1/2}$ with
845: $\Delta\theta=|\Vang_1-\Vang_2|$ and
846: $\bar\chi=(\chi_1+\chi_2)/2$, and the angle subtended by the comoving
847: separation is
848: $\gamma=\cos\Delta\theta=(\chi_2-\chi_1)/\rtd$. $P_n(x)$ and $j_n(x)$
849: are the $n$-th order Legendre polynomial and spherical Bessel function,
850: respectively. We assumed
851: the distant observer approximation such that $k_z$ is the line-of-sight
852: component of the wavenumber $k$, but it can be relaxed by replacing
853: $k_z^4$ by $\left[(\Vang_1\cdot\bdv{k})(\Vang_2\cdot\bdv{k})\right]^2$.
854: The linear matter power spectrum
855: is computed by $\Pm(k;z_1,z_2)=D(z_1)D(z_2)\Pm(k)$, while
856: we use $\Pm(k;z_1,z_2)=\Pm(k;\bar z)$ with $\bar z=(z_1+z_2)/2$ when we
857: compute the effect of the nonlinear matter power spectrum using the
858: \citet{SMPEET03} approximation. The power spectrum of the redshift-space
859: distortion bias can be readily read off
860: from Eq.~(\ref{eq:pzz}) and its power is boosted along the line-of-sight by
861: $f_1f_2\mu_k^4$ with $\mu_k=k_z/k$.
862:
863: Next we consider the correlation of the evolution bias. The observed redshift
864: $\zobs$ is different from the true redshift $z$ due to the generalized
865: Sachs-Wolfe effect and the redshift distribution
866: of the source mean number density
867: gives rise to the evolution bias. The evolution bias $\devo$ is proportional
868: to $\varepsilon(z)\simeq V(z)$ and it is typically smaller than $\zdist$ by
869: the ratio of a correlation scale $1/k$ to the Hubble distance $1/H$. However,
870: beyond the mean redshift of source populations, the mean number density
871: changes exponentially and the evolution bias can be substantially boosted by
872: the prefactor
873: \beeq
874: E(z;f)=-{1+z\over z}\left[\alpha-\beta\left({z\over z_0}\right)^\beta\right],
875: \label{eq:eboo}
876: \eneq
877: defined such that Eq.~(\ref{eq:bevo}) becomes $\devo=E(z)~\varepsilon(z)$.
878: While the exact functional form of $E(z)$ depends on the assumed redshift
879: distribution, it captures the general trend of the enhancement in $\devo$
880: beyond the mean redshift. The correlation of the evolution bias is therefore
881: \bear
882: \xievo&=&\langle\devo(\Vang_1,z_1)~\devo(\Vang_2,z_2)\rangle\\
883: &=&(HfaE)_1(HfaE)_2\!\!\int\!\!\!
884: {d^3\bdv{k}\over(2\pi)^3}e^{i\bdv{k}\cdot(\bdv{x_1-x_2})}
885: \Pm(k;z_1,z_2){k_z^2\over k^4} \nonumber \\
886: &=&(HfaE)_1(HfaE)_2\int_0^\infty{dk\over k}{k\over 2\pi^2}~\Pm(k;z_1,z_2)
887: \nonumber \\
888: &\times&\left[{1\over3}j_0(k\rtd)P_0(\gamma)-{2\over3}j_2(k\rtd)
889: P_2(\gamma)\right],
890: \nonumber
891: \enar
892: where the subscripts in the round brackets represent that the products
893: $(HfaE)$ are
894: computed at $z_1$ and $z_2$. Its power spectrum is also anisotropic and has
895: structure similar to the redshift-space distortion bias.
896:
897: Finally, the inhomogeneous matter fluctuations along the two lines-of-sight
898: result in the correlation of the magnification bias
899: \bear
900: \label{eq:xiMB}
901: \xiMB&=&\langle\dMB(\Vang_1,z_1)~\dMB(\Vang_2,z_2)\rangle \\
902: &=&(5p_1-2)(5p_2-2)\left({3H_0^2\over2}\OM\right)^2
903: \int_0^{\chi_1}d\chi\left[{r(\chi)\over a(\chi)}\right]^2 \nonumber \\
904: &\times&{r(\chi_1-\chi)\over r_1}{r(\chi_2-\chi)\over r_2}~
905: w_p\left[r(\chi)\Delta\theta;z\right], \nonumber
906: \enar
907: where we used the Limber approximation \citep{LIMBE54} (see
908: Appendix~\ref{app:limber}). Without loss of generality,
909: we assumed $z_1\leq z_2$. The projected correlation function
910: $w_p(R)$ is obtained by integrating the 3D matter correlation function
911: $\xim(x)$ along the line-of-sight at a fixed redshift $z(\chi)$ and 2D
912: transverse separation $R$,
913: \bear
914: w_p\left[R;z\right]&=&\int_{-\infty}^\infty dr_\parallel~
915: \xim\left[\rtd=\sqrt{R^2+r_\parallel^2};z\right] \\
916: &=&\int_0^\infty{k~dk\over2\pi}\Pm(k;z)J_0(kR),\nonumber
917: \enar
918: where $J_n(x)$ is the $n$-th order Bessel function of the first kind.
919: Assuming that the source redshifts are sufficiently high and hence $\xiMB$
920: is independent of $z_1$ and $z_2$, the power spectrum of the magnification
921: bias is
922: $P_\up{mb}=(2\pi)\delta^D(k_z)(5p_1-2)(5p_2-2)r^2(\bar\chi)
923: C^\up{\kappa\kappa}_{l=k_\perp r(\bar\chi)}$,
924: where the angular power spectrum of the convergence is
925: \beeq
926: \label{eq:akk}
927: \clcor{\up{\kappa\kappa}}=\left({3H_0^2\over2}\OM\right)^2
928: \int_0^{\bar\chi} d\chi~\left[{r(\bar\chi-\chi)\over a(\chi)~r(\bar\chi)}
929: \right]^2\Pm\left[k={l\over r(\chi)};z\right].
930: \eneq
931: The Dirac delta function results from our assumption
932: that $\xiMB$ is a function of transverse direction only, but it can be
933: somewhat relaxed
934: by replacing $(2\pi)\delta^D(k_z)$ by a survey window function
935: \citep{HUGALO08}. Note that while we are interested in how the magnification
936: bias affects the 3D correlation of the intrinsic source fluctuations, the
937: magnification bias arises from the matter fluctuations along the line-of-sight
938: (not at a single redshift plane) and thereby angular correlation function and
939: its angular
940: power spectrum are better suited for quantifying its statistics. Indeed,
941: the correlation function of the magnification bias is identical to the angular
942: correlation function,
943: $\xiMB(\Vang_1,z_1,\Vang_2,z_2)=w_\up{mb}(\Delta\theta;z_1,z_2)$, and we relate
944: 2D angular power spectrum to 3D power spectrum by
945: $P(k)=(2\pi)\delta^D(k_z)r^2(\bar\chi)C_{l=k_\perp r(\bar\chi)}$
946: (see Appendix~\ref{app:limber}).
947:
948: With all the additional contributions of the gravitational lensing and the
949: generalized Sachs-Wolfe effects in hand, the correlation of the intrinsic
950: fluctuation of sources is modeled using the linear bias model,
951: \bear
952: \label{eq:xint}
953: \xint&=&\langle\delta(\Vang_1,z_1)~\delta(\Vang_2,z_2)\rangle\\
954: &=&b_1~b_2
955: \int{d^3\bdv{k}\over(2\pi)^3}~e^{i\bdv{k}\cdot(\bdv{x_1-x_2})}~\Pm(k;z_1,z_2)
956: \nonumber \\
957: &=&b_1~b_2 \int_0^\infty{dk\over k}{k^3\over2\pi^2}~\Pm(k;z_1,z_2)~j_0(k\rtd),
958: \nonumber
959: \enar
960: where the constant linear bias factors $b_1$ and $b_2$ are the ratio of
961: the intrinsic source fluctuation to the underlying matter fluctuation at
962: $z_1$ and $z_2$.
963:
964: To complete our calculations of $\xiobs$,
965: we now compute the cross-correlation functions and power spectra between
966: the intrinsic source fluctuation and the fluctuations from the
967: gravitational lensing and the generalized Sachs-Wolfe effects. First, the
968: redshift-space distortion bias and the intrinsic source
969: fluctuation provide two cross-correlation functions
970: \bear
971: \label{eq:xidz}
972: \xi_{\delta\up{z}}&=&\langle\delta(\Vang_1,z_1)~\zdist(\Vang_2,z_2)\rangle\\
973: &=&b_1~f_2\int{d^3\bdv{k}\over(2\pi)^3}~e^{i\bdv{k}\cdot(\bdv{x_1-x_2})}
974: ~\Pm(k;z_1,z_2)
975: ~{k_z^2\over k^2} \nonumber \\
976: &=&b_1~f_2\int_0^\infty{dk\over k}~{k^3\over2\pi^2}~\Pm(k;z_1,z_2) \nonumber \\
977: &\times&\left[{1\over3}j_0(k\rtd)P_0(\gamma)-{2\over3}j_2(k\rtd)
978: P_2(\gamma)\right],
979: \nonumber
980: \enar
981: and similarly for
982: $\xi_{\up{z}\delta}=\langle\zdist(\Vang_1,z_1)~\delta(\Vang_2,z_2)\rangle$
983: with the two indices exchanged in Eq.~(\ref{eq:xidz}).
984: Combined with $\xizz$ in Eq.~(\ref{eq:pzz}), these two cross-correlation
985: functions constitute the standard redshift-space correlation function
986: \beeq
987: \xi_\up{z-dist}=\xint+\xizz+\xizd+\xidz
988: =\sum_{l=0,2,4}P_l(\gamma)~\xi_l(\rtd),
989: \label{eq:z-dist}
990: \eneq
991: which is often expressed in terms of the multipole components
992: \citep{HAMIL92,COFIWE94,HAMIL98}
993: \beeq
994: \xi_l=c_l(\beta_1,\beta_2)~b_1~b_2~i^l\int_0^\infty{dk\over k}~{k^3\over2\pi^2}
995: ~\Pm(k;z_1,z_2)~j_l(k\rtd),
996: \label{eq:zmul}
997: \eneq
998: with its coefficients
999: \beeq
1000: \label{eq:mulc}
1001: \left(\begin{array}{c}
1002: c_0\\[4pt]c_2\\[4pt]c_4
1003: \end{array}\right)=
1004: \left(\begin{array}{c}
1005: 1+{\beta_1+\beta_2\over3}+{\beta_1~\beta_2\over5}\\[4pt]
1006: {2\over3}(\beta_1+\beta_2)+{4\over7}\beta_1~\beta_2\\[4pt]
1007: {8\over35}\beta_1~\beta_2
1008: \end{array}\right),
1009: \eneq
1010: where $\beta=f/b$. Analogously, the redshift-space power spectrum is
1011: \bear
1012: \label{eq:pzf}
1013: P_\up{z-dist}&=&P_\up{int}+P_\up{z\delta}+P_\up{\delta z}+P_\up{zz}\\
1014: &=&\left[1+(\beta_1+\beta_2)~\mu_k^2+\beta_1~\beta_2~\mu_k^4\right]
1015: P_\up{int}(k) \nonumber \\
1016: &=&\sum_{l=0,2,4}P_l(\mu_k)~P^\up{z}_l(k), \nonumber
1017: \enar
1018: with the intrinsic source power spectrum $P_\up{int}(k)=b_1~b_2~\Pm(k;z_1,z_2)$,
1019: and its multipole components $P_l^\up{z}(k)$
1020: in Fourier space are related to the multipole components $\xi_l(\rtd)$
1021: in real space as
1022: \beeq
1023: P^\up{z}_l(k)=4\pi~i^l\int_0^\infty dx~x^2~\xi_l(x)~j_l(kx).
1024: \eneq
1025:
1026: \begin{figure*}
1027: \centerline{\psfig{file=f1.eps, width=7.0in}}
1028: \caption{Dissection of the observed two-point correlation function of galaxies.
1029: Solid, dotted, and long dashed lines represent correlation functions of the
1030: intrinsic source fluctuations $\xint/b^2$ and the redshift-space
1031: distortion bias $\xizz$, and their cross-correlation function
1032: $\xi_{\delta\up{z}}/b=\xi_{\up{z}\delta}/b$, respectively. Correlation
1033: functions of the magnification bias $\xiMB/(5p-2)^2$ and the evolution
1034: bias $\xievo/E^2$ are shown as short dashed and short dot-dashed lines.
1035: Note that while
1036: the galaxy bias factor $b$ and magnification bias factor $(5p-2)$ are
1037: of order unity, the evolution boost factor $E$ can be an order of magnitude
1038: larger
1039: (see Fig.~\ref{fig:evo}). The correlation functions are computed by using
1040: the linear ({\it thin}) and the nonlinear ({\it thick}) matter power spectra,
1041: and source galaxies are assumed to be at the same redshift indicated in the
1042: legend. The correlation functions of the intrinsic source fluctuations become
1043: negative at $\rtd\gtrsim128\hmpc$, where its absolute value is plotted.
1044: Panel~($d$) plots the cross-correlation function
1045: $\xidMB/b_1(5p_2-2)$ of the intrinsic fluctuation of source
1046: galaxies at $z_1=0.35$ and the magnification bias from source galaxies
1047: at $z_2=0.6$ as long dot-dashed lines.
1048: With large line-of-sight separation $600\hmpc$, only $\xiMB$ and
1049: $\xidMB$ that depend on projected separation $R$ rather than 3D
1050: separation $\rtd$ itself are appreciable, i.e.,
1051: $\xint\simeq\xi_{\delta\up{z}}\simeq\xi_{\up{z}\delta}\simeq\xizz\simeq
1052: \xievo\simeq0$.}
1053: \label{fig:corr}
1054: \end{figure*}
1055:
1056: Since the magnification bias arises from the matter fluctuations along the
1057: line-of-sight, it correlates with the intrinsic source fluctuation at
1058: lower redshift ($z_1<z_2$),
1059: \bear
1060: \label{eq:xidmb}
1061: \xidMB&=&\langle\delta(\Vang_1,z_1)~\dMB(\Vang_2,z_2)\rangle\\
1062: &=&b_1(5p_2-2)\left({3H_0^2\over2}\OM\right)
1063: {r(\chi_2-\chi_1)~r_1\over a_1~r_2}
1064: w_p\left[r_1\Delta\theta;z_1\right], \nonumber
1065: \enar
1066: but the correlation vanishes when the source is at higher redshift, i.e.,
1067: $\xi_{\up{mb}~\delta}=0$. The power spectrum is also related to
1068: the angular power spectrum of the cross-term
1069: \bear
1070: \label{eq:admb}
1071: \clcor{\delta~\up{mb}}&=&b_1~(5p_2-2)\left({3H_0^2\over2}\OM\right) \\
1072: &\times&{r(\chi_2-\chi_1)\over a_1~r_1~r_2}~
1073: \Pm\left[k_\perp={l\over r_1};z_1\right], \nonumber
1074: \enar
1075: as $P_{\delta~\up{mb}}=(2\pi)\delta^D(k_z)r^2_1
1076: C^{\delta~\up{mb}}_{l=k_\perp r_1}$.
1077: Finally, all the cross-correlations that involve $\devo$
1078: are zero, since $\devo$ is odd in the line-of-sight component $V$ of peculiar
1079: velocities and the universe has no preferred
1080: direction. Two remaining cross terms $\xi_\up{z~mb}$ and $\xi_\up{mb~z}$
1081: also vanish, since $\zdist$ is proportional to $k_z^2$ and the line-of-sight
1082: fluctuations are smoothed out in the Limber approximation.
1083:
1084: \begin{figure}
1085: \centerline{\psfig{file=f2.eps, width=3.5in}}
1086: \caption{Multipole components of the redshift-space correlation function
1087: $\xi_\up{z-dist}$ and the evolution bias $\xievo$ at $z=0.35$.
1088: We define multipole components of $\xievo$ as
1089: $\xievo=\xievo^0(\rtd)P_0(\gamma)+\xievo^2(\rtd)P_2(\gamma)$,
1090: in the same way
1091: multipole components of $\xi_\up{z-dist}$ is defined in Eq.~(\ref{eq:z-dist}).
1092: We assume that the galaxy bias factor is $b=2$ and the evolution boost factor
1093: is $E=100$ for a proper comparison.
1094: The correlation functions are computed by using the linear matter
1095: power spectrum only.
1096: $\xi_0(\rtd)$ becomes negative at $\rtd\gtrsim128\hmpc$, where its
1097: absolute value is plotted. The inset shows the correlation $\xint$
1098: of the intrinsic
1099: source fluctuations around the acoustic scale and it is related to the monopole
1100: by $\xi_0=\xint\cdot c_0$ in Eq.~(\ref{eq:zmul}), where $c_0=1.24$ in
1101: our fiducial model.}
1102: \label{fig:comp}
1103: \end{figure}
1104:
1105: \subsection{Comparison}
1106: \label{ssec:com}
1107: To compare the additional contributions to the observed correlation function
1108: $\xiobs$ and power spectrum $P_\up{obs}$,
1109: we consider near-future dark energy surveys that will
1110: target galaxies and quasars to measure their correlation function and power
1111: spectrum at high redshifts.
1112: For example, the baryonic oscillation spectroscopic survey (BOSS) will measure
1113: 1.5~million luminous red galaxies to determine the angular diameter distances
1114: at $z=0.35$ and~0.6, and use 160,000 quasars to measure the clustering of
1115: Lyman-$\alpha$ forests at $z=2.5$ \citep{WEBEET07,SCBLET07}.
1116: For the purpose of illustration, we show
1117: our calculations of galaxy two-point statistics at these redshifts.
1118: Here we adopt a flat $\Lambda$CDM universe: the cosmological parameters are
1119: the matter density $\OM h^2=0.137$, the baryon density $\OB h^2=0.0227$, the
1120: Hubble constant $h=0.70$, the spectral index $n_s=0.96$, the optical depth to
1121: the last scattering surface $\tau=0.084$, and the primordial
1122: curvature perturbation amplitude $\Delta_\zeta^2=2.457\times10^{-9}$
1123: at $k=0.002~\mpci$ (corresponding to the
1124: matter power spectrum normalization $\rms=0.817$), consistent with the
1125: recent results (e.g., \citep{TESTET06,SPBEET07,KODUET08}).
1126: The matter transfer function is computed by using
1127: {\scriptsize CMBFAST} \citep{SEZA96}.
1128:
1129:
1130: Figure~\ref{fig:corr} examines the separate contributions of the gravitational
1131: lensing and the generalized Sachs-Wolfe effects to the observed two-point
1132: correlation function of galaxies. We show the correlation functions of the
1133: intrinsic galaxy fluctuations ($\xint/b^2;~solid$) and the
1134: redshift-space distortion bias ($\xizz;~dotted$), and their cross-correlation
1135: function ($\xi_{\delta\up{z}}/b=\xi_{\up{z}\delta}/b;~long~dashed$).
1136: The correlation functions are computed by
1137: using the linear ($thin$) and the nonlinear
1138: ($thick$) matter power spectrum. Note that they only differ on small scales
1139: and the nonlinear effect decreases at high redshift as shown in
1140: Fig.~\ref{fig:corr}$a$ to Fig.~\ref{fig:corr}$c$, going from $z=0.35$ to
1141: $z=2.5$. The source galaxies are assumed to be at the same redshift
1142: $(z=z_1=z_2)$ shown in the figure legend, and thus 3D separation $\rtd$ is
1143: equal to 2D projected separation $R=r(\bar\chi)\Delta\theta$. However,
1144: two galaxy populations are separately placed at $z_1=0.35$ and $z_2=0.6$
1145: in Fig.~\ref{fig:corr}$d$, and the $x$-axis represents
1146: projected separation $R$, rather than 3D separation $\rtd$.
1147:
1148: The solid lines $\xint/b^2$ are identical to the matter correlation
1149: function $\xi_m$ and the linear bias factor $b$ is constant. However,
1150: the nonlinear evolution and galaxy formation process complicate
1151: the relation between galaxies and underlying matter fluctuations, and
1152: galaxy bias becomes scale-dependent on small scales, even when the nonlinear
1153: matter power spectrum is used (e.g., \citep{ZEZE05,YTWZKD06}).
1154: While we plot the correlation functions at $\rtd\simeq0.5-200\hmpc$ for
1155: completeness, the validity of our calculation is limited to the linear regime.
1156: The solid lines at $\rtd=151~\mpc$ ($=106\hmpc$)
1157: show prominent enhancement in the clustering amplitude,
1158: known as the baryonic acoustic peak \citep{PEYU70,SUZE70b}.
1159: The baryon-photon plasma in the early universe propagates as sound waves
1160: and these periodic oscillations in Fourier space translate into one peak in
1161: real space with its width deviating from a sharp delta function due to the
1162: termination of the harmonic series,
1163: determined by the horizon size at the cosmological recombination epoch.
1164: Note that the correlation function becomes negative at $\rtd\simeq128\hmpc$,
1165: beyond which we plot its absolute value.
1166:
1167: The correlation functions ($\xizz$, $\xidz$, and
1168: $\xizd$) of the redshift-space distortion bias have the overall
1169: shape similar to $\xint$. However, since
1170: $\xizz$, $\xidz$, and $\xizd$
1171: in Eqs.~(\ref{eq:pzz}) and~(\ref{eq:xidz})
1172: have additional functional dependence on spherical Bessel functions $j_2(x)$
1173: and $j_4(x)$ compared to $\xint$ in Eq.~(\ref{eq:xint}),
1174: it puts more weight on higher $k$ and hence
1175: the nonlinear effects persist up to
1176: $\rtd\simeq10\hmpc$ in Fig.~\ref{fig:corr}$a$,
1177: larger than $3\hmpc$ for $\xint$. However,
1178: the incoherent superposition of the additional Bessel
1179: functions washes out the acoustic peak in the correlation functions of the
1180: redshift-space distortion bias,
1181: leaving little structure in $\xizz$, $\xidz$, and $\xizd$ at the acoustic
1182: scale. Since the observed correlation function $\xiobs$ is
1183: the sum of all the contributions and it is hard in practice to separate
1184: each contribution from one another, it may
1185: look as if $\xint$ is swamped by $\xizz$, $\xidz$ and $\xizd$
1186: at the acoustic scale, but note that we plot $\xint/b^2$ and
1187: $\xidz/b$: the linear bias factor
1188: of luminous red galaxies is $b_0\simeq1.5-2.0$
1189: \citep{EIBLET05,TEEIST06,PASCET07,BLCOET07}.
1190: Assuming that galaxies
1191: have no velocity bias $\bdv{v}_g=\bdv{v}$, the linear bias factor
1192: at high redshift is $b(z)-1=(b_0-1)/D(z)$, sufficient for $\xint$
1193: to show its structure, when combined with $\xizz$, $\xidz$, and $\xizd$, yet
1194: the plot without $b$ captures the main structure of the correlation
1195: functions, since the linear bias factor is still of order unity.
1196:
1197: Note that since the source galaxies are at the same redshift
1198: in Fig.~\ref{fig:corr}$a$ to Fig.~\ref{fig:corr}$c$, the cosine angle of the
1199: comoving separation is $\gamma=(\chi_2-\chi_1)/\rtd=0$, i.e., the
1200: redshift-space correlation function (the sum of the solid, dotted, and dashed
1201: lines) is $\xi_\up{z-dist}=\xint+\xizz+\xizd+\xidz=
1202: \xi_0-(1/2)~\xi_2+(3/8)~\xi_4$,
1203: different from the angle-averaged
1204: (monopole) correlation $\xi_0$ often used in the literature
1205: \citep{EIZEET05,ROSHET08}.
1206: Figure~\ref{fig:comp} illustrates the multipole components of the
1207: redshift-space correlation function at $z=0.35$. The monopole ($solid$) is
1208: identical to $\xint$ in shape but differs in normalization by
1209: $\xi_0=\xint\cdot c_0$ with the multipole coefficient $c_0$ in
1210: Eq.~(\ref{eq:mulc}). The quadrupole $\xi_2$ ($dotted$) is negative by the
1211: sign convention and the hexadecapole $\xi_4$ ($dashed$) is positive in the
1212: figure, while the monopole $\xi_0$ changes its sign as $\xint$ changes at
1213: $\rtd\gtrsim128\hmpc$ (see the inset). As noted before, the spherical Bessel
1214: functions $j_2(k\rtd)$ and $j_4(k\rtd)$ in $\xi_2$ and $\xi_4$ peak at scales
1215: different from the typical scale $k\sim1/\rtd$ for $\xi_0$, and thus
1216: the acoustic structure seen in $\xi_0$ is smoothed out in $\xi_2$ and $\xi_4$.
1217:
1218: In practice, galaxy
1219: redshift surveys have a narrow but nonzero radial window function and galaxy
1220: pairs in the same redshift bin often have the line-of-sight separation
1221: comparable to the transverse separation, i.e., $\gamma\ne0$.
1222: The angular dependence of the redshift-space correlation function,
1223: therefore, complicates the interpretation of its
1224: measurements, which are further plagued by low
1225: signal-to-noise ratios in estimates of $\xi_2$ and $\xi_4$. While the
1226: monopole $\xi_0$ can be used to ease the
1227: theoretical and/or observational challenge, full analysis of the anisotropic
1228: structure of
1229: $\xi_\up{z-dist}$ could in principle bring more information than $\xi_0$
1230: measurements (see \citep{OKMAET08,GACAHU08} for recent analysis).
1231: We analyze the full anisotropic structure of the observed correlation function
1232: $\xiobs$ below. On small scales,
1233: virial motions of galaxies result in additional anisotropic
1234: structure in $\xi_\up{z-dist}$, known as the Finger-of-God (FoG) effect.
1235: Note that since this effect involves galaxy motions in nonlinear objects,
1236: it is not considered in our calculation and
1237: linear theory provides an inaccurate description of the FoG effect:
1238: while a simple dispersion model \citep{BAPEHE96} is often adopted to extract
1239: additional information contained in the anisotropic structure,
1240: it is demonstrated \citep{SCOCC04} that this model leads to an unphysical
1241: distribution of pairwise velocities. However, this difficulty could
1242: be tackled by recent approach based on modeling nonlinear galaxy bias
1243: in redshift-space
1244: \citep{SELJA01,WHITE01,TINKE07}.
1245:
1246: The short dashed and short dot-dashed lines in Fig.~\ref{fig:corr}
1247: show the correlation functions
1248: of the magnification bias $\xiMB/(5p-2)^2$ and the evolution bias
1249: $\xievo/E^2$, respectively. The magnification bias $\xiMB$ is typically
1250: smaller than $\xint$ by the ratio of the transverse correlation scale
1251: $1/k_\perp$ to the Hubble distance $1/H$, and the magnification bias factor
1252: is of order unity, $(5p-2)=-1.0\sim2.0$ \citep{SCMEET05} for galaxies and
1253: quasars, while it can be further suppressed by the source effect canceling the
1254: volume effect $(5p-2)\simeq0$. Since $\xiMB$ in Eq.~(\ref{eq:xiMB})
1255: is proportional to the projected correlation function $w_p$, its overall
1256: shape is similar to $\xint$ but $\xiMB$ is positive due to
1257: projection of $\xint$.
1258: As the source population is located at higher redshift,
1259: longer line-of-sight distance increases the gravitational lensing effect
1260: and $\xiMB$ increases in redshift,
1261: as opposed to $\xint\propto D^2\propto 1/(1+z)^2$
1262: decreasing in redshift. For example,
1263: $\xiMB$ at $z=2.5$ in Fig.~\ref{fig:corr}$c$ can grow up to a few percent
1264: of $\xint$ at the acoustic scale \citep{HUGALO07,VADOET07}.
1265:
1266: \begin{figure}
1267: \centerline{\psfig{file=f3.eps, width=3.5in}}
1268: \caption{Redshift distribution of source populations and boost factor
1269: $E(z)$ of the evolution bias. Assuming the standard functional form
1270: $\bar n(z)dz\propto z^\alpha\exp\left[-(z/z_0)^\beta\right]dz$, the redshift
1271: distribution of galaxies ($dashed$) and quasars ($solid$) are shown in the
1272: upper panel with $(\alpha,\beta,z_0)=(4,4,0.4)$ for galaxies and $(3,13,2)$
1273: for quasars,
1274: respectively \citep{SCMEET05,JIFAET06,PASCET07}. The bottom panel shows the
1275: evolution bias boost factor computed by using Eq.~(\ref{eq:eboo}).
1276: The number density of sources changes exponentially beyond the mean
1277: redshift and the evolution bias is substantially enhanced
1278: in proportion to $E(z)$.}
1279: \label{fig:evo}
1280: \end{figure}
1281:
1282: The evolution bias $\devo$ is often ignored in the literature compared to the
1283: redshift-space distortion bias $\zdist$, since
1284: $\devo\propto\varepsilon\simeq V$ and $V\ll\zdist$.
1285: However, the evolution bias can be
1286: significantly enhanced when the mean number density of sources changes
1287: rapidly in redshift. To estimate the
1288: evolution boost factor $E(z)$ in Eq.~(\ref{eq:eboo}), we assume
1289: the standard functional form of a redshift distribution
1290: $\bar n(z)dz\propto z^\alpha\exp\left[-(z/z_0)^\beta\right]dz$ and take
1291: two source populations as illustrative examples: galaxies and quasars
1292: characterized by $(\alpha,\beta,z_0)=(4,4,0.4)$ and $(3,13,2)$, respectively
1293: \citep{SCMEET05,JIFAET06,PASCET07}. While the bright samples of luminous red
1294: galaxies in the BOSS will have a redshift distribution flatter than the
1295: assumed here, the faint samples with larger number density and volume
1296: will have a non-flat redshift distribution \citep{EIANET01,ZEEIET05}.
1297: The clustering of Lyman-$\alpha$ forests at $z=2.5$ will
1298: be measured by the spectrum of quasars
1299: at $z>2.5$, not by quasars themselves at $z=2.5$.
1300: However, we simply assume that $\xiobs$ is measured from the galaxy samples
1301: at $z=0.35$ and $z=0.6$, and from the quasar samples at $z=2.5$.
1302:
1303: The upper panel of Fig.~\ref{fig:evo} illustrates the redshift distribution
1304: of the galaxy ($dashed$) and quasar ($solid$) samples, with its peak at
1305: $z=0.4$ and~1.8, and the bottom panel shows the evolution boost factor $E(z)$
1306: of each sample. For the assumed redshift distribution, the evolution boost
1307: factor is typically a factor $\sim10$, and it vanishes at the peak redshift.
1308: However, a sharp decline in the mean number density of source populations
1309: beyond the mean redshift makes the evolution bias $\devo$ sensitive to the
1310: change in observed redshift $\zobs$ due the generalized Sachs-Wolfe effect,
1311: and $E(z)$ can be further enhanced by another factor of ten.
1312: With significant boost of
1313: $E^2(z)\simeq100-10000$, the correlation $\xievo$
1314: of the evolution bias
1315: should be given a proper consideration, especially when the mean
1316: number of the source population changes rapidly.
1317: Note that the evolution bias
1318: $\xievo/E^2$ ({\it short dot-dashed}) in Fig.~\ref{fig:corr}$a$ is
1319: comparable to the magnification bias $\xiMB/(5p-2)^2$ ({\it short dashed})
1320: and is larger at the acoustic peak scale, and the evolution boost factor $E(z)$
1321: can be significantly larger than the magnification bias factor $(5p-2)$.
1322: Therefore, it is of particular importance
1323: to select samples of source populations that
1324: have relatively flat redshift distribution in number density ($E\simeq10$),
1325: and to limit the redshift range of measurements below the peak redshift.
1326: The short and the long dot-dashed lines in Fig.~\ref{fig:comp} show the
1327: multipole components $\xievo^l$ of the evolution bias, defined as
1328: $\xievo=\sum_{l=0,2}\xievo^l(\rtd)P_l(\gamma)$. Both components are
1329: smooth and change little over $\rtd=1-200\hmpc$.
1330: With $\xievo\propto H^2a^2D^2\propto1/(1+z)$,
1331: it decreases slowly in redshift,
1332: and the nonlinear effect is relatively small compared to $\xint$,
1333: since less weight is given to short wavelength modes.
1334:
1335: \begin{figure*}
1336: \centerline{\psfig{file=f4.eps, width=7.0in}}
1337: \caption{(color online)
1338: Anisotropic structure of the observed correlation function $\xiobs$
1339: at $z=0.35$. We plot the correlation function $\xint$
1340: of the intrinsic source fluctuations
1341: in Panel~($a$), and show the individual effects of the magnification bias
1342: $\xint+\xiMB+\xidMB$ ({\it Panel~b}), the evolution bias $\xint+\xievo$
1343: ({\it Panel~c}), and the redshift-space distortion bias
1344: $\xi_\up{z-dist}=\xint+\xizz+\xizd+\xidz$
1345: ({\it Panel~d}) on the anisotropic structure.
1346: The observed correlation function $\xiobs$,
1347: the sum of all the correlation functions, is shown in Panel~($e$).
1348: We assume that the galaxy bias factor
1349: is $b=2$, the magnification bias factor is $(5p-2)=2$, and the evolution
1350: boost factor is $E=100$. The color maps are linearly proportional to the value
1351: of correlation function $\xi$ in each panel
1352: at $\xi<4\times10^{-4}$ and are logarithmically proportional
1353: to $\xi$ at $\xi>4\times10^{-4}$.
1354: The solid contours are logarithmically spaced at
1355: $\xi\geq1.5\times10^{-3}$ and their thickness increases with the value of $\xi$
1356: (contour values are labeled in the color bar), while the thickest solid
1357: curves represent the $\xi=0$ contour. The dot-dashed and dotted curves
1358: represent the contours with $\xi=-4.5\times10^{-4}$ and $-9.0\times10^{-4}$,
1359: respectively. The acoustic scale in $\xint$ ($\rtd=106\hmpc$)
1360: is shown as dashed lines in each panel for reference.
1361: Note that two regions are underrepresented in the color maps, as
1362: the region with $\xi>1.0$ is highly concentrated at $\rtd\ll20\hmpc$
1363: and there is no distinctive feature in the anisotropic structure around
1364: the region with $\xi<0$.}
1365: \label{fig:aniso}
1366: \end{figure*}
1367:
1368: Now recall that there
1369: are four terms of $\zdist$ in Eq.~(\ref{eq:zfive}) that are ignored in our
1370: calculation, and they are comparable to $\varepsilon\simeq V$, albeit smaller
1371: than the dominant term in
1372: $\zdist\simeq-{1+z\over H}{\partial V\over\partial\chi}$.
1373: While there is no additional boost factor like $E(z)$ in $\devo$,
1374: their contributions
1375: to $\xiobs$ are typically of order $\xievo/E^2$ and are as large as
1376: $\xiMB$ at $z=0.35$ in Fig.~\ref{fig:corr}$a$.
1377: Though $\xiMB$ is larger at higher redshift, their impact on $\xiobs$
1378: also increases in redshift: approximately at the sub-percent level for each
1379: contribution at $z=2.5$. This level of accuracy would be appropriate given the
1380: statistical errors present in current samples, but further calculations of
1381: the ignored terms may be needed in future surveys.
1382:
1383: In Fig.~\ref{fig:corr}$d$, we consider the correlation functions of two
1384: source populations, separately located
1385: at $z_1=0.35$ and $z_2=0.6$ as a function of 2D projected
1386: separation $R=r(\bar\chi)\Delta\theta$.
1387: Note that given a large line-of-sight separation $\sim600\hmpc$
1388: between $z_1$ and $z_2$, all the correlation functions that depend on 3D
1389: comoving separation $\rtd$ are nearly zero, i.e.,
1390: $\xint\simeq\xi_{\delta\up{z}}\simeq\xi_{\up{z}\delta}\simeq\xizz\simeq
1391: \xievo\simeq0$. The two non-vanishing contributions in
1392: Fig.~\ref{fig:corr}$d$ are the auto-correlation of the magnification bias
1393: $\xiMB/(5p_1-2)(5p_2-2)$ ({\it short dashed}) and the cross-correlation of the
1394: intrinsic source fluctuation and the magnification bias
1395: $\xidMB/b_1(5p_2-2)$ ({\it long dot-dashed}) that depend on the
1396: projected separation, rather than 3D separation itself. Note that the
1397: cross-correlation in Fig.~\ref{fig:corr}$a$ to Fig.~\ref{fig:corr}$c$ is
1398: identically zero: $\xidMB=0$ at $z_1=z_2$ with the Limber
1399: approximation we adopted here, but it is in general smaller than $\xiMB$
1400: unless $z_1\neq z_2$ (see \citep{VADOET07} for a somewhat different derivation).
1401: Note that while both $\xiMB$ and $\xidMB$ in
1402: Eqs.~(\ref{eq:xiMB}) and~(\ref{eq:xidmb}) depend on the projected separation
1403: via $w_p$, $\xidMB$ has additional linear dependence on the comoving
1404: line-of-sight separation $\chi_2-\chi_1=\Delta\chi$, and it increases with
1405: $\Delta\chi$, as opposed to $\xiMB$ with little dependence on $\Delta\chi$.
1406: With the large $\Delta\chi\sim600\hmpc$ in Fig.~\ref{fig:corr}$d$,
1407: $\xidMB$ is substantially larger than $\xiMB$.
1408:
1409: Figure~\ref{fig:aniso} examines the anisotropic structure of the observed
1410: correlation function $\xiobs$, evaluated at $\bar z=0.35$.
1411: The $x$-axis represents the transverse separation $R=r(\bar\chi)\Delta\theta$
1412: and the $y$-axis represents the line-of-sight separation
1413: $\Delta\chi=\chi_2-\chi_1$ with fixed $\bar z=(z_1+z_2)/2=0.35$,
1414: $\bar\chi=(\chi_1+\chi_2)/2=980\hmpc$, and $\chi(z_1)\leq\chi(z_2)$.
1415: The color maps are linearly proportional to the value of the correlation
1416: function $\xi$ plotted in each panel below the adopted threshold
1417: $\xi=4\times10^{-4}$, and they are logarithmically proportional
1418: to $\xi$ above the threshold. The solid contours are also logarithmically
1419: spaced with increasing thickness at $\xi\geq1.5\times10^{-3}$ to emphasize
1420: the structure shown as the color maps, and their contour values are labeled in
1421: the color bar. The thickest solid contours separate the regions with $\xi>0$
1422: from those with $\xi<0$, and the dot-dashed and dotted curves represent the
1423: contours with
1424: $\xi=-4.5\times10^{-4}$ and $-9.0\times10^{-4}$, respectively. For reference,
1425: we also plot the acoustic scale in $\xint$ ($\rtd=106\hmpc$)
1426: as dashed lines in each panel.
1427: In Fig.~\ref{fig:aniso}$a$, we plot the correlation function $\xint$ of the
1428: intrinsic source fluctuations, assuming the galaxy bias factor $b=2$.
1429: As the rings of the concentric contours show, $\xint$ is spherically
1430: symmetric and depends only on 3D separation $\rtd$.
1431: The acoustic peak shows its structure as a circular ring
1432: at $\rtd=106\hmpc$ ({\it dashed}) and beyond $\rtd\sim128\hmpc$ $\xint$
1433: becomes negative without further distinctive feature in its structure.
1434:
1435:
1436: The gravitational lensing effects, the magnification bias $\xiMB$ and
1437: its cross-correlation $\xidMB$, break the spherical symmetry in
1438: $\xint$, and its impact on the anisotropic structure is shown in
1439: Fig.~\ref{fig:aniso}$b$, assuming the magnification bias factor $(5p-2)=2$.
1440: $\xiMB$ depends on the line-of-sight separation $\Delta\chi$ only
1441: through $\chi_1$ and $\chi_2$ in Eq.~(\ref{eq:xiMB}), and
1442: $\Delta\chi$ is small compared to the line-of-sight distance, i.e.,
1443: $\Delta\chi\ll\chi_1\simeq\chi_2$. Thus $\xiMB$ is virtually
1444: independent of $\Delta\chi$ and is just a function of transverse separation
1445: $R$, decreasing with increasing $R$.
1446: As is seen in Fig.~\ref{fig:corr}$a$, $\xiMB$ is
1447: in general orders of magnitude smaller than $\xint$ at $z=0.35$,
1448: but $\xint$ becomes negative at large $\rtd$ and smaller than $\xiMB$, e.g.,
1449: $\xint=-8.4\times10^{-4}<\xiMB=4.8\times10^{-6}$ at $R=10.0\hmpc$ and
1450: $\Delta\chi=140\hmpc$ ($\rtd=140.3\hmpc$). Since $\xiMB$ changes slowly with
1451: $R$, the demarcation curve between the regions with $\xiMB>\xint$ and
1452: $\xiMB<\xint$ roughly corresponds to the $\xint=0$ contour
1453: ({\it thickest solid}) in
1454: Fig.~\ref{fig:aniso}$a$. However, since $|\xint|>\xiMB$ in general
1455: except at a narrow strip around the $\xint=0$ contour, the impact of
1456: $\xiMB$ on the anisotropic structure is negligible at $\bar z=0.35$.
1457: Note that the impact of $\xiMB$ is substantially enhanced at higher redshift,
1458: where longer line-of-sight distance results in more fluctuations
1459: and the clustering amplitude of $\xint$ is lower.
1460:
1461: While both $\xiMB$ and $\xidMB$ are proportional to $w_p$, $\xidMB$ depends
1462: on $w_p$ itself, rather than the integral of $w_p$ along the line-of-sight,
1463: on which $\xiMB$ depends:
1464: $\xidMB$ becomes negative at large transverse separation
1465: $R\simeq110\hmpc$ as $\xint$ becomes negative at large 3D separation $\rtd$.
1466: Note that for an observable angular separation $\Delta\theta$, the transverse
1467: separation $R=r(\bar\chi)\Delta\theta$ in the figure is slightly different
1468: from $r(\chi_1)\Delta\theta$ for $w_p$ in Eq.~(\ref{eq:xidmb}), and this
1469: difference tilts otherwise a vertical line with $\xidMB=0$ at $R\simeq110\hmpc$
1470: toward larger $R$ at large $\Delta\chi$ ($\xidMB=0$ at $R\simeq128\hmpc$
1471: and $\Delta\chi=140\hmpc$), because $\Delta\chi=\chi_2-\chi_1$ and
1472: $z_1\le z_2$ with fixed $\bar z=(z_1+z_2)/2=0.35$.
1473: Furthermore, since $\xidMB$ linearly
1474: increases with $\Delta\chi$ (hence it vanishes at $\Delta\chi=0$), the
1475: absolute value of $\xidMB$ is larger than $\xiMB$ except at the regions
1476: at $\Delta\chi=0$ and around the nearly vertical strip with $\xidMB=0$, and
1477: it is also comparable to $\xint$ at large $\Delta\chi$ and small $R$, e.g.,
1478: $\xidMB=4.3\times10^{-4}$ at $R=10\hmpc$ and $\Delta\chi=140\hmpc$.
1479: Therefore, when $\xint$ is combined with $\xiMB$ and $\xidMB$ as
1480: shown in Fig.~\ref{fig:aniso}$b$,
1481: $\xiMB$ has little impact but $\xidMB$ distorts the symmetric contours of
1482: $\xint$ at large $\Delta\chi$ and small $R$. At higher redshift, the amplitude
1483: of $\xidMB$ decreases with that of $\xint$, and $\xiMB$ becomes a more
1484: dominant contribution than $\xidMB$.
1485: Note that the amplitude at the acoustic scale ($dashed$) is not significantly
1486: altered by
1487: the gravitational lensing effects, even along the line-of-sight
1488: direction at $\bar z=0.35$.
1489:
1490: The impact of the evolution bias $\xievo$ is shown in Fig.~\ref{fig:aniso}$c$,
1491: where we assume the evolution boost factor $E=100$.
1492: While $\xi^0_\up{evo}$ and $\xi^2_\up{evo}$
1493: change slowly with separation, they have the opposite sign as shown in
1494: Fig.~\ref{fig:comp}. Therefore, they tend to cancel out along
1495: the line-of-sight direction ($\gamma=1$), reducing the amplitude of $\xievo$,
1496: and the largest contribution of $\xievo$ arises along the
1497: transverse direction ($\gamma=0$), where the absolute values of the
1498: monopole $\xievo^0$ and the quadrupole $\xievo^2$ add up. For example,
1499: $\xievo=1.6\times10^{-3}>\xint=-8.4\times10^{-4}$ at $R=140\hmpc$ and
1500: $\Delta\chi=10\hmpc$. As noted in Fig.~\ref{fig:corr}$a$, the amplitude of
1501: $\xievo$ with $E=100$ is smaller than $\xint$ at $\rtd<50\hmpc$ and
1502: its impact is appreciable only at $\rtd\geq100\hmpc$
1503: along the transverse direction.
1504: As the angular separation becomes small with fixed 3D separation, the
1505: impact of $\xievo$ decreases, because the second
1506: order Legendre polynomial is a monotonic function of angle. Note that
1507: compared to $\xint$, $\xievo$ changes slowly in redshift and its impact is
1508: larger at higher redshift for a fixed $E$. The overall shape of the acoustic
1509: scale ($dashed$) also remains unaffected by the evolution bias.
1510:
1511: \begin{figure}
1512: \centerline{\psfig{file=f5.eps, width=3.5in}}
1513: \caption{Dissection of the observed galaxy power spectrum $P_\up{obs}$.
1514: Solid and dashed
1515: lines represent power spectra of the intrinsic source fluctuations
1516: $P_\up{int}/b^2$ and the evolution bias $P_\up{evo}/E^2$. Power spectrum
1517: of the magnification bias $P_\up{mb}/(5p-2)^2$ is shown as dotted
1518: lines for a survey window of width $200\hmpc$ (see the text).
1519: Note that we omit power spectra of the
1520: redshift-space distortion bias $P_\up{zz}$, $P_{\delta\up{z}}$, and
1521: $P_{\up{z}\delta}$, since they have the same shape as $P_\up{int}$
1522: for the line-of-sight component of the
1523: wavenumber $k_z=k$ up to numerical factors of order unity. Thin and thick
1524: lines represent power spectra computed by using the linear and nonlinear
1525: matter power spectrum. {\it Left panel:} power spectra are computed at two
1526: different redshifts, only the power spectrum of the magnification bias
1527: increases at higher $z$, while $P_\up{int}$ and $P_\up{evo}$ decrease.
1528: {\it Right panel:} cross-power spectra are computed
1529: for two galaxy populations at $z_1=0.35$ and $z_2=0.6$, and dot-dashed
1530: lines show the cross-power spectrum of the intrinsic source fluctuation
1531: at $z_1$ and the magnification bias of source galaxies at $z_2$.
1532: Large line-of-sight separation $\sim600\hmpc$ corresponds to
1533: $k_z\simeq0.002\hmpci$ and it has little impact on power spectrum
1534: ($k\simeq k_\perp\gg k_z$).}
1535: \label{fig:pow}
1536: \end{figure}
1537:
1538: Figure~\ref{fig:aniso}$d$ examines the redshift-space correlation function
1539: $\xi_\up{z-dist}=\xint+\xizz+\xidz+\xizd$. Though its angular structure is
1540: similar to $\xievo$, it differs in two aspects:
1541: $\xi_\up{z-dist}$ has the additional hexadecapole $\xi_4$,
1542: and the quadrupole $\xi_2$ becomes dominant over the monopole $\xi_0$ at
1543: $\rtd\simeq50\hmpc$ smaller than $\rtd\simeq200\hmpc$, where
1544: $\xievo^0\simeq\xievo^2$, shown in Fig.~\ref{fig:comp}. Therefore, the angular
1545: structure changes more dramatically than that seen in Fig.~\ref{fig:aniso}$c$.
1546: Note that while the fourth
1547: order Legendre polynomial is not a monotonic function
1548: of angle, the hexadecapole $\xi_4$ is generally smaller than $\xi_0$ and
1549: $\xi_2$, and its contribution is minor.
1550: The contours exhibits the well known Kaiser effect \citep{KAISE87}
1551: that coherent infall toward the overdense regions squashes the clustering
1552: amplitude and the underdense regions inflate along the line-of-sight.
1553: The large region with negative values at $\Delta\chi\simeq50-100\hmpc$ and
1554: $R\leq60\hmpc$ is the characteristic feature of this effect, largely due to
1555: the negative quadrupole $\xi_2$, and this structure has been recently
1556: measured with high signal-to-noise ratio \citep{OKMAET08,GACAHU08}. Even
1557: along the line-of-sight direction, the monopole $\xi_0$ briefly takes over the
1558: negative quadrupole around the acoustic scale, because the clustering
1559: amplitude is enhanced. Furthermore, while the clustering amplitude increases
1560: with angle at the acoustic scale ($dashed$), its structure manifests itself
1561: as ridges \citep{MATSU04}.
1562:
1563: Figure~\ref{fig:aniso}$e$ puts together our discussion,
1564: plotting the observed correlation function $\xiobs$ at $\bar z=0.35$,
1565: the sum of all the correlation functions shown in each
1566: panel. The redshift-space distortion affects the anisotropic structure by far
1567: the most among the other effects considered here. The gravitational
1568: lensing effects, mostly from $\xidMB$ at low redshift, become important
1569: only at a small transverse but large line-of-sight separation. With the
1570: same dependence on quadrupole, $\xievo$ follows the similar angular pattern
1571: of $\xi_\up{z-dist}$, boosting their contributions along the transverse
1572: direction, but its sole impact shows up at $R\geq110\hmpc$ due to the lower
1573: amplitude. The clustering amplitude of $\xiobs$ at the acoustic scale
1574: ($dashed$) is
1575: no longer constant, nor a monotonic function of angle.
1576: The peak location, imprinted in $\xint$ with local enhancement in clustering
1577: amplitude, remains largely
1578: unaffected as the gravitational lensing and generalized
1579: Sachs-Wolfe effects distorts the anisotropic structure. However, note that
1580: it is beyond our scope of the current investigation to what accuracy the
1581: acoustic peak remains unaffected by these effects (see, e.g.,
1582: \citep{EIWH04,SEEI05,EISEWH07,VADOET07,GUBESM07,HUGALO07,CRSC08,LOHUGA08}
1583: for recent work on the robustness of the baryonic acoustic peak).
1584:
1585: Figure~\ref{fig:pow} shows the equivalent dissection of the contributions
1586: of the gravitational lensing and the generalized Sachs-Wolfe effects to the
1587: observed galaxy power spectrum $P_\up{obs}$
1588: in Fourier space. The solid, dashed, and dotted
1589: lines represent $P_\up{int}/b^2$, $P_\up{evo}/E^2$, and $P_\up{mb}/(5p-2)^2$,
1590: respectively. The power spectra are also computed by using the linear ($thin$)
1591: and the nonlinear ($thick$) matter power spectrum as in Fig.~\ref{fig:corr}.
1592: In Fig.~\ref{fig:pow}$a$, the source galaxies are assumed to be
1593: at the same redshift ($z=z_1=z_2$), and the two sets of lines
1594: show the power spectra at
1595: $z=0.35$ and $z=2.5$, which decrease in redshift except that
1596: $P_\up{mb}$ increases as we have seen $\xiMB$ increase in redshift.
1597: The power spectrum of the intrinsic source fluctuations
1598: $P_\up{int}/b^2$ ($solid$)
1599: exhibits two characteristic scales in its structure: a series of the acoustic
1600: oscillations starting at $k\simeq 0.085\hmpci$, and the turnover in the
1601: overall shape at $k\simeq0.015\hmpci$ imprinted by
1602: the horizon size at the matter-radiation equality $z=3300$.
1603:
1604: For simplicity, the wavenumber is set equal to the line-of-sight
1605: direction $k=k_z$ for plotting $P_\up{evo}$ and to the transverse direction
1606: $k=k_\perp$ for plotting $P_\up{mb}$. The power spectrum of the evolution bias
1607: $P_\up{evo}/E^2$ ($dashed$) is typically many orders-of-magnitude smaller than
1608: $P_\up{int}/b^2$ at $k\geq0.03\hmpci$ in Fourier space, but its contribution
1609: can be at the few percent level of $P_\up{int}$ at the acoustic scale and
1610: comparable to $P_\up{int}$ at the matter-radiation equality scale,
1611: with the evolution boost factor $E^2\simeq10000$.
1612: Note that the power spectra of the redshift-distortion bias
1613: $P_\up{zz}$, $P_{\delta\up{z}}$, and $P_{\up{z}\delta}$ are omitted in the figure,
1614: because they have the same shape as
1615: $P_\up{int}$ up to numerical factors of order unity when $k=k_z$.
1616:
1617: To plot the power spectrum of the magnification bias $P_\up{mb}/(5p-2)^2$
1618: ($dotted$),
1619: we replace $(2\pi)\delta^D(k_z)$ by a flat window function of width
1620: $200\hmpc$, typical value in redshift surveys, hence the dotted line de facto
1621: delineates the angular power spectrum of the magnification bias
1622: $(5p-2)^2~C^{\kappa\kappa}_{l=k_\perp r(\bar\chi)}$ with a dimensional
1623: coefficient $r^2(\bar\chi)\times(200\hmpc)$ (see Appendix~\ref{app:limber}).
1624: While the magnification bias is negligible at $z=0.35$, its effect increases
1625: with larger line-of-sight distance at higher redshift, amounting to a few
1626: percent at the acoustic scale and larger at the matter-radiation equality
1627: scale at $z=2.5$. However, note that even with relatively
1628: large contributions to $P_\up{obs}$, the shift in the peak positions can
1629: be at the sub-percent level or smaller \citep{VADOET07,HUGALO07}.
1630:
1631: Figure~\ref{fig:pow}$b$ plots the cross power spectra of two source
1632: populations at $z_1=0.35$ and $z_2=0.6$. As opposed to the correlation
1633: functions shown in Fig.~\ref{fig:corr}$d$, all the power spectra that depend
1634: on 3D wavenumber remains virtually unaffected by the large line-of-sight
1635: separation $\sim600\hmpc$, because it corresponds to very small wavenumber
1636: $k_z\simeq0.002\hmpci$ and $k\simeq k_\perp\gg k_z$. The dot-dashed lines
1637: show the cross power spectrum of the intrinsic source fluctuation and the
1638: magnification bias $P_{\delta~\up{mb}}$. Since it is proportional to the
1639: line-of-sight separation, its contribution is larger than $P_\up{mb}$ in
1640: Fig.~\ref{fig:pow}$b$, but it is absent in Fig.~\ref{fig:pow}$a$.
1641:
1642: The anisotropic structure of the observed power spectrum $P_\up{obs}$ has been
1643: well studied with main focus on the effect of the redshift-space distortion
1644: bias \citep{KAISE87}, and the redshift-space power spectrum $P_\up{z-dist}$
1645: in Eq.~(\ref{eq:pzf}) has the multipole components that are identical in shape
1646: but only differ in normalization. The evolution bias results in the similar
1647: angular pattern: two multipole components that share its shape with
1648: $P_\up{int}$ with different normalization. However, note that since the
1649: magnification bias and its cross term are intrinsically 2D quantities,
1650: their impact on the anisotropic structure of $P_\up{obs}$ is small, even
1651: with realistic survey window functions \citep{HUGALO08}.
1652:
1653: \section{Discussion}
1654: \label{sec:discuss}
1655: Galaxy two-point statistics, correlation function in real space and power
1656: spectrum in Fourier space, have been extensively used in cosmology
1657: to characterize the underlying matter fluctuations.
1658: We have presented a coherent theoretical framework based on the linearized
1659: Friedmann-Lema{\^\i}tre-Robertson-Walker (FLRW) metric for computing the
1660: gravitational lensing and the generalized Sachs-Wolfe effects. Within this
1661: framework, the metric perturbations are sourced by the underlying matter
1662: fluctuations, and they naturally give rise to perturbations in the observable
1663: redshift of source galaxies and their angular position on the sky. The time
1664: component of the photon geodesic equations can be used to show the former,
1665: the generalized Sachs-Wolfe effect \citep{SAWO67} that generalizes the
1666: standard redshift-space distortion by peculiar velocities in a cosmological
1667: context, including the Sachs-Wolfe and the integrated Sachs-Wolfe effects.
1668: The spatial components of the photon geodesic equations can be used to
1669: derive the latter, the gravitational lensing effect that includes the weak
1670: lensing distortion, magnification, and time delay effects. This unified
1671: treatment provides a complete description of the relation between
1672: these seemingly different effects and the underlying matter fluctuations.
1673:
1674: Furthermore, it becomes transparent in this treatment how the gravitational
1675: lensing and the generalized Sachs-Wolfe effects affect the observed
1676: fluctuation field of source galaxies. To the linear order in perturbations,
1677: we have computed all the additional contributions to the intrinsic source
1678: fluctuation, arising from the gravitational lensing and the generalized
1679: Sachs-Wolfe effects. We can gain more insight on the impact of these effects
1680: by separating them as two physically distinct origins:
1681: the volume and the source effects.
1682: The former effect that involves the change of volume is independent of
1683: source galaxy populations and hence regardless thereof the volume effect
1684: is always present in galaxy two-point statistics. By contraries,
1685: the latter effect depends on the intrinsic properties of source galaxy
1686: populations and may vanish for a certain population. All of
1687: these contributions to the intrinsic source fluctuations result in numerous
1688: additional auto and cross terms in the observed galaxy two-point statistics,
1689: and therefore proper account should be taken into these additional terms
1690: in interpreting measurements of galaxy two-point statistics from
1691: upcoming dark energy surveys.
1692:
1693: With the complete list of the contributions of the gravitational lensing and
1694: the generalized Sachs-Wolfe effects, separated as two physically distinct
1695: origins, we have identified several contributions in the volume effect
1696: and one contribution
1697: in the source effect, which are ignored in the
1698: standard treatment: the evolution bias in the source effect arises from the
1699: generalized Sachs-Wolfe effect, when the mean number density of sources
1700: changes rapidly in redshift, and its impact on the observed galaxy two-point
1701: statistics can be substantially larger than that of the gravitational lensing
1702: magnification bias.
1703: The ignored contributions in the volume
1704: effect are typically of order peculiar velocities and hence they are
1705: subdominant, compared to the standard redshift-space distortion effect.
1706: However, their impact is comparable to the magnification bias at low redshift.
1707: While the cross term of the magnification bias and the intrinsic source
1708: fluctuation is more important at low redshift than the contribution of
1709: the magnification bias itself in the gravitational lensing effect,
1710: further calculations of the additional contributions associated with
1711: the volume effect may be needed, if higher
1712: accuracy of theoretical modeling is required from observation.
1713:
1714: We have investigated the impact of the additional contributions to the
1715: anisotropic structure of the observed galaxy two-point statistics, after
1716: simplifying some of the contributions to the intrinsic source fluctuations.
1717: The redshift-space distortion affects the observed galaxy two-point statistics
1718: most, imprinting its well-known feature in the anisotropic structure
1719: \citep{KAISE87,STWI95,HAMIL98}.
1720: The gravitational lensing effect is small but non-negligible at a percent
1721: level, particularly
1722: along the line-of-sight separation and at high redshift, since their
1723: contribution increases with longer line-of-sight distance to the source
1724: galaxies and the clustering amplitude of the intrinsic source fluctuations
1725: decreases in redshift. The evolution bias has an angular pattern similar to
1726: the redshift-space distortion, but its impact becomes appreciable,
1727: only at fairly large transverse separation. While it is challenging to
1728: analyze the observed anisotropic structure of galaxy two-point statistics,
1729: its full analysis from upcoming dark energy surveys
1730: can provide a great opportunity to separately
1731: identify each contribution from the gravitational lensing and the generalized
1732: Sachs-Wolfe effects, increasing the leverage to understand the underlying
1733: physical mechanism.
1734:
1735: However, we note that constraining the underlying cosmological model
1736: will require not only accurate theoretical predictions, but also model
1737: fitting to measurements, which results in further distortion
1738: in galaxy two-point statistics, known as Alcock-Paczy\'nski
1739: effect \citep{ALPA79}.
1740: Furthermore, our current investigation has focused on the linear theory
1741: predictions and its additional contributions: nonlinearity and scale-dependent
1742: galaxy bias can affect our results, though its impact is expected to
1743: be less than at the percent level
1744: around the acoustic scale (see, e.g., \citep{EIWH04,SEEI05,EISEWH07}).
1745: However, additional leverage can be gained by modeling scale-dependent
1746: galaxy bias on nonlinear scales \citep{YOWEET08}.
1747:
1748: \acknowledgments
1749: We acknowledge useful discussions with Jordi Miralda-Escud\'e.
1750: We are very grateful to Matias Zaldarriaga and Liam Fitzpatrick
1751: for discussions of the effect of volume distortion.
1752: J.~Y. is supported by the Harvard College Observatory under the
1753: Donald~H. Menzel fund.
1754:
1755: \appendix
1756:
1757: \section{2D and 3D Statistics}
1758: \label{app:limber}
1759: Here we derive the relation between 2D and 3D fluctuations and their
1760: two-point statistics.
1761: Consider a fluctuation field $\delta^\up{2D}(\Vang;z_s)$ on the sky from
1762: a source population at $z_s$. In general, it can be expressed in terms of
1763: the convolution of a window function $W(\chi)$ and its 3D fluctuation
1764: $\delta^\up{3D}(\bdv{x})$
1765: \beeq
1766: \delta^\up{2D}(\Vang;z_s)=\int_0^\infty d\chi ~W(\chi_s-\chi)~\delta^\up{3D}
1767: \left[r(\chi)\Vang,\chi;z\right].
1768: \eneq
1769: When the window function is appreciable only around $z_s$ representing a
1770: narrow selection function in redshift surveys,
1771: $\delta^\up{2D}(\Vang;z_s)\simeq\delta^\up{3D}(\bdv{x};z_s)$ with its functional
1772: dependence
1773: $\bdv{x}=\left[r(\chi_s)\Vang,\chi_s\right]$. However, contributions to
1774: $\delta^\up{2D}(\Vang;z_s)$ can come from the fluctuations
1775: $\delta^\up{3D}(\bdv{x};z)$ at $z<z_s$ and
1776: $\delta^\up{2D}(\Vang;z_s)$ may be substantially different from
1777: $\delta^\up{3D}(\bdv{x};z_s)$, when the window function is broad. For
1778: example,
1779: the convergence field $\kappa(\Vang;z_s)$ in Eq.~(\ref{eq:conv}) has
1780: the window function
1781: \beeq
1782: W^\kappa(\chi_s-\chi)=\left({3H_0^2\over2}\OM\right)
1783: {r(\chi_s-\chi)~r(\chi)\over a(\chi)~r(\chi_s)},
1784: \label{eq:window}
1785: \eneq
1786: which peaks roughly at a half of $r(\chi_s)$.
1787:
1788: In a sufficiently small patch of the sky, the Fourier mode of
1789: $\delta^\up{2D}(\Vang;z_s)$ is
1790: \bear
1791: \delta^\up{2D}_\bdv{l}(z_s)&=&\int d^2\Vang~e^{-i\bdv{l}\cdot\Vang}~
1792: \delta^\up{2D}(\Vang;z_s) \\
1793: &=&\int_0^\infty d\chi~ {W(\chi_s-\chi)\over r^2(\chi)} \nonumber \\
1794: &\times&\int{dk_z\over2\pi}~e^{ik_z\chi}
1795: ~\delta^\up{3D}\left[k_z,\bdv{k}_\perp={\bdv{l}\over r};z\right], \nonumber
1796: \enar
1797: and its (angular) power spectrum is
1798: \bear
1799: \label{eq:2dps}
1800: C_l(z_1,z_2)&=&\int{d^2\bdv{l}'\over(2\pi)^2}~\langle\delta^\up{2D}_\bdv{l}(z_1)~
1801: \delta^{\up{2D}*}_{\bdv{l}'}(z_2)\rangle \\
1802: &=&\int d\chi_a\int d\chi_b~{W(\chi_1-\chi_a)W(\chi_2-\chi_b)
1803: \over r(\chi_a)^2} \nonumber \\
1804: &\times&\int{dk_z\over2\pi}~e^{ik_z(\chi_a-\chi_b)}~
1805: P\left[k_z,k_\perp={l\over r(\chi_a)};z_a,z_b\right] \nonumber \\
1806: &=&\int d\chi ~{W(\chi_1-\chi)W(\chi_2-\chi)\over r^2(\chi)}
1807: P\left[k={l\over r(\chi_a)};z\right]. \nonumber
1808: \enar
1809: The last equality is obtained by adopting the Limber approximation, in which
1810: fluctuations along the line-of-sight are smoothed out and only long wavelength
1811: modes ($k_z\simeq0$) can contribute to the integral \citep{LIMBE54,KAISE92}.
1812: With the Limber approximation,
1813: the angular correlation function is
1814: \bear
1815: \label{eq:2dcc}
1816: w(\Delta\theta;z_1,z_2)&=&\langle\delta^\up{2D}(\Vang_1;z_1)~\delta^\up{2D}
1817: (\Vang_2;z_2)\rangle \\
1818: &=&\int_0^\infty d\chi~W(\chi_1-\chi)W(\chi_2-\chi) \nonumber \\
1819: &\times&w_p\left[r(\chi)\Delta\theta;z\right], \nonumber
1820: \enar
1821: where the projected correlation function is
1822: \bear
1823: w_p\left[R;z\right]&=&\int_{-\infty}^\infty dr_\parallel~
1824: \xi\left[r=\sqrt{R^2+r_\parallel^2};z\right] \\
1825: &=&\int_0^\infty{k~dk\over2\pi}P(k;z)J_0(kR).\nonumber
1826: \enar
1827:
1828: The angular correlation function and power spectrum in Eqs.~(\ref{eq:xiMB})
1829: and~(\ref{eq:akk}) can be readily obtained by substituting the window
1830: function $W^\kappa(\chi)$ for the convergence in Eq.~(\ref{eq:window})
1831: with $W(\chi)$ in Eqs.~(\ref{eq:2dps}) and~(\ref{eq:2dcc}). The cross
1832: correlation function and power spectrum in Eqs.~(\ref{eq:xidmb})
1833: and~(\ref{eq:admb}) can be computed in a similar manner, since
1834: $W(\chi_s-\chi)=\delta^D(\chi_s-\chi)$ gives
1835: $\delta^\up{2D}(\Vang;z_s)=\delta^\up{3D}(\bdv{x};z_s)$.
1836:
1837: In Sec.~\ref{sec:two}, we associated the angular power spectrum
1838: $C_l^{\kappa\kappa}$ to a 3D power spectrum to compare its impact with other
1839: 3D power spectra. A 3D fluctuation field can be constructed from
1840: $\delta^\up{2D}(\Vang;z_s)$ by
1841: \bear
1842: \delta(\bdv{k};z_s)&=&\int d^3\bdv{x}~e^{-i\bdv{k}\cdot\bdv{x}}~
1843: \delta^\up{2D}(\Vang;z_s) \\
1844: &=&\int d\chi_s ~r^2(\chi_s)~e^{-ik_z\chi_s}~\delta^\up{2D}_{\bdv{l}=\bdv{k}_\perp
1845: r_s}(z_s)\nonumber \\
1846: &=&(2\pi)~\delta^D(k_z)~r^2(\chi_s)~\delta^\up{2D}_{\bdv{l}=\bdv{k}_\perp r_s}(z_s).
1847: \nonumber
1848: \enar
1849: We assumed $\delta^\up{2D}_\bdv{l}$ is independent of $z_s$ in the last
1850: equality. For high redshift source populations, this approximation is
1851: accurate, since the growth of the comoving angular diameter distance flattens
1852: at high $z$ and it becomes nearly constant. Within this approximation
1853: $r(\chi_1)=r(\chi_2)=r(\bar\chi)$,
1854: the 3D power spectrum is anisotropic and it is related to the angular
1855: power spectrum by
1856: \bear
1857: P(k_z,k_\perp;z_1,z_2)&=&\int{d^3\bdv{k}'\over(2\pi)^3}~\langle\delta(\bdv{k};
1858: z_1)~\delta^*(\bdv{k}';z_2)\rangle \\
1859: &=&(2\pi)~\delta^D(k_z)~r^2(\bar\chi)~C_{l=k_\perp r(\bar\chi)}(\bar z). \nonumber
1860: \enar
1861: In practice, $\delta^D(k_z)$ need to be replaced by a survey window function
1862: \citep{HUGALO08}, but note that it is crucial to assume the independence of
1863: source redshift, when computing the power spectrum.
1864: We also note that the Limber approximation breaks down when the radial
1865: window function of a survey is narrow compared to the correlation length scale
1866: (see, e.g., \citep{SIMON07,LOAF08}). However, the use of the Limber
1867: approximation is readily
1868: justified in galaxy surveys, in which the radial window
1869: function has width of $\Delta z\simeq0.1-0.2$, corresponding to several
1870: hundred Mpc.
1871:
1872: \section{Gauge-Invariant Form of Observed Number Density}
1873: \label{app:ngal}
1874: Here we provide a rigorous derivation of the observed number density
1875: in Sec.~\ref{ssec:vol}. For simplicity, we assume a flat universe.
1876:
1877: In the observer's frame, local coordinates $p^\alpha$ are used
1878: to describe the observed positions of galaxies and their true positions are
1879: related to the observed positions by
1880: the photon geodesic $x^a(\lambda)$.
1881: The total number of observed galaxies can be computed by
1882: considering a covariant volume integral \citep{WEINB72}:
1883: \beeq
1884: N_\up{gal}=\int\sqrt{-g}~n_\up{phy}~u^d~dS_d~,
1885: \label{Aeq:ngal}
1886: \eneq
1887: where the (oriented) hyper-surface element is
1888: \beeq
1889: dS_d=\epsilon_{abcd}~
1890: {\partial x^a\over\partial p^1}{\partial x^b\over\partial p^2}
1891: {\partial x^c\over\partial p^3}~dp^1dp^2dp^3~,
1892: \eneq
1893: $g$ is the determinant of the space-time metric, $n_\up{phy}$ is the physical
1894: number density of galaxies, and $\epsilon_{abcd}=\epsilon_{[abcd]}$
1895: is the Levi-Civita tensor density. We take the Newtonian gauge variables
1896: $(\ztobs,\tobs,\pobs)$ as the observed local coordinates $p^\alpha$.
1897: For notational simplicity, tilde is used to represent observed quantities
1898: (e.g., $\ztobs=\zobs$).
1899: By imposing the number conservation, the observed number density is then
1900: related to the total number of galaxies by
1901: \beeq
1902: N_\up{gal}=\int~\ntobs~{\chi^2(\ztobs)\over H(\ztobs)}~\sin\tobs
1903: ~d\ztobs d\tobs d\pobs~.
1904: \label{Aeq:nobs}
1905: \eneq
1906:
1907: In a homogeneous universe, the local coordinates are identical to the
1908: true coordinates $(\ztobs,\tobs,\pobs)=(z,\theta,\phi)$, and
1909: the photon geodesic is simply
1910: $x^a(\lambda)=(y,y~e^\alpha)$, where we choose the normalization of the affine
1911: parameter as $\lambda=(a/\nu)y$. Noting that the four velocity of a comoving
1912: observer is $u^a=(1/a,0)$, equation~(\ref{Aeq:ngal})
1913: can be readily solved as
1914: \bear
1915: N_\up{gal}&=&\int a^4~{n_\up{phy}\over a}~\epsilon_{\alpha\beta\gamma0}~
1916: {\partial x^\alpha\over\partial \ztobs}{\partial x^\beta\over\partial \tobs}
1917: {\partial x^\gamma\over\partial \pobs}~d\ztobs d\tobs d\pobs \nonumber \\
1918: &=&\int a^3~n_\up{phy}~{\chi^2(z)\over H(z)}~\sin\theta~dz d\theta d\phi.
1919: \enar
1920: Therefore, we recover the standard relation for $N_\up{gal}$
1921: and $\ntobs=a^3~n_\up{phy}=n(z,\theta,\phi)$.
1922:
1923: In an inhomogeneous universe, the photon geodesic deviates from the null path
1924: and the local coordinates are different from the true coordinates.
1925: Perturbations to the photon geodesic in an inhomogeneous universe
1926: can be computed by integrating the null vector $k^a(\lambda)=dx^a/d\lambda$,
1927: \beeq
1928: x^a(\lambda)=(y,y~e^\alpha)+\int_0^ydy'~(\delta\nu,\delta e^\alpha)~.
1929: \eneq
1930: Note that to the first order in perturbations
1931: the integration is performed
1932: along the null path, ranging from the observer
1933: at $y=0$ to the source galaxies at $y$.
1934:
1935: With $u^a=((1-\psi)/a,v^\alpha/a)$, the integrand of equation~(\ref{Aeq:ngal})
1936: is
1937: \bear
1938: u^d~dS_d&=&{1-\psi\over a}~\epsilon_{\alpha\beta\gamma0}~
1939: {\partial x^\alpha\over\partial \ztobs}{\partial x^\beta\over\partial \tobs}
1940: {\partial x^\gamma\over\partial \pobs} \nonumber \\
1941: &+&{v^\alpha\over a}~\epsilon_{abc\alpha}~
1942: {\partial x^a\over\partial \ztobs}{\partial x^b\over\partial \tobs}
1943: {\partial x^c\over\partial \pobs}~.
1944: \label{Aeq:int}
1945: \enar
1946: The last two terms, proportional to $\psi$ and $v^\alpha$, contribute to the
1947: first order in perturbations and the partial derivatives need to be computed,
1948: only to the zeroth order. The first term has two sources of perturbations
1949: from the partial derivatives: perturbations in the photon geodesic and
1950: the relation between the local and true coordinates. The former is non-zero,
1951: only when the derivative is taken with respect to $\ztobs$, i.e.,
1952: \beeq
1953: {1\over a}~\epsilon_{\alpha\beta\gamma0}\left(
1954: {\partial x^\beta\over\partial \theta}{\partial x^\gamma\over\partial\phi}
1955: \right)_0{\delta e^\alpha\over H(z)}~,
1956: \eneq
1957: and the latter is
1958: \beeq
1959: {1\over a}~\epsilon_{\alpha\beta\gamma0}\left({\partial x^\alpha\over\partial z}
1960: {\partial x^\beta\over\partial \theta}{\partial x^\gamma\over\partial\phi}
1961: \right)_0
1962: \left({\partial z\over\partial\ztobs}+{\partial\theta\over\partial\tobs}
1963: +{\partial\phi\over\partial\pobs}\right)_1~,
1964: \eneq
1965: where the subscripts denote the order in perturbations,
1966: to which quantities in the bracket need to be computed.
1967: When combined together, equation~(\ref{Aeq:int}) is
1968: \bear
1969: u^d~dS_d&=&{1\over a}{\chi^2(z)\over H(z)}\sin\theta\\
1970: &\times&\left[1+\delta e^\alpha e_\alpha+\left(
1971: {\partial z\over\partial\ztobs}+{\partial\theta\over\partial\tobs}
1972: +{\partial\phi\over\partial\pobs}\right)_1
1973: -\psi+v^\alpha e_\alpha\right]~. \nonumber
1974: \enar
1975:
1976: Finally, the determinant in equation~(\ref{Aeq:ngal}) gives
1977: $\sqrt{-g}=a^4~(1+\psi+3~\phi)$ and the total number of observed galaxies is
1978: \bear
1979: N_\up{gal}&=&\int a^3~n_\up{phy}~
1980: {\chi^2(z)\over H(z)}~\sin\theta ~d\ztobs d\tobs d\pobs \\
1981: &\times&\left[1+2~\phi+\varepsilon+\left(
1982: {\partial z\over\partial\ztobs}+{\partial\theta\over\partial\tobs}
1983: +{\partial\phi\over\partial\pobs}\right)_1\right]~. \nonumber
1984: \enar
1985: From equation~(\ref{Aeq:nobs}), we obtain our final result,
1986: \beeq
1987: \ntobs=n~{\chi^2(z)\over\chi^2(\ztobs)}~{H(\ztobs)\over H(z)}~
1988: \left[1+2~\phi-(1+z){d\varepsilon\over dz}-2~\kappa\right]~,
1989: \eneq
1990: which includes the distortion of volume element $\delta V$ and gravitational
1991: lensing magnification. This expression is manifestly gauge-invariant and
1992: valid on all scales. In special relativity, the volume element of the local
1993: Lorentz frame of matter is distorted by $\gamma=\sqrt{1-v^2}$, and hence is
1994: identical to an observer at rest, to the first order in perturbations.
1995: However, in our case the volume element can be measured, only by observing
1996: light rays of photons, of which time component is related to the spatial
1997: component, giving rise to the first order distortion in volume element.
1998:
1999: \bibliography{ms.bbl}
2000:
2001: \end{document}
2002: