1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: % Perturbation Theory Reloaded II:
4: % NON-LINEAR BIAS AND MILLENNIUM SIMULATION
5: %
6: % First Written: DJ, Nov 1, 2007
7: % revision: EK, May 3, 2008
8: % revision: EK, May 7, 2008 [removed ns]
9: %
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: %\documentclass[12pt,preprint]{aastex}
12: %\documentclass[manuscript]{aastex}
13: \documentclass[apj,twocolumn]{emulateapj}
14: %\documentclass[apjl]{emulateapj}
15: \newcommand{\vdag}{(v)^\dagger}
16: \newcommand{\myemail}{djeong@astro.as.utexas.edu}
17: \def\mathbi#1{\textbf{\em #1}}
18: \newcommand{\simgt}%
19: {\,\hbox{\lower0.6ex\hbox{$\sim$}\llap{\raise0.6ex\hbox{$>$}}}\,}
20: \newcommand{\simlt}%
21: {\,\hbox{\lower0.6ex\hbox{$\sim$}\llap{\raise0.6ex\hbox{$<$}}}\,}
22: \shorttitle{Non-Linear Bias, BAOs and Millennium Simulation}
23: \shortauthors{JEONG $\&$ KOMATSU}
24:
25:
26: \begin{document}
27: \title{%
28: Perturbation Theory Reloaded II:
29: Non-linear Bias, Baryon Acoustic Oscillations and Millennium Simulation
30: In Real Space
31: }%
32: \author{Donghui Jeong and Eiichiro Komatsu}
33: \affil{Department of Astronomy, University of Texas at Austin,
34: 1 University Station, C1400, Austin, TX, 78712, USA}
35: \email{djeong@astro.as.utexas.edu}
36:
37: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
38: \begin{abstract}
39: We calculate the non-linear galaxy power spectrum in real space, including non-linear
40: distortion of the Baryon
41: Acoustic Oscillations, using the standard 3rd-order
42: perturbation theory (PT). The calculation is based upon the assumption that
43: the number density of galaxies is a local function of the underlying,
44: non-linear density field. The galaxy bias is allowed to be both
45: non-linear and stochastic. We show that the PT calculation agrees with
46: the galaxy power spectrum estimated from the Millennium Simulation, in
47: the weakly non-linear regime (defined by the matter power spectrum) at
48: high redshifts, $1\le z\le6$. We also
49: show that, once 3 free parameters characterizing galaxy bias are
50: marginalized over, the PT power spectrum fit to the Millennium
51: Simulation data yields unbiased estimates of the distance scale,
52: $D$, to within the statistical error. This distance scale
53: corresponds to the angular diameter distance, $D_A(z)$, and
54: the expansion rate, $H(z)$, in real galaxy surveys.
55: Our results presented
56: in this paper are still restricted to real space. The future work should
57: include the effects of non-linear redshift space
58: distortion. Nevertheless, our results indicate that non-linear galaxy
59: bias in the weakly non-linear
60: regime at high redshifts is reasonably under control.
61: \end{abstract}
62: \keywords{cosmology : theory --- large-scale structure of universe}
63: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64: \section{Introduction}
65: Surveys of galaxies are the oldest way of mapping cosmological
66: fluctuations. Over the last three decades they have been used for
67: measuring cosmological parameters,
68: such as the matter density of the universe, $\Omega_m$
69: \citep[see][for a review]{peebles:POPC}.
70:
71: The galaxy surveys are largely complementary to CMB, as they allow us to
72: determine the important cosmological parameters that remain poorly
73: constrained by the CMB data alone
74: \citep[e.g.,][]{takada/komatsu/futamase:2006}: e.g., the mass of
75: neutrinos,
76: the shape of the primordial power spectrum, and the properties of
77: dark energy.
78:
79: The latest data sets, Two Degree Field Galaxy Redshift
80: Survey \citep[2dFGRS,][]{cole/etal:2005} and Sloan Digital Sky
81: Survey \citep[SDSS,][]{tegmark/etal:2006}, have enabled us to determine
82: most of the cosmological parameters to
83: better than 5\% accuracy,
84: when combined with the Cosmic Microwave Background (CMB) data from the Wilkinson
85: Microwave Anisotropy Probe \citep{bennett/etal:2003,spergel/etal:2003,spergel/etal:2007,hinshaw/etal:prep,dunkley/etal:prep,komatsu/etal:prep}.
86:
87: The galaxy power spectrum, the Fourier transform of the galaxy two point
88: correlation function, has been used widely for extracting cosmological
89: information from the galaxy survey data.
90: The amplitude, overall shape, as well as oscillatory features (called
91: the Baryon Acoustic Oscillations, or BAOs) contain a wealth of
92: cosmological information \citep[see][for a recent review]{weinberg:COS}.
93: In order to extract this information correctly,
94: we must understand how the observed galaxy power spectra are related to
95: the underlying cosmological models.
96:
97: How do we model the galaxy power spectrum?
98: We may use the cosmological perturbation theory (PT).
99: The accuracy of the linear PT has been verified
100: observationally by the
101: temperature and polarization data of CMB measured by WMAP
102: \citep{hinshaw/etal:2003,hinshaw/etal:2007,kogut/etal:2003,page/etal:2007,nolta/etal:prep}.
103: However, we cannot use the linear PT for
104: the galaxy power spectrum, as the matter density field grows non-linearly
105: due to gravitational instability. One must therefore use the {\it
106: non-linear} PT.
107:
108: There are three sources of non-linearities:
109: \begin{itemize}
110: \item[(1)] Non-linear evolution of the underlying matter density
111: field, which alters the matter power spectrum away from the linear prediction.
112: \item[(2)] Non-linear galaxy bias, or non-linear mapping between the underlying matter density field
113: and the distribution of collapsed objects such as dark matter halos and
114: galaxies, which alters the galaxy power spectrum
115: away from the matter power spectrum.
116: \item[(3)] Non-linear redshift space distortion, which arises as
117: the observed redshifts of galaxies used for measuring
118: locations of galaxies along the line of sight contain both the Hubble
119: expansion and the peculiar velocity of galaxies. This leads to the
120: systematic shifts in the line-of-sight positions of galaxies, altering
121: the galaxy power spectrum in redshift space away from that in real
122: space.
123: \end{itemize}
124:
125: Using the 3rd-order PT
126: \citep[see][for a review]{bernardeau/etal:2002} we have shown that
127: the first effect can be modeled accurately in the weakly non-linear
128: regime \citep[][hereafter Paper I]{jeong/komatsu:2006}.
129: In this paper we address the second effect, the non-linear
130: galaxy bias, using the 3rd-order PT. We will address
131: the third effect, the non-linear redshift space distortion, in the
132: future work.
133:
134: Our study is motivated by recently proposed high redshift galaxy
135: surveys such as Cosmic Inflation Probe (CIP)\footnote{\sf
136: http://cfa-www.harvard.edu/cip},
137: Hobby-Eberly Dark Energy Experiment \citep[HETDEX;][]{hill/etal:2004},
138: Baryon Oscillation Spectroscopic Survey (BOSS)\footnote{\sf http://howdy.physics.nyu.edu/index.php/BOSS},
139: and Wide-field Fiber-fed Multi Object Spectrograph survey
140: \citep[WFMOS;][]{glazebrook/etal:prep}, to mention a few.
141: These proposed surveys will observe the galaxy power spectra to the
142: unprecedented precision, which demands the precision modeling of the
143: galaxy power spectrum at 1\% accuracy or better.
144:
145: Over the last decade, the non-linear PT, including modeling of
146: non-linear galaxy power spectra, had been studied actively
147: \citep[see][for a review]{bernardeau/etal:2002}.
148: However, PT had never been applied to the real data such as 2dFGRS or
149: SDSS, as non-linearities are too strong for PT to be valid at low
150: redshifts, $z<1$ \citep[e.g.,][]{meiksin/white/peacock:1999}.
151: At high redshifts, i.e., $z>1$, however, PT is expected to perform better
152: because of weaker non-linearity.
153: In Paper I we have shown that the matter power spectrum
154: computed from the 3rd-order PT describes that from $N$-body simulations
155: accurately.\footnote{See also \citet{jain/bertschinger:1994} for
156: the earlier, pioneering work.}
157:
158: But, what about the {\it galaxy} power spectrum?
159: One may generally expect that, since non-linearities were milder in a
160: high-$z$ universe, there should be a plenty of room for PT to be a good
161: approximation. On the other hand, galaxies were more highly biased at
162: higher redshifts for a given mass, and therefore one might suspect,
163: somewhat naively, that
164: non-linear bias could compromise the success of PT.
165: In this paper we shall show that is not the case, and PT does provide a good
166: approximation to the galaxy power spectrum at high redshifts.
167:
168: This paper is organized as follows.
169: In \S~\ref{sec:PT} we give the formula for the 3rd-order PT galaxy power
170: spectrum.
171: In \S~\ref{sec:DM} we compare the 3rd-order PT matter power spectrum
172: with the matter power spectrum estimated from the Millennium
173: Simulation \citep{springel/etal:2005}, in order to confirm our previous
174: results (Paper I) with the Millennium Simulation.
175: In \S~\ref{sec:galaxy} we show that the PT calculation of the galaxy
176: power spectrum agrees with the galaxy power spectrum estimated from the
177: Millennium Simulation in the weakly non-linear regime (defined by the
178: matter power spectrum) at high redshifts, $1\le z\le 6$.
179: In \S~\ref{sec:bias} we extract the distance scale from
180: the Millennium Simulation, which is related to the angular diameter distance
181: and the expansion rate of the universe in real surveys.
182: In \S~\ref{sec:conclusion} we give discussion and conclusions.
183:
184:
185: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
186: \section{Non-linear galaxy power spectrum from perturbation theory}
187: \label{sec:PT}
188: \subsection{Locality Assumption}
189: Galaxies are biased tracers of the underlying density field
190: \citep{kaiser:1984}, which implies that the distribution of galaxies
191: depends on the underlying matter density fluctuations in a complex
192: way. This relation must depend
193: upon the detailed galaxy formation processes, which are not yet
194: understood completely.
195:
196: However, on large enough scales, one may approximate this function as
197: a local function of the underlying density fluctuations, i.e.,
198: the number density of galaxies at a given position in the universe is
199: given solely by the underlying matter density at the same position.
200: With this approximation, one may expand the density fluctuations of
201: galaxies, $\delta_g$, in terms of the underlying matter density
202: fluctuations, as \citep{fry/gaztanaga:1993,mcdonald:2006}
203: \begin{equation}\label{eq:Taylor_expansion}
204: \delta_g(\mathbf{x})
205: =
206: \epsilon+
207: b_1\delta(\mathbf{x}) + \frac12b_2\delta^2(\mathbf{x})+\frac16b_3\delta^3(\mathbf{x})+\dots,
208: \end{equation}
209: where $b_n$ are the galaxy bias parameters, and
210: $\epsilon$ is a random variable that represents the ``stochasticity'' of
211: the galaxy bias, i.e., the relation between $\delta_g(\mathbf{x})$ and
212: $\delta(\mathbf{x})$ is not deterministic, but contains some noise
213: \citep[e.g.,][and references therein]{yoshikawa/etal:2001}.
214: We assume that the stochasticity is white noise, and is uncorrelated
215: with the density fluctuations, i.e., $\langle \epsilon\delta
216: \rangle=0$. While both of these assumptions should be violated at some small
217: scales, we assume that these are valid assumptions on the scales that we
218: are interested in -- namely, on the scales where the 3rd-order PT
219: describes the non-linear matter power spectrum with 1\% accuracy.
220: Since both bias parameters and stochasticity evolve in time
221: \citep{fry:1996,tegmark/peebles:1998}, we allow them to depend on redshifts.
222:
223: One obtains the traditional ``linear bias model'' when the Taylor series
224: expansion
225: given in Eq.~(\ref{eq:Taylor_expansion}) is truncated at the first
226: order and the stochasticity is ignored.
227:
228: The precise values of the galaxy bias parameters depend on the galaxy
229: formation processes, and different types of galaxies have different
230: galaxy bias parameters.
231: However, we are {\it not} interested in the precise values of the galaxy bias
232: parameters, but only interested in extracting
233: cosmological parameters from the observed galaxy power spectra with {\it
234: all the bias parameters marginalized over}.
235:
236: \subsection{3rd-order PT galaxy power spectrum}
237: The analysis in this paper adopts the framework of \cite{mcdonald:2006},
238: and we briefly summarize the result for clarity.
239: We shall use the 3rd-order PT; thus,
240: we shall keep the terms up to the 3rd order in $\delta$.
241: The resulting power spectrum can be written in terms of
242: the linear matter power spectrum, $P_L(k)$, and the 3rd order matter
243: power spectrum, $P_{\delta\delta}(k)$, as
244: %\citep{mcdonald:2006}
245: \begin{equation}\label{eq:3rd_PT_Pk}
246: P_{g}(k)
247: =
248: P_0 + \tilde{b}_1^2
249: \biggl[
250: P_{\delta\delta}(k) + \tilde{b}_2 P_{b2}(k) + \tilde{b}_2^2 P_{b22}(k)
251: \biggl],
252: \end{equation}
253: where $P_{b2}$ and $P_{b22}$ are given by
254: \begin{displaymath}\label{eq:P_b2}
255: P_{b2}=
256: 2
257: \int \frac{d^3 \mathbf{q}}{(2\pi)^3}
258: P_L(q) P_L(|\mathbf{k}-\mathbf{q}|)
259: F_2^{(s)}(\mathbf{q},\mathbf{k}-\mathbf{q}),
260: \end{displaymath}
261: and
262: \begin{displaymath}\label{eq:P_b22}
263: P_{b22}=
264: \frac{1}{2}\int \frac{d^3 \mathbf{q}}{(2\pi)^3}
265: P_L(q)
266: \biggl[
267: P_L(|\mathbf{k}-\mathbf{q}|) -P(q)
268: \biggl],
269: \end{displaymath}
270: respectively, with $F_2^{(2)}$ given by
271: \begin{displaymath}
272: F_2^{(s)}(\mathbf{q}_1,\mathbf{q}_2)
273: =
274: \frac{5}{7}+\frac{2}{7}
275: \frac{(\mathbf{q}_1\cdot\mathbf{q}_2)^2}{q_1^2q_2^2}
276: +
277: \frac{\mathbf{q}_2\cdot\mathbf{q}_2}{2}
278: \left(
279: \frac{1}{q_1^2}+\frac{1}{q_2^2}
280: \right).
281: \end{displaymath}
282:
283: We use the standard formula for $P_{\delta\delta}$
284: (see Eq.~(14) of Paper I and references therein).
285: Here, $\tilde{b}_1$, $\tilde{b}_2$, and $P_0$ are the
286: non-linear bias parameters\footnote{These parameters correspond to
287: $b_1$, $b_2$, and $N$ in the original paper
288: by \citet{mcdonald:2006}.}, which are given in terms of the original
289: coefficients for the Taylor expansion as
290: \begin{eqnarray}
291: \label{eq:tb1b2}
292: \tilde{b}_1^2
293: &=&b_1^2+b_1b_3\sigma^2+\frac{68}{21}b_1b_2\sigma^2,\nonumber\\
294: \tilde{b}_2
295: &=&\frac{b_2}{\tilde{b}_1},\\
296: P_0
297: &=& \langle \epsilon^2\rangle + \frac12b_2^2\int
298: \frac{k^2dk}{2\pi^2}P_L^2(k)\nonumber,
299: \end{eqnarray}
300: where $\sigma$ is the r.m.s. of density fluctuations.
301:
302: We will never have to deal with the original coefficients, $b_1$, $b_2$,
303: $b_3$, or $\epsilon$.\footnote{For the expression of $P_g(k)$ with the
304: original coefficients, see \citet{heavens/matarrese/verde:1998,smith/scoccimarro/sheth:2007}.}
305: Instead, we will only use the re-parametrized bias
306: parameters, $\tilde{b}_1$, $\tilde{b}_2$, and $P_0$, as these are
307: related more directly to the observables.
308: As shown by \citet{mcdonald:2006},
309: in the large-scale limit, $k\rightarrow 0$, one finds
310: \begin{equation}
311: P_{g}(k)
312: \rightarrow
313: P_0 + \tilde{b}_1^2 P_L(k).
314: \end{equation}
315: Therefore, in the large-scale limit one recovers the traditional linear
316: bias model plus the constant term.
317: Note that $\tilde{b}_1$ is the same as what is called the ``effective
318: bias'' in \cite{heavens/matarrese/verde:1998}.
319:
320: Throughout this paper we shall use Eq.~(\ref{eq:3rd_PT_Pk}) for calculating
321: the non-linear galaxy power spectra.
322:
323: \subsection{Why we do not care about the precise values of bias parameters}
324: The precise values of the galaxy bias parameters depend on the details of the
325: galaxy formation and evolution, as well as on galaxy types,
326: luminosities, and so on.
327:
328: However, our goal is to extract the cosmological information from the
329: observed galaxy power spectra, without having to worry about which
330: galaxies we are using as tracers of the underlying density field.
331:
332: Therefore, we will marginalize the likelihood function over the bias
333: parameters, without ever paying attention to their precise values.
334: Is this approach sensible?
335:
336: One might hope that one should be able to calculate the bias parameters
337: for given properties of galaxies from the first principles using, e.g.,
338: sophisticated numerical simulations.
339:
340: Less numerically expensive way of doing the same thing would be to use
341: the semi-analytical halo model approach, calibrated with a smaller set of
342: numerical simulations \citep[see][for a
343: review]{cooray/sheth:2002}.
344: Using the peak-background split method \citep{sheth/tormen:1999}
345: based upon the excursion set approach
346: \citep{Bond/etal:1991},
347: one can calculate $b_1$, $b_2$, $b_3$, etc., the coefficients of the
348: Taylor series expansion given in Eq.~(\ref{eq:Taylor_expansion}), for the
349: density of dark matter halos.
350: Once the bias parameters for dark matter halos are specified,
351: the galaxy bias parameters may be calculated using the so-called Halo
352: Occupation Distribution (HOD) \citep{seljak:2000}.
353:
354: \citet{smith/scoccimarro/sheth:2007} have attempted this approach, and
355: shown that it is difficult to calculate even the power spectrum of
356: dark matter halos that matches $N$-body simulations.
357: The halo-model predictions for bias parameters are not yet accurate
358: enough, and we do not yet have a correct model for $P_0$.
359:
360: The situation would be even worse for the galaxy power spectrum, as
361: we would have to model the HOD in addition to the halo bias.
362: At the moment the form of HOD is basically a free empirical function.
363: We therefore feel that it is dangerous to rely on our limited understanding
364: of these complications for computing the bias parameters.
365:
366: This is the reason why we have decided to give up predicting the precise
367: values of bias parameters entirely.
368: Instead, we shall treat 3 bias parameters, $\tilde{b}_1$, $\tilde{b}_2$,
369: and $P_0$, as free parameters, and fit them to the observed galaxy power
370: spectra simultaneously with the cosmological parameters.
371:
372: The most important question that we must ask is the following, ``using the
373: 3rd-order PT
374: with 3 bias parameters, can we extract the correct cosmological
375: parameters from the galaxy power spectra?'' If the answer is yes, we
376: will not have to worry about the precise values of bias parameters
377: anymore.
378:
379:
380:
381: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
382: \section{Dark Matter Power spectrum from Millennium Simulation}
383: \label{sec:DM}
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: In this section we show that the matter power spectrum computed from the
386: 3rd-order PT agrees with that estimated from the Millennium
387: Simulation \citep{springel/etal:2005}. This result confirms our previous
388: finding (Paper I).
389:
390: Using the result obtained in this
391: section we define the maximum wavenumber, $k_{max}$, below which the
392: 3rd-order PT may be trusted. The matter power spectrum gives an unambiguous
393: definition of $k_{max}$, which will then be used thereafter when we
394: analyze power spectra of halos and galaxies in \S~\ref{sec:galaxy}.
395:
396: \subsection{Millennium Simulation}
397: The Millennium simulation \citep{springel/etal:2005} is a large $N$-body simulation with the box
398: size of $(500~\textrm{Mpc}/h)^3$ and $2160^3$ dark matter particles. The
399: cosmological parameters used
400: in the simulation are
401: $(\Omega_{dm},\Omega_{b},\Omega_\Lambda,h)
402: =(0.205,~0.045,~0.75,~0.73)$.
403:
404: The primordial power spectrum used in the simulation is
405: the scale-invariant Peebles-Harrison-Zel'dovich spectrum, $n_s=1.0$, and
406: the linear r.m.s. density fluctuation smoothed with a top-hat filter of
407: radius $8~h^{-1}\mathrm{Mpc}$ is $\sigma_8=0.9$.
408: Note that these values are significantly larger than
409: the latest values found from the WMAP 5-year data,
410: $\sigma_8\simeq 0.8$ and $n_s\simeq 0.96$
411: \citep{dunkley/etal:prep,komatsu/etal:prep}, which implies that
412: non-linearities in the
413: Millennium Simulation should be stronger than those in our Universe.
414:
415: The Millennium Simulation was carried out using the {\sf GADGET} code
416: \citep{springel/yoshida/white:2001,springel:2005}.
417: The {\sf GADGET} uses the tree Particle Mesh (tree-PM) gravity solver,
418: which tends to have a larger dynamic range than the traditional PM
419: solver for the same box size
420: and the same number of particles (and meshes)\citep{heitmann/etal:prep}.
421: Therefore, the matter power spectrum from the Millennium Simulation does
422: not suffer from an artificial suppression of power as much as those from
423: the PM codes.
424:
425: The initial particle distribution was generated at the initial redshift
426: of $z_{ini}=127$ using the standard Zel'dovich approximation.
427: While the initial conditions generated from the standard Zel'dovich
428: approximation tend to produce an artificial suppression of power at
429: later times, and the higher-order scheme such as the second-order
430: Lagrangian perturbation theory usually produces better results
431: \citep{scoccimarro:1998,crocce/pueblas/scoccimarro:2006},
432: the initial redshift of the Millennium Simulation, $z_{ini}=127$, is
433: reasonably high for the resulting power spectra to have converged in the
434: weakly non-linear regime.
435:
436: The mass of each dark matter particle in the simulation is
437: $M_{dm}=8.6\times 10^8 M_\odot/h$. They require at least 20 particles
438: per halo for their halo finder, and thus the minimum mass resolution of
439: halos is given by
440: $M_{halo}\ge20M_{dm}\simeq 1.7 \times 10^{10}~M_\odot/h$.
441: Therefore, the Millennium Simulation covers the mass range that is
442: relevant to real galaxy surveys that would detect galaxies with masses
443: in the range of $M\simeq 10^{11}-10^{12}~M_\odot$.
444: This property distinguishes our study from the previous studies on
445: non-linear distortion of BAOs due to galaxy bias
446: \citep[e.g.,][]{smith/scoccimarro/sheth:2007,huff/etal:2007}, whose
447: mass resolution was greater than $\sim 10^{12}~M_\odot$.
448:
449: In addition to the dark matter halos, the Millennium database\footnote{
450: \textsf{http://www.g-vo.org/MyMillennium2/}}
451: also provides galaxy catalogues from two different
452: semi-analytic galaxy formation models
453: \citep{delucia/blaizot:2007,croton/etal:2006,bower/etal:2006,benson/etal:2003,cole/etal:2000}.
454: These catalogues give us an excellent opportunity for testing
455: validity of the non-linear
456: galaxy power spectrum model based upon the 3rd-order PT
457: with the unprecedented precision.
458:
459: \subsection{3rd-order PT versus Millennium Simulation: Dark Matter Power
460: Spectrum}
461: First, we compare the matter power spectrum from the Millennium
462: simulation with the 3rd-order PT calculation. The matter power
463: spectrum we use here was measured directly from the Millennium
464: simulation on the fly.\footnote{We thank Volker Springel for providing
465: us with the matter power spectrum data.}
466:
467: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468: \begin{figure}
469: \centering
470: \rotatebox{90}{%
471: \includegraphics[width=6.5cm]{fig1.ps}
472: }%
473: \caption{%
474: Matter power spectrum at $z=0$, 1, 2, 3, 4, 5 and 6
475: (\textit{from top to bottom}) derived from the Millennium
476: Simulation (\textit{dashed lines}),
477: the 3rd-order PT (\textit{solid lines}), and the linear PT
478: (\textit{dot-dashed lines}).
479: }%
480: \label{fig1}
481: \end{figure}
482: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
483: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
484: \begin{figure}
485: \centering
486: \rotatebox{90}{%
487: \includegraphics[width=6.5cm]{fig2.ps}
488: % \includegraphics[width=9cm]{figs/new.ps}
489: }%
490: \caption{%
491: Dimensionless matter power spectrum, $\Delta^2(k)$,
492: at $z=1$, 2, 3, 4, 5, and 6.
493: The dashed and solid lines show the Millennium Simulation data and the
494: 3rd-order PT calculation, respectively.
495: The dot-dashed lines show the linear power spectrum.
496: }%
497: \label{fig2}
498: \end{figure}
499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
500: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501: \begin{figure}
502: \centering
503: \rotatebox{90}{
504: \includegraphics[width=6.5cm]{fig3.ps}
505: }
506: \caption{
507: Fractional difference between the matter power spectra from the 3rd-order PT
508: and that from the Millennium Simulation, $P_m^{sim}(k)/P_m^{PT}-1$
509: (dots with errorbars). The solid lines show the perfect match,
510: while the dashed lines show $\pm 2\%$ accuracy.
511: We also show $k_{max}(z)$, below which we trust the prediction from the
512: 3rd-order PT, as a vertical dotted line.
513: }%
514: \label{fig3}
515: \end{figure}
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
517:
518: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
519: %% The values (usually only l,r and c) in the last part of
520: %% \begin{deluxetable}{} command tell LaTeX how many columns
521: %% there are and how to align them.
522: \begin{deluxetable}{ccc}
523:
524: %% Keep a portrait orientation
525:
526: %% Over-ride the default font size
527: %% Use 10pt
528: \tabletypesize{\footnotesize}
529:
530: %% Use \tablewidth{?pt} to over-ride the default table width.
531: %% If you are unhappy with the default look at the end of the
532: %% *.log file to see what the default was set at before adjusting
533: %% this value.
534:
535: %% This is the title of the table.
536: \tablecaption{Maximum wavenumbers, $k_{max}$, for the Millennium Simulation}
537: %% This command over-rides LaTeX's natural table count
538: %% and replaces it with this number. LaTeX will increment
539: %% all other tables after this table based on this number
540: \tablenum{1}
541: \label{table:kmax}
542:
543: %% The \tablehead gives provides the column headers. It
544: %% is currently set up so that the column labels are on the
545: %% top line and the units surrounded by ()s are in the
546: %% bottom line. You may add more header information by writing
547: %% another line between these lines. For each column that requries
548: %% extra information be sure to include a \colhead{text} command
549: %% and remember to end any extra lines with \\ and include the
550: %% correct number of &s.
551: \tablehead{\colhead{$z$} & \colhead{$k_{max}$} & \colhead{$\tilde{k}_{max}$} \\
552: \colhead{} & \colhead{($h/\mathrm{Mpc}$)} & \colhead{$(h/\mathrm{Mpc})$} }
553:
554: %% All data must appear between the \startdata and \enddata commands
555: \startdata
556: 6 & 1.5 & 1.99 \\
557: 5 & 1.3 & 1.37 \\
558: 4 & 1.2 & 1.02 \\
559: 3 & 1.0 & 0.60 \\
560: 2 & 0.25 & 0.35 \\
561: 1 & 0.15 & 0.20
562: \enddata
563:
564: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
565: %% or \tablecomments{text} between the \enddata and
566: %% \end{deluxetable} commands
567:
568: %% General table comment marker
569: \tablecomments{$z$: redshift\\
570: $k_{max}$: the maximum wavenumber for the simulated $P_m(k)$ to
571: agree with the PT calculation at 2\% accuracy within the statistical
572: error of the Millennium Simulation\\
573: $\tilde{k}_{max}$:
574: $\tilde{k}_{max}$ is defined by $\Delta^2_m(\tilde{k}_{max})=0.4$
575: which is the criteria recommended in Paper I.}
576: %% General table references marker
577: %%\tablerefs{table:kmax}
578:
579: \end{deluxetable}
580: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
581: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
582: \begin{figure}
583: \centering
584: \rotatebox{90}{
585: \includegraphics[width=6.5cm]{fig4.ps}
586: }
587: \caption{
588: Distortion of BAOs due to non-linear matter clustering. All of the power
589: spectra have been divided by
590: a smooth power spectrum without baryonic oscillations from eq.
591: (29) of \citet{eisenstein/hu:1998}. The error bars show the simulation
592: data, while the solid lines show the PT calculations. The dot-dashed
593: lines show the linear theory calculations.
594: The power spectrum data shown here have been taken from Figure~6 of
595: \citet{springel/etal:2005}.
596: }%
597: \label{fig4}
598: \end{figure}
599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
600:
601:
602: Figure \ref{fig1} shows the matter power spectrum from the Millennium
603: simulation (dashed lines), the 3rd-order PT calculation
604: (solid lines), and the linear PT (dot-dashed lines)
605: for seven different redshifts, $z=0$, 1, 2, 3, 4, 5, and 6.
606: The analytical calculation of the 3rd-order
607: PT reproduces the non-linear matter power spectrum from the
608: Millennium Simulation accurately at high redshifts, i.e., $z>1$,
609: up to certain maximum wavenumbers, $k_{max}$, that will be specified below.
610: To facilitate the comparison better, we show the dimensionless
611: matter power spectrum, $\Delta_m^2(k)\equiv k^3P_m(k)/2\pi^2$,
612: in Figure \ref{fig2}.
613:
614: We find the maximum wavenumber, $k_{max}(z)$, below which we trust
615: the prediction from the 3rd-order PT, by comparing the matter
616: power spectrum from PT and the Millennium Simulation.
617: The values of $k_{max}$ found here will be used later
618: when we analyze the halo/galaxy power spectra.
619:
620: In Paper I we have defined $k_{max}$ such that the fractional difference
621: between PT and the average of $\sim 100$ simulations is 1\%.
622: Here, we have only one realization, and thus the results are subject to
623: statistical fluctuations that might be peculiar to this particular
624: realization.
625: Therefore, we relax our criteria for $k_{max}$:
626: we define $k_{max}$ such that the fractional difference
627: between PT and the Millennium Simulation is 2\%.
628:
629: Figure \ref{fig3} shows the fractional differences
630: at $z=1$, 2, 3, 4, 5, and 6. Since we have only one realization,
631: we cannot compute statistical errors from the standard deviation of
632: multiple realizations.
633: Therefore, we derive errors from the leading-order 4-point function
634: assuming Gaussianity of the underlying density fluctuations (see
635: Appendix~\ref{sec:appA}),
636: $\sigma_{P(k)}=P(k)/\sqrt{N_k}$,
637: where $N_k$ is the number of independent Fourier modes per bin at a
638: given $k$ shown in Figure~\ref{fig3}.
639:
640: We give the values of $k_{max}$ in Table \ref{table:kmax}. We shall use
641: these values when we fit the halo/galaxy power spectrum in the
642: next section. Note that $k_{max}$ decreases rapidly below $z=2$.
643: %%revision%%
644: It is because $P(k)/P_{PT}(k)-1$ is not a monotonic function of $k$.
645: The dip in $P(k)/P_{PT}(k)-1$ is larger than $2\%$ at lower redshift,
646: $z<2$, while it is inside of the $2\%$ range at $z\ge3$.
647: Therefore, our criteria of $2\%$ make that sudden change.
648: This feature is due to the limitation of the standard 3rd order PT.
649: However, we can remove this feature by using the improved perturbation theory,
650: e.g. using renormalization group techniques.
651: (See, Figure 9 of \citet{matarrese/pietroni:2007}.)
652: %%end of revision%%
653:
654: We also give the values of $\tilde{k}_{max}$, for which
655: $\Delta^2_m(\tilde{k}_{max})=0.4$ (criteria recommended in Paper I).
656: The difference between $k_{max}$ and $\tilde{k}_{max}$ is probably due
657: to the fact that we have only one realization of the Millennium
658: Simulation, and thus estimation of $k_{max}$ is noisier.
659: Note that the values of $\tilde{k}_{max}$ given in Table
660: \ref{table:kmax}
661: are smaller than those given in Paper I.
662: This is simply because $\sigma_8$ of the Millennium Simulation
663: ($\sigma_8=0.9$) is larger than that of Paper I ($\sigma_8=0.8$).
664:
665: In Figure \ref{fig4} we show
666: the matter power spectra divided a smooth spectra without BAOs
667: \citep[Eq.~(29) of][]{eisenstein/hu:1998}.
668: The results are consistent with what we have found in Paper I:
669: although BAOs in the matter power spectrum are distorted heavily
670: by non-linear evolution of matter fluctuations,
671: the analytical predictions from the 3rd-order PT
672: capture the distortions very well at high redshifts, $z>2$.
673:
674: At lower redshifts, $z\sim 1$, the 3rd-order PT is clearly
675: insufficient, and one needs to go beyond the standard PT.
676: This is a subject of recent studies
677: \citep{crocce/scoccimarro:2008,matarrese/pietroni:2007,taruya/hiramatsu:2008,valageas:2007,matsubara:2008,mcdonald:2007}.
678:
679:
680:
681:
682:
683:
684:
685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
686: \section{HALO/GALAXY POWER SPECTRUM AND THE NON-LINEAR BIAS MODEL}
687: \label{sec:galaxy}
688: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
689: In this section we compare the 3rd-order PT galaxy power spectrum
690: with the power spectra of dark matter halos and galaxies
691: estimated from the Millennium Simulation.
692: After briefly describing the analysis method in \S~\ref{sec:analysis},
693: we analyze the halo bias and galaxy bias in \S~\ref{sec:halo}
694: and \S~\ref{sec:gal}, respectively. We then study
695: the dependence of bias parameters on halo/galaxy mass in
696: \S~\ref{sec:mass_dependence}.
697:
698: \subsection{Analysis method}
699: \label{sec:analysis}
700: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
701: %% The values (usually only l,r and c) in the last part of
702: %% \begin{deluxetable}{} command tell LaTeX how many columns
703: %% there are and how to align them.
704: \begin{deluxetable*}{cccccccc}
705:
706: %% Keep a portrait orientation
707:
708: %% Over-ride the default font size
709: %% Use 10pt
710: \tabletypesize{\footnotesize}
711:
712: %% Use \tablewidth{?pt} to over-ride the default table width.
713: %% If you are unhappy with the default look at the end of the
714: %% *.log file to see what the default was set at before adjusting
715: %% this value.
716:
717: %% This is the title of the table.
718: \tablecaption{Summary of six snapshots from the Millennium Simulation}
719:
720: %% This command over-rides LaTeX's natural table count
721: %% and replaces it with this number. LaTeX will increment
722: %% all other tables after this table based on this number
723: \tablenum{2}
724: \label{table:Msummary}
725: %% The \tablehead gives provides the column headers. It
726: %% is currently set up so that the column labels are on the
727: %% top line and the units surrounded by ()s are in the
728: %% bottom line. You may add more header information by writing
729: %% another line between these lines. For each column that requries
730: %% extra information be sure to include a \colhead{text} command
731: %% and remember to end any extra lines with \\ and include the
732: %% correct number of &s.
733: \tablehead{\colhead{$z$} & \colhead{$z_\mathrm{show}$} & \colhead{$N_h$} & \colhead{$1/n_h$} & \colhead{$N_{Mg}$} & \colhead{$1/n_{Mg}$} & \colhead{$N_{Dg}$} & \colhead{$1/n_{Dg}$} \\
734: \colhead{} &\colhead{} & \colhead{} & \colhead{($[\mathrm{Mpc}/h]^3$)} & \colhead{} & \colhead{($[\mathrm{Mpc}/h]^3$)} & \colhead{} & \colhead{($[\mathrm{Mpc}/h]^3$)} }
735:
736: %% All data must appear between the \startdata and \enddata commands
737: \startdata
738: 5.724&6 & 5,741,720 & 21.770 & 6,267,471 & 19.944 & 4,562,368 & 27.398 \\
739: 4.888&5 & 8,599,981 & 14.535 & 9,724,669 & 12.854 & 7,604,063 & 16.439 \\
740: 4.179&4 & 11,338,698 & 11.024 & 13,272,933 & 9.418 & 10,960,404 & 11.405 \\
741: 3.060&3 & 15,449,221 & 8.091 & 19,325,842 & 6.468 & 17,238,935 & 7.251 \\
742: 2.070&2 & 17,930,143 & 6.972 & 23,885,840 & 5.233 & 22,962,129 & 5.444 \\
743: 1.078&1 & 18,580,497 & 6.727 & 26,359,329 & 4.742 & 27,615,058 & 4.527
744: \enddata
745:
746: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
747: %% or \tablecomments{text} between the \enddata and
748: %% \end{deluxetable} commands
749:
750: %% General table comment marker
751: \tablecomments{$z$: the exact redshift of each snapshot\\
752: $z_\mathrm{show}$: the redshift we quote in this paper\\
753: $N_h$: the number of MPA halos in each snapshot; $1/n_h$: the
754: corresponding Poisson shot noise\\
755: $N_{Mg}$: the number of MPA galaxies in each snapshot; $1/n_{Mg}$: the
756: corresponding Poisson shot noise\\
757: $N_{Dg}$: the number of Durham galaxies in each snapshot; $1/n_{Dg}$: the
758: corresponding Poisson shot noise}
759:
760: %% No \tablerefs indicated
761:
762: \end{deluxetable*}
763: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
764:
765: We choose six redshifts between $1\le z\le6$ from 63 snapshots of
766: the Millennium Simulation, and use all the available catalog of halos
767: (MPA Halo (MHalo), hereafter `halo') and two galaxy catalogues
768: (MPA Galaxies, hereafter `Mgalaxy'; Durham Galaxies,
769: hereafter `Dgalaxy') at each redshift. The
770: exact values of redshifts and the other relevant information
771: of chosen snapshots are summarized in Table \ref{table:Msummary}.
772:
773: %%revised%%
774: Halos are the groups of matter particles found directly from
775: the Millennium Simulation. First,
776: the dark matter groups (called FOF group)
777: are identified by using Friends-of-Friends (FoF) algorithm with
778: a linking length equal to 0.2 of the mean particle separation.
779: Then, each FoF group is divided into the gravitationally
780: bound local overdense regions, which we call halos here.
781:
782: Mgalaxies and Dgalaxies are the galaxies assigned to the halos
783: using two different semi-analytic galaxy formation codes:
784: L-Galaxies \citep[Mgalaxies,][]{delucia/blaizot:2007,croton/etal:2006}
785: and GALFORM \citep[Dgalaxies,][]{bower/etal:2006,benson/etal:2003,cole/etal:2000}.
786:
787: While both models successfully explain a number of observational
788: properties of galaxies like the break shape of
789: the galaxy luminosity function, star formation rate, etc, they differ
790: in detailed implementation.
791: For example, while the L-Galaxies code uses the halo merger tree constructed
792: by MHalos, the GALFORM code uses different
793: criteria for identifying subhalos inside the FOF group,
794: and thus uses a different merger tree.
795: Also, two models use different gas cooling prescriptions
796: and different initial mass functions (IMF) of star formation:
797: L-Galaxies and GALFORM define the cooling radius,
798: within which gas has a sufficient time to cool,
799: by comparing the cooling time with halo dynamical time and the
800: age of the halo, respectively.
801: Cold gas turns into stars with two different IMFs:
802: the L-Galaxies code ueses IMF from \citet{chabrier:2003} and
803: the GALFORM code uses \citet{kennicutt:1983}.
804: In addition to that, they treat AGN (Active Galactic Nucleus)
805: feedback differently: the L-Galaxies code
806: introduces a parametric model of AGN feedback depending
807: on the black hole mass and the virial velocity of halo,
808: and the GALFORM code imposes the condition
809: that cooling flow is quenched when the energy released by radiative cooling
810: (cooling luminosity) is less than some fraction
811: (which is modeled by a parameter, $\epsilon_{\mathrm{SMBH}}$)
812: of Eddington luminosity of the black hole.
813: For more detailed comparison of the two model, we refer readers to the
814: original papers cited above.
815: %%end of revision%%
816:
817: We compute the halo/galaxy power spectra from the Millennium Simulation
818: as follows:
819: \begin{itemize}
820: \item [(1)] Use the Cloud-In-Cell (CIC)
821: mass distribution scheme to
822: calculate the density field on $1024^3$
823: regular grid points from each catalog.
824: \item [(2)] Fourier-transform the discretized density field using {\sf
825: FFTW}\footnote{{\sf http://www.fftw.org}}.
826: \item [(3)] Deconvolve the effect of the CIC pixelization and aliasing
827: effect. We divide $P(\mathbf{k},z)\equiv|\delta(\mathbf{k},z)|^2$
828: at each cell by the following
829: window function \citep{Jing:2005}:
830: \begin{equation}\label{eq:window}
831: W(\mathbf{k})=\prod_{i=1}^3
832: \left[
833: 1-\frac{2}{3}\sin^2 \left(\frac{\pi k_i}{2 k_{N}}\right)
834: \right],
835: \end{equation}
836: where $\mathbf{k}=(k_1,k_2,k_3)$, and $k_N\equiv\pi/H$
837: is the Nyquist frequency,
838: ($H$ is the physical size of the grid).\footnote{Note that
839: Eq.~(\ref{eq:window}) is strictly valid
840: for the flat (white noise) power spectrum, $P(k)={\rm constant}$.
841: Nevertheless, it is still accurate for our purposes because, on small
842: scales, both the halo and galaxy power spectra are dominated by
843: the shot noise, which is also given by $P(k)={\rm constant}$.}
844: \item [(4)] Compute $P(k,z)$ by taking the angular average of
845: CIC-corrected $P(\mathbf{k},z)\equiv|\delta(\mathbf{k},z)|^2$ within a
846: spherical shell defined by
847: $k-\Delta k/2 < |\mathbf{k}| < k+\Delta k/2$.
848: Here, $\Delta k=2\pi/500~[h/\rm{Mpc}]$ is the
849: fundamental frequency that corresponds to the box size of the Millennium
850: Simulation.
851: \end{itemize}
852:
853: From the measured power spectra we find the maximum likelihood
854: values of the bias parameters using the likelihood
855: function approximated as a Gaussian:
856: \begin{equation}\label{eq:likelihood}
857: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,P_0)
858: =\prod_{k_i<k_{max}}
859: \frac{1}{\sqrt{2\pi\sigma_{Pi}^2}}
860: \exp
861: \left[-
862: \frac{
863: (P_{obs,i}-P_{g,i})^2
864: }{2\sigma_{Pi}^2}
865: \right],
866: \end{equation}
867: where $k_i$'s are integer multiples of the fundamental frequency
868: $\Delta k$, $P_{obs,i}$ is the measured power spectrum at $k=k_i$,
869: $P_{g,i}$ is the theoretical model given by Eq.~(\ref{eq:3rd_PT_Pk}),
870: and $\sigma_{Pi}$ is the statistical error in the measured power
871: spectrum.
872:
873: We estimate $\sigma_{Pi}$ in the same way as in \S~\ref{sec:DM}
874: (see also Appendix~\ref{sec:appA}). However, the power spectrum of
875: the point-like particles like halos and galaxies includes the Poisson shot
876: noise, $1/n$, where $n$ is the number density of objects, on top of the
877: power spectrum due to clustering.
878: Therefore, $\sigma_{Pi}$ must also include the shot-noise
879: contribution. We use
880: \begin{equation}\label{eq:varpk}
881: \sigma_{Pi}=\sigma_P(k_i)=\sqrt{\frac{1}{N_{ki}}}\left[P_g(k_i)+\frac{1}{n}\right],
882: \end{equation}
883: where
884: \begin{equation}
885: N_{ki}=2\pi\left(\frac{k}{\Delta k}\right)^2
886: \end{equation}
887: is the number of independent Fourier modes used for estimating the
888: power spectrum and $P_g(k_i)$ is the halo/galaxy power spectrum at $k=k_i$.
889: Here, $\Delta k=2\pi/(500~h^{-1}~\rm{Mpc})$ is the fundamental wavenumber
890: of the Millennium Simulation.
891: Note that we subtract the Poisson shot noise contribution,
892: $P_{shot}=1/n$, from the observed power spectrum
893: before the likelihood analysis.
894:
895: Eq.~(\ref{eq:varpk}) shows that the error on $P_{obs}(k)$ depends upon
896: the underlying $P_g(k)$. For the actual data analysis one should vary
897: $P_g(k)$
898: in the numerator of Eq.~(\ref{eq:likelihood}) as well as that in
899: $\sigma_{Pi}$, simultaneously. However, to simplify the analysis, we evaluate the
900: likelihood function in an iterative way: we first find the best-fitting
901: $P_g(k)$ using $\sigma_{Pi}$ with $P_g(k)$ in Eq.~(\ref{eq:varpk})
902: replaced by $P_{obs}(k)$. Let us call this $\tilde{P}_g(k)$. We then use
903: $\tilde{P}_g(k)$ in Eq.~(\ref{eq:varpk}) for finding the best-fitting
904: $P_g(k)$ that we shall report in this paper.
905: Note that we iterate this procedure only once for current study.
906:
907: Finally, we compute the 1-d marginalized 1-$\sigma$ interval (or the
908: marginalized $68.27\%$ confidence
909: interval) of each bias parameter by integrating the likelihood function
910: (Eq.~(\ref{eq:likelihood})), assuming a flat prior on the
911: bias parameters (see also Appendix \ref{sec:appB}).
912:
913: We first analyze the power spectrum of halos (in \S~\ref{sec:halo}) as
914: well as that of galaxies (in \S~\ref{sec:gal})
915: using all the halos and all the galaxies in the Millennium halo/galaxy
916: catalogues. We then study the mass dependence of bias parameters
917: in \S~\ref{sec:mass_dependence}.
918:
919: In order to show that the non-linear bias model (Eq.~\ref{eq:3rd_PT_Pk})
920: provides a much better fit than the linear bias model,
921: we also fit the measured power spectra with two linear bias models:
922: (i) linear bias with the linear matter power spectrum, and
923: (ii) linear bias with the non-linear matter power spectrum from the
924: 3rd-order PT. When fitting with the linear model, we use
925: $k_{max}=0.15~[h/\mathrm{Mpc}]$ for all redshift bins.
926:
927: \subsection{Halo power spectra}
928: \label{sec:halo}
929: \subsubsection{Measuring non-linear halo bias parameters}
930: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
931: \begin{figure*}
932: \centering
933: \rotatebox{90}{
934: \includegraphics[width=10cm]{fig5.ps}
935: }
936: \caption{
937: Halo power spectra from the Millennium Simulation at
938: $z=1$, 2, 3, 4, 5, and 6. Also shown in smaller panels
939: are the residual of fits. The points with errorbars show
940: the measured halo power spectra, while the solid, dashed, and
941: dot-dashed lines
942: show the best-fitting non-linear bias model (Eq.~(\ref{eq:3rd_PT_Pk})),
943: the best-fitting linear bias with the non-linear matter power spectrum,
944: and the best-fitting linear bias with the linear matter power spectrum,
945: respectively. Both linear models have been fit for
946: $k_{max,linear}=0.15~[h~\mathrm{Mpc}^{-1}]$, whereas
947: $k_{max}(z)$ given in Table~\ref{table:kmax} (also marked in each panel)
948: have been used for the non-linear bias model.
949: }%
950: \label{fig5}
951: \end{figure*}
952: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
953: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
954: %% The values (usually only l,r and c) in the last part of
955: %% \begin{deluxetable}{} command tell LaTeX how many columns
956: %% there are and how to align them.
957: \begin{deluxetable*}{ccclcccc}
958: %% Keep a portrait orientation
959: %
960: %% Over-ride the default font size
961: %% Use 10pt
962: \tabletypesize{\footnotesize}
963: %% Use \tablewidth{?pt} to over-ride the default table width.
964: %% If you are unhappy with the default look at the end of the
965: %% *.log file to see what the default was set at before adjusting
966: %% this value.
967: %
968: %% This is the title of the table.
969: \tablecaption{Non-linear halo bias parameters and the corresponding 68\%
970: interval estimated from the MPA halo power spectra
971: }
972: %
973: %% This command over-rides LaTeX's natural table count
974: %% and replaces it with this number. LaTeX will increment
975: %% all other tables after this table based on this number
976: \tablenum{3}
977: \label{table:halobias}
978: %
979: %% The \tablehead gives provides the column headers. It
980: %% is currently set up so that the column labels are on the
981: %% top line and the units surrounded by ()s are in the
982: %% bottom line. You may add more header information by writing
983: %% another line between these lines. For each column that requries
984: %% extra information be sure to include a \colhead{text} command
985: %% and remember to end any extra lines with \\ and include the
986: %% correct number of &s.
987: \tablehead{\colhead{$z$} & \colhead{$\tilde{b}_1$} & \colhead{$\tilde{b}_2$} & \colhead{$P_0$} & \colhead{$b_1^L$} & \colhead{$b_1^{LL}$} & \colhead{$b_1^{ST}$} & \colhead{$\tilde{b}_2^{ST}$} \\
988: \colhead{} & \colhead{} & \colhead{} & \colhead{($[\mathrm{Mpc}/h]^3$)} & \colhead{} & \colhead{} & \colhead{} & \colhead{} }
989: %
990: %% All data must appear between the \startdata and \enddata commands
991: \startdata
992: 6 & 3.41$\pm$0.01 & 1.52$\pm$0.03 & 141.86$\pm$3.73 & 3.50$\pm$0.03 & 3.51$\pm$0.03 & 3.69 & 2.10 \\
993: 5 & 2.76$\pm$0.01 & 0.91$\pm$0.03 & 57.77$\pm$2.84 & 2.79$\pm$0.03 & 2.80$\pm$0.03 & 3.16 & 1.70 \\
994: 4 & 2.27$\pm$0.01 & 0.52$\pm$0.03 & 22.65$\pm$1.88 & 2.28$\pm$0.02 & 2.29$\pm$0.02 & 2.77 & 1.40 \\
995: 3 & 1.52$\pm$0.01 & -1.94$\pm$0.05 & 329.42$\pm$10.6 & 1.62$\pm$0.01 & 1.63$\pm$0.01 & 2.23 & 1.07 \\
996: 2 & 1.10$\pm$0.06 & -2.12$\pm$0.65 & 507.25$\pm$214.7 & 1.19$\pm$0.01 & 1.20$\pm$0.01 & 1.84 & 0.76 \\
997: 1 & 0.74$\pm$0.09 & -3.05$\pm$1.49 & 1511.46$\pm$526.7 & 0.88$\pm$0.01 & 0.90$\pm$0.01 & 1.54 & 0.58
998:
999: \enddata
1000: %
1001: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
1002: %% or \tablecomments{text} between the \enddata and
1003: %% \end{deluxetable} commands
1004: %
1005: %% General table comment marker
1006: \tablecomments{$z$: redshift\\
1007: $\tilde{b}_1$, $\tilde{b}_2$, $P_0$: non-linear bias parameters\\
1008: $b_1^{L}$: linear bias parameter for the linear bias model with the 3rd-order
1009: matter power spectrum\\
1010: $b_1^{LL}$: linear bias parameter for the linear bias model with the
1011: linear power spectrum\\
1012: $b_1^{ST}$, $\tilde{b}_2^{ST}$: non-linear bias parameters calculated
1013: from the Sheth-Tormen model, $\tilde{b}_2^{ST}$=$b_2^{ST}/\tilde{b}_1$\\
1014: \textit{
1015: Caution: We estimate 1-$\sigma$ ranges for the low redshift ($z\le3$)
1016: only for the peak which involves the maximum likelihood value.
1017: If two peaks in maginalized likelihood function are blended, we use
1018: only unblended side of the peak to estimate the 1-$\sigma$ range.
1019: }
1020: }
1021: %
1022: %% No \tablerefs indicated
1023: %
1024: \end{deluxetable*}
1025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1026: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1027: \begin{figure}
1028: \centering
1029: \rotatebox{90}{
1030: \includegraphics[width=7cm]{fig6.ps}
1031: }
1032: \caption{
1033: One-dimensional marginalized distribution of non-linear bias parameters
1034: at $z=6$: from top to bottom panels, $P_0$, $\tilde{b}_2$, and $\tilde{b}_1$.
1035: Different lines show the different values of $k_{max}$ used for the
1036: fits. The dashed and solid lines correspond to
1037: $0.3\le k_{max}/[h~\mathrm{Mpc}^{-1}]\le1.0$ and $1.0<
1038: k_{max}/[h~\mathrm{Mpc}^{-1}]\le1.5$, respectively. The double-peak
1039: structure disappears for higher $k_{max}$.
1040: }%
1041: \label{fig6}
1042: \end{figure}
1043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1045: \begin{figure}
1046: \centering
1047: \rotatebox{90}{
1048: \includegraphics[width=7cm]{fig7.ps}
1049: }
1050: \caption{
1051: Same as Figure~\ref{fig6}, but for a Monte Carlo simulation
1052: of a galaxy survey with a bigger box size, $L_{box}=1.5~\mathrm{Gpc}/h$.
1053: }%
1054: \label{fig7}
1055: \end{figure}
1056: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1057: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1058: \begin{figure}
1059: \centering
1060: \rotatebox{90}{
1061: \includegraphics[width=7cm]{fig8.ps}
1062: }
1063: \caption{
1064: One-dimensional marginalized constraints and two-dimensional
1065: joint marginalized constraint of 2-$\sigma$ ($95.45\%$ CL)
1066: range for bias parameters
1067: ($\tilde{b}_1$,$\tilde{b}_2$,$P_0$).
1068: Covariance matrices are calculated from
1069: the Fisher information matrix
1070: (Eq. (\ref{eq:fisher_bias})) with the best-fitting bias
1071: parameters for halo at $z=4$.
1072: }%
1073: \label{fig8}
1074: \end{figure}
1075: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1076:
1077:
1078: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1079: \begin{figure*}
1080: \centering
1081: \rotatebox{90}{
1082: \includegraphics[width=10cm]{fig9.ps}
1083: }
1084: \caption{
1085: Distortion of BAOs due to non-linear matter clustering and non-linear halo bias.
1086: All of the power spectra have been divided by a smooth power
1087: spectrum without baryonic oscillations from equation~(29) of
1088: \cite{eisenstein/hu:1998}. The errorbars show the Millennium Simulation,
1089: while the solid lines show the PT calculations.
1090: The dashed lines show the linear bias model with the non-linear
1091: matter power spectrum, and the dot-dashed lines show the linear
1092: bias model with the linear matter power spectrum.
1093: Therefore, the difference between the solid lines and the dashed lines
1094: shows the distortion solely due to non-linear halo bias.
1095: }%
1096: \label{fig9}
1097: \end{figure*}
1098: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1099:
1100: Figure \ref{fig5} shows the best-fitting non-linear ({\it solid
1101: lines}) and linear bias
1102: models ({\it dashed and dot-dashed lines}), compared with the halo
1103: spectra estimated from the Millennium Simulation ({\it points with
1104: errorbars}). The smaller panels show the residuals of fits.
1105: The maximum wavenumber used in the fits, $k_{max}(z)$, are also marked with
1106: the arrows (bigger panels), and the vertical lines (smaller panels).
1107: We find that the non-linear bias model provides substantially better
1108: fits than the linear bias models.
1109:
1110: We find that all of non-linear bias parameters, $\tilde{b}_1$,
1111: $\tilde{b}_2$, and $P_0$, are strongly degenerate, when the maximum
1112: wavenumbers used in the fits, $k_{max}$, are small.
1113: In Figure~\ref{fig6} we show the one-dimensional marginalized
1114: distribution of bias parameters at $z=6$, as a function of
1115: $k_{max}$. For lower $k_{max}$, $0.3\le
1116: k_{max}/[h~\mathrm{Mpc}^{-1}]\le1.0$, the marginalized distribution has two
1117: peaks ({\it dashed lines}), indicating strong degeneracy with the other
1118: parameters. The double-peak structure disappears for
1119: $1.0< k_{max}/[h~\mathrm{Mpc}^{-1}]\le1.5$ ({\it solid lines}).
1120:
1121: We find that the origin of degeneracy is simply due to the small box size of the
1122: Millennium Simulation, i.e., the lack of statistics, or too a large
1123: sampling variance.
1124: To show this,
1125: we have generated a mock Monte Carlo realization of
1126: halo power spectra, assuming a much bigger box size,
1127: $L_{box}=1.5~h^{-1}~\mathrm{Gpc}$, which gives the fundamental frequency of
1128: $\Delta k=5.0\times10^{-4}~h~\mathrm{Mpc}^{-1}$.
1129: Note that this volume roughly corresponds to that would be surveyed by
1130: the HETDEX survey \citep{hill/etal:2004}.
1131: We have used the
1132: same non-linear matter power spectrum and the best-fitting bias
1133: parameters from the Millennium Simulation (MPA halos) when creating Monte Carlo
1134: realizations. The resulting marginalized likelihood function at $z=6$
1135: is shown in Figure~\ref{fig7}.
1136: The double-peak structure has disappeared even for
1137: low $k_{max}$, $k_{max}= 0.3~h~\mathrm{Mpc}^{-1}$.
1138: Therefore, we conclude that the double-peak problem
1139: can be resolved simply by increasing the survey volume.
1140:
1141: The best-fitting non-linear halo bias parameters and the corresponding
1142: 1-$\sigma$ intervals are summarized in Table \ref{table:halobias}.
1143: Since we know that the double-peak structure is spurious, we pick one
1144: peak that corresponds to the maximum likelihood value,
1145: and quote the 1-$\sigma$ interval.
1146: At $z\le 2$, the bias parameters are not constrined very well
1147: because of lower $k_{max}$ and the limited statistics of the
1148: Millennium Simulation, and hence the two peaks are blended;
1149: thus, we estimate 1-$\sigma$ range only from the unblended side of the
1150: marginalized likelihood function.
1151: %only when $k_{max}$ (given in
1152: %Table~\ref{table:kmax}) is high enough for the double-peak to disappear.
1153: %At $z\le 3$ the bias parameters are not constrained very well because of
1154: %lower $k_{max}$ and the
1155: %limited statistics of the Millennium Simulation; thus, we only
1156: %quote the best-fitting values.
1157: Two linear bias parameters, one with the linear matter power
1158: spectrum and another with the non-linear PT matter power spectrum, are
1159: also presented with their 1-$\sigma$ intervals.
1160:
1161: \subsubsection{Degeneracy of bias parameters}
1162: In order to see how strongly degenerate bias parameters are,
1163: we calculate the covariance matrix of each pair of bias
1164: parameters.
1165: We calculate the covariance matrix of each pair of bias parameters
1166: by using the Fisher information matrix, which is the inverse
1167: of the covariance matrix.
1168: The Fisher information matrix for the galaxy power spectrum
1169: can be approximated as
1170: \citep{tegmark:1997c}
1171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1172: \footnote{Eq. (\ref{eq:fisher_bias}) is equivalent to Eq. (6)
1173: in \citet{tegmark:1997c}. The number of $k$ mode in real space
1174: power spectrum from a survey of volume $V$ is
1175: (See Appendix A for notations.)
1176: $$
1177: N_{k_n}=
1178: \frac{4\pi k_n^2 \delta k_n}{2 (\delta k_n)^3}
1179: =\frac{Vk_n^2\delta k_n}{4\pi^2}.
1180: $$
1181: Then, the variance of power spectrum (Eq. (\ref{eq:varpk})) becomes
1182: $$
1183: \sigma_P^2(k_n)
1184: =
1185: \frac{4\pi^2}{Vk_n^2\delta k_n}\left[P(k_n)+\frac{1}{n}\right]^2
1186: =
1187: \frac{4\pi^2P(k_n)^2}{k_n^2\delta k_n}\frac{1}{V_{\mathrm{eff}}(k_n)},
1188: $$
1189: where $V_{\mathrm{eff}}$
1190: is the constant density version of Eq. (5) of \citet{tegmark:1997c}.
1191: Finally, the elements of Fisher matrix are given by
1192: \begin{eqnarray*}
1193: F_{ij}
1194: &=&\sum_n \frac{1}{\sigma_P^2(k_n)}
1195: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_i}
1196: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_j}\\
1197: &=&
1198: \frac{1}{4\pi^2}
1199: \sum_n
1200: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_i}
1201: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_j}
1202: \frac{V_{\mathrm{eff}}(k_n)k_n^2\delta k_n}{P(k_n)^2}
1203: \end{eqnarray*}
1204: which is the same as Eq. (6) in \citet{tegmark:1997c}.
1205: }
1206: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1207: \begin{equation}
1208: \label{eq:fisher_bias}
1209: F_{ij}=\sum_n \frac{1}{\sigma_P^2(k_n)}
1210: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_i}
1211: \frac{\partial P(k_n,\mathbf{\theta})}{\partial \theta_j}
1212: \end{equation}
1213: where $\mathbf{\theta}$ is a vector in the parameter space,
1214: $\theta_i=\tilde{b}_1$, $\tilde{b}_2$ ,$P_0$, for
1215: $i=1$, $2$, $3$, respectively.
1216: We calculate the marginalized errors on the bias parameters as following.
1217: We first calculate the full Fisher matrix and invert it to
1218: estimate the covariance matrix. Then, we get the
1219: the covariance matrices of any pairs of bias parameters by taking
1220: the $2$ by $2$ submatrix of the full covariance matrix.
1221: Figure \ref{fig8} shows the resulting 2-$\sigma$
1222: ($95.45\%$ interval) contour for the bias parameters at $z=4$.
1223: We find the strong degeneracy between $\tilde{P}_0$
1224: and $\tilde{b}_2$. We also find that
1225: $\tilde{b}_1$ is degenerate with the other two parameters.
1226: On top of the error contours for the Millennium Simulation,
1227: we show the expected contour from the HETDEX like survey
1228: ($1.5~\mathrm{Gpc}/h$). Since the volume of HETDEX like survey
1229: is 27 times bigger, the likelihood functions and the error-contours
1230: are about a factor of 5 smaller than those from the Millennium Simulation.
1231: Other than that, two contours follow the same trend.
1232: Results are the same for the other redshifts.
1233:
1234: \subsubsection{Comparison with the halo model predictions}
1235: The effective linear bias, $\tilde{b}_1$, is larger at
1236: higher redshifts.
1237: This is the expected result, as halos of mass greater than
1238: $\sim 10^{10} M_\odot$ were rarer in the earlier time,
1239: resulting in the larger bias.
1240:
1241: From the same reason, we expect that
1242: the non-linear bias parameters, $\tilde{b}_2$ and $P_0$, are also
1243: larger at higher $z$.
1244: While we observe the expected trend at $z\ge 4$, the results from
1245: $z\le 3$ are somewhat peculiar.
1246: This is probably due to the large sampling variance making
1247: the fits unstable: for $z\le 3$ the maximum wavenumbers inferred
1248: from the matter power spectra are less than $1.0~h~\mathrm{Mpc}^{-1}$
1249: (see Table~\ref{table:kmax}), which makes the likelihood function
1250: double-peaked and leaves the bias parameters poorly constrained.
1251:
1252: How do these bias parameters compare with the expected values?
1253: We use the halo model for computing the mass-averaged bias parameters,
1254: $b_1^{ST}$ and $b_2^{ST}$, assuming that the minimum mass
1255: is given by the minimum mass of the MPA halo catalogue,
1256: $M_{min}=1.72\times10^{10} M_\odot/h$:
1257: \begin{equation}
1258: b_i^{ST}=\frac
1259: {\int_{M_{min}}^{M_{max}}\frac{dn}{dM}M b_i(M) dM}
1260: {\int_{M_{min}}^{M_{max}}\frac{dn}{dM}MdM},
1261: \end{equation}
1262: where $dn/dM$ is the Sheth-Tormen mass function and $b_i(M)$ is the
1263: $i$-th order bias parameter from \citet{scoccimarro/etal:2001}.
1264:
1265: There is one subtlety. The halo model predicts the coefficients
1266: of the Taylor series (Eq.~(\ref{eq:Taylor_expansion})), whereas what we
1267: have measured are the re-parametrized bias parameters given by
1268: Eq.~(\ref{eq:tb1b2}).
1269: However, the formula for $\tilde{b}_1$ includes the mass variance,
1270: $\sigma^2$, which depends on our choice of a smoothing scale that is not
1271: well defined. This shows how difficult it is to actually compute the
1272: halo power spectrum from the halo model. While the measured values of
1273: $\tilde{b}_1$ and the predicted $b_1^{ST}$ compare reasonably well, it
1274: is clear that we cannot use the predicted bias values for doing cosmology.
1275:
1276: For $\tilde{b}_2$, we compute $\tilde{b}_2^{ST}=b_2^{ST}/\tilde{b}_1$
1277: where $\tilde{b}_1$ is the best-fitting value from the Millennium
1278: Simulation. This would give us a semi apple-to-apple comparison.
1279: Nevertheless, while the agreement is reasonable at $z\ge 4$, the halo model
1280: predictions should not be used for predicting $\tilde{b}_2$ either.
1281:
1282: \subsubsection{Comments on the bispectrum}
1283: \label{sec:bispectrum}
1284: While the degeneracy between bias parameters may appear to be a serious
1285: issue, there is actually a powerful way of breaking degeneracy:
1286: the bispectrum, the Fourier transform of the 3-point correlation
1287: function \citep{matarrese/verde/heavens:1997}. The reduced bispectrum, which is the bispectrum normalized
1288: properly by the power spectrum, depends primarily on two bias
1289: parameters, $\tilde{b}_1$ and $\tilde{b}_2$, nearly independent of the
1290: cosmological parameters \citep{sefusatti/etal:2006}.
1291: Therefore, one can use this property to fix the bias parameters,
1292: and use the power spectrum for determining the cosmological parameters
1293: and the remaining bias parameter, $P_0$. \citet{sefusatti/komatsu:2007}
1294: have shown that the planned high-$z$ galaxy surveys would be able to
1295: determine $\tilde{b}_1$ and $\tilde{b}_2$ with a few percent accuracy.
1296:
1297: We have begun studying the bispectrum of the Millennium Simulation.
1298: Our preliminary results show that we can indeed obtain
1299: better constraints on $\tilde{b}_1$ and $\tilde{b}_2$ from the
1300: bispectrum than from the power spectrum, provided that we use the same
1301: $k_{max}$ for both the bispectrum and power spectrum analyses.
1302: Therefore, even when the non-linear bias parameters are poorly
1303: constrained by the power spectrum alone, or have the double-peak
1304: likelihood function from the power spectrum for lower $k_{max}$, we can
1305: still find tight constraints on $\tilde{b}_1$ and $\tilde{b}_2$ from the
1306: bispectrum. These results will be reported elsewhere.
1307:
1308: \subsubsection{Effects on BAOs}
1309: In Figure~\ref{fig9} we show the distortion of BAO features
1310: due to non-linear matter clustering and non-linear bias.
1311: To show only the distortions of BAOs at each redshift,
1312: we have divided the halo power spectra by smooth power spectra without
1313: baryonic oscillations from equation (29) of \citet{eisenstein/hu:1998}
1314: with $\tilde{b}_1^2$ multiplied.
1315: Three theoretical models are shown:
1316: the non-linear bias model (\textit{solid line}),
1317: a linear bias model with the 3rd-order matter power spectrum
1318: (\textit{dashed line}), and a linear bias model with the linear matter
1319: power spectrum (\textit{dot-dashed line}).
1320: Therefore, the difference between the solid lines and the dashed lines
1321: is solely due to non-linear halo bias.
1322:
1323: The importance of non-linear bias affecting BAOs grows with $z$;
1324: however, as the matter clustering is weaker at higher $z$, the
1325: 3rd-order PT still performs better than at lower $z$.
1326: In other words, the higher bias at higher $z$ does not mean that surveys
1327: at higher $z$ are worse at measuring BAOs; on the contrary, it is still
1328: easier to model the halo power spectrum at higher $z$ than at lower
1329: $z$.
1330: For $z\ge3$, where $k_{max}$ is larger than the BAO scale,
1331: the distortion of BAOs is modeled very well by
1332: the non-linear bias model, while the linear bias models fail badly.
1333:
1334: The sampling variance of the Millennium Simulation at $k\lesssim
1335: 0.15~h~{\rm Mpc}^{-1}$ is too large for us to study the distortion on
1336: the first two BAO peaks. Since the PT performs well at higher $k$,
1337: we expect that the PT describes the first two peaks even better.
1338: However, to show this explicitly one would need to run a bigger
1339: simulation with a bigger volume with the same mass resolution as the
1340: Millennium Simulation, which should be entirely doable with the existing
1341: computing resources.
1342:
1343:
1344:
1345:
1346: \subsection{Galaxy power spectra}
1347: \label{sec:gal}
1348: \subsubsection{Measuring non-linear galaxy bias parameters}
1349: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1350: \begin{figure*}
1351: \centering
1352: \rotatebox{90}{
1353: \includegraphics[width=10cm]{fig10.ps}
1354: }
1355: \caption{
1356: Same as Figure~\ref{fig5}, but for the MPA galaxy catalogue (Mgalaxy).
1357: }%
1358: \label{fig10}
1359: \end{figure*}
1360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1361: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1362: \begin{figure*}
1363: \centering
1364: \rotatebox{90}{
1365: \includegraphics[width=10cm]{fig11.ps}
1366: }
1367: \caption{
1368: Same as Figure~\ref{fig5}, but for the Durham galaxy catalogue (Dgalaxy).
1369: }%
1370: \label{fig11}
1371: \end{figure*}
1372: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1373: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1374: %% The values (usually only l,r and c) in the last part of
1375: %% \begin{deluxetable}{} command tell LaTeX how many columns
1376: %% there are and how to align them.
1377: \begin{deluxetable*}{ccclcccc}
1378:
1379: %% Keep a portrait orientation
1380:
1381: %% Over-ride the default font size
1382: %% Use 10pt
1383: \tabletypesize{\footnotesize}
1384: %% Use \tablewidth{?pt} to over-ride the default table width.
1385: %% If you are unhappy with the default look at the end of the
1386: %% *.log file to see what the default was set at before adjusting
1387: %% this value.
1388:
1389: %% This is the title of the table.
1390: \tablecaption{Non-linear halo bias parameters and the corresponding 68\%
1391: interval estimated from the MPA galaxy power spectra
1392: }
1393:
1394: %% This command over-rides LaTeX's natural table count
1395: %% and replaces it with this number. LaTeX will increment
1396: %% all other tables after this table based on this number
1397: \tablenum{4}
1398: \label{table:Mgbias}
1399:
1400: %% The \tablehead gives provides the column headers. It
1401: %% is currently set up so that the column labels are on the
1402: %% top line and the units surrounded by ()s are in the
1403: %% bottom line. You may add more header information by writing
1404: %% another line between these lines. For each column that requries
1405: %% extra information be sure to include a \colhead{text} command
1406: %% and remember to end any extra lines with \\ and include the
1407: %% correct number of &s.
1408: \tablehead{\colhead{$z$} & \colhead{$\tilde{b}_1$} & \colhead{$\tilde{b}_2$} & \colhead{$P_0$} & \colhead{$b_1^L$} & \colhead{$b_1^{LL}$} & \colhead{$b_1^{ST}$} & \colhead{$\tilde{b}_2^{ST}$} \\
1409: \colhead{} & \colhead{} & \colhead{} & \colhead{($[h/\mathrm{Mpc}]^3$)} & \colhead{} & \colhead{} & \colhead{} & \colhead{} }
1410:
1411: %% All data must appear between the \startdata and \enddata commands
1412: \startdata
1413: 6 & 3.55$\pm$0.01 & 1.70$\pm$0.03 & 194.23$\pm$4.45 & 3.67$\pm$0.03 & 3.68$\pm$0.03 & 3.10 & 1.03 \\
1414: 5 & 2.93$\pm$0.01 & 1.08$\pm$0.03 & 94.08$\pm$3.71 & 2.97$\pm$0.03 & 2.98$\pm$0.03 & 2.55 & 0.59 \\
1415: 4 & 2.46$\pm$0.01 & 0.68$\pm$0.03 & 47.79$\pm$2.84 & 2.47$\pm$0.02 & 2.48$\pm$0.02 & 2.13 & 0.28 \\
1416: 3 & 1.69$\pm$0.01 & -2.12$\pm$0.04 & 486.69$\pm$12.7 & 1.83$\pm$0.02 & 1.83$\pm$0.02 & 1.58 & -0.12 \\
1417: 2 & 1.28$\pm$0.08 & -2.16$\pm$0.64 & 738.22$\pm$291.3 & 1.40$\pm$0.01 & 1.40$\pm$0.01 & 1.19 & -0.34 \\
1418: 1 & 0.89$\pm$0.11 & -2.97$\pm$1.60 & 2248.35$\pm$786.13 & 1.09$\pm$0.01 & 1.10$\pm$0.01 & 0.91 & -0.45
1419:
1420:
1421: \enddata
1422:
1423: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
1424: %% or \tablecomments{text} between the \enddata and
1425: %% \end{deluxetable} commands
1426:
1427: %% General table comment marker
1428: \tablecomments{$z$: redshift\\
1429: $\tilde{b}_1$, $\tilde{b}_2$, $P_0$: non-linear bias parameters\\
1430: $b_1^{L}$: linear bias parameter for the linear bias model with the 3rd-order
1431: matter power spectrum\\
1432: $b_1^{LL}$: linear bias parameter for the linear bias model with the
1433: linear power spectrum\\
1434: $b_1^{ST}$, $\tilde{b}_2^{ST}$: non-linear bias parameters calculated
1435: from the Sheth-Tormen model, $\tilde{b}_2^{ST}$=$b_2^{ST}/\tilde{b}_1$
1436: \\
1437: \textit{
1438: Caution: We estimate 1-$\sigma$ ranges for the low redshift ($z\le3$)
1439: only for the peak which involves the maximum likelihood value.
1440: If two peaks in maginalized likelihood function are blended, we use
1441: only unblended side of the peak to estimate the 1-$\sigma$ range.
1442: }
1443: }
1444:
1445: %% No \tablerefs indicated
1446: \end{deluxetable*}
1447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1448: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1449: %% The values (usually only l,r and c) in the last part of
1450: %% \begin{deluxetable}{} command tell LaTeX how many columns
1451: %% there are and how to align them.
1452: \begin{deluxetable*}{ccclcccc}
1453:
1454: %% Keep a portrait orientation
1455:
1456: %% Over-ride the default font size
1457: %% Use 10pt
1458: \tabletypesize{\footnotesize}
1459: %% Use \tablewidth{?pt} to over-ride the default table width.
1460: %% If you are unhappy with the default look at the end of the
1461: %% *.log file to see what the default was set at before adjusting
1462: %% this value.
1463:
1464: %% This is the title of the table.
1465: \tablecaption{Non-linear halo bias parameters and the corresponding 68\%
1466: interval estimated from the Durham galaxy power spectra
1467: }
1468:
1469:
1470: %% This command over-rides LaTeX's natural table count
1471: %% and replaces it with this number. LaTeX will increment
1472: %% all other tables after this table based on this number
1473: \tablenum{5}
1474: \label{table:Dgbias}
1475:
1476: %% The \tablehead gives provides the column headers. It
1477: %% is currently set up so that the column labels are on the
1478: %% top line and the units surrounded by ()s are in the
1479: %% bottom line. You may add more header information by writing
1480: %% another line between these lines. For each column that requries
1481: %% extra information be sure to include a \colhead{text} command
1482: %% and remember to end any extra lines with \\ and include the
1483: %% correct number of &s.
1484: \tablehead{\colhead{$z$} & \colhead{$\tilde{b}_1$} & \colhead{$\tilde{b}_2$} & \colhead{$P_0$} & \colhead{$b_1^L$} & \colhead{$b_1^{LL}$} & \colhead{$b_1^{ST}$} & \colhead{$\tilde{b}_2^{ST}$} \\
1485: \colhead{} & \colhead{} & \colhead{} & \colhead{($[h/\mathrm{Mpc}]^3$)} & \colhead{} & \colhead{} & \colhead{} & \colhead{} }
1486:
1487: %% All data must appear between the \startdata and \enddata commands
1488: \startdata
1489: 6 & 3.73$\pm$0.01 & 1.96$\pm$0.03 & 288.39$\pm$5.82 & 3.90$\pm$0.04 & 3.90$\pm$0.04 & 3.10 & 0.98 \\
1490: 5 & 3.07$\pm$0.01 & 1.26$\pm$0.03 & 143.15$\pm$4.81 & 3.15$\pm$0.03 & 3.15$\pm$0.03 & 2.55 & 0.56 \\
1491: 4 & 2.57$\pm$0.01 & 0.83$\pm$0.03 & 78.97$\pm$3.93 & 2.60$\pm$0.02 & 2.61$\pm$0.02 & 2.13 & 0.26 \\
1492: 3 & 1.75$\pm$0.01 & -2.26$\pm$0.04 & 604.65$\pm$13.8 & 1.92$\pm$0.02 & 1.93$\pm$0.02 & 1.58 & -0.11 \\
1493: 2 & 1.36$\pm$0.08 & -2.14$\pm$0.65 & 843.49$\pm$331.4 & 1.49$\pm$0.01 & 1.50$\pm$0.01 & 1.19 & -0.32 \\
1494: 1 & 0.96$\pm$0.11 & -2.94$\pm$1.62 & 2640.20$\pm$960.32 & 1.18$\pm$0.01 & 1.20$\pm$0.01 & 0.91 & -0.42
1495: \enddata
1496:
1497: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
1498: %% or \tablecomments{text} between the \enddata and
1499: %% \end{deluxetable} commands
1500:
1501: %% General table comment marker
1502: \tablecomments{$z$: redshift\\
1503: $\tilde{b}_1$, $\tilde{b}_2$, $P_0$: non-linear bias parameters\\
1504: $b_1^{L}$: linear bias parameter for the linear bias model with the 3rd-order
1505: matter power spectrum\\
1506: $b_1^{LL}$: linear bias parameter for the linear bias model with the
1507: linear power spectrum\\
1508: $b_1^{ST}$, $\tilde{b}_2^{ST}$: non-linear bias parameters calculated
1509: from the Sheth-Tormen model, $\tilde{b}_2^{ST}$=$b_2^{ST}/\tilde{b}_1$
1510: \\
1511: \textit{
1512: Caution: We estimate 1-$\sigma$ ranges for the low redshift ($z\le3$)
1513: only for the peak which involves the maximum likelihood value.
1514: If two peaks in maginalized likelihood function are blended, we use
1515: only unblended side of the peak to estimate the 1-$\sigma$ range.
1516: }
1517: }
1518:
1519: %% No \tablerefs indicated
1520: \end{deluxetable*}
1521: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1522: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1523: \begin{figure*}
1524: \centering
1525: \rotatebox{90}{
1526: \includegraphics[width=10cm]{fig12.ps}
1527: }
1528: \caption{
1529: Same as Figure~\ref{fig9}, but for the MPA galaxy power spectrum
1530: (Mgalaxy).
1531: }%
1532: \label{fig12}
1533: \end{figure*}
1534: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1536: \begin{figure*}
1537: \centering
1538: \rotatebox{90}{
1539: \includegraphics[width=10cm]{fig13.ps}
1540: }
1541: \caption{
1542: Same as Figure~\ref{fig9}, but for the Durham galaxy power spectrum
1543: (Dgalaxy).
1544: }%
1545: \label{fig13}
1546: \end{figure*}
1547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1548:
1549: Figures \ref{fig10} and \ref{fig11} show the galaxy power spectra
1550: estimated from the MPA (Mgalaxy) and Durham (Dgalaxy) galaxy catalogues,
1551: respectively.
1552: Here, we basically find the same story as we have found for the halo
1553: power spectra (\S~\ref{sec:halo}): for $k<k_{max}$
1554: the non-linear bias model fits both galaxy power spectra (Mgalaxy and
1555: Dgalaxy), whereas the linear bias models fit neither.
1556:
1557: The galaxy bias parameters extracted from Mgalaxy and Dgalaxy are summarized in
1558: Table \ref{table:Mgbias} and \ref{table:Dgbias}, respectively.
1559: While the bias parameters are different for halo, Mgalaxy and Dgalaxy,
1560: they follow the same trend: (i) $\tilde{b}_1$ becomes lower as the redshift
1561: becomes lower, and (ii) $\tilde{b}_2$ also becomes lower as the redshift
1562: becomes lower when $z>3$, but suddenly changes to large negative values at
1563: $z\le 3$. As we have already pointed out in \S~\ref{sec:halo}, this
1564: sudden peculiar change is most likely caused by the double-peak nature
1565: of the likelihood function, owing to the poor statistical power for
1566: lower $k_{max}$ at lower $z$. In order to study $\tilde{b}_2$ further
1567: with better statistics, one needs a bigger simulation.
1568:
1569: \subsubsection{Comparison with the simplest HOD predictions}
1570: To give a rough theoretical guide for the galaxy bias parameters,
1571: we assume that each dark matter halo hosts one galaxy above a certain
1572: minimum mass.
1573: This specifies the form of the HOD completely:
1574: $\langle N|M\rangle=1$, with the same lower mass cut-off as the
1575: minimum mass of the halo, $M_{min}=1.72\times10^{10} M_\odot/h$.
1576:
1577: This is utterly simplistic, and is probably not correct for describing
1578: Mgalaxy or Dgalaxy. Nevertheless, we give the resulting values in
1579: Table \ref{table:Mgbias} and \ref{table:Dgbias}, which have been
1580: computed from
1581: \begin{equation}
1582: b_i^{ST}=\frac
1583: {\int_{M_{min}}^{M_{max}}\frac{dn}{dM}b_i(M)\langle N|M\rangle dM}
1584: {\int_{M_{min}}^{M_{max}}\frac{dn}{dM}\langle N|M\rangle dM},
1585: \end{equation}
1586: where $dn/dM$ is the Sheth-Tormen mass function and $b_i(M)$ is the
1587: $i$-th order bias parameter from \citet{scoccimarro/etal:2001}.
1588: To compare with the non-linear bias parameters, we also calculate
1589: $\tilde{b}_2=b_2^{ST}/\tilde{b}_1$.
1590:
1591: While these ``predictions'' give values that are reasonably close to the
1592: ones obtained from the fits, they are many $\sigma$ away from the
1593: best-fitting values. The freedom in the choice of the HOD may be used to
1594: make the predicted values match the best-fitting values; however, such
1595: an approach would require at least as many free parameters as the
1596: non-linear bias parameters. Also, given that the {\it halo} bias
1597: prediction fails to fit the halo power spectra, the HOD approach, which
1598: is still based upon knowing the halo bias, is bound to fail as well.
1599:
1600: \subsubsection{Effects on BAOs}
1601: In Figures \ref{fig12} and \ref{fig13} we show
1602: how non-linear galaxy bias distorts the structure of BAOs.
1603: Again, we find the same story as we have found for the halo bias:
1604: the galaxy bias distorts BAOs more at higher $z$ because, for a given
1605: mass, galaxies were rarer at higher redshifts and thus more highly
1606: biased, while the quality of the fits is better at higher $z$ because of
1607: less non-linearity in the matter clustering.
1608:
1609: In all cases (halo, Mgalaxy and Dgalaxy) the non-linear bias model given
1610: by Eq.~(\ref{eq:3rd_PT_Pk}) provides very good fits, and describes how
1611: bias modifies BAOs.
1612:
1613:
1614:
1615:
1616:
1617:
1618:
1619: \subsection{Mass dependence of bias parameters and effects on
1620: BAOs}\label{sec:mass_dependence}
1621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1622: \begin{figure*}
1623: \centering
1624: \rotatebox{90}{
1625: \includegraphics[width=10cm]{fig14.ps}
1626: }
1627: \caption{
1628: Mass dependence of distortion of BAOs due to non-linear bias.
1629: Four mass bins, $M<5\times10^{10}M_\odot/h$,
1630: $5\times10^{10}M_\odot/h<M<10^{11}M_\odot/h$,
1631: $10^{11}M_\odot/h<M<5\times10^{11}M_\odot/h$, and
1632: $5\times10^{11}M_\odot/h<M<10^{12}M_\odot/h$, are shown.
1633: ($M_{10}$ stands for $M/(10^{10}M_\sun)$.)
1634: All of the power spectra have been divided by a smooth power spectrum
1635: without baryonic oscillations from equation~(29) of \cite{eisenstein/hu:1998}.
1636: The errorbars show the Millennium Simulation data, while the solid
1637: lines show the PT calculation.
1638: }%
1639: \label{fig14}
1640: \end{figure*}
1641: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1642: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1643: \begin{figure*}
1644: \centering
1645: \rotatebox{90}{
1646: \includegraphics[width=10cm]{fig15.ps}
1647: }
1648: \caption{
1649: Same as Figure~\ref{fig14}, but for the MPA galaxy catalogue (Mgalaxy).
1650: }%
1651: \label{fig15}
1652: \end{figure*}
1653: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1654:
1655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1656: \begin{deluxetable*}{cccccccc}
1657:
1658: %% Keep a portrait orientation
1659:
1660: %% Over-ride the default font size
1661: %% Use Default (12pt)
1662:
1663: %% Use \tablewidth{?pt} to over-ride the default table width.
1664: %% If you are unhappy with the default look at the end of the
1665: %% *.log file to see what the default was set at before adjusting
1666: %% this value.
1667:
1668: %% This is the title of the table.
1669: \tablecaption{Mass dependence of non-linear halo bias parameters (MPA halos)}
1670: \tabletypesize{\footnotesize}
1671:
1672: %% This command over-rides LaTeX's natural table count
1673: %% and replaces it with this number. LaTeX will increment
1674: %% all other tables after this table based on this number
1675: \tablenum{6}
1676: \label{table:Mhbias_mdep}
1677:
1678: %% The \tablehead gives provides the column headers. It
1679: %% is currently set up so that the column labels are on the
1680: %% top line and the units surrounded by ()s are in the
1681: %% bottom line. You may add more header information by writing
1682: %% another line between these lines. For each column that requries
1683: %% extra information be sure to include a \colhead{text} command
1684: %% and remember to end any extra lines with \\ and include the
1685: %% correct number of &s.
1686: \tablehead{\colhead{$z$} & \colhead{$\mathrm{M_{min}}$} & \colhead{$\mathrm{M_{max}}$} & \colhead{$\tilde{b}_1$} & \colhead{$\tilde{b}_2$} & \colhead{$P_0$} & \colhead{$b_1^{ST}$} & \colhead{$\tilde{b}_2^{ST}$} \\
1687: \colhead{} & \colhead{($M_\odot/h$)} & \colhead{($M_\odot/h$)} & \colhead{} & \colhead{} & \colhead{($[h/\mathrm{Mpc}]^3$)} & \colhead{} & \colhead{} }
1688:
1689: %% All data must appear between the \startdata and \enddata commands
1690: \startdata
1691: 6 & 1.7E+10 & 5.0E+10 & 3.19$\pm$0.01 & 1.28$\pm$0.03 & 88.76$\pm$2.97 & 2.96 & 0.93 \\
1692: & 5.0E+10 & 1.0E+11 & 3.90$\pm$0.02 & 1.91$\pm$0.04 & 288.18$\pm$8.02 & 3.52 & 1.36 \\
1693: & 1.0E+11 & 5.0E+11 & 4.66$\pm$0.03 & 3.04$\pm$0.05 & 1029.19$\pm$18.84 & 4.41 & 2.28 \\
1694: & 5.0E+11 & 1.0E+12 & 6.41$\pm$0.14 & 5.76$\pm$0.21 & 6910.17$\pm$200.74 & 5.95 & 3.59 \\
1695: \hline
1696: 5 & 1.7E+10 & 5.0E+10 & 2.55$\pm$0.01 & 0.71$\pm$0.03 & 31.51$\pm$2.06 & 2.41 & 0.48 \\
1697: & 5.0E+10 & 1.0E+11 & 3.09$\pm$0.01 & 1.19$\pm$0.04 & 120.84$\pm$5.69 & 2.84 & 0.81 \\
1698: & 1.0E+11 & 5.0E+11 & 3.78$\pm$0.02 & 1.79$\pm$0.04 & 402.11$\pm$12.40 & 3.55 & 1.48 \\
1699: & 5.0E+11 & 1.0E+12 & 5.14$\pm$0.07 & 3.55$\pm$0.11 & 2805.48$\pm$94.53 & 4.71 & 2.44 \\
1700: \hline
1701: 4 & 1.7E+10 & 5.0E+10 & 2.08$\pm$0.01 & 0.38$\pm$0.04 & 10.90$\pm$1.19 & 2.01 & 0.15 \\
1702: & 5.0E+10 & 1.0E+11 & 2.51$\pm$0.01 & 0.66$\pm$0.04 & 42.34$\pm$3.52 & 2.33 & 0.40 \\
1703: & 1.0E+11 & 5.0E+11 & 3.05$\pm$0.01 & 1.08$\pm$0.04 & 161.22$\pm$8.11 & 2.90 & 0.92 \\
1704: & 5.0E+11 & 1.0E+12 & 3.80$\pm$0.05 & -4.08$\pm$0.09 & 3431.19$\pm$64.81 & 3.79 & 1.77 \\
1705: \hline
1706: 3 & 1.7E+10 & 5.0E+10 & 1.39$\pm$0.01 & -1.83$\pm$0.05 & 241.59$\pm$9.58 & 1.47 & -0.25 \\
1707: & 5.0E+10 & 1.0E+11 & 1.75$\pm$0.01 & 0.11$\pm$0.05 & 2.48$\pm$0.29 & 1.67 & -0.10 \\
1708: & 1.0E+11 & 5.0E+11 & 2.09$\pm$0.01 & 0.35$\pm$0.04 & 20.95$\pm$3.22 & 2.04 & 0.19 \\
1709: & 5.0E+11 & 1.0E+12 & 2.78$\pm$0.02 & 0.82$\pm$0.06 & 171.31$\pm$21.03 & 2.57 & 0.60 \\
1710: \hline
1711: 2 & 1.7E+10 & 5.0E+10 & 1.01$\pm$0.05 & -1.98$\pm$0.68 & 373.60$\pm$149.24 & 1.11 & -0.46 \\
1712: & 5.0E+10 & 1.0E+11 & 1.14$\pm$0.07 & -2.30$\pm$0.63 & 627.69$\pm$204.69 & 1.23 & -0.40 \\
1713: & 1.0E+11 & 5.0E+11 & 1.31$\pm$0.08 & -2.34$\pm$0.63 & 869.30$\pm$272.06 & 1.44 & -0.28 \\
1714: & 5.0E+11 & 1.0E+12 & 1.62$\pm$0.11 & -2.53$\pm$0.69 & 1566.40$\pm$476.07 & 1.75 & -0.05 \\
1715: \hline
1716: 1 & 1.7E+10 & 5.0E+10 & 0.68$\pm$0.09 & -3.12$\pm$1.46 & 1315.40$\pm$447.14 & 0.86 & -0.58 \\
1717: & 5.0E+10 & 1.0E+11 & 0.75$\pm$0.10 & -3.24$\pm$1.46 & 1699.76$\pm$571.75 & 0.92 & -0.56 \\
1718: & 1.0E+11 & 5.0E+11 & 0.85$\pm$0.09 & -2.80$\pm$1.65 & 1783.07$\pm$683.57 & 1.02 & -0.53 \\
1719: & 5.0E+11 & 1.0E+12 & 0.99$\pm$0.11 & -2.82$\pm$1.95 & 2443.76$\pm$972.16 & 1.17 & -0.47
1720: \enddata
1721:
1722: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
1723: %% or \tablecomments{text} between the \enddata and
1724: %% \end{deluxetable} commands
1725:
1726: %% No \tablecomments indicated
1727: \tablecomments{$z$: redshift\\
1728: $\mathrm{M_{min}}$: minimum mass for a given bin\\
1729: $\mathrm{M_{max}}$: maximum mass for a given bin\\
1730: $\tilde{b}_1$, $\tilde{b}_2$, $P_0$: non-linear bias parameters\\
1731: $b_1^{ST}$, $\tilde{b}_2^{ST}$: bias parameters from the Sheth-Tormen model, $\tilde{b}_2^{ST}$=$b_2^{ST}/\tilde{b}_1$
1732: \\
1733: \textit{
1734: Caution: We estimate 1-$\sigma$ ranges for the low redshift ($z\le3$)
1735: only for the peak which involves the maximum likelihood value.
1736: If two peaks in maginalized likelihood function are blended, we use
1737: only unblended side of the peak to estimate the 1-$\sigma$ range.
1738: }
1739: }
1740:
1741: %% No \tablerefs indicated
1742: \end{deluxetable*}
1743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1744:
1745: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1746: \begin{deluxetable*}{cccccccc}
1747:
1748: %% Keep a portrait orientation
1749:
1750: %% Over-ride the default font size
1751: %% Use 8pt
1752: \tabletypesize{\footnotesize}
1753:
1754: %% Use \tablewidth{?pt} to over-ride the default table width.
1755: %% If you are unhappy with the default look at the end of the
1756: %% *.log file to see what the default was set at before adjusting
1757: %% this value.
1758:
1759: %% This is the title of the table.
1760: \tablecaption{Mass dependence of non-linear galaxy bias parameters (MPA
1761: galaxies)}
1762:
1763: %% This command over-rides LaTeX's natural table count
1764: %% and replaces it with this number. LaTeX will increment
1765: %% all other tables after this table based on this number
1766: \tablenum{7}
1767: \label{table:Mgbias_mdep}
1768:
1769: %% The \tablehead gives provides the column headers. It
1770: %% is currently set up so that the column labels are on the
1771: %% top line and the units surrounded by ()s are in the
1772: %% bottom line. You may add more header information by writing
1773: %% another line between these lines. For each column that requries
1774: %% extra information be sure to include a \colhead{text} command
1775: %% and remember to end any extra lines with \\ and include the
1776: %% correct number of &s.
1777: \tablehead{\colhead{$z$} & \colhead{$\mathrm{M_{min}}$} & \colhead{$\mathrm{M_{max}}$} & \colhead{$\tilde{b}_1$} & \colhead{$\tilde{b}_2$} & \colhead{$P_0$} & \colhead{$b_1^{ST}$} & \colhead{$\tilde{b}_2^{ST}$} \\
1778: \colhead{} & \colhead{($M_\odot/h$)} & \colhead{($M_\odot/h$)} & \colhead{} & \colhead{} & \colhead{($[h/\mathrm{Mpc}]^3$)} & \colhead{} & \colhead{} }
1779:
1780: %% All data must appear between the \startdata and \enddata commands
1781: \startdata
1782: 6 & 1.7E+10 & 5.0E+10 & 3.37$\pm$0.01 & 1.50$\pm$0.03 & 136.39$\pm$3.69 & 2.91 & 0.82 \\
1783: & 5.0E+10 & 1.0E+11 & 3.96$\pm$0.02 & 2.00$\pm$0.04 & 325.38$\pm$8.44 & 3.49 & 1.31 \\
1784: & 1.0E+11 & 5.0E+11 & 4.69$\pm$0.03 & 3.09$\pm$0.05 & 1078.72$\pm$19.25 & 4.23 & 2.01 \\
1785: & 5.0E+11 & 1.0E+12 & 6.43$\pm$0.14 & 5.79$\pm$0.20 & 7046.28$\pm$201.94 & 5.89 & 3.49 \\
1786: \hline
1787: 5 & 1.7E+10 & 5.0E+10 & 2.77$\pm$0.01 & 0.93$\pm$0.03 & 63.17$\pm$2.99 & 2.38 & 0.40 \\
1788: & 5.0E+10 & 1.0E+11 & 3.16$\pm$0.01 & 1.27$\pm$0.04 & 144.20$\pm$6.16 & 2.82 & 0.77 \\
1789: & 1.0E+11 & 5.0E+11 & 3.81$\pm$0.02 & 1.84$\pm$0.04 & 432.51$\pm$12.80 & 3.41 & 1.28 \\
1790: & 5.0E+11 & 1.0E+12 & 5.15$\pm$0.07 & 3.60$\pm$0.11 & 2897.95$\pm$95.17 & 4.67 & 2.37 \\
1791: \hline
1792: 4 & 1.7E+10 & 5.0E+10 & 2.33$\pm$0.01 & 0.58$\pm$0.03 & 32.25$\pm$2.25 & 1.98 & 0.11 \\
1793: & 5.0E+10 & 1.0E+11 & 2.59$\pm$0.01 & 0.74$\pm$0.04 & 56.91$\pm$4.08 & 2.32 & 0.37 \\
1794: & 1.0E+11 & 5.0E+11 & 3.09$\pm$0.02 & 1.13$\pm$0.04 & 179.81$\pm$8.52 & 2.79 & 0.77 \\
1795: & 5.0E+11 & 1.0E+12 & 3.83$\pm$0.05 & -4.09$\pm$0.09 & 3507.05$\pm$64.85 & 3.76 & 1.71 \\
1796: \hline
1797: 3 & 1.7E+10 & 5.0E+10 & 1.62$\pm$0.01 & -2.07$\pm$0.05 & 431.79$\pm$12.04 & 1.45 & -0.22 \\
1798: & 5.0E+10 & 1.0E+11 & 1.84$\pm$0.01 & 0.19$\pm$0.04 & 7.04$\pm$1.04 & 1.66 & -0.10 \\
1799: & 1.0E+11 & 5.0E+11 & 2.14$\pm$0.01 & 0.38$\pm$0.04 & 27.64$\pm$3.67 & 1.96 & 0.12 \\
1800: & 5.0E+11 & 1.0E+12 & 2.80$\pm$0.02 & 0.84$\pm$0.06 & 191.24$\pm$21.65 & 2.55 & 0.57 \\
1801: \hline
1802: 2 & 1.7E+10 & 5.0E+10 & 1.26$\pm$0.07 & -2.09$\pm$0.66 & 683.11$\pm$240.40 & 1.10 & -0.37 \\
1803: & 5.0E+10 & 1.0E+11 & 1.21$\pm$0.08 & -2.35$\pm$0.62 & 738.09$\pm$231.05 & 1.22 & -0.38 \\
1804: & 1.0E+11 & 5.0E+11 & 1.35$\pm$0.09 & -2.32$\pm$0.63 & 919.65$\pm$288.79 & 1.40 & -0.29 \\
1805: & 5.0E+11 & 1.0E+12 & 1.65$\pm$0.11 & -2.50$\pm$0.69 & 1602.17$\pm$480.23 & 1.74 & -0.06 \\
1806: \hline
1807: 1 & 1.7E+10 & 5.0E+10 & 0.91$\pm$0.11 & -2.96$\pm$1.59 & 2344.13$\pm$802.55 & 0.86 & -0.43 \\
1808: & 5.0E+10 & 1.0E+11 & 0.79$\pm$0.11 & -3.28$\pm$1.51 & 1956.42$\pm$657.89 & 0.92 & -0.53 \\
1809: & 1.0E+11 & 5.0E+11 & 0.87$\pm$0.10 & -2.88$\pm$1.63 & 1964.64$\pm$738.71 & 1.00 & -0.51 \\
1810: & 5.0E+11 & 1.0E+12 & 1.01$\pm$0.11 & -2.80$\pm$2.01 & 2550.21$\pm$1007.23 & 1.16 & -0.46
1811: \enddata
1812:
1813: %% Include any \tablenotetext{key}{text}, \tablerefs{ref list},
1814: %% or \tablecomments{text} between the \enddata and
1815: %% \end{deluxetable} commands
1816:
1817: %% No \tablecomments indicated
1818: \tablecomments{$z$: redshift\\
1819: $\mathrm{M_{min}}$: minimum mass for a given bin\\
1820: $\mathrm{M_{max}}$: maximum mass for a given bin\\
1821: $\tilde{b}_1$, $\tilde{b}_2$, $P_0$: non-linear bias parameters\\
1822: $b_1^{ST}$, $\tilde{b}_2^{ST}$: bias parameters from the Sheth-Tormen model, $\tilde{b}_2^{ST}$=$b_2^{ST}/\tilde{b}_1$
1823: \\
1824: \textit{
1825: Caution: We estimate 1-$\sigma$ ranges for the low redshift ($z\le3$)
1826: only for the peak which involves the maximum likelihood value.
1827: If two peaks in maginalized likelihood function are blended, we use
1828: only unblended side of the peak to estimate the 1-$\sigma$ range.
1829: }
1830: }
1831: %% No \tablerefs indicated
1832:
1833: \end{deluxetable*}
1834: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1835:
1836: So far, we have used all the available halos and galaxies in the
1837: Millennium catalogues for computing the halo and galaxy power spectra.
1838: In this section we divide the samples into different mass bins given by
1839: $M<5\times10^{10}M_\odot/h$,
1840: $5\times10^{10}M_\odot/h<M<10^{11}M_\odot/h$,
1841: $10^{11}M_\odot/h<M<5\times10^{11}M_\odot/h$,
1842: $5\times10^{11}M_\odot/h<M<10^{12}M_\odot/h$,
1843: and study how the derived bias parameters depend on mass.
1844:
1845: The power spectra of the selected halos and galaxies in a given mass bin
1846: are calculated and fit in the exactly same manner as before.
1847: Note that we shall use only the halo and Mgalaxy, as we expect
1848: that Dgalaxy would give similar results to Mgalaxy.
1849:
1850: Figures \ref{fig14} and \ref{fig15}
1851: show the results for the halo and galaxies, respectively.
1852: To compare the power spectra of different mass bins in the same
1853: panel, and highlight the effects on BAOs at the same time,
1854: we have divided the power spectra by a non-oscillating matter power
1855: spectrum from equation (29) of \citet{eisenstein/hu:1998} with
1856: the best-fitting $\tilde{b}_1^2$ from each mass bin multiplied.
1857: These figures show the expected results: the larger the mass is, the
1858: larger the non-linear bias becomes. Nevertheless, the 3rd-order PT
1859: calculation captures the dependence on mass well, and there is no
1860: evidence for failure of the PT for highly biased objects.
1861:
1862: In Tables \ref{table:Mhbias_mdep} and \ref{table:Mgbias_mdep} we give
1863: values of the measured bias parameters as well as the ``predicted'' values.
1864: For all redshifts we see the expected trend again: the higher the mass
1865: is, the larger the effective linear bias ($\tilde{b}_1$) is.
1866: The same is true for $\tilde{b}_2$ for $z>3$,
1867: while it is not as apparent for lower redshifts,
1868: and eventually becomes almost fuzzy for $z=1$. Again, these are probably
1869: due to the lack of statistics due to lower values of $k_{max}$ at lower
1870: $z$, and we need a bigger simulation to handle these cases with more
1871: statistics.
1872:
1873: The high values of bias do not mean failure of PT. The PT galaxy power
1874: spectrum model fails only when $\Delta^2_m(k,z)$ exceeds $\sim 0.4$ (Paper
1875: I), or the locality of bias is violated.
1876: Overall, we find that the non-linear bias model given by
1877: Eq.~(\ref{eq:3rd_PT_Pk}) performs well for halos and galaxies with
1878: all mass bins, provided that we use the data only up to $k_{max}$
1879: determined from the matter power spectra. This implies that the locality
1880: assumption is a good approximation for $k<k_{max}$; however, is it good
1881: enough for us to extract cosmology from the observed galaxy power
1882: spectra?
1883:
1884:
1885:
1886:
1887:
1888:
1889:
1890: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1891: \section{Cosmological parameter estimation with the non-linear bias
1892: model}
1893: \label{sec:bias}
1894: In the previous sections we have shown that the 3rd-order PT galaxy
1895: power spectrum given by
1896: Eq.~(\ref{eq:3rd_PT_Pk}) provides good fits to the galaxy power spectrum
1897: data from the Millennium Simulation.
1898:
1899: However, we must not forget that Eq.~(\ref{eq:3rd_PT_Pk}) contains {\it
1900: 3 free parameters}, $\tilde{b}_1$, $\tilde{b}_2$, and $P_0$. With 3
1901: parameters it may seem that it should not be so difficult to fit smooth
1902: curves like those shown in, e.g., Figure~\ref{fig10}.
1903:
1904: While the quality of fits is important, it is not the end of story. We
1905: must also show that Eq.~(\ref{eq:3rd_PT_Pk}) can be used for extracting the
1906: {\it correct cosmological parameters} from the observed galaxy power
1907: spectra.
1908:
1909: In this section we shall extract the distance scale from the
1910: galaxy power spectra of the Millennium Simulation, and compare them
1911: with the input values that were used to generate the simulation.
1912: If they do not agree, Eq.~(\ref{eq:3rd_PT_Pk}) must be discarded. If
1913: they do, we should proceed to the next level by including non-linear
1914: redshift space distortion.
1915:
1916: \subsection{Measuring Distance Scale}
1917: \label{sec:da}
1918: \subsubsection{Background}
1919: Dark energy influences the expansion rate of the universe as well as the
1920: growth of structure \citep[see][for a recent
1921: review]{copeland/sami/tsujikawa:2006}.
1922:
1923: The cosmological distances, such as the luminosity distance, $D_L(z)$, and angular
1924: diameter distance, $D_A(z)$, are powerful tools for measuring the expansion rates
1925: of the universe, $H(z)$, over a wide range of redshifts.
1926: Indeed, it was $D_L(z)$ measured out to high-$z$ ($z\le
1927: 1.7$) Type Ia
1928: supernovae that gave rise to the first compelling evidence for the
1929: existence of dark energy \citep{riess/etal:1998,perlmutter/etal:1999}.
1930: The CMB power spectrum provides us with a high-precision measurement of $D_A(z_*)$
1931: out to the photon decoupling epoch,
1932: $z_*\simeq 1090$ \citep[see][for the latest determination from the WMAP
1933: 5-year data]{komatsu/etal:prep}.
1934:
1935: The galaxy power spectrum can be used for measuring $D_A(z)$
1936: as well as $H(z)$ over a wider
1937: range of redshifts. From galaxy surveys we find three-dimensional
1938: positions of galaxies by measuring their angular positions on the sky as
1939: well as their redshifts. We can then estimate the two-point correlation
1940: function of galaxies as a function of the angular separation,
1941: $\Delta\theta$, and the redshift separation, $\Delta z$.
1942: To convert $\Delta\theta$ and $\Delta z$ into the comoving separations
1943: perpendicular to the line of sight, $\Delta r_\perp$, and those along the line
1944: of sight, $\Delta r_\parallel$, one needs to know $D_A(z)$ and $H(z)$,
1945: respectively, as
1946: \begin{eqnarray}
1947: \Delta r_\perp &=& (1+z)D_A(z)\Delta\theta,\\
1948: \Delta r_\parallel &=& \frac{c\Delta z}{H(z)},
1949: \end{eqnarray}
1950: where $(1+z)$ appears because $D_A(z)$ is the proper (physical) angular
1951: diameter distance, whereas $\Delta r_\perp$ is the comoving separation.
1952: Therefore, if we know $\Delta r_\perp$ and $\Delta r_\parallel$ {\it a
1953: priori}, then we may use the above equations to measure $D_A(z)$ and
1954: $H(z)$.
1955:
1956: The galaxy power spectra contain at least three distance scales which
1957: may be used in the place of $\Delta r_\perp$ and $\Delta r_\parallel$:
1958: (i) the sound horizon size at the so-called baryon drag epoch,
1959: $z_{drag}\simeq 1020$, at which
1960: baryons were released from the baryon-photon plasma, (ii) the photon
1961: horizon size at the matter-radiation equality, $z_{eq}\simeq 3200$, and
1962: (iii) the Silk damping scale \citep[see, e.g.,][]{eisenstein/hu:1998}.
1963:
1964: In Fourier space, we may write the observed power spectrum as
1965: \citep{seo/eisenstein:2003}
1966: \begin{eqnarray}\label{eq:pkDaHz}
1967: P_{obs}&(&k_\parallel,k_\perp,z)
1968: =
1969: \left(\frac{D_A(z)}{D_{A,\mathrm{true}}(z)}\right)^2
1970: \left(\frac{H_{\mathrm{true}}(z)}{H(z)}\right)
1971: \nonumber
1972: \\
1973: &\times&
1974: P_{\mathrm{true}}
1975: \left(
1976: \frac{D_{A,\mathrm{true}}(z)}{D_A(z)}k_\perp,
1977: \frac{H(z)}{H_{\mathrm{true}}(z)}k_\parallel,z
1978: \right),
1979: \end{eqnarray}
1980: where $k_\perp$ and $k_\parallel$ are the wavenumbers perpendicular to and
1981: parallel to the line of sight, respectively, and
1982: $P_{true}(k)$, $D_{A,true}(z)$, and $H_{true}(z)$ are the true,
1983: underlying values. We then vary $D_A(z)$ and $H(z)$, trying to estimate
1984: $D_{A,true}(z)$ and $H_{true}(z)$.
1985:
1986: There are two ways of measuring $D_A(z)$ and $H(z)$
1987: from the galaxy power spectra:
1988: \begin{itemize}
1989: \item[(1)] Use BAOs. The BAOs contain the information of one of the
1990: standard rulers, the sound horizon size at $z_{drag}$. This
1991: method relies on measuring only the phases of BAOs, which are
1992: markedly insensitive to all the non-linear
1993: effects (clustering, bias, and redshift space distortion)
1994: \citep{seo/eisenstein:2005,eisenstein/seo:2007,nishimichi/etal:2007,
1995: smith/scoccimarro/sheth:2008,angulo/etal:2008,sanchez/baugh/angulo:prep,
1996: seo/etal:prep,shoji/jeong/komatsu:prep},
1997: despite the fact that the amplitude is
1998: distorted by non-linearities (see Figures~\ref{fig4},
1999: \ref{fig9}, \ref{fig12}, and
2000: \ref{fig13}). Therefore, BAOs provide a robust means to
2001: measure $D_A(z)$ and $H(z)$, and they have been used for
2002: determining $D_A^2H^{-1}$ out to $z=0.2$ from the SDSS main
2003: galaxy sample and 2dFGRS, as well as to $z=0.35$ from the
2004: SDSS Luminous Red Galaxy (LRG) sample
2005: \citep{eisenstein/etal:2005,percival/etal:2007c}; however,
2006: since they use only one
2007: standard ruler, the constraints on $D_A(z)$ and $H(z)$ from
2008: the BAO-only analysis are weaker than the full analysis
2009: \citep{shoji/jeong/komatsu:prep}.
2010: \item[(2)] Use the \textit{entire} shpae of the power spectrum. This
2011: approach gives the best determination (i.e., the smallest
2012: error) of $D_A(z)$ and $H(z)$, as it uses all the standard
2013: rulers encoded in the galaxy power spectrum; however, one must
2014: understand the distortions of the shape of the power spectrum
2015: due to non-linear effects. The question is, ``is the
2016: 3rd-order (or higher) PT good enough for correcting the
2017: key non-linear effects?''
2018: \end{itemize}
2019: In this paper we show, for the first time, that we can extract
2020: the distance scale
2021: using the 3rd-order PT galaxy power spectrum in real space. While we have not
2022: yet included the effects of redshift space distortion, this is a
2023: significant step towards extracting $D_A(z)$ and $H(z)$ from the entire
2024: shape of the power spectrum of galaxies.
2025: We shall address the effect of non-linear redshift space distortion in
2026: the future work.
2027:
2028: \subsubsection{Method: Measuring ``Box Size'' of the Millennium Simulation}
2029: In real space simulations (as opposed to redshift space ones),
2030: there is only one distance scale in the problem:
2031: the box size of the simulation, $L_{\mathrm{box}}$, which
2032: is $L_{\mathrm{box}}^{\mathrm{(true)}}=500~\mathrm{Mpc}/h$
2033: for the Millennium simulation.
2034: Then, ``estimating
2035: the distance scale from the Millennium Simulation'' becomes
2036: equivalent to ``estimating $L_{\rm box}$ from the Millennium
2037: Simulation.
2038: Eq.~(\ref{eq:pkDaHz}) now leads:
2039: \begin{equation}\label{eq:pkLbox}
2040: P_{obs}(k,L_{\mathrm{box}})
2041: =
2042: \left(
2043: \frac{L_{\mathrm{box}}}{L_{\mathrm{box}}^{\mathrm{(true)}}}
2044: \right)^3
2045: P_{\mathrm{true}}
2046: \left(
2047: \frac{L_{\mathrm{box}}^{\mathrm{(true)}}}{L_{\mathrm{box}}}k
2048: \right).
2049: \end{equation}
2050:
2051: As we estimate the variance of power spectrum from the
2052: observed power spectrum,
2053: we need to rescale the variance
2054: when the normalization of the observed power spectrum changes :
2055: \begin{equation}
2056: \label{eq:sigpkLbox}
2057: \sigma_{Pi}^2(L_{\mathrm{box}})=
2058: \left(
2059: \frac{L_{\mathrm{box}}}{L_{\mathrm{box}}^{\mathrm{(true)}}}
2060: \right)^6
2061: \sigma_{Pi}^2(L_{\mathrm{box}}^{\mathrm{(true)}})
2062: \end{equation}
2063:
2064: We estimate $L_{\mathrm{box}}$ using
2065: the likelihood function given by
2066: \begin{eqnarray}
2067: \label{eq:LLbox}
2068: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,&P_0&,L_{\mathrm{box}})
2069: =\prod_{k_i<k_{max}}
2070: \frac{1}{\sqrt{2\pi\sigma_{Pi}^2(L_{\mathrm{box}})}}
2071: \nonumber
2072: \\
2073: &\times&
2074: \exp
2075: \left[-
2076: \frac{
2077: \left\{
2078: P_{obs}(k_i/\alpha)
2079: -
2080: P_{g}(k_i/\alpha)/\alpha^3
2081: \right\}^2
2082: }{2\sigma_{P}^2(k_i/\alpha)}
2083: \right],
2084: \end{eqnarray}
2085: where
2086: $\alpha=L_{\mathrm{box}}/L_{\mathrm{box}}^{\mathrm{(true)}}$.
2087:
2088: The likelihood function, Eq.~(\ref{eq:LLbox}), still depends upon the
2089: bias parameters that we wish to eliminate. Therefore we marginalize
2090: the likelihood function over all the bias parameters with flat
2091: priors.\footnote{Note that this is the most conservative analysis one can do. In reality we can use the
2092: bispectrum for measuring $\tilde{b}_1$ and $\tilde{b}_2$, which would
2093: give appropriate priors on them (see \S~\ref{sec:bispectrum}). We shall
2094: report on the results from this analysis elsewhere.}
2095: We obtain (see also Appendix~\ref{sec:appB}):
2096: \begin{equation}
2097: \mathcal{L}(L_{\mathrm{box}})
2098: =
2099: \int_0^\infty d\tilde{b}_1^2
2100: \int_{-\infty}^\infty d\tilde{b}_2
2101: \int_{-\infty}^\infty dP_0~
2102: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,P_0,L_{\mathrm{box}}).
2103: \end{equation}
2104:
2105: Hereafter, we shall simply call $L_{\mathrm{box}}$ as $D$ for
2106: `distance scale'. $D$ is closely related to the angular diameter distance,
2107: $D_A(z)$, and the expansion rate, $H(z)$, in real surveys. (See, \S 5.1.1)
2108:
2109:
2110: \subsubsection{Results: Unbiased Extraction of the distance scale from the Millennium Simulation}
2111: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2112: \begin{figure*}
2113: \centering
2114: \rotatebox{90}{
2115: \includegraphics[width=10cm]{fig16.ps}
2116: }
2117: \caption{
2118: Distance scale extracted from the Millennium Simulation
2119: using the 3rd-order PT galaxy power spectrum given by
2120: Eq.~(\ref{eq:3rd_PT_Pk}), divided by the true value.
2121: The mean of the likelihood ({\it stars}), and the maximum likelihood values
2122: ({\it filled circles}) and the corresponding
2123: 1-$\sigma$ intervals ({\it errorbars}), are shown as a function of
2124: maximum wavenumbers used in the fits, $k_{max}$.
2125: We find
2126: $D/D_{true}=1$
2127: to within the 1-$\sigma$ errors from all the
2128: halo/galaxy catalogues (``halo,'' ``Mgalaxy,'' and ``Dgalaxy'') at all
2129: redshifts, provided that we use
2130: $k_{max}$ estimated from the matter power spectra,
2131: $k_{max}=0.15$, 0.25, 1.0, 1.2, 1.3, and 1.5 at $z=1$, 2, 3, 4, 5, and
2132: 6, respectively (see Table~\ref{table:kmax}).
2133: Note that the errors on $D$ do not decrease as $k_{max}$ increases
2134: due to degeneracy between $D$ and the bias parameters.
2135: See Figure (\ref{fig18})
2136: and (\ref{fig19}) for further analysis.
2137: }%
2138: \label{fig16}
2139: \end{figure*}
2140: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2142: \begin{figure*}
2143: \centering
2144: \rotatebox{90}{
2145: \includegraphics[width=10cm]{fig17.ps}
2146: }
2147: \caption{
2148: Same as Figure \ref{fig8}, but including the
2149: distance scale $D/D_{true}$.
2150: }%
2151: \label{fig17}
2152: \end{figure*}
2153: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2154:
2155: In Figure \ref{fig16} we show
2156: $D(z)/D_{true}(z)$
2157: estimated from the halo,
2158: Mgalaxy, and Dgalaxy catalogues at $z=1$, 2, 3, 4, 5, and 6.
2159: The maximum likelihood values (filled circles) and the corresponding
2160: 1-$\sigma$ intervals (errorbars), as well as the mean of the likelihood
2161: ({\it stars}) are shown.
2162: We find
2163: $D(z)/D_{true}(z)=1$
2164: to within the 1-$\sigma$ errors from
2165: {\it all of the halo/galaxy catalogues} at {\it all redshifts},
2166: provided that we use $P_{obs}(k)$ only up to $k_{max}$ that has been
2167: determined unambiguously from the matter power spectrum (see
2168: Table~\ref{table:kmax}).
2169: Not only does this provide a strong support for the validity of
2170: Eq.~(\ref{eq:3rd_PT_Pk}), but also it provides a practical means for
2171: extracting
2172: $D$ from the full shape of the observed galaxy power spectra.
2173:
2174: Despite a small volume of the Millennium Simulation and the use of flat
2175: priors on the bias parameters upon marginalization, we could determine
2176: $D$ to about 2.5\% accuracy.
2177:
2178: In addition, we also find that the error on $D$ hardly decreases
2179: even though $k_{max}$ increases.
2180: It is because of the degeneracy between $D$ and the bias parameters.
2181: In order to see how strongly degenerate they are,
2182: we calculate correlations between pairs of parameters
2183: ($\tilde{b}_1$,$\tilde{b}_2$,$P_0$,$D/D_{true}$)
2184: by the Fisher information matrix from Eq. (\ref{eq:fisher_bias}).
2185:
2186: Figure \ref{fig17} shows both one-dimensional
2187: marginalized constraints and two-dimensional joint marginalized
2188: constraints of 2-$\sigma$ range ($95.45\%$ CL) for
2189: the bias parameters and the distance scale.
2190: This figure indicates that when we include the distance scale, the
2191: correlations between bias parameters become milder. It is mainly
2192: due to the correlation between the distance scale and $\tilde{b}_1$
2193: making the constraint on $\tilde{b}_1$ weaker.
2194: On the other hand, the one-dimensional marginalized
2195: likelihood functions for $\tilde{b}_2$ and $P_0$ are hardly changed.
2196: The remaining degeneracies are those between ($\tilde{b}_2$,$P_0$) and
2197: ($\tilde{b}_1$,$D/D_{true}$).
2198: These degeneracies would be broken when we
2199: include the information from the bispectrum, as the bispectrum
2200: will measure $\tilde{b}_1$ and $\tilde{b}_2$.
2201:
2202:
2203: \subsubsection{Optimal estimation of the distance scale}
2204: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2205: \begin{figure*}
2206: \centering
2207: \rotatebox{90}{
2208: \includegraphics[width=10cm]{fig18.ps}
2209: }
2210: \caption{
2211: Same as Figure \ref{fig16}, but
2212: with $\tilde{b}_1$ and $\tilde{b}_2$ fixed at the best-fitting values.
2213: The 1-$\sigma$ ranges for $D$ are
2214: $1.5 \%$
2215: and
2216: $0.15 \%$
2217: for $k_{max}=0.2~h/\mathrm{Mpc}$ and $k_{max}=1.5~h/\mathrm{Mpc}$,
2218: respectively.
2219: The errors on $D$ decrease as $k_{max}$ increases, but the scaling is
2220: still milder than $1/\sqrt{\sum_{k<k_{max}} N_k}$.
2221: }%
2222: \label{fig18}
2223: \end{figure*}
2224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2226: \begin{figure*}
2227: \centering
2228: \rotatebox{90}{
2229: \includegraphics[width=10cm]{fig19.ps}
2230: }
2231: \caption{
2232: Same as Figure \ref{fig16}, but
2233: with $\tilde{b}_1$, $\tilde{b}_2$ and $P_0$ fixed at the best-fitting values.
2234: The 1-$\sigma$ ranges for $D$ are
2235: $0.8 \%$
2236: and
2237: $0.05 \%$
2238: for $k_{max}=0.2~h/\mathrm{Mpc}$ and $k_{max}=1.5~h/\mathrm{Mpc}$,
2239: respectively.
2240: The errors on $D$ decrease as $k_{max}$ increases as
2241: $1/\sqrt{\sum_{k<k_{max}} N_k}$.
2242: }%
2243: \label{fig19}
2244: \end{figure*}
2245: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2246: The constraint we find from the previous subsection
2247: will get better when we include the bispectrum,
2248: as the reduced bispectrum provides independent and strong
2249: constraints on $\tilde{b}_1$ and $\tilde{b}_2$
2250: \citep{sefusatti/etal:2006}.
2251:
2252: How much will it be better?
2253: First, let us assume that we know the exact values of
2254: $\tilde{b}_1$ and $\tilde{b}_2$.
2255: In this case, we get the error on $D$ by marginalizing only over $P_0$
2256: while setting $\tilde{b}_1$ and $\tilde{b}_2$ to be the best-fitting values,
2257: i.e.
2258: \begin{equation}\label{eq:Lalpha_b1b2_fixed}
2259: \mathcal{L}^{\mathrm{fix}~\tilde{b}_1\tilde{b}_2}(D)
2260: =\int_{-\infty}^{\infty} dP_0
2261: \mathcal{L}
2262: (
2263: \tilde{b}_1^{\mathrm{bf}},
2264: \tilde{b}_2^{\mathrm{bf}},
2265: P_0,D
2266: )
2267: \end{equation}
2268: where $\tilde{b}_1^{\mathrm{bf}}$ and $\tilde{b}_2^{\mathrm{bf}}$
2269: denote the best-fitting values of $\tilde{b}_1$ and $\tilde{b}_2$ for
2270: each $k_{max}$, respectively.
2271: In Figure \ref{fig18}, we show $D/D_{true}$ estimated
2272: from Eq. (\ref{eq:Lalpha_b1b2_fixed}).
2273: This figure shows that we can extract $D$ to about $1.5\%$ accuracy even
2274: for the low $k_{max}=0.2~h/\mathrm{Mpc}$, and the error decreases further
2275: to $0.15\%$ for $k_{max}=1.5~h/\mathrm{Mpc}$.
2276: Note that the uncertainties on $D/D_{true}$ decrease as $k_{max}$ increases
2277: as expected. The reason is because
2278: fixing $\tilde{b}_1$ and $\tilde{b}_2$
2279: breaks the degeneracy between them and the distance scale.
2280:
2281: In reality, the bias parameters estimated from the bispectrum
2282: have finite errors, and thus the accuracy of extracting $D$ will be
2283: somewhere in between Figure \ref{fig16}
2284: and Figure \ref{fig18}.
2285: The result of the full analysis including both power spectrum and
2286: bispectrum of Millennium Simulation will be reported elsewhere.
2287:
2288: In the ideal situation where we completely understand the complicated
2289: halo/galaxy formation, we may be able to calculate the
2290: three bias parameters from the first principle.
2291: This \textit{ideal} determination of bias parameters will provide
2292: more accurate constraints on the distance scale $D$.
2293: In this case, we get the likelihood function
2294: by fixing all the bias parameters to their best-fitting values :
2295: \begin{equation}\label{eq:Lalpha_bias_fixed}
2296: \mathcal{L}^{\mathrm{fix~bias}}(D)
2297: =
2298: \mathcal{L}
2299: (
2300: \tilde{b}_1^{\mathrm{bf}},
2301: \tilde{b}_2^{\mathrm{bf}},
2302: P_0^{\mathrm{bf}},D
2303: )
2304: \end{equation}
2305: By knowing all the bias parameters, we can extract the distance
2306: scale $D$ to $0.8\%$ accuracy for $k_{max}=0.2~h/\mathrm{Mpc}$.
2307: The error decreases further to $0.05\%$ for $k_{max}=1.5~h/\mathrm{Mpc}$.
2308: (See Figure (\ref{fig19}))
2309:
2310:
2311: \subsubsection{Forecast for a HETDEX-like survey}
2312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2313: \begin{figure}
2314: \centering
2315: \rotatebox{90}{
2316: \includegraphics[width=6.5cm]{fig20.ps}
2317: }
2318: \caption{
2319: Projected constraints on
2320: $D$ %$D_A$
2321: at $z=3$ from a HETDEX-like survey with
2322: the survey volume of $(1.5~\mathrm{Gpc/h})^3$. We have used the
2323: best-fitting 3rd-order PT power spectrum of MPA halos in the Millennium
2324: Simulation for generating a mock simulation
2325: data. We show the results for the number of objects of
2326: $N_{galaxy}=2\times 10^5$, $10^6$, $2\times 10^6$, and $10^9$, from the
2327: top to bottom panels, respectively, for which we find the projected 1-$\sigma$
2328: errors of 2.5\%, 1.5\%, 1\%, and 0.3\%, respectively.
2329: }%
2330: \label{fig20}
2331: \end{figure}
2332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2333:
2334: The planned future surveys would cover a larger volume than the
2335: Millennium Simulation. Also, since the real surveys would be limited by
2336: their continuum/flux sensitivity, they would not be able to detect
2337: all galaxies that were resolved in the Millennium Simulation.
2338: In this subsection we explore how the constraints would be
2339: affected by the volume and the number of objects.
2340:
2341: To simulate the mock data, we take a simplified approach: we take
2342: our best-fitting power spectrum at $z=3$, i.e.,
2343: Eq.~(\ref{eq:3rd_PT_Pk}) fit to the power spectrum of MPA halos in the
2344: Millennium Simulation at $z=3$, and add random Gaussian noise to it with
2345: the standard deviation given by Eq.~(\ref{eq:varpk}).
2346: To compute the standard deviation we need to specify the survey volume,
2347: which determines the fundamental wavenumber, $\Delta k$, as $\Delta
2348: k=2\pi/V_{survey}^{1/3}$. We use the volume that would be surveyed by
2349: the HETDEX survey \citep{hill/etal:2004},
2350: $V_{survey}=(1.5~\mathrm{Gpc}/h)^3$, which is 27 times as large as the
2351: volume of the Millennium Simulation. We then vary the number of galaxies,
2352: $N_{galaxy}$, which determines the shot noise as
2353: $P_{shot}=1/n=V_{survey}/N_{galaxy}$.
2354: We have generated only one realization, and repeated the same analysis
2355: as before to extract $D_A$ from the mock HETDEX data.
2356:
2357: In Figure \ref{fig20} we show $D/D_{true}$ as a
2358: function of $k_{max}$ and $N_g$. For $N_{galaxy}=10^9$, which gives the
2359: same number density as the Millennium Simulation, the projected
2360: error on $D$ is 0.3\%, or 8 times better than the original result
2361: presented in Figure~\ref{fig16}.
2362: Since the volume is 27 times bigger, the statistics alone would reduce
2363: the error by a factor of about 5.
2364:
2365: The other factor of about 1.5 comes from the fact that the variance
2366: of the distance scale estimated from the Millennium Simulation
2367: lies on the tail of the distribution of the variance of the
2368: distance scale, (See, appendix C)
2369: while the error estimated from the HETDEX volume mock is close to the peak of
2370: PDF of the variance.
2371:
2372: However, real surveys will not get as high the number density as the
2373: Millennium Simulation. For example, the HETDEX survey will detect about
2374: one million Ly$\alpha$ emitting galaxies, i.e., $N_{galaxy}=10^6$.
2375: In Figure~\ref{fig20} we show that
2376: the errors on $D$ increase from 0.3\% for $N_{galaxy}=10^9$ to
2377: 1\%, 1.5\%, and 3\% for $N_{galaxy}=2\times 10^6$, $10^6$, and $2\times 10^5$, respectively.
2378:
2379: Finally, we note that these forecasts are not yet final, as we have not
2380: included the effect of non-linear redshift space distortion. Also,
2381: eventually one needs to repeat this analysis using the ``super Millennium
2382: Simulation'' with a bigger volume.
2383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2384:
2385: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2386: \section{Discussion and Conclusions}
2387: \label{sec:conclusion}
2388: Two main new results that we have presented in this paper are:
2389: \begin{itemize}
2390: \item The 3rd-order PT galaxy power spectrum given by
2391: Eq.~(\ref{eq:3rd_PT_Pk}), which is based upon the assumption that
2392: the number density of galaxies at a given location is a local
2393: function of the underlying matter density at the same location
2394: \citep{fry/gaztanaga:1993} plus stochastic noise
2395: \citep{mcdonald:2006}, fits the halo as well as galaxy power spectra
2396: estimated from the Millennium Simulation at high
2397: redshifts, $1\le z\le 6$, up to the maximum wavenumber,
2398: $k_{max}$, that has been determined from the matter power spectrum.
2399: \item When 3 galaxy bias parameters, $\tilde{b}_1$, $\tilde{b}_2$, and
2400: $P_0$, are marginalized over, the
2401: 3rd-order PT galaxy power spectrum fit to the Millennium
2402: Simulation yields the correct (unbiased)
2403: distance scale to within the statistical error of the simulation,
2404: $\sim 3\%$.
2405: \end{itemize}
2406: These results suggest that the 3rd-order PT provides us with a practical
2407: means to extract the cosmological information from the observed galaxy
2408: power spectra at high redshifts, i.e., $z>1$, accurately.
2409:
2410: We would like to emphasize that our approach does not require
2411: simulations to calibrate the model. The 3rd-order PT is based upon the
2412: solid physical framework, and the only assumption made for the galaxy
2413: formation is that it is a local process, at least on the scales where
2414: the 3rd-order PT is valid, i.e., $k<k_{max}$.
2415: The only serious drawback so far is that the 3rd-order PT breaks down at
2416: low redshifts, and thus it cannot be applied to the current generation
2417: of survey data such as 2dFGRS and SDSS. However, the planned future
2418: high-$z$ surveys would benefit immensely from the PT approach.
2419:
2420: The practical application of our approach may proceed as follows:
2421: \begin{itemize}
2422: \item [(1)] Measure the galaxy power spectra at various
2423: redshifts. When we have $N$ redshift bins, the number of bias
2424: parameters is $3N$, as the bias parameters evolve with $z$.
2425: \item [(2)] Calculate $k_{max}(z)$ from the condition,
2426: $\Delta_m^2(k_{max},z)=0.4$, where
2427: $\Delta_m^2(k,z)=k^3P_{\delta\delta}(k,z)/(2\pi^2)$
2428: is computed from the fiducial cosmology, e.g., the WMAP 5-year
2429: best-fitting parameters. The results should not be sensitive to
2430: the exact values of $k_{max}(z)$.
2431: \item [(3)] Fit Eq.~(\ref{eq:3rd_PT_Pk}) to the observed galaxy spectra
2432: up to $k_{max}(z)$
2433: at all $z$ simultaneously for extracting the cosmological parameters.
2434: \end{itemize}
2435: In addition to this, we should be able to improve upon the accuracy of
2436: parameter determinations by including the bispectrum as well, as the
2437: bispectrum basically fixes $\tilde{b}_1$ and $\tilde{b}_2$
2438: \citep{sefusatti/komatsu:2007}.
2439: Therefore, the step (3) may be replaced by
2440: \begin{itemize}
2441: \item [(3')] Fit Eq.~(\ref{eq:3rd_PT_Pk}) to the observed galaxy
2442: spectra up to $k_{max}(z)$, and fit the PT bispectrum to the
2443: observed galaxy bispectra up to the same $k_{max}(z)$,
2444: at all $z$ simultaneously for extracting the cosmological parameters.
2445: \end{itemize}
2446: We are currently performing a joint analysis of the galaxy power spectra
2447: and bispectra on the Millennium Simulation. The results will be reported
2448: elsewhere.
2449:
2450: There are limitations in our present study, however. First, a relatively small
2451: volume of the Millennium Simulation does not allow us to make a
2452: precision test of the 3rd-order PT. Also, this limitation does not
2453: allow us to study constraints on more than one cosmological
2454: parameter. We have picked $D$ as the representative example because
2455: this parameter seems the most interesting one in light of the future
2456: surveys whose primary goal is to constrain the properties of dark
2457: energy. In the future we must use larger simulations to show
2458: convincingly that the bias in cosmological parameters is much lower than
2459: 1\% level.
2460: Second, we have found that,
2461: due to the limited statistics of a
2462: small volume and the smaller $k_{max}$ due
2463: to stronger non-linearities, the
2464: bias parameters are not determined very well from the galaxy power
2465: spectra alone at $z\le 3$. This issue should disappear by including the
2466: bispectrum in the joint analysis.
2467: Last and foremost, our study has been
2468: restricted to the real space power spectra: we have not addressed the
2469: non-linearities in redshift space distortion. This is a subject of the
2470: future study.
2471:
2472:
2473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2474:
2475:
2476: \acknowledgments
2477: We would like to thank Volker Springel for providing us with the matter
2478: power spectrum data from the Millennium Simulation shown in
2479: \S~\ref{sec:DM}, and Gerard Lemson for his help on the
2480: Millennium database.
2481: We would like to thank Paul Shapiro and Ilian Iliev for their
2482: contribution during the initial stage of this project.
2483: This material is based in part upon work supported by the Texas Advanced
2484: Research Program under Grant No. 003658-0005-2006.
2485: E.K. acknowledges support from an Alfred P. Sloan Fellowship.
2486: The Millennium Simulation databases used in this paper and the web application
2487: providing online access to them were constructed as part of the activities
2488: of the German Astrophysical Virtual Observatory.
2489:
2490: \appendix
2491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2492: \section{Error on power spectrum}\label{sec:appA}
2493: Besides the normalization,
2494: an estimator for the power spectrum may be written as
2495: \begin{equation}
2496: P_{obs}(k)
2497: =
2498: \frac{1}{N_k} \sum_{i=1}^{N_k}
2499: |\delta(\mathbf{k}_i)|^2
2500: \Biggl|_{|\mathbf{k}_i-k|\le\Delta k}
2501: \end{equation}
2502: where $\delta(\mathbf{k}_i)$ is a Fourier transform of the density field
2503: in position space, $\Delta k$ is the fundamental wavenumber of either
2504: survey volume or simulation box, and $N_k$ is the number of
2505: independent $k$-modes available per bin.
2506: This estimator is unbiased because
2507: \begin{equation}
2508: \langle P_{obs}(k) \rangle
2509: =
2510: \frac{1}{N_k} \sum_{i=1}^{N_k}
2511: \langle
2512: |\delta(\mathbf{k}_i)|^2
2513: \rangle
2514: \Biggl|_{|\mathbf{k}_i-k|\le\Delta k}
2515: =
2516: \langle
2517: |\delta(k)|^2
2518: \rangle
2519: =P(k),
2520: \end{equation}
2521: where $P(k)$ is the underlying power spectrum.
2522: The variance of this estimator is given by
2523: \begin{equation}
2524: \left\langle
2525: \left( \frac{P_{obs}(k)-P(k)}{P(k)} \right)^2
2526: \right\rangle
2527: =1-2 \frac{\langle P_{obs} \rangle}{P(k)}
2528: +\frac{1}{N_k^2P(k)^2}\sum_{i=1}^{N_k}\sum_{j=1}^{N_k}
2529: \langle
2530: \delta^*(\mathbf{k}_i)\delta(\mathbf{k}_i)
2531: \delta^*(\mathbf{k}_j)\delta(\mathbf{k}_j)
2532: \rangle.
2533: \end{equation}
2534: Assuming that the density field is a Gaussian random variable
2535: with its variance given by $P(k)$, i.e.,
2536: \begin{equation}
2537: \langle\delta_i^*\delta_j\rangle =P(k)\delta_{ij},
2538: \end{equation}
2539: we use Wick's theorem for evaluating the last double summation:
2540: \begin{eqnarray}
2541: \label{eq:wick}
2542: \sum_{i=1}^{N_k}\sum_{j=1}^{N_k}
2543: \langle
2544: \delta^*_i\delta_i
2545: \delta^*_j\delta_j
2546: \rangle
2547: &=&
2548: \sum_{i=1}^{N_k}\sum_{j=1}^{N_k}
2549: \left[
2550: \langle
2551: \delta^*_i\delta_i
2552: \rangle
2553: \langle
2554: \delta^*_j\delta_j
2555: \rangle
2556: +
2557: \langle
2558: \delta^*_i\delta_j
2559: \rangle
2560: \langle
2561: \delta^*_j\delta_i
2562: \rangle
2563: +
2564: \langle
2565: \delta^*_i\delta^*_j
2566: \rangle
2567: \langle
2568: \delta_i\delta_j
2569: \rangle
2570: \right]\nonumber\\
2571: &=&
2572: N_k^2 [P(k)]^2 + N_k [P(k)]^2.
2573: \end{eqnarray}
2574: Therefore, the variance is given by
2575: \begin{equation}
2576: \left\langle
2577: \left[P_{obs}(k)-P(k) \right]^2
2578: \right\rangle
2579: = \frac{[P(k)]^2}{N_k},
2580: \end{equation}
2581: and the standard deviation is given by
2582: \begin{equation}\label{eq:errpk}
2583: \sigma_{P(k)}
2584: \equiv
2585: \left\langle
2586: \left[P_{obs}(k)-P(k) \right]^2
2587: \right\rangle^{1/2}
2588: =
2589: \sqrt{\frac{1}{N_k}}P(k).
2590: \end{equation}
2591: Note that this formula is valid only when $\delta$ is a Gaussian random
2592: field. When $\delta$ is non-Gaussian due to, e.g., non-linear evolution,
2593: primordial non-Gaussianity, non-linear bias, etc., we must add the
2594: connected four-point function to Eq.~(\ref{eq:wick}).
2595: See also \citet{takahashi/etal:prep} for the study of finite box size effects on the four-point function.
2596:
2597: How do we calculate $N_k$?
2598: As the Fourier transformation of a real-valued field has symmetry given by
2599: $\delta^*(\mathbf{k})=\delta(-\mathbf{k})$, the number of independent
2600: $k$-modes is exactly a half of the number of modes available in a
2601: spherical shell at a given $k$.
2602: We find
2603: \begin{equation}
2604: N_k=\frac{1}{2} \frac{4\pi k^2\delta k}{(\delta k)^3}
2605: =2\pi\left(\frac{k}{\delta k}\right)^2,
2606: \end{equation}
2607: where $\delta k$ is the fundamental wavenumber given by $\delta k =
2608: 2\pi/L$, where $L$ is the survey size or simulation box size.
2609:
2610: In the literature one may often find a different formula such as
2611: \begin{equation}\label{eq:errpk_litereature}
2612: \sigma_{P(k)}^{literature}
2613: =
2614: \sqrt{\frac{2}{N^{literature}_k}}P(k).
2615: \end{equation}
2616: Here, there is an extra factor of $\sqrt{2}$, as
2617: $N^{literature}_k$ is the number of modes available in a
2618: spherical shell at a given $k$, {\it without symmetry,
2619: $\delta^*(\mathbf{k})=\delta(-\mathbf{k})$, taken into account},
2620: i.e., $N^{literature}_k=2N_k$. Both formulas give the same results,
2621: provided that we understand what we mean by $N_k$ in these formulas.
2622:
2623: We have tested the formula Eq.~(\ref{eq:errpk}) by comparing it to
2624: the standard deviation estimated the ensemble of dark matter simulations
2625: used in Paper I. (See Paper I for details of the simulations.)
2626: Figure \ref{figA1} and \ref{figA2}
2627: show the result of this comparison.
2628: The formula Eq.~(\ref{eq:errpk}) and the simulation data agree well.
2629:
2630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2631: \begin{figure}
2632: \centering
2633: \rotatebox{90}{
2634: \includegraphics[width=10cm]{figA1.ps}
2635: }
2636: \caption{
2637: Standard deviation of the matter power spectrum:
2638: analytical versus simulations.
2639: The symbols show the standard deviations directly measured from
2640: 120 independent $N$-body simulations whose box sizes are
2641: $L=512~\mathrm{Mpc}/h$ (60 realizations for $k<0.24 h/\mathrm{Mpc}$)
2642: and
2643: $L=256~\mathrm{Mpc}/h$ (60 realizations for $0.24<k<0.5 h/\mathrm{Mpc}$)
2644: . Each simulation contains $256^3$ particles.
2645: The solid and dot-dashed lines show the analytical
2646: formula (Eq.~(\ref{eq:errpk})) with the 3rd-order PT non-linear $P(k)$
2647: and the linear $P(k)$, respectively.
2648: Note that the graph is discontinuous at $k=0.24 h/\mathrm{Mpc}$
2649: because the number of $k$ modes, $N_k$, for a given wavenumber $k$
2650: is different for different box sizes.
2651: }%
2652: \label{figA1}
2653: \end{figure}
2654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2656: \begin{figure}
2657: \centering
2658: \rotatebox{90}{
2659: \includegraphics[width=10cm]{figA2.ps}
2660: }
2661: \caption{
2662: Residuals. We divide both analytical estimation and simulation results
2663: by the analytical formula (Eq.~(\ref{eq:errpk})) with
2664: the 3rd-order PT nonlinear $P(k)$.
2665: }%
2666: \label{figA2}
2667: \end{figure}
2668: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2669:
2670:
2671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2672: \section{Analytical marginalization of the likelihood function over
2673: $\tilde{b}_1^2$ and $P_0$}\label{sec:appB}
2674: In this appendix we derive the analytical formulas for the
2675: likelihood function marginalized over $\tilde{b}_1^2$ and $P_0$.
2676:
2677: The likelihood function, Eq.~(\ref{eq:likelihood}), is given by
2678: \begin{equation}
2679: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,P_0,\theta_n)
2680: =
2681: \left(\prod_{i}\frac{1}{\sqrt{2\pi\sigma_{Pi}^2}}\right)
2682: \exp
2683: \left[
2684: -
2685: \sum_{i}
2686: \frac
2687: {\left(P_{obs,i}
2688: -
2689: \tilde{b}_1^2(P_{\delta\delta,i}+\tilde{b}_2P_{b2,i}
2690: +\tilde{b}_2^2P_{b22,i})-P_0\right)^2}
2691: {2\sigma_{Pi}^2}
2692: \right],
2693: \end{equation}
2694: where $\theta_n$ are the cosmological parameters that do not depend on
2695: any of the bias parameters.
2696: The subscript $i$ denotes bins, $k_i$.
2697:
2698: Integrating the likelihood function over $P_0$, we find
2699: \begin{eqnarray}\label{eq:marN}
2700: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,\theta_n)
2701: &
2702: =&
2703: \int_{-\infty}^{\infty} d P_0 \mathcal{L}(\tilde{b}_1,\tilde{b}_2,P_0,\theta_n)
2704: \nonumber
2705: \\
2706: &
2707: =&
2708: \mathcal{N}\sqrt{\frac{2 \pi}{\sum_i w_i}}
2709: \exp\left[
2710: -\frac{1}{2}
2711: \frac{
2712: \sum_{i>j} w_iw_j(a_j-a_i)^2
2713: }
2714: {\sum_i w_i}
2715: \right],
2716: \end{eqnarray}
2717: where we have defined new variables
2718: \begin{eqnarray}
2719: \mathcal{N}
2720: &\equiv&\prod_{i}
2721: \frac{1}{\sqrt{2\pi\sigma_{Pi}^2}}\\
2722: \label{eq:ai}a_i&\equiv&
2723: P_{obs,i}
2724: -
2725: \tilde{b}_1^2(P_{\delta\delta,i}+\tilde{b}_2P_{b2,i}+\tilde{b}_2^2P_{b22,i})\\
2726: w_i&\equiv&
2727: \frac{1}{\sigma_{Pi}^2}.
2728: \end{eqnarray}
2729:
2730: We then integrate this function over $\tilde{b}_1^2$.
2731: Introducing new variables given by
2732: \begin{eqnarray}
2733: \bar{\mathcal{N}}
2734: &\equiv&\mathcal{N}\sqrt{\frac{2 \pi}{\sum_i w_i}},\\
2735: P_{th,i}&\equiv&P_{\delta\delta,i}+\tilde{b}_2P_{b2,i}+\tilde{b}_2^2P_{b22,i},
2736: \end{eqnarray}
2737: and $a_i=P_{obs,i}-\tilde{b}_1^2P_{th,i}$, we rewrite Eq.~(\ref{eq:marN}) as
2738: \begin{eqnarray}
2739: \mathcal{L}(\tilde{b}_1,\tilde{b}_2,\theta_n)
2740: &=&
2741: \bar{\mathcal{N}}
2742: \exp
2743: \left[
2744: -\frac{1}{2}
2745: \frac{
2746: \sum_{i>j} w_iw_j \left\{(P_{th,i}-P_{th,j})\tilde{b}_1^2-(P_{obs,i}-P_{obs,j})\right\}^2
2747: }
2748: {\sum_i w_i}
2749: \right]
2750: \nonumber\\
2751: &=&
2752: \bar{\mathcal{N}}
2753: \exp
2754: \left[
2755: -\frac{1}{2}\left(A \tilde{b}_1^4-2B \tilde{b}_1^2 +C\right)
2756: \right],
2757: \end{eqnarray}
2758: where
2759: \begin{eqnarray}
2760: A&\equiv&\frac{\sum_{i>j}w_iw_j(P_{th,i}-P_{th,j})^2}{\sum_iw_i}\\
2761: B&\equiv&\frac{\sum_{i>j}w_iw_j(P_{th,i}-P_{th,j})(P_{obs,i}-P_{obs,j})}{\sum_iw_i}\\
2762: C&\equiv&\frac{\sum_{i>j}w_iw_j(P_{obs,i}-P_{obs,j})^2}{\sum_iw_i}.
2763: \end{eqnarray}
2764: Assuming a flat prior on $\tilde{b}_1^2$, we integrate
2765: the likelihood function to find the desired result:
2766: \begin{eqnarray}\label{eq:marNb1}
2767: \mathcal{L}(\tilde{b}_2,\theta_n)
2768: &
2769: =&
2770: \bar{\mathcal{N}}
2771: \int_0^{\infty}\exp
2772: \left[
2773: -\frac{1}{2}\left(A \tilde{b}_1^4-2B \tilde{b}_1^2 +C\right)
2774: \right]
2775: d(\tilde{b}_1^2)
2776: \nonumber\\
2777: &
2778: =&
2779: \bar{\mathcal{N}}
2780: \exp\left[
2781: \frac{B^2-AC}{2A}
2782: \right]
2783: \sqrt{\frac{\pi}{2A}}\left\{1+\rm{erf}\left(\frac{B}{\sqrt{2A}}\right)\right\}.
2784: \end{eqnarray}
2785: Note that the convergence of the likelihood function is ensured
2786: by Cauchy's inequality, $B^2-AC<0$.
2787:
2788: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2789: \section{distribution of errors on the distance scale}
2790: \label{sec:appC}
2791: We find that the error on $D$ extracted from the halo power spectrum
2792: of Millennium Simulation is about $2.17\%$ for $k_{max}=1.5~h/\mathrm{Mpc}$
2793: at $z=6$. (See Figure \ref{fig16}.)
2794: On the other hand, the error on $D$ calculated from the Fisher information
2795: matrix is $1.57\%$
2796: Are they consistent?
2797:
2798: In order to test whether it is possible to get the error on $D$ far from
2799: the value derived from the Fisher matrix, we generate $1000$ realizations
2800: of mock power spectra with the best-fitting bias parameters for halo
2801: with $k_{max}=1.5~h/\mathrm{Mpc}$ at $z=6$.
2802: Then, we calculate the best-fitting value of $D$ as well as the 1-$\sigma$
2803: ($68.27\%$ CL) range from the one-dimensional marginalized likelihood
2804: function of $D$ for each realization.
2805:
2806: We find that the mean 1-$\sigma$ error on $D$ calculated
2807: from these realizations is $1.66\%$, and their standard deviation is $0.43\%$.
2808: Figure \ref{figC1} shows the distribution of the fractional
2809: 1-$\sigma$ error on $D$ compared with $D_{true}$.
2810: While the error derived from the Fisher matrix is close to the mean,
2811: the error calculated from the Millennium Simulation is on the tail of the
2812: distribution. The probability of having an error on $D$ greater
2813: than that from the Millennium Simulation is about $6 \%$,
2814: which is acceptable.
2815:
2816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2817: \begin{figure}
2818: \centering
2819: \rotatebox{90}{
2820: \includegraphics[width=10cm]{figC1.ps}
2821: }
2822: \caption{
2823: Histogram for the 1-$\sigma$ errors on $D$ calculated from
2824: $1000$ Monte Carlo realizations generated
2825: with the best-fitting bias parameters of halo power spectrum of
2826: Millennium Simulation with $k_{max}=1.5~h/\mathrm{Mpc}$ at $z=6$.
2827: The error derived from the Fisher matrix is close to the mean,
2828: while the error from the marginalized one-dimensional likelihood function
2829: of Millennium Simulation is on the tail of the distribution.
2830: The probability of having an error on $D$ greater than that from the
2831: Millennium Simulation is about $6 \%$.
2832: }%
2833: \label{figC1}
2834: \end{figure}
2835: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2836:
2837:
2838:
2839:
2840: \begin{thebibliography}{74}
2841: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2842:
2843: \bibitem[{{Angulo} et~al.(2008){Angulo}, {Baugh}, {Frenk}, \&
2844: {Lacey}}]{angulo/etal:2008}
2845: {Angulo}, R.~E., {Baugh}, C.~M., {Frenk}, C.~S., \& {Lacey}, C.~G. 2008,
2846: \mnras, 383, 755
2847:
2848: \bibitem[{{Bennett} et~al.(2003)}]{bennett/etal:2003}
2849: {Bennett}, C.~L., et~al. 2003, \apj, 583, 1
2850:
2851: \bibitem[{{Benson} et~al.(2003){Benson}, {Bower}, {Frenk}, {Lacey}, {Baugh}, \&
2852: {Cole}}]{benson/etal:2003}
2853: {Benson}, A.~J., {Bower}, R.~G., {Frenk}, C.~S., {Lacey}, C.~G., {Baugh},
2854: C.~M., \& {Cole}, S. 2003, \apj, 599, 38
2855:
2856: \bibitem[{{Bernardeau} et~al.(2002){Bernardeau}, {Colombi}, {Gazta{\~n}aga}, \&
2857: {Scoccimarro}}]{bernardeau/etal:2002}
2858: {Bernardeau}, F., {Colombi}, S., {Gazta{\~n}aga}, E., \& {Scoccimarro}, R.
2859: 2002, \physrep, 367, 1
2860:
2861: \bibitem[{Bond et~al.(1991)Bond, Efstathiou, Lubin, \&
2862: Meinhold}]{Bond/etal:1991}
2863: Bond, J.~R., Efstathiou, G., Lubin, P.~M., \& Meinhold, P.~R. 1991, \prl, 66,
2864: 2179
2865:
2866: \bibitem[{{Bower} et~al.(2006){Bower}, {Benson}, {Malbon}, {Helly}, {Frenk},
2867: {Baugh}, {Cole}, \& {Lacey}}]{bower/etal:2006}
2868: {Bower}, R.~G., {Benson}, A.~J., {Malbon}, R., {Helly}, J.~C., {Frenk}, C.~S.,
2869: {Baugh}, C.~M., {Cole}, S., \& {Lacey}, C.~G. 2006, \mnras, 370, 645
2870:
2871: \bibitem[{{Chabrier}(2003)}]{chabrier:2003}
2872: {Chabrier}, G. 2003, \pasp, 115,763
2873:
2874: \bibitem[{{Cole} et~al.(2000){Cole}, {Lacey}, {Baugh}, \&
2875: {Frenk}}]{cole/etal:2000}
2876: {Cole}, S., {Lacey}, C.~G., {Baugh}, C.~M., \& {Frenk}, C.~S. 2000, \mnras,
2877: 319, 168
2878:
2879: \bibitem[{{Cole} et~al.(2005)}]{cole/etal:2005}
2880: {Cole}, S., et~al. 2005, \mnras, 362, 505
2881:
2882: \bibitem[{{Cooray} \& {Sheth}(2002)}]{cooray/sheth:2002}
2883: {Cooray}, A. \& {Sheth}, R. 2002, \physrep, 372, 1
2884:
2885: \bibitem[{Copeland et~al.(2006)Copeland, Sami, \&
2886: Tsujikawa}]{copeland/sami/tsujikawa:2006}
2887: Copeland, E.~J., Sami, M., \& Tsujikawa, S. 2006, Int. J. Mod. Phys., D15, 1753
2888:
2889: \bibitem[{{Crocce} et~al.(2006){Crocce}, {Pueblas}, \&
2890: {Scoccimarro}}]{crocce/pueblas/scoccimarro:2006}
2891: {Crocce}, M., {Pueblas}, S., \& {Scoccimarro}, R. 2006, \mnras, 373, 369
2892:
2893: \bibitem[{{Crocce} \& {Scoccimarro}(2008)}]{crocce/scoccimarro:2008}
2894: {Crocce}, M. \& {Scoccimarro}, R. 2008, \prd, 77, 023533
2895:
2896: \bibitem[{{Croton} et~al.(2006)}]{croton/etal:2006}
2897: {Croton}, D.~J., et~al. 2006, \mnras, 365, 11
2898:
2899: \bibitem[{{De Lucia} \& {Blaizot}(2007)}]{delucia/blaizot:2007}
2900: {De Lucia}, G. \& {Blaizot}, J. 2007, \mnras, 375, 2
2901:
2902: \bibitem[{{Dunkley} et~al.(2008)}]{dunkley/etal:prep}
2903: {Dunkley}, J., et~al. 2008, ArXiv e-prints, 803
2904:
2905: \bibitem[{{Eisenstein} \& {Hu}(1998)}]{eisenstein/hu:1998}
2906: {Eisenstein}, D.~J. \& {Hu}, W. 1998, \apj, 496, 605
2907:
2908: \bibitem[{{Eisenstein} et~al.(2007){Eisenstein}, {Seo}, \&
2909: {White}}]{eisenstein/seo:2007}
2910: {Eisenstein}, D.~J., {Seo}, H.-J., \& {White}, M. 2007, \apj, 664, 660
2911:
2912: \bibitem[{{Eisenstein} et~al.(2005)}]{eisenstein/etal:2005}
2913: {Eisenstein}, D.~J., et~al. 2005, \apj, 633, 560
2914:
2915: \bibitem[{{Fry}(1996)}]{fry:1996}
2916: {Fry}, J.~N. 1996, \apjl, 461, L65+
2917:
2918: \bibitem[{{Fry} \& {Gaztanaga}(1993)}]{fry/gaztanaga:1993}
2919: {Fry}, J.~N. \& {Gaztanaga}, E. 1993, \apj, 413, 447
2920:
2921: \bibitem[{{Glazebrook} et~al.(2005){Glazebrook}, {Eisenstein}, {Dey}, {Nichol},
2922: \& {The WFMOS Feasibility Study Dark Energy Team}}]{glazebrook/etal:prep}
2923: {Glazebrook}, K., {Eisenstein}, D., {Dey}, A., {Nichol}, B., \& {The WFMOS
2924: Feasibility Study Dark Energy Team}. 2005, ArXiv Astrophysics e-prints
2925:
2926: \bibitem[{{Heavens} et~al.(1998){Heavens}, {Matarrese}, \&
2927: {Verde}}]{heavens/matarrese/verde:1998}
2928: {Heavens}, A.~F., {Matarrese}, S., \& {Verde}, L. 1998, \mnras, 301, 797
2929:
2930: \bibitem[{{Heitmann} et~al.(2007)}]{heitmann/etal:prep}
2931: {Heitmann}, K., et~al. 2007, ArXiv e-prints, 706
2932:
2933: \bibitem[{{Hill} et~al.(2004){Hill}, {Gebhardt}, {Komatsu}, \&
2934: {MacQueen}}]{hill/etal:2004}
2935: {Hill}, G.~J., {Gebhardt}, K., {Komatsu}, E., \& {MacQueen}, P.~J. 2004, in
2936: American Institute of Physics Conference Series, Vol. 743, The New Cosmology:
2937: Conference on Strings and Cosmology, ed. R.~E. {Allen}, D.~V. {Nanopoulos},
2938: \& C.~N. {Pope}, 224--233
2939:
2940: \bibitem[{{Hinshaw} et~al.(2003)}]{hinshaw/etal:2003}
2941: {Hinshaw}, G., et~al. 2003, \apjs, 148, 135
2942:
2943: \bibitem[{{Hinshaw} et~al.(2007)}]{hinshaw/etal:2007}
2944: ---. 2007, \apjs, 170, 288
2945:
2946: \bibitem[{{Hinshaw} et~al.(2008)}]{hinshaw/etal:prep}
2947: ---. 2008, ArXiv e-prints, 803
2948:
2949: \bibitem[{{Huff} et~al.(2007){Huff}, {Schulz}, {White}, {Schlegel}, \&
2950: {Warren}}]{huff/etal:2007}
2951: {Huff}, E., {Schulz}, A.~E., {White}, M., {Schlegel}, D.~J., \& {Warren}, M.~S.
2952: 2007, Astroparticle Physics, 26, 351
2953:
2954: \bibitem[{{Jain} \& {Bertschinger}(1994)}]{jain/bertschinger:1994}
2955: {Jain}, B. \& {Bertschinger}, E. 1994, \apj, 431, 495
2956:
2957: \bibitem[{{Jeong} \& {Komatsu}(2006)}]{jeong/komatsu:2006}
2958: {Jeong}, D. \& {Komatsu}, E. 2006, \apj, 651, 619
2959:
2960: \bibitem[{{Jing}(2005)}]{Jing:2005}
2961: {Jing}, Y.~P. 2005, \apj, 620, 559
2962:
2963: \bibitem[{{Kaiser}(1984)}]{kaiser:1984}
2964: {Kaiser}, N. 1984, \apjl, 284, L9
2965:
2966: \bibitem[{{Kennicutt}(1983)}]{kennicutt:1983}
2967: {Kennicutt}, R. C., Jr. 1983, \apj, 272, 54
2968:
2969: \bibitem[{{Kogut} et~al.(2003)}]{kogut/etal:2003}
2970: {Kogut}, A., et~al. 2003, \apjs, 148, 161
2971:
2972: \bibitem[{{Komatsu} et~al.(2008)}]{komatsu/etal:prep}
2973: {Komatsu}, E., et~al. 2008, ArXiv e-prints, 803
2974:
2975: \bibitem[{{Matarrese} \& {Pietroni}(2007)}]{matarrese/pietroni:2007}
2976: {Matarrese}, S. \& {Pietroni}, M. 2007, Journal of Cosmology and Astro-Particle
2977: Physics, 6, 26
2978:
2979: \bibitem[{{Matarrese} et~al.(1997){Matarrese}, {Verde}, \&
2980: {Heavens}}]{matarrese/verde/heavens:1997}
2981: {Matarrese}, S., {Verde}, L., \& {Heavens}, A.~F. 1997, \mnras, 290, 651
2982:
2983: \bibitem[{{Matsubara}(2008)}]{matsubara:2008}
2984: {Matsubara}, T. 2008, \prd, 77, 063530
2985:
2986: \bibitem[{{McDonald}(2006)}]{mcdonald:2006}
2987: {McDonald}, P. 2006, \prd, 74, 103512
2988:
2989: \bibitem[{{McDonald}(2007)}]{mcdonald:2007}
2990: ---. 2007, \prd, 75, 043514
2991:
2992: \bibitem[{{Meiksin} et~al.(1999){Meiksin}, {White}, \&
2993: {Peacock}}]{meiksin/white/peacock:1999}
2994: {Meiksin}, A., {White}, M., \& {Peacock}, J.~A. 1999, \mnras, 304, 851
2995:
2996: \bibitem[{{Nishimichi} et~al.(2007)}]{nishimichi/etal:2007}
2997: {Nishimichi}, T., et~al. 2007, \pasj, 59, 1049
2998:
2999: \bibitem[{{Nolta} et~al.(2008)}]{nolta/etal:prep}
3000: {Nolta}, M.~R. et~al. 2008, \apjs
3001:
3002: \bibitem[{{Page} et~al.(2007)}]{page/etal:2007}
3003: {Page}, L., et~al. 2007, \apjs, 170, 335
3004:
3005: \bibitem[{Peebles(1993)}]{peebles:POPC}
3006: Peebles, P. J.~E. 1993, Principles of Physical Cosmology (Princeton, NJ:
3007: Princeton University Press)
3008:
3009: \bibitem[{{Percival} et~al.(2007){Percival}, {Cole}, {Eisenstein}, {Nichol},
3010: {Peacock}, {Pope}, \& {Szalay}}]{percival/etal:2007c}
3011: {Percival}, W.~J., {Cole}, S., {Eisenstein}, D.~J., {Nichol}, R.~C., {Peacock},
3012: J.~A., {Pope}, A.~C., \& {Szalay}, A.~S. 2007, \mnras, 381, 1053
3013:
3014: \bibitem[{{Perlmutter} et~al.(1999)}]{perlmutter/etal:1999}
3015: {Perlmutter}, S., et~al. 1999, \apj, 517, 565
3016:
3017: \bibitem[{{Riess} et~al.(1998)}]{riess/etal:1998}
3018: {Riess}, A.~G., et~al. 1998, \aj, 116, 1009
3019:
3020: \bibitem[{{Sanchez} et~al.(2008){Sanchez}, {Baugh}, \&
3021: {Angulo}}]{sanchez/baugh/angulo:prep}
3022: {Sanchez}, A.~G., {Baugh}, C.~M., \& {Angulo}, R. 2008, ArXiv e-prints, 804
3023:
3024: \bibitem[{{Scoccimarro}(1998)}]{scoccimarro:1998}
3025: {Scoccimarro}, R. 1998, \mnras, 299, 1097
3026:
3027: \bibitem[{{Scoccimarro} et~al.(2001){Scoccimarro}, {Feldman}, {Fry}, \&
3028: {Frieman}}]{scoccimarro/etal:2001}
3029: {Scoccimarro}, R., {Feldman}, H.~A., {Fry}, J.~N., \& {Frieman}, J.~A. 2001,
3030: \apj, 546, 652
3031:
3032: \bibitem[{{Sefusatti} et~al.(2006){Sefusatti}, {Crocce}, {Pueblas}, \&
3033: {Scoccimarro}}]{sefusatti/etal:2006}
3034: {Sefusatti}, E., {Crocce}, M., {Pueblas}, S., \& {Scoccimarro}, R. 2006, \prd,
3035: 74, 023522
3036:
3037: \bibitem[{{Sefusatti} \& {Komatsu}(2007)}]{sefusatti/komatsu:2007}
3038: {Sefusatti}, E. \& {Komatsu}, E. 2007, \prd, 76, 083004
3039:
3040: \bibitem[{{Seljak}(2000)}]{seljak:2000}
3041: {Seljak}, U. 2000, \mnras, 318, 203
3042:
3043: \bibitem[{{Seo} \& {Eisenstein}(2003)}]{seo/eisenstein:2003}
3044: {Seo}, H.-J. \& {Eisenstein}, D.~J. 2003, \apj, 598, 720
3045:
3046: \bibitem[{Seo \& Eisenstein(2005)}]{seo/eisenstein:2005}
3047: Seo, H.-J. \& Eisenstein, D.~J. 2005, Astrophys. J., 633, 575
3048:
3049: \bibitem[{{Seo} et~al.(2008){Seo}, {Siegel}, {Eisenstein}, \&
3050: {White}}]{seo/etal:prep}
3051: {Seo}, H.-J., {Siegel}, E.~R., {Eisenstein}, D.~J., \& {White}, M. 2008, ArXiv
3052: e-prints, 805
3053:
3054: \bibitem[{{Sheth} \& {Tormen}(1999)}]{sheth/tormen:1999}
3055: {Sheth}, R.~K. \& {Tormen}, G. 1999, \mnras, 308, 119
3056:
3057: \bibitem[{{Shoji} et~al.(2008){Shoji}, {Jeong}, \&
3058: {Komatsu}}]{shoji/jeong/komatsu:prep}
3059: {Shoji}, M., {Jeong}, D., \& {Komatsu}, E. 2008, To be submitted
3060:
3061: \bibitem[{{Smith} et~al.(2007){Smith}, {Scoccimarro}, \&
3062: {Sheth}}]{smith/scoccimarro/sheth:2007}
3063: {Smith}, R.~E., {Scoccimarro}, R., \& {Sheth}, R.~K. 2007, \prd, 75, 063512
3064:
3065: \bibitem[{{Smith} et~al.(2008){Smith}, {Scoccimarro}, \&
3066: {Sheth}}]{smith/scoccimarro/sheth:2008}
3067: ---. 2008, \prd, 77, 043525
3068:
3069: \bibitem[{{Spergel} et~al.(2003)}]{spergel/etal:2003}
3070: {Spergel}, D.~N., et~al. 2003, \apjs, 148, 175
3071:
3072: \bibitem[{{Spergel} et~al.(2007)}]{spergel/etal:2007}
3073: ---. 2007, \apjs, 170, 377
3074:
3075: \bibitem[{{Springel}(2005)}]{springel:2005}
3076: {Springel}, V. 2005, \mnras, 364, 1105
3077:
3078: \bibitem[{{Springel} et~al.(2001){Springel}, {Yoshida}, \&
3079: {White}}]{springel/yoshida/white:2001}
3080: {Springel}, V., {Yoshida}, N., \& {White}, S.~D.~M. 2001, New Astronomy, 6, 79
3081:
3082: \bibitem[{{Springel} et~al.(2005)}]{springel/etal:2005}
3083: {Springel}, V., et~al. 2005, \nat, 435, 629
3084:
3085: \bibitem[{{Takada} et~al.(2006){Takada}, {Komatsu}, \&
3086: {Futamase}}]{takada/komatsu/futamase:2006}
3087: {Takada}, M., {Komatsu}, E., \& {Futamase}, T. 2006, \prd, 73, 083520
3088:
3089: \bibitem[{{Takahashi} et~al.(2008)}]{takahashi/etal:prep}
3090: {Takahashi}, R., et~al. 2008, ArXiv e-prints, 802
3091:
3092: \bibitem[{{Taruya} \& {Hiramatsu}(2008)}]{taruya/hiramatsu:2008}
3093: {Taruya}, A. \& {Hiramatsu}, T. 2008, \apj, 674, 617
3094:
3095: \bibitem[{{Tegmark}(1997)}]{tegmark:1997c}
3096: {Tegmark}, M. 1997, Physical Review Letters, 79, 3806
3097:
3098: \bibitem[{{Tegmark} \& {Peebles}(1998)}]{tegmark/peebles:1998}
3099: {Tegmark}, M. \& {Peebles}, P.~J.~E. 1998, \apjl, 500, L79+
3100:
3101: \bibitem[{{Tegmark} et~al.(2006)}]{tegmark/etal:2006}
3102: {Tegmark}, M., et~al. 2006, \prd, 74, 123507
3103:
3104: \bibitem[{{Valageas}(2007)}]{valageas:2007}
3105: {Valageas}, P. 2007, \aap, 465, 725
3106:
3107: \bibitem[{Weinberg(2008)}]{weinberg:COS}
3108: Weinberg, S. 2008, Cosmology (Oxford, UK: Oxford University Press)
3109:
3110: \bibitem[{{Yoshikawa} et~al.(2001){Yoshikawa}, {Taruya}, {Jing}, \&
3111: {Suto}}]{yoshikawa/etal:2001}
3112: {Yoshikawa}, K., {Taruya}, A., {Jing}, Y.~P., \& {Suto}, Y. 2001, \apj, 558,
3113: 520
3114:
3115: \end{thebibliography}
3116:
3117:
3118:
3119: \end{document}
3120:
3121: