1: \documentclass[11pt,preprint]{aastex}
2:
3: %\documentclass[manuscript]{aastex}
4:
5: \input epsf.tex
6: \voffset=-1cm
7:
8: \newcommand {\Mpc} {\mbox{$h^{-1}$ Mpc \,}}
9: \newcommand {\kpc} {\mbox{$h^{-1}$ kpc \,}}
10: \newcommand{\mincir}{\raise -2.truept\hbox{\rlap{\hbox{$\sim$}}\raise5.truept
11: \hbox{$?$}\ }}
12: \newcommand{\gr}{\kern 2pt\hbox{}^\circ{\kern -2pt K}} % ====? GRADI KELVIN
13: \newcommand{\magcir}{\raise -2.truept\hbox{\rlap{\hbox{$\sim$}}\raise5.truept
14: \hbox{$?$}\ }}
15: \newcommand{\oml}{\Omega_{\Lambda}}
16: \newcommand{\Om}{\Omega}
17: \newcommand{\om}{\omega}
18: \newcommand{\si}{\sigma}
19: \newcommand{\be}{\begin{equation}}
20: \newcommand{\ee}{\end{equation}}
21: \newcommand{\bea}{\begin{eqnarray}}
22: \newcommand{\eea}{\end{eqnarray}}
23: \newcommand{\De}{\Delta}
24: \newcommand{\La}{\Lambda}
25: \newcommand{\bn}{{\bf n}}
26: \newcommand{\bx}{{\bf x}}
27: \newcommand{\bk}{{\bf k}}
28: \newcommand{\etal}{{et al.}}
29: \newcommand{\ie}{{\it i.e.}}
30: \newcommand{\eg}{{\it e.g.}}
31:
32:
33: \shorttitle{Acoustic peaks and dips in the CMB power spectrum:
34: theory and observations}
35: \shortauthors{Durrer et al.}
36:
37: \begin{document}
38:
39: \title{Acoustic peaks and dips in the CMB power spectrum:\\
40: observational data and cosmological constraints}
41:
42: \author{R.~Durrer\altaffilmark{1}}
43: \affil{School of Natural Sciences, Institute for
44: Advanced Study, Einstein Drive, Princeton, NJ 08540}
45: \author{B.~Novosyadlyj \altaffilmark{2} and S. Apunevych}
46: \affil{Astronomical Observatory of Ivan Franko L'viv National University, Kyryla and
47: Mephodia str.8, 79005, L'viv, Ukraine}
48: \altaffiltext{1}{Department de Physique Th\'eorique, Universit\'e de Gen\`eve,
49: Quai Ernest Ansermet 24, CH-1211 Gen\`eve 4, Switzerland}
50: \altaffiltext{2}{on leave from Department de Physique Th\'eorique,
51: Universit\'e de Gen\`eve, Switzerland}
52:
53:
54:
55: \begin{abstract}
56: The locations and amplitudes of three acoustic peaks and two dips
57: in the Boomerang, MAXIMA and DASI measurements of
58: the cosmic microwave background (CMB) anisotropy power spectra
59: as well as their statistical confidence levels are determined in a
60: model-independent way. It is shown that the Boomerang-2001
61: data \cite{boom01} fixes the location and amplitude of the first
62: acoustic peak at more than $3\sigma$ confidence. The next
63: two peaks and dips are determined at a confidence level above
64: $1\sigma$ but below $2\sigma$.
65: The locations and amplitudes of the first three peaks and two dips are
66: $\ell_{p_1}=212\pm 17$, $A_{p_1}=5426\pm 1218\;\mu K^2$,
67: $\ell_{d_1}=413\pm 50$, $A_{d_1}=1960\pm 503\;\mu K^2$,
68: $\ell_{p_2}=544\pm 56$, $A_{p_2}=2266\pm 607\;\mu K^2$,
69: $\ell_{d_2}=746\pm 89$, $A_{d_2}=1605\pm 650\;\mu K^2$,
70: $\ell_{p_3}=843\pm 35$, $A_{p_3}=2077\pm 876\;\mu K^2$
71: respectively ($1\sigma$ errors include statistical and systematic errors).
72: The MAXIMA and DASI experiments give similar values
73: for the extrema which they determine. For MAXIMA these are the 1st and
74: 3rd peaks, for DASI the 1st and 2nd peaks and the 1st dip. Moreover,
75: the locations and amplitudes of the extrema determined from the combined
76: data of all experiments are quite close to the
77: corresponding values extracted from the Boomerang data alone.
78:
79: In order to use these data in a fast search for cosmological parameters
80: an accurate analytic approximation to calculate CMB peak and dip positions
81: and amplitudes in mixed dark matter models with cosmological constant
82: and curvature is derived and tested.
83:
84: The determined cosmological parameters from the CMB acoustic
85: extrema data show good agreement with other determinations, especially
86: with the baryon content as deduced from standard nucleosynthesis
87: constraints \cite{burles}.
88: These data supplemented by constraints from direct measurements of
89: some cosmological parameters and data on large scale structure (LSS)
90: lead to a best-fit model which agrees with practically all the used
91: experimental data within $1\sigma$. The best-fit parameters are:
92: $\Omega_{\Lambda}=0.64^{+0.14}_{-0.27}$, $\Omega_{m}=0.36^{+0.21}_{-0.11}$,
93: $\Omega_b=0.047^{+0.093}_{-0.024}$, $n_s=1.0^{+0.59}_{-0.17}$, $h=0.65^{+0.35}_{-0.27}$
94: and $\tau_c=0.15^{+0.95}_{-0.15}$. The best-fit values of $\Omega_{\nu}$ and $T/S$
95: are close to zero, their 1$\sigma$ upper limits are $0.17$ and $1.7$ respectively.
96:
97: \end{abstract}
98:
99: \keywords{cosmology: microwave background anisotropies
100: -- acoustic peaks -- cosmological parameters }
101:
102: \section{Introduction}
103: The new data on the cosmic microwave background (CMB) temperature
104: anisotropy obtained in the Boomerang (de Bernardis et al. 2000; Netterfield
105: et al. 2001), Maxima~I
106: \cite{maxima00,maxima01} and DASI \cite{dasi} experiments
107: provide relatively accurate measurements of the CMB anisotropies up to
108: $\ell \sim 1000$. Boomerang is a long duration balloon (LDB) flight
109: around the South Pole, MAXIMA is a balloon flight from Palestine,
110: Texas, and DASI is an interferometer experiment. The mutual
111: agreement of such divers experiments within statistical uncertainties
112: is very reassuring.
113:
114: After a correction of the first results from Boomerang \cite{boom00} by
115: Netterfield et al. (2001), these
116: measurements are in astounding agreement with the simplest
117: flat adiabatic purely scalar model of structure formation. The best fit
118: cosmological parameters obtained coincide with other, completely
119: independent determinations, like \eg~ the baryon density parameter
120: predicted by nucleosynthesis \cite{burles}.
121:
122: CMB anisotropies can be calculated within linear perturbation theory in a
123: multi-component universe. These calculations are very well established
124: and allow accurate predictions of the CMB power spectrum for a given
125: model of initial perturbations and given cosmological parameters. All
126: the calculations are linear and very well controlled. Publicly available
127: codes, \eg~ CMBfast \cite{cmbfast} provide 1\% accurate results for a
128: given model within two minutes of CPU time on an ordinary PC. Due to
129: these advantages, CMB temperature fluctuation data are extremely
130: valuable for testing theoretical models of structure formation and for the
131: determination of cosmological parameters.
132:
133: Nevertheless, the efficiency of parameter determination using codes like
134: CMBfast to compute the temperature anisotropy spectrum for each model
135: has several problems:
136: 1) The complete set of observational data of the current state and the
137: early history of the Universe is described by models with at least six
138: parameters. The implementation of CMBfast-like codes into search
139: procedures for best fits in high dimensional parameter spaces consumes
140: too much CPU time even for the most advanced computers due to the
141: necessity to carry out numerical integration of the Einstein-Boltzmann
142: system of equations which describe the evolution of temperature and
143: density perturbations of each component through the decoupling epoch.
144: 2) The CMB power spectrum alone has several more or less exact
145: degeneracies in parameter space (see \eg ~Efstathiou \& ~Bond (1999))
146: which can only
147: be reduced substantially or removed completely if other data sets,
148: \eg~ galaxy clustering data, corresponding to different scales and
149: redshifts, are combined with CMB measurements. The results and
150: especially the error bars which are obtained from search procedures
151: using different classes of cosmological observations with different
152: quality and different statistical properties are difficult to
153: interpret.
154:
155: Several groups have overcome the first problem by computing a grid of CMB
156: anisotropy spectra in the space of models and interpolating between them
157: to obtain the spectra for intermediate values of the parameters (see
158: Tegmark et al. (2001); Lange et al. (2001); Balbi et al. (2000); de
159: Bernardis et al. (2001); Wang et al. (2001) and references therein).
160:
161: Here we propose an alternative method:
162: The CMB angular power spectrum obtained by COBE \cite{cobe92},
163: Boomerang, MAXIMA-1 and DASI has well defined statistical and systematic
164: errors in the range of scales from quadrupole up to the spherical harmonic
165: $\ell\sim 1000$ and the present data can be represented by a few dozen
166: uncorrelated measurements. Practically the same information
167: is contained in a few characteristics such as the amplitude and
168: inclination of the power spectrum at COBE scale and the amplitudes and
169: locations of the observed acoustic peaks and dips.
170: Indeed, it was shown (see \cite{hs95,hw96,efs99,dl01} and references therein)
171: that the sensitivity of the locations and amplitudes of the extrema,
172: especially of the peaks, to cosmological parameters is the same as their
173: sensitivity to the flat band powers, the $C_l$'s, from presently available
174: data. This is not surprising: the extrema have an obvious physical
175: interpretation and have the largest weight among data points on
176: CMB power spectrum as a result of a minimal ratio of error to value.
177: Hu et al. (2001) have shown that most of the information from the Boomerang
178: and MAXIMA-1 data sets can be compressed into four observables: amplitude
179: and location of 1-st peak, and amplitudes of 2-nd and 3-rd peak.
180:
181: The first three
182: acoustic peaks and the two dips indicated by the above mentioned
183: experiments and the COBE large scale data can be presented by not more
184: than 12 experimental points. If we use the approach by
185: Bunn \& White (1997) for the four year COBE data and the data on
186: acoustic peaks, we have 7 experimental values to compare with
187: theoretical models. Each of them can be calculated by analytical or
188: semi-analytical methods. This enables us to study present CMB
189: data in a very fast search procedure for multicomponent models.
190:
191: Even though, in principle, a measurement of the entire $C_\ell$ spectrum
192: contains of course more informations than the position and amplitudes
193: of its peaks and dips, with present errorbars, this additional informations
194: seems to be quite modest.
195:
196: The goal of this paper is to use these main characteristics of the CMB
197: power spectrum to determine cosmological parameters. To do this we have to
198: accomplish the following steps: 1) to locate the positions and
199: amplitudes of three peaks and two dips as well as determining their
200: error bars from experimental data,
201: 2) to derive accurate analytical approximations to calculate these
202: positions and amplitudes and test them by full numerical
203: calculations. We also derive an accurate and fast semi-analytical method
204: to normalize the power spectrum to the 4-year COBE data.
205: Such analytical approximations have been derived in the past for the
206: matter power spectrum \cite{eh98,eh99,ndl99} and for the Sachs-Wolfe
207: part of the CMB anisotropy spectrum (Kofman \& Starobinsky 1985; Apunevych
208: \& Novosyadlyj 2000). Here we derive
209: an analogous approximation for the acoustic part of the CMB anisotropy
210: spectrum by improving an approximation proposed by Efstathiou \& ~Bond
211: (1999).
212:
213: The outline of the paper is as follows. In Section 2 we determine the
214: locations and amplitudes of the 1st, 2nd and 3rd acoustic peak as well as
215: 1st and 2nd dip and their confidence levels using
216: the published data on the CMB angular power spectrum from
217: Boomerang \cite{boom01}, MAXIMA-1 \cite{maxima01} and DASI (Halverson et
218: al., 2001).
219: Analytical approximations for the positions and amplitudes of the
220: acoustic peaks and dips are described in Section 3. A new method for an
221: accurate and fast COBE normalization is also presented in this section.
222: Details are given in two Appendices.
223: Our search procedure to determine cosmological parameters along with
224: the discussion of the results is presented in Section 4. In Section 5.
225: we draw conclusions.
226:
227: \section{Peaks and dips in the CMB power spectrum: experimental data}
228:
229: We have to determine the locations and amplitudes of acoustic peaks and
230: dips as well as their uncertainties in the data of the
231: angular power spectrum of CMB temperature fluctuations obtained in the
232: Boomerang \cite{boom01}, MAXIMA-1 \cite{maxima01} and DASI (Halverson et
233: al. 2001).
234: We carry out a model-independent analysis of the peaks and dips in the power
235: spectra for each experiment separately, as well as using all data points
236: jointly.
237:
238: \subsection{Boomerang-2001}
239:
240: A model-independent determination of peak and dip locations and amplitudes in
241: the Boomerang data \cite{boom01} has been carried out recently by de
242: Bernardis \etal ~(2001). Our approach is based on a conceptually somewhat
243: different method, especially in the determination of statistical errors.
244:
245: At first, mainly for comparison of our results with de Bernardis \etal
246: ~(2001),
247: we fit the peaks in the Boomerang CMB power spectrum by curves of second
248: order (parabolas) as shown in Fig.~\ref{peaks}. The six experimental points
249: ($N_{exp}=6$) in the range $100\le \ell \le 350$, which trace the first
250: acoustic peak, are well approximated by a three parameter curve
251: ($N_{par}=3$). Hence, the number of degrees of freedom
252: $N_f=N_{exp}-N_{par}$
253: for the determination of these parameters is $N_f= 3$. The best-fit
254: parabola has $\chi_{min}^2=2.6$. Its extremum, located at
255: $\ell_{p_1}=212$ and
256: $A_{p_1}=5426\;\mu K^2$, is the best-fit location and
257: amplitude of the first acoustic peak. We estimate the statistical error
258: in the following way. Varying the 3 parameters of the fitting
259: curve so that $\chi^2-\chi_{min}^2\le 3.53$ the maxima of the parabolas
260: define the $1\sigma$ range of positions and amplitudes of the first
261: acoustic peak in the plane $\left(\ell, \ell(\ell+1)C_{\ell}/2\pi\right)$.
262: The boundary of this region determines the statistical $1\sigma$ errors
263: for the location and amplitude of the first acoustic peak. We obtain
264: $$\ell_{p_1}=212^{+13}_{-20}, ~~ A_{p_1}=5426^{+540}_{-539}\;\mu K^2.$$
265: In Fig.~\ref{peaks} the $1\sigma$,
266: $2\sigma$ (the boundary of the region with $\chi^2-\chi_{min}^2\le
267: 8.02$) and $3\sigma$ ($\chi^2-\chi_{min}^2\le 14.2$) contours are shown
268: in the plane $\left(\ell, \ell(\ell+1)C_{\ell}/2\pi\right)$. All contours
269: for the first
270: acoustic peak are closed. This shows that the Boomerang-2001 data
271: prove the existence of a first peak at a confidence level higher than
272: $3\sigma$. The values $\Delta\chi^2=3.53$, 8.02 and 14.2, given by the
273: incomplete Gamma function, $Q(N_f,\Delta\chi^2/2)=1-0.683,~ 1-0.954$ and
274: $1-0.9973$, correspond to 68.3\%, 94.5\% and 99.73\% confidence levels
275: respectively for a Gaussian likelihood of $N_f=3$ degrees of freedom.
276: These levels which depend on $N_f$ and thus on the number of independent
277: data points (which we just took at face value from Netterfield \etal
278: ~(2001)) define the regions within which the maxima of parabolas leading
279: to the data points lie with a probability $\ge p$, where $p=0.68,~0.954$
280: and $0.9973$ for $1$-,
281: $2$- and $3$-$\sigma$ contours respectively. The same method for the
282: second peak using the eight experimental points in the range
283: $400\le \ell \le750$, hence $N_f=5$, gives
284: $$ \ell_{p_2}=541^{+40}_{-102},~~ A_{p_2}=2225^{+231}_{-227}\;\mu K^2.$$
285: For the third peak ($750\le \ell\le 1000$, 6 experimental points,
286: $N_f=3$) we obtain
287: $$ \ell_{p_3}=843^{+25}_{-42},~~ A_{p_3}=2077^{+426}_{-412}\;\mu K^2.$$
288:
289: For the second and third peaks, the $2$- and $3$-$\sigma$ contours are open
290: as shown in Fig.~\ref{peaks}. This means that the Boomerang experiment
291: indicates the 2nd and 3rd acoustic peaks at a confidence level higher than
292: $1$- but lower than $2$-$\sigma$. This is in disagreement with the result
293: obtained in \cite{ber01}. Formally the disagreement
294: consists in the fact that de Bernardis \etal~ (2001) set $N_f=2$ for all
295: peaks and dips (see paragraph 3.1.2 of their paper) leading to different
296: values of $\Delta\chi^2$ for the $1$-, $2$- and $3$-$\sigma$ contours. They
297: argue that there are two free parameters, namely the height and the
298: position of the peak. The 'philosophy' of the two approaches is somewhat
299: different:
300: While our contours limit the probability that the given data is measured if
301: the correct theoretical curve has the peak position and amplitude inside
302: the contour, in their approach the contours limit the probability that the
303: given best fit parabola leads to data with peak position and amplitude
304: inside the contour. In other words, while they compare a given parabola
305: to the best fit curve, we compare it to the data. In that sense we think
306: that the closed $2$-$\sigma$ contours of de Bernardis \etal (2001) do not
307: prove the existence of the secondary peaks at the 2-$\sigma$ level.
308: Of course our approach has a problem as well: It relies
309: on the data points being independent. If they are not, the number of
310: degrees of freedom should be reduced.
311:
312: The same procedure can be applied for the amplitudes and positions of the
313: two dips between the peaks.
314:
315: There is also a slight logical problem in
316: the results presented in Fig.~2 of de Bernardis \etal (2001):
317: If 2nd and 3rd peaks are established at $2\sigma$ C.L., then the 2nd dip
318: should be determined at the same C.L.; but even the
319: $1\sigma$ contour for the position of the second dip is not closed.
320: Also the position of the first dip is actually fixed by a single data point
321: at $\ell=450$ as can be seen from Fig.~3 in \cite{ber01}.
322: In order to remove these problems we approximate the experimental points
323: in the range $250\le \ell\le 850$ (12 experimental points) by one single
324: fifth order polynomial (6 free parameters). The number of degrees of
325: freedom $N_f=6$. The best-fit curve with $\chi_{min}^2=3.26$ is presented
326: in Fig.~\ref{dips}. Its 6 coefficients are
327: $a_0=1.13\cdot 10^5$, $a_1=-944.7$, $a_2=3.106$, $a_3=-4.92\cdot 10^{-3}$,
328: $a_4=3.75\cdot 10^{-6}$, $a_5=-1.099\cdot 10^{-9}$. This method allows us
329: to determine the locations and amplitudes of both dips and of the second
330: peak by taking into account the relatively prominent raises to the third and
331: especially to the first peak. The local extrema of the polynomial best-fit
332: give the following locations and amplitudes of the 1st and 2nd dip
333: (positive curvature extrema) and the 2nd peak (negative
334: curvature) between them. The $1\sigma$ error bars are determined as above:
335: $$ \ell_{d_1}=413^{+54}_{-27},\;\;\; A_{d_1}=1960^{+272}_{-282}\;\mu K^2,$$
336: $$ \ell_{p_2}=544^{+56}_{-52},\;\;\; A_{p_2}=2266^{+275}_{-274}\;\mu K^2,$$
337: $$ \ell_{d_2}=746^{+114}_{-63},\;\;\; A_{d_2}=1605^{+373}_{-436}\;\mu K^2.$$
338: The results discussed here are presented in Table~\ref{tab_boom} and
339: shown in Figs.~\ref{peaks} and \ref{dips}.
340: In Fig.~\ref{dips} also the $1\sigma$ ($\Delta\chi^2=7.04$),
341: $2\sigma$ ($\Delta\chi^2=12.8$) and $3\sigma$ ($\Delta\chi^2=20.1$)
342: confidence contours are shown.
343: The 1-$\si$ contours for all peaks and dips are now closed.
344: The $2\sigma$ contour for the 2nd peak
345: has a 'corridor' connecting it with the 3rd peak. As we have noted before
346: (see Fig.~\ref{peaks}), the $2\sigma$ confidence contour for the 3rd
347: peak is also open towards low $\ell$. This implies that we can not
348: establish the second peak at $2\sigma$ C.L. The probability of its
349: location is spread out over the entire range $450\le \ell \le 920$.
350: Therefore at $2\sigma$ C.L. we can not state whether the
351: Boomerang-2001 results indicate a second peak without a third, or third
352: without a second or both. We only can
353: state at $2\sigma$ C.L. that there are one or two negative curvature
354: extrema of the function $\ell(\ell+1)C_\ell/2\pi$ situated in the range
355: $450\le \ell \le 920$ with amplitude in the range
356: $1500\le \ell(\ell+1)C_\ell/2\pi \le 2700 \mu K^2$. Furthermore, there is
357: one secondary peak present in this $\ell$ range at $2\sigma$ confidence.
358: Now, the contours
359: for the dips are in logical agreement with the information about the
360: second and third peaks. If at $2\sigma$ C.L. the 2nd peak can be at
361: the range of the location of 3rd one, then the 1st dip will move to
362: $\ell\sim 520$. Its $2\sigma$ contour is closed since the 3rd peak
363: has a closed $2\sigma$ C.L. contour at the high $\ell$ side. On the
364: contrary, the 2nd dip is open at high $\ell$ as it disappears when the
365: 'second' peak disappears and the 'third' peak becomes the second.
366: The $3\sigma$ contours for the 2nd and 3rd peak as well as for 1st and
367: 2nd dips are open in the direction of high $\ell$.
368:
369: So far we discussed only the statistical errors. The Boomerang LDB
370: measurements have two systematic errors: $20\%$ calibration uncertainty
371: and beam width uncertainty leading to scale-dependent correlated
372: uncertainties in the determination of the power spectrum (Netterfield
373: et al. 2001).
374: The calibration error results in the same relative error for all data
375: points and can be taken into account easily. The beam width uncertainty
376: which induces an error which becomes larger at higher values of $\ell$,
377: needs more care.
378:
379: We estimate the beam width uncertainty as follows:
380: Using the data of Netterfield \etal~ (2001) for the $1\sigma$ dispersion
381: of the CMB
382: power spectrum due to beam width uncertainty, we have to estimate
383: its effect on determination of peak and dip locations and amplitudes. To
384: take into account the effect of a $1\sigma$ overestimated beam
385: width we have lowered the central points of the CMB power spectrum
386: presented in Table~3 of \cite{boom01} by multiplying them by the
387: $\ell$--dependent factor
388: $$f_{o}(\ell)=1-1.1326\cdot 10^{-4}\ell-2.72\cdot 10^{-7}\ell^2.$$
389: To take into account the effect of a $1\sigma$ underestimated beam width
390: we raise the central points of CMB power spectrum all by multiplication
391: with the factor
392: $$f_{u}(\ell)=1-6.99\cdot 10^{-5}\ell+5.53\cdot 10^{-7}\ell^2.$$
393: ($f_{o}(\ell)$ and $f_{u}(\ell)$ are best fits to the
394: Boomerang-2001 data and are shown in Fig.~\ref{beam}). For both cases we have
395: repeated the peak and dip determination procedure.
396: Best-fit values determined for the central points of CMB power spectrum
397: give us the $1\sigma$ errors of the peak/dip characteristics due to beam
398: width uncertainties. The results are presented in Table~\ref{tab_boom}. They
399: show that error bars of all peak and dip {\em locations} caused by the
400: beam width uncertainty are substantially less than the statistical
401: errors. But they dominate for the {\em amplitude} of the 3rd
402: peak and are comparable with the statistical error for the amplitudes
403: of the 1st dip, the 2nd peak and the 2nd dip. However, the beam size
404: errors are significantly smaller than statistical errors for the 1st
405: acoustic peak.
406:
407: Since all sources of errors have different nature and are statistically
408: independent they add in quadrature. The resulting symmetrized total errors
409: are shown in the before last and last columns of
410: Table~\ref{tab_boom}. They are used in the cosmological parameter search
411: procedure described in Section 4.
412:
413: \begin{figure}
414: %\epsfxsize=8truecm
415: \plotone{f1.eps}
416: \caption{Best-fit parabolas for the acoustic peaks of the Boomerang-2001 CMB
417: power spectrum ($\chi_{min}^2=2.59$ for the 1st peak, $\chi_{min}^2=0.92$ for
418: the 2nd peak and $\chi_{min}^2=0.65$ for the 3rd peak) as well as
419: $1$ (solid), $2$ (dotted) and $3\sigma$ (dashed) contours for their
420: locations and amplitudes are shown. The crosses indicate the top of the
421: best-fit parabolas. The contours limit the regions in the
422: $\left(\ell, \ell(\ell+1)C_{\ell}/2\pi\right)$ plane which contain the
423: tops of parabolas with
424: $\Delta\chi^2=3.53, \;8.02, \;14.2$ for the 1st and 3rd peaks ($N_f=3$) and
425: $\Delta\chi^2=5.89, \;11.3, \;18.2$ for the 2nd peak ($N_f=5$). }
426: \label{peaks}
427: \end{figure}
428:
429: \begin{figure}
430: %\epsfxsize=8truecm
431: \plotone{f2.eps}
432: \caption{The best polynomial fit for the Boomerang-2001 CMB power spectrum
433: in the range of the 1st dip, 2nd peak and 2nd dip ($\chi_{min}^2=3.26$), and
434: the $1,\;2\;{\rm and}\;3\sigma$ contours for their locations and amplitudes.
435: The crosses indicate the positive (dips) and negative (peak) curvature
436: extrema. The contours limit the regions in the $\left(\ell,
437: \ell(\ell+1)C_{\ell}/2\pi\right)$ plane containing the corresponding
438: extrema of polynomial fits with
439: $\Delta\chi^2=7.04,\;12.8,\;20.1$ ($N_f=6$).}
440: \label{dips}
441: \end{figure}
442:
443: \begin{figure}[ht]
444: %\epsfxsize=8truecm
445: \plotone{f3.eps}
446: \caption{The correction factors $f_{o}(\ell)$ (upward pointing triangles)
447: and $f_{u}(\ell)$ (downward pointing triangles) for the correlated CMB
448: power spectrum error caused by the uncertainty of the effective beam
449: width of the Boomerang experiment as given in Netterfield \etal~ (2001).
450: The solid lines are the fitting functions $f_o$ and $f_u$ given in the
451: text.}
452: \label{beam}
453: \end{figure}
454:
455:
456:
457: \begin{deluxetable}{crrrr}
458: \tabletypesize{\scriptsize}
459: \tablecaption{Best fit values for locations ($\ell_p$) and amplitudes
460: ($A_p$, $[\mu K^2]$) of the peaks and dips in the CMB temperature
461: fluctuation power spectrum measured by Boomerang
462: (Netterfield et al. 2001). Statistical errors (1st upper/lower values)
463: and errors caused by beam width uncertainties (2nd upper/lower values)
464: are shown in columns 2 and 3. The 20\% calibration uncertainty is
465: included in the symmetrized total errors presented in the last column.
466: \label{tab_boom}}
467: \tablewidth{0pt}
468: \tablehead{
469: \colhead{Features} & \colhead{$\ell_p$}&\colhead{$A_p$}&
470: \colhead{ $\ell_p$}&\colhead{$A_p$} \\ [4pt] }
471: \startdata
472: 1st peak &$212^{+13+2}_{-20-3}$ &$5426^{+540+112}_{-539-135}$&$212\pm 17$ &$5426\pm 1218$\\[4pt]
473: 1st dip &$413^{+54+6}_{-27-6}$ &$1960^{+272+142}_{-282-158}$&$413\pm 50$ &$1960\pm 503$\\[4pt]
474: 2nd peak &$544^{+56+14}_{-52-14}$ &$2266^{+275+309}_{-274-283}$&$544\pm 56$ &$2266\pm 607$\\[4pt]
475: 2nd dip &$746^{+114+9}_{-63-9}$ &$1605^{+373+422}_{-436-362}$&$746\pm 89$ &$1605\pm 650$\\[6pt]
476: 3rd peak &$843^{+26+5}_{-42-7}$ &$2077^{+426+720}_{-411-573}$&$843\pm 35$ &$2077\pm 876$\\[4pt]
477: \enddata
478: \end{deluxetable}
479:
480:
481:
482:
483:
484: \subsection{Adding DASI and MAXIMA-1 data}
485:
486: We have repeated the determination of peak and dip locations and amplitudes
487: with the data of two other experiments, DASI (Halverson et al. 2001) and
488: MAXIMA-1 \cite{maxima01}, released simultaneously with Boomerang-2001. Both
489: confirm the main features of the Boomerang CMB power spectrum: a
490: dominant first acoustic peak at $\ell\sim 200$, DASI shows a second peak
491: at $\ell\sim 540$ and MAXIMA-1 exhibits mainly a 'third peak' at
492: $\ell\sim 840$. The results presented in Figs.~\ref{all1},\ref{all3}
493: and \ref{all2} are quantitative figures of merit for their mutual
494: agreement and/or disagreement. In Fig.~\ref{all1} the
495: $1,\;2\;{\rm and}\;3\sigma$ contours for the first peak location and
496: amplitude for each experiment as well as the contours
497: for the combined data are presented.
498:
499: \begin{figure}
500: %\epsfxsize=8truecm
501: \plotone{f4.eps}
502: \caption{The location of the first acoustic peak in the plane
503: $\left(\ell, \ell(\ell+1)C_{\ell}/2\pi\right)$ for the Boomerang-2001
504: (blue diamond),
505: DASI (green triangle), MAXIMA-1 (red square) data and for all experiments
506: together (black cross) determined as maxima of corresponding best-fit
507: parabola. The $1,\;2,\;3\sigma$ confidence contours are also shown.
508: $\chi^2_{min}$, $N_f$ and $\Delta\chi^2$ for the first peak of the
509: Boomerang-2001 data are given in the caption of Fig.~\ref{peaks}, for
510: the other cases we have: DASI -- $\chi^2_{min}=1.8$, $N_f=1$ and
511: $\Delta\chi^2=1,\;4,\;9$ for the
512: $1,\;2,\;3\sigma$ confidence contours accordingly, MAXIMA-1 --
513: $\chi^2_{min}=4.22$, $N_f=2$ and $\Delta\chi^2=2.3,\;6.17,\;11.8$, all
514: experiments together -- $\chi^2_{min}=15.3$, $N_f=9$ and
515: $\Delta\chi^2=10.43,\;17.18,\;25.26$. The dominant contribution
516: to $\chi^2_{min}$ comes from the MAXIMA-1 data.}
517: \label{all1}
518: \end{figure}
519:
520:
521: In Fig.~\ref{all1} one sees that MAXIMA, like Boomerang,
522: indicates on the existence of the first acoustic peak at
523: approximately 3$\sigma$ C.L. But its $1\sigma$
524: contour for location of this peak in the $\left(\ell,
525: \ell(\ell+1)C_{\ell}/2\pi\right)$ plane does not intersect the Boomerang
526: $1\sigma$ contour, though their projections on
527: $\ell$ and $\ell(\ell+1)C_{\ell}/2\pi$ axes do. This can be caused by
528: systematic (normalization) errors inherent in both experiments.
529: Approximately a quarter of the area outlined by the Boomerang $2\sigma$
530: contour falls within the MAXIMA $2\sigma$ contour.
531: The experiments show the same level of agreement in the data on 3rd
532: acoustic peak (Fig. \ref{all3}). In the range of the 1st dip - 2nd peak
533: - 2nd dip, the MAXIMA data have no significant extrema,
534: even $1\sigma$ contours are open in both directions of the $\ell$ axis.
535:
536:
537: \begin{figure}
538: %\epsfxsize=8truecm
539: \plotone{f5.eps}
540: \caption{The location of the third acoustic peak in the plane $\left(\ell,
541: \ell(\ell+1)C_{\ell}/2\pi\right)$ for the Boomerang-2001 (blue diamond)
542: and MAXIMA-1 (red square) data and for both experiments together
543: (black cross), determined as maxima of the corresponding best-fit
544: parabola. The $1,\;2,\;3\sigma$ confidence contours
545: are also shown. $\chi^2_{min}$, $N_f$ and $\Delta\chi^2$ for the
546: Boomerang-2001 third acoustic peak is given in the caption of
547: Fig.~\ref{peaks}, for the other cases we have: MAXIMA-1 --
548: $\chi^2_{min}=0.67$, $N_f=1$ and $\Delta\chi^2=1,\;4,\;9$, all
549: experiments together -- $\chi^2_{min}=3.0$, $N_f=6$ and
550: $\Delta\chi^2=7.04,\;12.82, \;20.06$. The dominant contribution to
551: $\chi^2_{min}$ comes from the MAXIMA-1 data.}
552: \label{all3}
553: \end{figure}
554:
555:
556: The DASI experiment establishes the location and amplitude of the first
557: acoustic peak at somewhat more than $1\sigma$ but less than $2\sigma$.
558: The remarkable feature is the intersection of the $1\sigma$ contours of
559: DASI and Boomerang.
560: Approximately 1/5 of the area outlined by the MAXIMA $2\sigma$ contour
561: is within the corresponding DASI contour.
562:
563: Our analysis has also shown that the DASI data on the second acoustic
564: peak agree very well with Boomerang; the $1\sigma$ contours
565: nearly superimpose. The agreement of these two experiments is
566: impressive.
567:
568: We have repeated the
569: determination of peak and dip locations and amplitudes using the data of all
570: experiments jointly. The contours for the combined data are shown by the
571: thicker black lines in Figs. \ref{all1} and \ref{all3}. The C.L.
572: contours for the 1st dip, 2nd peak and 2nd dip determined as
573: regions of locations of negative and positive extrema of a 5-th order
574: polynomial fit are shown
575: in Fig.~\ref{all2}. The comparison with the corresponding figure for
576: the Boomerang data alone (Fig. \ref{dips}) shows their agreement.
577:
578:
579: \begin{figure}
580: %\epsfxsize=8truecm
581: \plotone{f6.eps}
582: \caption{The locations of the second acoustic peak and the two dips in the
583: $\left(\ell,\ell(\ell+1)C_{\ell}/2\pi\right)$ plane are shown together
584: with the best-fit 5-th order polynomial.
585: All points in this range from Boomerang, MAXIMA-1 and DASI have been
586: used jointly. The second peak and the dips are determined as
587: extrema of negative and positive curvature of the corresponding best fit
588: polynomial. The $1,\;2,\;3\sigma$ confidence contours are also shown.
589: For all experiments together, the best-fit gives $\chi^2_{min}=17.8$ for $
590: N_f=18$ degrees of freedom.}
591: \label{all2}
592: \end{figure}
593:
594:
595: The best-fit values of $\ell_{p_i}$, $A_{p_i}$
596: ($i=1,\; 2,\; 3$) and $\ell_{d_k}$, $A_{d_k}$ ($k=1,\; 2$) as well
597: as their $1\sigma$ statistical errors are given in
598: Table~\ref{tab_all}.
599:
600:
601: Other methods of model-independent determinations
602: of acoustic oscillation extrema were proposed by \cite{df01,mn01}.
603: Their analysis like ours finds low statistical
604: significance (less than $2\sigma$) for the detection of second and
605: third peaks.
606:
607:
608:
609:
610: \begin{deluxetable}{crrrrrr}
611: \tabletypesize{\scriptsize}
612: \tablecaption{Best fit values for the locations and amplitudes
613: of peaks and dips in the CMB temperature fluctuation power spectrum
614: from the DASI and MAXIMA-1 experiments. Statistical errors are
615: determined as described in the text. In the last column the results
616: obtained from the data of all three experiments together are presented.
617: \label{tab_all}}
618: \tablewidth{0pt}
619: \tablehead{
620: \colhead{}&\multicolumn{2}{c}{DASI}&\multicolumn{2}{c}{MAXIMA-1}
621: &\multicolumn{2}{c}{All three experiments}\\
622: \cline{2-3} \cline{4-5} \cline{6-7}\\
623: \colhead{Features} & \colhead{$ \ell_p$}&\colhead{$A_p$}&
624: \colhead{ $ \ell_p$}&\colhead{$A_p$}&\colhead{ $ \ell_p$}&
625: \colhead{$A_p$}\\ [4pt]
626: }
627: \startdata
628: 1st peak &$193^{+24}_{-45}$ &$4716^{+376}_{-351}$
629: &$236^{+20}_{-17}$ & $4438^{+743}_{-743}$
630: &$213^{+35}_{-59}$ &$5041^{+1017}_{-1196}$ \\[4pt]
631: 1st dip \tablenotemark{a}
632: &$378^{+15}_{-11}$ &$1578^{+170}_{-178}$
633: &$475^{+264}_{-83}$ &$1596^{+427}_{-443}$
634: &$406^{+97}_{-32}$ &$1843^{+385}_{-405}$ \\[4pt]
635: 2nd peak \tablenotemark{a}
636: &$536^{+30}_{-24}$ &$2362^{+176}_{-176}$
637: &$435-739$\tablenotemark{b} &$1500-2800$ \tablenotemark{b}
638: &$545^{+204}_{-89}$ &$2266^{+397}_{-609}$ \\[4pt]
639: 2nd dip \tablenotemark{a}
640: &$709^{+45}_{-46}$ &$1799^{+221}_{-308}$
641: &$435-739$ \tablenotemark{b}&$1000-2700$ \tablenotemark{b}
642: &$736^{+163}_{-117}$ &$1661^{+517}_{-663}$ \\[4pt]
643: 3rd peak & -- & -- & $813^{+286}_{-112}$ &$2828^{+1880}_{-1584}$
644: &$847^{+252}_{-146}$ &$2175^{+897}_{-836}$ \\[4pt]
645: \enddata
646: \tablenotetext{a}{The extrema were determined by
647: approximating the experimental CMB power spectrum by 5-th order polynomial}
648:
649: \tablenotetext{b}{Just the ranges where the probability to find the peak or
650: dip is $>68.3\%$ are indicated}
651: \end{deluxetable}
652:
653:
654: The peak locations and amplitudes from the
655: Boomerang-2001 CMB data presented in the Table~\ref{tab_boom}
656: show good quantitative agreement in the locations and, somewhat less
657: good, in the amplitudes obtained from the corresponding data of the
658: other experiments and all the data together. The agreement can be
659: improved when other error sources (calibration, beam width uncertainty,
660: cosmic variance etc) of each experiment are taken into account. With
661: some luck, the new mission MAP, which has been launched successfully
662: last June, will remove many of the current problems
663: and will considerably improve the data on CMB power spectrum.
664:
665: Clearly, the existence of the first peak in the spectrum is very well
666: established in the present experiments. It has already been established
667: (with less accuracy) before (see, eg. \cite{kp00,nd00}).
668: Finally we have also assessed approximately the probability for each of the
669: experiments to show no secondary peaks structure whatsoever. For the Boomerang
670: data this probability is less than 4.6\%, while for MAXIMA it is on the order
671: of 15\% and for DASI about 8\%.
672:
673:
674: \section{Analytic determinations of the locations and amplitudes of
675: the acoustic peaks and dips}
676:
677: In order to use the data in Tables~\ref{tab_boom} and~\ref{tab_all} to
678: determine cosmological parameters, we need a fast algorithm
679: to calculate the peak and dip positions for a given model. Here we
680: improve the analytical approximations of peak/dip positions
681: and amplitudes which have been derived in several papers
682: \cite{efs99,hu01,dn01,dl01}. We start by discussing the normalization
683: procedure.
684:
685: \subsection{Normalization of the density power spectrum}
686:
687: The 4-year COBE data, which establish the amplitude and the form of the
688: CMB power spectrum at the largest angular scales ($\ell\le 20$), are
689: taken into account via the approximation for $C_{10}$
690: proposed by~Bunn \& White (1997). This requires accurate calculations of
691: $C_\ell$ in the range $\ell\le 12$. The dominant contribution
692: on these angular scales is given by ordinary Sachs-Wolfe (SW) effect.
693: However, the Doppler (D) effect and the cross-correlation term Sachs-Wolfe --
694: adiabatic (SW-A) in the general expression for the correlation
695: function
696: $\langle{\Delta T\over T}({\bf n_1})\cdot {\Delta T\over T}({\bf n_2})\rangle$
697: have to be taken into account as well if we want to achieve an accuracy
698: better than 20\% (see Appendix A). For $\Lambda$ dark matter models
699: and models with non-zero 3-curvature, also the integrated Sachs-Wolfe
700: effect (ISW) contributes. We use the factors $K_\ell$ ($\ge 1$)
701: introduced and calculated by Kofman \&~ Starobinsky (1985) and improved
702: by Apunevych \&~ Novosyadlyj (2000), so that
703: $C_\ell^{\rm SW+ISW}= K_\ell^2C_\ell^{\rm SW}$ (for details see
704: Appendix~A).
705:
706: The normalization of the power spectrum of scalar perturbations then
707: consist in two steps:
708:
709: i) We calculate
710: \be
711: C_\ell=C_\ell^{\rm SW+ISW}+C_\ell^{\rm D}+C_\ell^{\rm A}+C_\ell^{\rm SW-A}
712: \label{f_cl}
713: \ee
714: (for $\ell=2,\;3,\;5,\;7,\;10,\;11,\;12$) by the analytical formulae
715: given in Appendix~A with arbitrary normalization.
716: This determines the shape of the CMB power spectrum in the range of
717: the COBE data, and hence the best-fit parameter $C_{10}^{\rm COBE}$
718: to 4-year COBE and the first and second derivatives as defined in
719: Bunn \& White (1997) for models with given cosmological parameters;
720:
721: ii) Since each term in the expression (\ref{f_cl}) is
722: $\propto \delta_h^2$,
723: where $\delta_h$ is the present matter density perturbations at horizon
724: scale, we can now determine $\delta_h$ and along with it the value of
725: the normalization constant for scalar perturbations
726: $A_s=2\pi^{2}\delta_{h}^{2}(3000{\rm Mpc}/h)^{3+n_s}$ for a model with
727: given cosmological parameters. Here $n_s$ is the spectral index for
728: primordial scalar density perturbations
729: and $h$ is dimensionless Hubble parameter (in units of 100 km/sec/Mpc).
730:
731: Both these steps are also performed in CMBfast. Hence, our
732: normalization procedure for the power spectrum is equivalent to
733: normalization with CMBfast. Calculations show that our value
734: $C_{10}^{\rm COBE}$ never differs from the result of CMBfast by more
735: than $3\%$. The accuracy of the overall normalization constant
736: $\delta_h$ for $\Lambda$DM models with appropriate values of
737: parameters is better then $5\%$. This has been controlled by comparing
738: the value of $\sigma_8$ from CMBfast with our semi-analytical
739: approach. This error simply reflects the accuracy of the analytical
740: approximation of the transfer function for
741: density fluctuations by Eisenstein \&~ Hu (1999) which we have used.
742:
743:
744: \subsection{Positions and amplitudes of CMB extrema: analytic approach}
745:
746: One of the main ingredients for our search procedure is a fast and
747: accurate calculation of the positions and amplitudes of the acoustic
748: peaks and dips, which depend on cosmological parameters.
749:
750: The dependence of the position and amplitude of the first acoustic
751: peak of the CMB power spectrum on cosmological
752: parameters has been investigated using CMBfast.
753: As expected, the results are, within reasonable
754: accuracy, independent of the hot dark matter
755: contribution ($\Om_{\nu}$). This was also shown by Novosyadlyj \etal~
756: (2000). For the remaining parameters, $n_s$, $h$, $\Omega_b$,
757: $\Omega_{cdm}$ and $\Omega_{\Lambda}$, we determine the resulting
758: values $\ell_{p_1}$ and $A_{p_1}$ using the analytical approximation
759: given by Efstathiou \&~ Bond (1999) and Durrer \&~ Novosyadlyj (2001).
760: In these papers the CMB anisotropy spectrum is approximated in the
761: vicinity of the first acoustic peak by
762: \bea
763: \nonumber
764: {\ell(\ell+1)\over 2\pi}C_{\ell}={\ell(\ell+1)\over 2\pi}(C^{SW}_{\ell}+0.838C^{SW}_2\cdot\\
765: {\cal A}(\Omega_b,\Omega_{cdm},\Omega_k,n_s,h)\exp
766: \left[ -{(\ell-\ell_{p_1})^2\over 2(\Delta\ell_{p_1})^2}\right])~,
767: \eea
768: where $\Delta\ell_{p_1}=0.42\ell_{p_1}$, $C^{SW}_{\ell}$ is the Sachs-Wolfe
769: approximation for the $C_{\ell}$s derived in the Appendix ~A,
770: Eq.~(A9), and
771: \bea
772: \nonumber
773: {\cal A}(\Omega_b,\Omega_{cdm},\Omega_k,n_s,h) = \\
774: \nonumber
775: \exp{[a_1+a_2\omega_{cdm}^2+a_3\omega_{cdm}+a_4\omega_b^2+a_5\omega_b+}\\
776: {+a_6\omega_b\omega_{cdm}+a_7\omega_k + a_8\omega^2_k+a_9(n_s-1)]}.
777: \eea
778: Here
779: $\omega_b\equiv\Omega_bh^2$, $\omega_{cdm}\equiv\Omega_{cdm}h^2$,
780: $\omega_k\equiv(1-\Omega_m-\Omega_{\Lambda})h^2$.
781: The position of the acoustic peaks is determined as in \cite{efs99} for
782: open and flat models and in \cite{dn01} for closed models.
783: The coefficients $a_i$ are defined by fitting to the numerical CMBfast
784: amplitudes of the first acoustic peak on a sufficiently wide grid of
785: parameters. We find: $a_1=2.503$, $a_2=8.906$, $a_3=-7.733$,
786: $a_4=-115.6$, $a_5=35.66$, $a_6=-7.225$, $a_7=1.96$, $a_8=-11.16$,
787: $a_9=4.439$. The accuracy of the approximation is better than 5\% in
788: the parameter range $0.2\le\Omega_m\le 1.2$,
789: $0\le\Omega_{\Lambda}\le 0.8$, $0.015\le \Omega_b\le 0.12$,
790: $0.8\le n_s\le 1.2$ and $0.4\le h\le 1.0$. The approximation for the
791: amplitude breaks down in the models with large curvature
792: ($\Omega_k\le -0.2$ and $\Omega_k\ge 0.6$) and low baryon density, ($\omega_b
793: \ll 0.006$).
794:
795: To calculate the amplitudes of the 2nd and 3rd peaks, we use the
796: analytic relations for the relative heights of these peaks w.r.t
797: the first peak as given by Hu et al. (2001).
798: \bea
799: \nonumber
800: A_{p_2}=A_{p_1}H_2(\Omega_m,\Omega_b,n_s), \\
801: A_{p_3}=A_{p_1}H_3(\Omega_m,\Omega_b,n_s),
802: \eea
803: where the functions $H_2$ and $H_3$ are given in Eqs. (B16) and (B17)
804: respectively. For the locations of 2nd and 3rd peaks we
805: use the analytic approximations given by Hu \etal~ (2001)
806: and Doran \& Lilley (2001) (see also Appendix B).
807:
808: Unfortunately, we have no analytic approximation for the dip amplitudes
809: and hence we can not use their experimental values to determine
810: cosmological parameters. But a sufficiently accurate analytic
811: approximation for the location of the 1st dip is given in Doran \&~ Lilley
812: (2001). We use it here.
813:
814: Hence, we have analytical approximations for the dependences of the
815: positions and amplitudes of three acoustic peaks and the location of the
816: 1st dip on cosmological parameters
817: \bea
818: &\ell_{p_i}(\Omega_m,\Omega_{\Lambda},\Omega_k,\Omega_b,h),&\nonumber \\
819: &A_{p_i}(\Omega_m,\Omega_{\Lambda},\Omega_k,\Omega_b,n_s,h),&\;\;\; (i=1,2,3) \nonumber \\
820: &\ell_{d_1}(\Omega_m,\Omega_{\Lambda},\Omega_k,\Omega_b,h).& \nonumber
821: \eea
822: Comparing the analytical values for different sets of parameters
823: with numerical calculations using CMBfast, shows that the accuracy is
824: about 5\% for all locations
825: and amplitudes of 1st and 3rd peaks in the ranges of cosmological parameters
826: indicated above.
827: The accuracy for amplitude of the 2nd peak is always
828: better than $9$\% in the same ranges.
829: For some parameter values the second peak is underestimated.
830:
831: Of course, in principle the half-widths of peaks and dips, their "smoothness",
832: "sharpness" or their inflection points may contain additional information
833: on cosmological parameters. But when one compares CMB spectra obtained in the
834: different experiments up to date, one concludes that their accuracy is not
835: sufficient to influence the resulting cosmological parameters at the
836: present precision. Therefore,
837: we only use the most prominent observable patterns of the CMB power spectrum -
838: locations and amplitudes of acoustic peaks.
839:
840:
841:
842: For the convenience, we present all analytic approximations
843: used here in Appendix B.
844:
845:
846: \section{Cosmological parameters from the CMB peak amplitudes and
847: locations}
848:
849: We now use the results of Sections 2 and 3 to determine the cosmological
850: parameters $\Omega_m$, $\Omega_{\Lambda}$, $\Omega_{\nu}$ (one sort of
851: massive neutrino), $\Omega_b$, $n_s$, $h$, $T/S$
852: ($\equiv C^{tensor}_{10}/C^{scalar}_{10}$) and $\tau_c$
853: (optical depth to decoupling). We use the method described in detail in a
854: previous paper \cite{dn01}. We include 8 experimental points from the CMB
855: power spectrum (COBE $C_{10}$, the amplitudes and locations of three acoustic
856: peaks and the location of the first dip). In order to have a positive number
857: of degrees of freedom, $N_f\ge 1$ we add a weak constraint for the Hubble
858: constant, $0.5\le h \le 0.8$ when searching for 8 parameters.
859:
860: In order to establish $1\sigma$ confidence intervals for each parameter we
861: have applied the marginalization procedure described in
862: (Durrer \& Novosyadlyj 2001). The results are presented in
863: Table~\ref{tab_par}.
864:
865:
866: At first we check how the best fit parameters depend on the accuracy of
867: the peak/dip locations and amplitudes.
868: For this we compare the resulting cosmological parameters from the data
869: given in Table \ref{tab_boom} if we consider only statistical errors
870: and with total errors. The results are given in the first two rows of
871: Table~\ref{tab_par}. In spite of the different errors of the experimental
872: values and their relations (the total errors of peak amplitudes increase
873: faster with the peak number than the statistical error) the best fit
874: cosmological parameters are similar.
875:
876: In order to estimate the sensitivity of cosmological parameters to experimental
877: values we substitute the Boomerang data on peak/dip locations and amplitudes
878: by the values from Table~\ref{tab_all}, obtained from all three experiments
879: combined. The comparison of the results in the third row with those above
880: shows that the results are practically unchanged. We believe that the
881: Boomerang data on peak/dip locations and amplitudes are best studied and
882: have well established statistics, hence we use them in the following
883: determinations.
884:
885: The neutrino contents for these three data sets can be rather large,
886: $\Omega_\nu \sim 0.3$. This is due to the low sensitivity of
887: the CMB anisotropy spectrum to $\Omega_{\nu}$. When
888: $\Omega_{\nu}=0$ is fixed, the best-fit values for the remaining parameters
889: stay practically unchanged and also $\chi^2$ increases only very little.
890: The CMB doesn't care whether dark matter should be hot or cold. In order to
891: distinguish between cold and hot dark matter data which is
892: sensitive to the density power spectrum on smaller scales needs to be
893: added. In Table~\ref{tab_par} we therefore exclude $\Omega_{\nu}$ from
894: the determination procedure and fix its value to $0$ in the first five rows.
895: Contrary, for $T/S$ and $\tau_c$ we obtain $0$, but with large
896: $1\sigma$ confidence limits due to the degeneracy in $T/S$, $\tau_c$ and $n_s$
897: \cite{efs99}.
898:
899:
900: A remarkable result is the good agreement of the best-fit content of
901: baryons $\Omega_bh^2\approx 0.02$ with the constraint from standard
902: nucleosynthesis and the observed intergalactic content of light
903: elements \cite{burles}.
904: The large $1\sigma$ confidence limits for $\Omega_m$ and $\Omega_{\Lambda}$ are due to
905: the well known degeneracy of the CMB power spectrum in these
906: parameters \cite{efs99}. However, the sum of their best fit values,
907: $\Omega_m+\Omega_\Lambda = 1 -\Omega_k$ is always very close to $1$ which
908: implies that spatial curvature is small for the best fit model.
909:
910: \begin{deluxetable}{crrrrrrrrr}
911: \tabletypesize{\scriptsize}
912: \tablecaption{Cosmological parameters from the extrema of the CMB angular
913: power spectrum in combination with other cosmological data sets.
914: The upper/low values show $1\sigma$ confidence limits which are obtained by maximizing
915: the (Gaussian) 68 percent confidence contours over all other parameters.
916: The LSS data set is the same as in \cite{dn01}. \label{tab_par}}
917: \tablewidth{0pt}
918: \tablehead{
919: \colhead{Observable data set} & \colhead{$\chi^2_{min}/N_f$}&
920: \colhead{$\Omega_{\Lambda}$}&\colhead{$\Omega_m$}&\colhead{$\Omega_{\nu}$}
921: &\colhead{$\Omega_b$}&\colhead{$n_s$}&\colhead{$h$}&\colhead{$T/S$}
922: &\colhead{$\tau_c$} \\ [4pt] }
923: \startdata
924: CMB$_{(Boom, stat.)}$&1.01/2&$0.69^{+0.23}_{-0.56}$&$0.31^{+0.61}_{-0.21}$&0$^{*)}$&$0.055^{+0.13}_{-0.028}$&$0.89^{+0.81}_{-0.08}$&$0.65^{+0.23}_{-0.24}$&0$^{+27}$&0$^{+1.65}$ \\[4pt]
925: CMB$_{(Boom, total)}$&0.95/2&$0.64^{+0.31}_{-1.42}$&$0.36^{+1.04}_{-0.35}$&0$^{*)}$&$0.057^{+0.18}_{-0.047}$&$0.89^{+0.97}_{-0.14}$&$0.65^{+0.23}_{-0.23}$&0$^{+44}$&0$^{+1.90}$ \\[4pt]
926: CMB$_{(All, stat.)} $&0.09/2&$0.63^{+0.35}_{-1.35}$&$0.37^{+1.04}_{-0.36}$&0$^{*)}$&$0.051^{+0.29}_{-0.05}$&$0.90^{+1.30}_{-0.11}$&$0.65^{+0.23}_{-0.24}$&0$^{+20}$&0$^{+1.75}$ \\[4pt]
927: CMB$_{(Boom, total)}$+&&&&&&&& \\[4pt]
928: $h$ $\&$ BBN\tablenotemark{a} &1.11/3&$0.69^{+0.26}_{-1.30}$&$0.31^{+1.05}_{-0.24}$&0$^{*)}$&$0.047^{+0.048}_{-0.018}$&$0.90^{+0.56}_{-0.10}$&$0.65^{+0.20}_{-0.19}$&$0^{+2.7}$&$0^{+0.90}$\\[4pt]
929: $h$, BBN $\&$ SNIa\tablenotemark{b} &1.11/4&$0.72^{+0.17}_{-0.21}$&$0.29^{+0.15}_{-0.13}$&0$^{*)}$&$0.047^{+0.048}_{-0.02}$&$0.90^{+0.60}_{-0.12}$&$0.65^{+0.22}_{-0.19}$&$0^{+3.5}$&$0^{+1.1}$\\[4pt]
930: $h$, BBN $\&$ LSS\tablenotemark{c} &8.22/11&$0.46^{+0.31}_{-0.46}$&$0.48^{+0.52}_{-0.22}$&$0.06^{+0.20}_{-0.06}$&$0.047^{+0.12}_{-0.026}$&$1.03^{+0.59}_{-0.23}$&$0.66^{+0.31}_{-0.31}$&$0^{+3.5}$&$0.15^{+0.95}_{-0.15}$\\[4pt]
931: $h$, BBN, SNIa $\&$ LSS\tablenotemark{d}&10.4/12&$0.64^{+0.14}_{-0.27}$&$0.36^{+0.21}_{-0.11}$&$0.00^{+0.17}$&$0.047^{+0.093}_{-0.024}$&$1.0^{+0.59}_{-0.17}$&$0.65^{+0.35}_{-0.27}$&$0^{+1.7}$&$0.15^{+0.95}_{-0.15}$\\[4pt]
932: the same\tablenotemark{e}&11.6/14&$0.61^{+0.16}_{-0.26}$&$0.37^{+0.21}_{-0.13}$&$0.00^{+0.11}$&$0.041^{+0.043}_{-0.023}$&$0.95^{+0.17}_{-0.14}$&$0.70^{+0.34}_{-0.20}$&0$^{*)}$&0$^{*)}$ \\[4pt] \enddata
933:
934: \tablenotetext{*)} {This parameter is fixed to $0$.}
935: \tablenotetext{a} {The big bang nucleosynthesis constraint on baryon content $\widetilde{\Omega_bh^2}=0.02\pm 0.001$ from Burles \etal~(2001) is included.}
936:
937: \tablenotetext{b} {The constraint on the $\Omega_{\Lambda}-\Omega_m$
938: relation from SNIa distance measurements \cite{per99},
939: $\widetilde{[\Omega_m-0.75\Omega_{\Lambda}]}=-0.25\pm 0.125$ is added.}
940:
941: \tablenotetext{c} {In addition to the parameters given in the different
942: columns, we have also to determine the Abell-ACO biasing parameter, $b_{cl}$.
943: The result is: $b_{cl}=2.64\pm 0.27$}
944:
945: \tablenotetext{d} {For this data set we obtain $b_{cl}=2.47\pm 0.19$}
946: \tablenotetext{e} {$b_{cl}=2.5\pm 0.2$}
947:
948: \end{deluxetable}
949:
950:
951: We repeat the search procedure for different combinations of the CMB power
952: spectrum extrema data (Table \ref{tab_boom}) with other cosmological
953: data sets. The LSS data set used here ranges from the Lyman alpha
954: forest, determining amplitude and spectral index of the matter power
955: spectrum at very small scales, to large scale bulk velocities, cluster
956: abundances and Abell cluster catalogs which determine $\sigma_8$ and
957: the position of the 'knee' in the matter power spectrum. All of this is
958: extensively discussed in \cite{dn01}.
959:
960: The results for cosmological parameters from different combinations
961: of observational data are shown in the lines 4 to 8 of Table~\ref{tab_par}.
962:
963: Adding a stronger constraint on the Hubble parameter, $h =0.65\pm 0.10$,
964: and the big bang nucleosynthesis (BBN) constraint
965: changes the best fit cosmological parameters only slightly. The
966: SNIa constraint on the relation between $\Omega_{\Lambda}$ and $\Omega_m$
967: \cite{per99} substantially reduces the errors
968: of these parameters (5th line in Table \ref{tab_par}) as
969: it removes the degeneracy between them. This degeneracy is also removed
970: when we combine CMB and LSS data.
971:
972: The cosmological parameters obtained from the Boomerang CMB power
973: spectrum extrema data combined with all other cosmological measurements
974: (a detailed list can be found in Durrer \& Novosyadlyj 2001)
975: are presented in lines 6th, 7th and 8th of Table~\ref{tab_par}. The best fit
976: values for the tensor mode amplitude $T/S$ defined as $C^T_{10}/C^S_{10}$
977: in the last two cases are practically zero but the $1\sigma$ confidence
978: limits are wide due to the degeneracy of the CMB extrema in $n_s,~T/S$ and
979: $\tau_c$. Even when combining the CMB with LSS data, the degeneracy in
980: $n_s$ and $T/S$ is not significantly removed. This is so, since a blue
981: spectrum, which allows for a high tensor contribution to the CMB, can be
982: compensated with a neutrino component which leads to damping of the matter
983: power spectrum on small scales. If massive neutrinos are not allowed, the
984: degeneracy between $n_s$ and $T/S$ is lifted as soon as small scale LSS
985: data is included.
986:
987: The best-fit values of spectral index
988: $n_s$ in all cases are in the $1\sigma$ range of the value obtained from the
989: COBE 4-year data, $n_s= 1.2\pm 0.3$ \cite{ben96,cobe96}.
990: When using the best fit model to calculate the data used to find it,
991: practically all results are within the $1\sigma$ error range of the
992: corresponding experimental data. Only two out of 31 experimental points are
993: slightly outside. Namely the best-fit value of $\Omega_m-0.75\Omega_{\Lambda}$
994: in the last determination is at $1.1\sigma$ lower of its experimental value
995: followed from SNIa test and $\sigma_8$ constraint established by \cite{bah98}
996: from the existence of three massive clusters of galaxies is at $1.18\sigma$
997: higher than model predicted value. But the value of $\sigma_8$ in our
998: best-fit model, $\sigma_8=0.91$, is in the range of current estimates
999: from the Sloan Digital Sky Survey $\sigma_8^{(SLOAN)}=0.915\pm 0.06$,
1000: \cite{sdss01}), which is not included in our data set.
1001: The high degree of consistency within completely independent cosmological data
1002: sets is very encouraging.
1003:
1004: Moreover, all parameters of our best-fit model agree well with those
1005: extracted from the full Boomerang data \cite{boom01} combined with
1006: LSS and SNIa priors (compare our values
1007: derived from CMB+LSS in the 7th row of Table 3 and theirs in
1008: the 4th row of Table 4). But our 1$\sigma$ ranges for
1009: most parameters are significantly wider. That is due to two facts. First
1010: of all, we allow also for a tensor component and neutrinos. This increase in
1011: the number of parameters also increases the degeneracies (e.g. between the
1012: tensor amplitude and the spectral index $n_s$) thereby enlarging the
1013: errors of the physical parameters. In this sense our parameter errors
1014: are rather to be compared with those of Wang et al. (2001) which allow
1015: for roughly the same degrees of freedom. But even their parameter estimation
1016: is somewhat more precise than ours. This is because
1017: we use more conservative errors for the peak and dip
1018: locations and amplitudes which include
1019: statistical, normalization and beam uncertainties.
1020: We think that with present data, and with the model assumptions made by our
1021: choice of parameters, our precision is realistic. Clearly, future
1022: experiments like the MAP satellite will improve this situation.
1023:
1024: Finally, to compare with the cosmological parameters obtained
1025: in our previous paper (Durrer \& Novosyadlyj 2001, Table 4),
1026: where the same LSS data set was used, we have repeated the search procedure
1027: fixing $T/S =\tau_c=0$. The best-fit values of the parameters
1028: with $1\sigma$ errors obtained by maximizing the confidence contours over all
1029: other parameters are given in the last row of Table \ref{tab_par}. Comparing
1030: them with values in the last column of Table 4 from \cite{dn01} shows that both
1031: determinations have best-fit values in the $1\sigma$ confidence limits
1032: of each other. There the best fit model has a slight positive curvature, here
1033: a slightly negative. The $1\sigma$ confidence ranges here are somewhat
1034: wider than those obtained in the previous determination. These differences
1035: are due to the different CMB observable data set and the different
1036: normalization procedure. Even though in our previous analysis we have only
1037: taken into account the first peak. The errors in its location and amplitude
1038: were significantly underestimated, leading to smaller error bars.
1039:
1040:
1041: \section{Conclusions}
1042:
1043: We have carried out a model-independent analysis of recent CMB power
1044: spectrum measurements in the Boomerang \cite{boom01}, DASI \cite{dasi} and
1045: MAXIMA (Lee et al. 2001) experiments and we have determined the locations and
1046: amplitudes of the first three acoustic peaks and two dips as well as their
1047: confidential levels (Table 1-2, Fig. 1-7).
1048:
1049: In the Boomerang experiment the second and third acoustic peaks are
1050: determined at a confidence level somewhat higher than $1\sigma$.
1051: Experimental errors which include statistics and systematics are
1052: still too large to establish the secondary peak locations and amplitudes
1053: at $2\sigma$ C.L. Only the position of one (the third ) secondary peak can
1054: be bounded from above $\ell_{p_3}\le 900$, at $2\sigma$ C.L. The same
1055: situation is encountered when determining the locations and amplitudes of
1056: the first and second dips. However, the location and amplitude of the first
1057: peak, are well established with confidence level, higher than $3\sigma$.
1058:
1059: The MAXIMA experiment also shows the existence of the first acoustic
1060: peak at approximately the same confidence level as Boomerang.
1061: But the $1\sigma$ contours for the peak position in the plane
1062: $\left(\ell, \ell(\ell+1)C_{\ell}/2\pi\right)$ do not intersect. However,
1063: their projections
1064: onto the $\ell$ and $\ell(\ell+1)C_{\ell}/2\pi$ axes do. Approximately one
1065: quarter of the area inside the Boomerang $2\sigma$ contour falls within the
1066: corresponding MAXIMA contour. The same level of agreement of
1067: these experiments is found in the data on the 3rd acoustic peak. In the
1068: range of 1st dip - 2nd peak - 2nd dip the MAXIMA data give no significant
1069: information. Even the $1\sigma$ contours are open in both directions of the
1070: $\ell$ axis.
1071:
1072: The DASI experiment establishes the location and amplitude of the first
1073: acoustic peak at somewhat higher than $1\sigma$ C.L. but less than
1074: $2\sigma$. The $1\sigma$ contours for the position of the first peak
1075: of the DASI and Boomerang experiments intersect.
1076: Approximately 1/5 of the area outlined by the MAXIMA $2\sigma$ contour is
1077: within the DASI $2\sigma$ contour. The DASI data on second acoustic peak
1078: is in excellent agreement with the Boomerang results - the $1\sigma$
1079: contours practically coincide.
1080:
1081: We have also determined the locations and amplitudes of the acoustic peaks
1082: and dips using the data of all three experiments. The results are very
1083: close to those from the Boomerang data alone.
1084:
1085: To determine cosmological parameters from these data, we
1086: have improved the analytical approximations for the peak positions and
1087: amplitudes to an accuracy (determined by comparing the approximations with
1088: the results of CMBfast) better than 5\% in a sufficiently wide range
1089: of parameters.
1090: % $0.2\le\Omega_m\le 1.2$, $0\le\Omega_{\Lambda}\le 0.8$,
1091: % $0.015\le \Omega_b\le 0.12$, $0.8\le n_s\le 1.2$, $0.4\le h\le 0.9$.
1092: We have also developed a fast and accurate analytical method to
1093: normalize the power spectrum to the 4-year COBE data
1094: on $C_{10}$. Our analytical approximation is accurate to a few percent
1095: (in comparison to CMBfast)
1096: when all main effects (ordinary Sachs-Wolfe effect, integrated Sachs-Wolfe
1097: effect, adiabatic term, Doppler term and their mutual cross-correlations)
1098: are taken into account. For example, in the
1099: model with parameters presented in the last row of Table~\ref{tab_par} the
1100: relation of contribution from these components at $\ell=10$ are
1101: $C_{10}^{\rm SW}:C_{10}^{\rm A}:C_{10}^{\rm SW-A}:C_{10}^{\rm D}=1:0.098:-0.24:0.42$.
1102:
1103: The cosmological parameters extracted from the data on locations and
1104: amplitudes of the first three peaks and the location of the first dip are in
1105: good agreement with other determinations (Netterfield et al. 2001; de
1106: Bernardis et al. 2001; Pryke et al. 2001; Wang et al. 2001; Durrer \&
1107: Novosyadlyj 2001).
1108: That shows also that present CMB data can essentially be compressed into the
1109: height and slope of the Sachs-Wolfe plateau (at $\ell=10$) and the positions
1110: and amplitudes of the first three acoustic peaks and the first two dips.
1111:
1112: A remarkable feature is the coincidence of the baryon content obtained from
1113: the CMB data, $\Omega_bh^2\approx 0.02$ with the value from
1114: standard nucleosynthesis ($0.02\pm 0.001$) \cite{burles}. Moreover, the
1115: CMB data together with constraints from direct measurements of the
1116: Hubble constant, the SNIa data, the baryon content and the large scale
1117: structure of the Universe (the power spectrum of rich clusters,
1118: the cluster mass function, the peculiar velocity field of galaxies,
1119: Ly-$\alpha$ absorption lines as seen in quasar spectra) select a best-fit
1120: model which gives predictions within about 1$\sigma$ error bars of all
1121: measurements. The cosmological parameters of this model are
1122: $\Omega_{\Lambda}=0.64^{+0.14}_{-0.27}$, $\Omega_{m}=0.36^{+0.21}_{-0.11}$,
1123: $\Omega_b=0.047^{+0.083}_{-0.024}$, $n_s=1.0^{+0.59}_{-0.17}$,
1124: $h=0.65^{+0.35}_{-0.27}$ and $\tau_c=0.15^{+0.95}_{-0.15}$. The
1125: best-fit values of $\Omega_{\nu}$ and $T/S$ are close to zero, their
1126: 1$\sigma$ upper limits are $\Omega_{\nu}\le 0.17$, $T/S\le 1.7$.
1127:
1128: The cosmological parameters determined from the CMB acoustic peak/dip
1129: locations and amplitudes data show good agreement with other cosmological
1130: measurements and indicate the existence of a simple (adiabatic) best-fit
1131: model for all the discussed cosmological data within the accuracy
1132: of present experiments.
1133:
1134:
1135: \acknowledgments
1136: It is a pleasure to acknowledge stimulating discussions with Alessandro
1137: Melchiorri, Roman Juszkiewicz and Roberto Trotta.
1138: BN is grateful to the Tomalla foundation for a
1139: visiting grant and to Geneva University for hospitality. RD thanks the
1140: Institute for Advanced Study for hospitality and acknowledges support
1141: from the Monell Foundation.
1142:
1143: \appendix
1144: \begin{center}
1145: {\Large\bf APPENDIX} \vspace{0.5cm}\\
1146: \end{center}
1147:
1148: \section{An analytic approximation for the CMB power spectrum at large
1149: angular scales}
1150: On sufficiently large angular scales (larger than the Silk damping scale)
1151: temperature fluctuations in the CMB
1152: can be related to density, velocity and metric perturbations at the
1153: last scattering surface and at later times by integrating the geodesic
1154: equation, similar to the classical paper by Sachs and Wolfe (1967).
1155: Here we discuss only scalar perturbations. Tensor perturbations can be
1156: simply added to the result and do not pose any significant difficulty.
1157: Scalar perturbations generate CMB temperature fluctuations which can
1158: be written in gauge-invariant form as a sum of four terms -- the
1159: ordinary Sachs-Wolfe effect, the integrated Sachs-Wolfe term, the Doppler term
1160: and the acoustic term \cite{d90}.
1161: \begin{eqnarray}
1162: \left({\De T\over T}\right)^{(s)}(\eta_0,\bx_0,\bn) &=&
1163: {1\over 4}D_r(\eta_{dec}, \bx_{dec})+
1164: V_i(\eta_{dec}, \bx_{dec})n^i
1165: +(\Phi-\Psi)(\eta_{dec}, \bx_{dec}) \nonumber \\
1166: && -\int_{\eta_{dec}}^{\eta_0}
1167: (\Phi'-\Psi')(\eta,\bx(\eta))d\eta~. \label{ani}
1168: \end{eqnarray}
1169: Here $\Phi$ and $\Psi$ are the Bardeen potentials \cite{bar80}, $V_i$ is
1170: the baryon velocity and $D_r$ is a gauge invariant variable for the radiation
1171: density fluctuations. A prime denotes the partial derivative w.r.t.
1172: conformal time $\eta$. For perfect fluids and for dust we have $\Psi=-\Phi$.
1173: In Newtonian limit the Bardeen potentials just reduce to the ordinary
1174: Newtonian potential. For adiabatic perturbations
1175: ${1\over 4}D_r = {1\over 3}\delta_m - {5\over 3}\Phi$ (see \eg~
1176: Durrer \& Straumann (1999)),
1177: where $\delta_m$ is the usual matter density perturbation., it
1178: corresponds to $\epsilon_m$ in Bardeen's notation. The variables
1179: $\eta$ and $\bx$ are conformal time and comoving position.
1180:
1181: In realistic models, cosmological recombination and decoupling of radiation
1182: from matter take place when
1183: $\rho_{m}>\rho_{r}$. Hence the large angular scale
1184: CMB power spectrum can be expressed in the terms of solutions of Einstein's
1185: equations for adiabatic linear perturbations in a dust Universe.
1186: The CMB anisotropies on angular scales $\theta\ge 10^o \;(\ell\le 20)$
1187: are generated mainly by the linear perturbations of matter density,
1188: velocity and the gravitational potential at scales much larger than the
1189: particle horizon at decoupling. Our approximation makes use of these facts.
1190:
1191: We use the solutions of Einstein's equations for linear density perturbations
1192: in flat models of a Universe with dust and a cosmological constant which can
1193: be found in \cite{ks85,an00}. The growing mode of density, velocity
1194: and gravitational potential perturbations, using the gauge-invariant
1195: variables introduced by Bardeen (1980) and normalizing the scale factor
1196: $a(t_0)=a_0=1$, are
1197: \be
1198: \Phi(t,k)=K_{\delta}(t)C(k),~\delta_m(t,k)=-{2C(k)k^2a(t) K_{\delta}(t)\over
1199: 3H_0^2\Omega_m},~
1200: V^{\alpha}(t,k)=-i{2C(k)k^{\alpha} a(t)\dot a(t)K_V(t)\over 3H_0^2\Omega_m}.
1201: \label{sol}
1202: \ee
1203: $C(k)$ is (up to the time dependent factor $K_\delta$) the Fourier
1204: transform of the Bardeen potential, so that
1205: $\Phi(t,{\bx})=(2\pi)^{-3/2}\int\Phi(t,\bk)e^{i{\bf kx}}d^3k$,
1206: $\delta_m(t,{\bf x})=(2\pi)^{-3/2}\int\delta(t,\bk)e^{i{\bf kx}}d^3k$ and
1207: $V^{\alpha}(t,{\bx})=(2\pi)^{-3/2}\int V^{\alpha}(t,\bk) e^{i{\bf kx}}d^3k$.
1208: The factors
1209: $K_{\delta}(t)\equiv {5\over 3}\left(1-\dot a/a^{2}\int_0^tadt \right)$ and
1210: $K_{V}(t)\equiv{5\over 3}\left(\dot a/a^{2}-\ddot a/a\dot a\right)\int_0^tadt$
1211: are both in the range $0< K_\bullet \le 1$ and reflect the reduction of growth
1212: of perturbations caused by the cosmological constant. The scale factor of the
1213: background model is given by
1214: $$ a(t)= \left({\Omega_{m} \over 1-\Omega_{m}} \right)^{1 \over 3}
1215: \sinh^{2 \over 3}\left({3H_{0}t\sqrt{1-\Omega_{m}}} \over 2 \right)$$
1216: Here $H_0\equiv (\dot a/a)(t_0) =\dot a (t_0)$ is the Hubble constant today.
1217: The $K_\bullet$-factors go to $1$ when $t\ll t_0$ or when
1218: $\Omega_m \rightarrow 1$ ($\Omega_{\Lambda} \rightarrow 0$). At decoupling
1219: $K_{\delta}=K_{V}=1$. An analytical approximation for $K_{\delta}(t)/\Omega_m$
1220: with sufficient accuracy can be found in \cite{car92} and for
1221: $K_{V}(t_0)/\Omega_m$ in \cite{la}.
1222:
1223: The power spectrum of density fluctuations is given by
1224: \bea
1225: P(k,t)&\equiv& <\delta(t,k) \delta^*(t,k)> = A_sk^{n_s}T_m^2(k;t)a^2(t)
1226: K^2_{\delta}(t)/\Omega_m^2,\\ \nonumber
1227: A_s &=& 2\pi^{2}\delta_{h}^{2}(3000{\rm Mpc}/h)^{3+n_s},
1228: \label{pkz}
1229: \eea
1230: where $T_m(k,t)$ is transfer function (divided by the growth factor) and
1231: $\delta_h$ is the present matter density perturbation at horizon scale. We use
1232: the analytical approximation of $T_m(k,t)$ in the space of cosmological
1233: parameters $h$, $\Omega_m$, $\Omega_b$, $\Omega_{\Lambda}$,
1234: $\Omega_{\nu}$ and $N_{\nu}$ (number of species of massive neutrino)
1235: by Eisenstein \& Hu (1999).
1236:
1237: From Eq.~(\ref{ani}), taking into account adiabaticity and setting
1238: $\bx_0={\bf 0}$, we obtain
1239: \be
1240: {\Delta T\over T}({\bf n})={1\over 3}\Phi(\eta_{dec},{\bf n}\eta_{0})+
1241: 2\int_{0}^{\omega_{e}}{\partial \Phi(\eta_0-\om,{\bf n}\om )
1242: \over\partial\eta}d\omega+
1243: n_{\alpha} V^{\alpha}(\eta_{dec},{\bf n}\eta_{0})+{1\over 3}\delta_m(\eta_{dec},{\bf n}\eta_{0}),
1244: \label{dt}
1245: \ee
1246: where {\bf n} is the unit vector in direction of the incoming photon and we
1247: have used $\bx(\eta)={\bf n}(\eta_0-\eta)$, $\bx_{dec}\simeq{\bf n}\eta_0$.
1248: The variable $\omega$ is the affine parameter along the geodesic
1249: which begins at the observer and ends in the emission point at the
1250: last scattering surface. The present value of conformal times, $\eta_0$
1251: gives also the present particle horizon or the distance to the last-scattering
1252: surface.
1253: The first term in (\ref{dt}) is the well known Sachs-Wolfe effect (SW),
1254: the second term is the integrated Sachs-Wolfe effect (ISW) which is important
1255: only at late times, where $K_{\delta}(t)$ starts to deviate from $1$ and
1256: ${\partial \Phi\over \partial \eta} \ne 0$,
1257: the third is the Doppler term (D) and the last is the acoustic term (A).
1258: At large angular scales ($\approx 10^{o})$, where anisotropies have been
1259: measured by COBE \cite{ben96}, the SW and ISW effects dominate. However,
1260: if we want to calculate $C_{10}$ with good accuracy, we must to also
1261: take into account the other terms.
1262: The angular correlation function of $\Delta T/T$ can be written
1263: symbolically as
1264: \bea
1265: \nonumber
1266: <{\Delta T\over T}({\bf n_1})\cdot {\Delta T\over T}({\bf n_2})>={\rm <SW\cdot SW>+2<SW\cdot ISW>+<ISW\cdot ISW>}\\
1267: {\rm +<A\cdot A>+2<SW\cdot A>+<D\cdot D>}~.
1268: \label{dtcf}
1269: \eea
1270: The cross-correlators ${\rm <D\cdot SW>}$ and ${\rm <D\cdot A>}$ are omitted
1271: because they are strongly suppressed on large angular scales.
1272: Indeed, if one uses Fourier presentations for the variables (\ref{sol}) in the
1273: equations (\ref{dt}-\ref{dtcf}) one finds that the $k$-integrand of
1274: these terms contains a spherical Bessel function $j_1(k\eta_0(\bn_1-\bn_2))$
1275: which oscillates for large angular separations, strongly reducing the integral
1276: if compared to the ${\rm <SW\cdot A>}$ term where the integrand has a
1277: definite sign.
1278: The terms ${\rm <ISW\cdot A>}$ and ${\rm <ISW\cdot D>}$ are also omitted
1279: because the ISW effect gives the maximal contribution to $\Delta T/T$ at
1280: the largest angular scales of the range of interest (at lowest spherical
1281: harmonics) where A and D are nearly zero.
1282: At 'smaller' angular scales ($\ell\approx 10$) where contribution of A
1283: and D are not negligible, the ISW effect is very small. Therefore, their
1284: cross-correlation terms are very small.
1285:
1286: We develop the ${\bf n}$-dependence of ${\Delta T\over T}({\bf n})$
1287: in spherical harmonics
1288: $$
1289: {\Delta T\over T}({\bf n})=\sum_{\ell,m}a_{\ell m}(\eta_0)Y_{\ell m}({\bf n}),
1290: \;\;< a_{\ell m} a_{\ell^{\prime}m^{\prime}}^*>= \delta_{\ell m}
1291: \delta_{\ell^{\prime}m^{\prime}}C_\ell~.
1292: $$
1293: The CMB power spectrum, $C_\ell$, has the same components as the
1294: correlation function:
1295: \be
1296: C_\ell=C_\ell^{\rm SW}+ C_\ell^{\rm SW-ISW}+ C_\ell^{\rm ISW}+ C_\ell^{\rm A}+ C_\ell^{\rm SW-A}+ C_\ell^{\rm D}
1297: \label{cl}
1298: \ee
1299: Each component on the right hand side comes from the corresponding
1300: contribution to $\Delta T$ above and is proportional to $ \delta_h^2$.
1301: Using the solutions (\ref{sol}) we obtain analytic approximations for them.
1302:
1303:
1304: We first approximate the SW and ISW contributions in the form
1305: \be
1306: C_\ell^{\rm SW+ISW}\equiv C_\ell^{\rm SW}+ C_\ell^{\rm SW-ISW}+
1307: C_\ell^{\rm ISW}=K^2_\ell C_\ell^{\rm SW},
1308: \ee
1309: where the factors $K_\ell$ ($\ge 1$) take into account the contribution
1310: of the ISW effect for each spherical harmonic. They have been calculated
1311: by Kofman \& Starobinsky (1985) and Apunevych \& Novosyadlyi (2000) for
1312: different $\Lambda$-models. Instead of
1313: the direct time consuming calculations of the ISW contribution, we use
1314: the following analytic approximations:
1315: \bea
1316: \nonumber
1317: K^2_2=1+8.20423\times\exp(-\Omega_m/0.01157)+
1318: 3.75518\times\exp(-\Omega_m/0.13073),\\
1319: \nonumber
1320: K^2_3=1+2.25571\times\exp(-\Omega_m/0.03115)+
1321: 2.35403\times\exp(-\Omega_m/0.15805),\\
1322: \nonumber
1323: K^2_4=1+1.80309\times\exp(-\Omega_m/0.0323)+
1324: 1.88325\times\exp(-\Omega_m/0.16163)\\
1325: \nonumber
1326: {\rm and}\;\;\; K^2_\ell=1+[23.46523\times\exp(-\Omega_m/0.0122)+
1327: 11.03227\times\exp(-\Omega_m/0.14558)]/(\ell+0.5)
1328: \eea
1329: for $\ell\ge 5$. These approximation formulae are determined from
1330: the data presented in the tables of \cite{ks85} and \cite{an00}.
1331:
1332: Using solutions (\ref{sol}) and the definition of the density power
1333: spectrum~(\ref{pkz}) we obtain the following general expression for the SW
1334: contribution to the CMB power spectrum:
1335: \be
1336: C_\ell^{\rm SW}={\pi\eta_0^{n_s-1}\delta_h^2\over 2^{n_s-1}D^2(t_0)}
1337: \int_0^\infty dk k^{n_s-2}T_m^2(t_{dec},k)j_\ell^2(k\eta_0),
1338: \label{clswi}
1339: \ee
1340: where $T_m^2(t_{dec},k)$ is the transfer function of matter density
1341: perturbations at decoupling, $D(t_0)\equiv K_{\delta}(t_0)/\Omega_m$ is
1342: the value of the
1343: growth factor at the current epoch, and $j_\ell$ is the spherical Bessel
1344: function of order $\ell$. For reasonable values of spectral index
1345: $-3\le n_s \le 3$ the main contribution to the integral~(\ref{clswi}) comes
1346: from very small $k$ where $T_m(t_{dec},k)\approx 1$ and can be
1347: omitted. Then integral can be performed analytically and the result can be
1348: expressed in terms of $\Gamma$-functions:
1349: \be
1350: C_\ell^{\rm SW}={\pi^2\delta_h^2\over 8D^2(t_0)}{\Gamma(3-n_s)
1351: \Gamma(\ell+{n_s-1\over 2})\over \Gamma^2(2-n_s/2)\Gamma(\ell+{5-n_s\over 2})}.
1352: \label{clsw}
1353: \ee
1354: In the same way we obtain the expressions for the other components of
1355: equation (\ref{cl}):
1356: \bea
1357: C_\ell^{\rm A}={\pi\eta_0^{n_s+3}\delta_h^2a^2(t_{dec})\over 18\cdot2^{n_s}D^2(t_0)\Omega_m^2}\int_0^\infty dk
1358: k^{n_s+2}T_b^2(t_{dec},k)j_\ell^2(k\eta_0),\\
1359: \label{cla}
1360: C_\ell^{\rm SW-A}=-{\pi\eta_0^{n_s+1}\delta_h^2a(t_{dec})\over 3\cdot2^{n_s-1}D^2(t_0)\Omega_m}\int_0^\infty dk
1361: k^{n_s}T_m(t_{dec},k)T_b(t_{dec},k)j_\ell^2(k\eta_0),\\
1362: \label{clswa}
1363: C_\ell^{\rm D}={\pi\eta_0^{n_s+1}\delta_h^2a(t_{dec})\over 2^{n_s-1}D^2(t_0)\Omega_m}\int_0^\infty dk
1364: k^{n_s}T_b^2(t_{dec},k)j_\ell^{\prime 2}(k\eta_0),
1365: \label{cld},
1366: \eea
1367: where $T_b(t_{dec})$ is transfer function for density perturbations of
1368: baryons \cite{eh98} and $(^{\prime})$ is the derivative w.r.t the argument
1369: $x=k\eta_0$. The minus sign in the expression for $C_\ell^{\rm SW-A}$ reflects
1370: the anti-correlation of the gravitational potential and density fluctuations:
1371: large positive density fluctuations generate deep negative potential wells.
1372: If we set $T_m=T_b=1$, the integrals (\ref{cla} --\ref{cld}) diverge for
1373: all $\ell$ for $n_s\ge 1$ because the main contribution to integrals of
1374: the D and A terms comes from small scales. Hence here the transfer functions
1375: must be kept and the integrals have to be calculated numerically. Fortunately,
1376: the integrands decay rapidly for large wave numbers and 99.9\% of the
1377: contribution comes from the range $0.001\le k\eta_{dec}\le 0.1$, so that the
1378: integration is not very time consuming.
1379:
1380: In Fig.\ref{fig_eff} the CMB power spectrum $C_\ell$ at large angular scales
1381: ($\theta\ge 20^o,\;\;\ell\le 20$) together with the contributions from
1382: the different terms given in (\ref{clsw}-\ref{cld}) is shown
1383: for a pure matter and a $\Lambda-$ dominated model.
1384: The relation of the contributions from different terms at $\ell=10$ are
1385: $$
1386: C_\ell^{\rm SW}:C_\ell^{\rm A}:C_\ell^{\rm SW-A}:C_\ell^{\rm D}=1:0.04:-0.11:0.22
1387: $$
1388: for the matter dominated flat model ($\Omega_m=1$) and
1389: $$
1390: C_\ell^{\rm SW}:C_\ell^{\rm A}:C_\ell^{\rm SW-A}:C_\ell^{\rm D}=1:0.08:-0.23:0.39.
1391: $$
1392: for the $\Lambda$ dominated model with the cosmological parameters shown in
1393: the figure. Therefore, a few percent accuracy of the normalization to
1394: 4-year COBE $C_{10}$ data can be achieved only if all these effects
1395: are taken into account.
1396:
1397: \begin{figure}
1398: %\epsfxsize=8truecm
1399: \plottwo{f7.eps}{f8.eps}
1400: \caption{The CMB power spectrum and the different contributions discussed
1401: in the text (formulae \ref{clsw}-\ref{cld}) for a pure matter model
1402: (left panel) and a $\Lambda$ dominated model (right panel). }
1403: \label{fig_eff}
1404: \end{figure}
1405:
1406:
1407: In Fig. \ref{fig_cl} the CMB power spectrum at large scales
1408: calculated using the analytic formulae (\ref{clsw}-\ref{cld}) and using
1409: CMBfast are shown for comparison.
1410: In the left panel we also present the power spectrum calculated by
1411: the analytical approach of \cite{hs95} (renormalized to the CMBfast value
1412: of $C_{10}$).
1413:
1414: The calculations show that value of $C_{10}$ calculated by our method
1415: deviates from the value calculated with CMBfast by 0.5\% for the matter
1416: dominated flat model ($\Omega_m=1$) and 2.7\% for the $\Lambda$ dominated
1417: model ($\Omega_m=0.2$). Therefore, our analytic approach is sufficient to
1418: normalize fast the power spectrum of scalar perturbations
1419: to the 4-year COBE data with virtually the same precision as CMBfast,
1420: the difference is less than $3\%$. (Remember, that the experimental
1421: errors of the COBE data are about $ 14\%$, so that
1422: the best-fit normalization parameter $C^{COBE}_{10}$ has the same error.)
1423:
1424: Another comparison of our normalization procedure with CMBfast comes from
1425: the value of $\sigma_8$. For the flat model (left panel of
1426: Fig.~\ref{fig_cl}) our approximation for the normalization together with
1427: the analytical transfer function of Eisenstein \& Hu (1998; 1999) leads to
1428: $\si_8=1.58$, the corresponding value calculated from CMBfast is 1.53.
1429: For the $\Lambda$ dark matter model (right panel) our $\sigma_8=0.62$, while
1430: CMBfast gives $\sigma_8=0.64$. The agreement of both approaches is
1431: quite well (the 5\% difference includes also the errors in the approximation
1432: of the transfer function which is actually of this order).
1433:
1434: \begin{figure}
1435: %\epsfxsize=8truecm
1436: \plottwo{f9.eps}{f10.eps}
1437: \caption{The CMB power spectrum at COBE scales calculated by CMBfast
1438: (solid line) and by our analytical formulae (\ref{clsw}-\ref{cld})
1439: (dotted line). For the pure matter model we also show the spectrum
1440: calculated with the analytic approach of \cite{hs95} (dashed line) (this
1441: approach does not allow a cosmological constant). All spectra are normalized
1442: to the best fit for $C_{10}$ from the 4-year COBE data given in \cite{bw97}.}
1443: \label{fig_cl}
1444: \end{figure}
1445:
1446: The slight deviation of $C_{10}$ as calculated by our code from the value
1447: obtained with CMBfast ($\le 3\%$) in spite of using the same analytic
1448: best-fit formula for $C_{10}^{COBE}$ by \cite{bw97} is due to a difference
1449: in the form of the spectra as shown in the Fig.~\ref{fig_cl}. This difference
1450: grows when $\Omega_m$ decreases. There are several possible reasons for this
1451: deviation in the form of the CMB power spectrum in our analytic approach
1452: from the exact numerical calculation:
1453: 1) We have used the solutions for the
1454: evolution of density, velocity and gravitational potential perturbations in
1455: the $\Lambda$- dust Universe. In reality, at decoupling the role
1456: of radiation is not completely negligible, this slightly influences the
1457: dynamics of the scale factor and the evolution of perturbations.
1458: It also results in additional time dependence of gravitational
1459: potential (early integrated Sachs-Wolfe effect) which is not taken into
1460: account here.
1461: 2) Our approach does not take into account the effects of the collisionless
1462: dynamics of photons and neutrinos after decoupling. Especially, the induced
1463: anisotropic stresses lead to $\approx 10\%$ difference of the gravitational
1464: potentials in the radiation-dominated epoch which results into a corrections
1465: of a few percent in the $C_\ell$'s.
1466: 3) Instantaneous recombination and tight coupling which were assumed, also
1467: cause slight inaccuracies. They
1468: should, however, be extremely small on the angular scales considered here.
1469: 4) To calculate the terms $C_\ell^{A}$, $C_\ell^{SW-A}$ and $C_\ell^{D}$
1470: we have used the analytic approximations for the transfer functions,
1471: $T_m(t_{dec},k)$ and $T_b(t_{dec},k)$,
1472: by Eisenstein \& Hu (1998; 1999) which have an accuracy $\sim 5\%$.
1473:
1474: More details on the theory of CMB anisotropies can be found in the
1475: reviews by Durrer \& Straumann (1999) and Durrer (2001).
1476:
1477: \newpage
1478:
1479: \section{Analytic formulae for the amplitudes and locations
1480: of acoustic peaks and dips in the CMB power spectrum}
1481:
1482:
1483: For completeness, we repeat here the formulas used in our parameter search
1484: which can also be found in the cited literature.
1485:
1486: We assume the standard recombination history and define the redshift of decoupling
1487: $z_{dec}$ as the redshift at which the optical depth of Thompson scattering is unity.
1488: A useful fitting formula for $z_{dec}$ is given by \cite{hs96}:
1489: \be
1490: z_{dec}=1048[1+0.00124\omega_b^{-0.738}][1+g_1\omega_m^{g_2}],
1491: \label{zdec}
1492: \ee
1493: where
1494: $$
1495: g_1=0.0783\omega_b^{-0.238}[1+39.5\omega_b^{0.763}]^{-1},\;\;\;
1496: g_2=0.56[1+21.1\omega_b^{1.81}]^{-1},
1497: $$
1498: $\omega_b\equiv\Omega_bh^2$ and $\omega_m\equiv\Omega_mh^2$.
1499:
1500:
1501: \subsection{Locations}
1502:
1503: The locations of the acoustic peaks in the CMB power spectrum depend on
1504: the value of sound horizon at decoupling
1505: epoch $r_s(\eta_{dec})\equiv \int_{0}^{\eta_{dec}}d\eta^{\prime}c_s$
1506: and the angular diameter distance to the last scattering surface,
1507: $d_A(z_{dec})$. Comparing with numerical calculations it was shown
1508: (see \cite{efs99,hu01,dl01} and references therein)
1509: that the spherical harmonic which corresponds to the $m$-th acoustic
1510: peak is well approximated by the relation
1511: \bea
1512: \ell_{p_m}= (m-\phi_m)\pi{d_A(z_{dec})\over r_s(z_{dec})},
1513: \label{lpm}
1514: \eea
1515: where $\phi_m$ take into account the shift of $m$-th peak
1516: from its location in the idealized model which is
1517: caused by driving effects from the decay of the gravitational potential.
1518: Doran and Lilley (2001) give an accurate analytic approximation in the form
1519: \be
1520: \phi_m={\bar\phi} - \delta\phi_m~,\ee
1521: where ${\bar\phi}$ is overall phase shift of the spectrum (or the first
1522: peak) and $\delta\phi_m$ is a relative shift
1523: of each peak and dip caused by the Doppler shift of the oscillating fluid.
1524: For the overall phase shift of the spectrum they find
1525: \be
1526: {\bar\phi}=(1.466-0.466n_s)a_1r_*^{a_2},
1527: \ee
1528: where
1529: $$r_*\equiv \rho_{rad}(z_{dec})/\rho_{m}(z_{dec})={0.0416\over \omega_m}
1530: \left({1+\rho_{\nu}/\rho_{\gamma}\over 1.6813}\right)
1531: \left({T_0\over 2.726}\right)^4\left({z_{dec}\over 1000}\right)$$
1532: is the ratio of radiation to matter at decoupling, and
1533: $$
1534: a_1=0.286+0.626\omega_b\;,\;\;a_2=0.1786-6.308\omega_b+174.9\omega_b^2-1168\omega_b^3
1535: $$
1536: are fitting coefficients. Here and below the numbers in the expressions are
1537: obtained for a present CMB temperature of $T_0=2.726$K and the ratio of
1538: densities of massless neutrinos and photons
1539: $\rho_{\nu}/\rho_{\gamma}=0.6813$ for three massless neutrino species
1540: (correspondingly $f_{\nu}\equiv \rho_{\nu}/(\rho_{\gamma}+\rho_{\nu})=0.405$).
1541: All values can be easily scaled to other values of $T_0$ and $f_{\nu}$.
1542:
1543: The relative shift of the 1st acoustic peak is zero, $\delta\phi_1=0$.
1544: For the 2nd one it is
1545: \be
1546: \delta\phi_2=c_0-c_1r_*-c_2/r_*^{c_3}+0.05(n_s-1)~,
1547: \ee
1548: with
1549: $$c_0=-0.1+0.213e^{-52\omega_b},\;\;\;c_1=0.015+0.063e^{-3500\omega_b^2},\;\;
1550: c_2=6\cdot 10^{-6}+0.137(\omega_b-0.07)^2,\;\;\;c_3=0.8+70\omega_b,$$
1551: and for the 3rd peak
1552: \be
1553: \delta\phi_3=10-d_1r_*^{d_2}+0.08(n_s-1)~,
1554: \ee
1555: with
1556: $$d_1=9.97+3.3\omega_b,\;\;\;d_2=0.0016+0.196\omega_b+2.25\cdot 10^{-5}\omega_b^{-1}.$$
1557:
1558: The formula (\ref{lpm}) is correct also for the location of dips if we set
1559: $m=3/2$ for the 1st dip and $m=5/2$ for the 2nd dip.
1560: The relative shift of the first dip given by \cite{dl01} is
1561: \be
1562: \delta\phi_{3/2}=b_0+b_1r_*^{1/3}\exp{b_2r_*}+0.158(n_s-1)~
1563: \ee
1564: with
1565: $$b_0=-0.086-2.22\omega_b-140\omega_b^2~,\;\;b_1=0.39-18.1\omega_b+440\omega_b^2,\;\;
1566: b_2=-0.57-3.8\exp({-2365\omega_b^2})~.$$
1567:
1568:
1569: The angular diameter distance to the last scattering surface is given by
1570: \be
1571: d_A(z_{dec})={c\over H_0\sqrt{|\Omega_k|}}\chi(\eta_0-\eta_{dec})~,
1572: \label{da}
1573: \ee
1574: where $\chi(x)=x,\;\;\sin{x}\;\;{\rm or}\;\;\sinh{x}$ for flat, closed or
1575: open models respectively, and
1576: \be
1577: \eta_0-\eta_{dec}=\sqrt{|\Omega_k|}\int_0^{z_{dec}}{dz\over \sqrt{\Omega_{rad}(z+1)^4+
1578: \Omega_m(z+1)^3+\Omega_{\Lambda}+\Omega_k(z+1)^2}}~.
1579: \ee
1580: Since, the sound speed in the pre-recombination plasma is
1581: \be
1582: c_s=c/\sqrt{3(1+R)}\;\;\;{\rm with}\;\;\;
1583: R\equiv 3\rho_b/4\rho_{\gamma}=30315(T_0/2.726)^{-4}\omega_ba
1584: \ee
1585: and scale factor is well approximated by
1586: \be
1587: a(\eta)=a_{eq}\left({\eta\over \eta_1}+({\eta\over 2\eta_1})^2\right),
1588: \ee
1589: with
1590: $$a_{eq}={4.16\cdot 10^{-5}\over \omega_m}\left({1+\rho_{\nu}/\rho_{\gamma}\over 1.6813}\right)
1591: \left({T_0\over 2.726}\right)^4,\;\;\;
1592: \eta_1\equiv {\eta_{eq}\over 2(\sqrt{2}-1)},$$
1593: the integral for sound horizon can be reduced to the analytic formula
1594: \be
1595: r_s(\eta_{dec})={19.9\over \sqrt{\omega_b\omega_m}}\left({T_0\over 2.726}\right)^2
1596: \ln{\sqrt{1+R_{dec}}+\sqrt{R_{dec}+R_{eq}}
1597: \over 1+\sqrt{R_{eq}}}\;\;\rm Mpc.
1598: \label{rs}
1599: \ee
1600:
1601:
1602: The deviation of the acoustic extrema locations calculated using formulae
1603: (\ref{lpm}-\ref{rs}) from the values obtained by CMBfast code is
1604: $< 3\%$ for a sufficiently wide range of parameters.
1605:
1606: \subsection{Amplitudes}
1607:
1608: The amplitude of the 1st acoustic peak can be approximated by the following
1609: expression
1610: \be
1611: A_{p_1}= {\ell_{p_1}(\ell_{p_1}+1)\over 2\pi} \left[C^{SW}_{\ell_{p_1}}
1612: +C^{SW}_2 {\tilde{\cal A}}(\Omega_b,\Omega_{cdm},\Omega_k,n_s,h)
1613: \right]~,
1614: \label{ap1}
1615: \ee
1616: where
1617: \be
1618: {\tilde {\cal A}}\equiv 0.838{\cal A}=
1619: \exp{[{\tilde a_1}+a_2\omega_{cdm}^2+
1620: a_3\omega_{cdm}+a_4\omega_b^2+a_5\omega_b+
1621: a_6\omega_b\omega_{cdm}+a_7\omega_k + a_8\omega^2_k+a_9(n_s-1)]}
1622: \label{A}
1623: \ee
1624: and $C^{SW}_{l_{p_1}}$ is given by (A9).
1625: We have re-determined the best-fit coefficients $a_{i}$ using the
1626: values of the 1st acoustic peak amplitudes from CMBfast for the grid of
1627: parameters given below. Their values are
1628: \bea
1629: \nonumber
1630: &{\tilde a_1}=2.326,\;\; a_2=8.906,\;\;a_3=-7.733,\;\;a_4=-115.6,\;\;a_5=35.66,&\\
1631: &a_6=-7.225,\;\;a_7=1.96,\;\;a_8=-11.16,\;\;a_9=4.439.&
1632: \label{ai}
1633: \eea
1634: The deviations of this approximation from the numerical value obtained
1635: by CMBfast are $\le 5\%$ within the range of cosmic parameters,
1636: $0.2\le\Omega_m\le 1.2$,
1637: $0\le\Omega_{\Lambda}\le 0.8$, $0.015\le\Omega_b\le 0.12$,
1638: $0.8\le n_s\le 1.2$ and $0.4\le h\le 1.0$.
1639:
1640: To calculate the amplitudes of the 2nd and 3rd peaks we use
1641: the relations $$H_2\equiv \left[\ell_{p_2}(\ell_{p_2}+1)C_{\ell_{p_2}}\right]/
1642: \left[\ell_{p_1}(\ell_{p_1}+1)C_{\ell_{p_1}}\right]\;\;\;{\rm and}\;\;\;
1643: H_3\equiv \left[\ell_{p_3}(\ell_{p_3}+1)C_{\ell_{p_3}}\right]/
1644: \left[\ell_{p_1}(\ell_{p_1}+1)C_{\ell_{p_1}}\right]$$
1645: given by Hu et al. (2001). This leads to the following amplitudes
1646: \bea
1647: A_{p_2}=A_{p_1}H_2(\Omega_m,\Omega_b,n_s),\;\;\;
1648: {\rm with}\;\;\; H_2={0.925\omega_m^{0.18}2.4^{n_s-1}\over \left
1649: [1+\left(\omega_b/0.0164\right)^{12\omega_m^{0.52}}\right]^{1/5}}~,\\
1650: A_{p_3}=A_{p_1}H_3(\Omega_m,\Omega_b,n_s),\;\;\;
1651: {\rm with}\;\;\;H_3={2.17\omega_m^{0.59}3.6^{n_s-1}\over
1652: \left[1+\left(\omega_b/0.044\right)^2\right]\left[1+1.63(1-
1653: \omega_b/0.071)\omega_m\right]}.
1654: \eea
1655:
1656: This approximation for $A_{p_3}$ deviates by less than 5\% from the value
1657: obtained with CMBfast for parameters within the range specified above. The
1658: accuracy of $A_{p_2}$ is better than 9\%. For some parameter values the
1659: second peak is under estimated leading to this somewhat poorer accuracy.
1660:
1661:
1662: \begin{thebibliography}{}
1663:
1664: \bibitem [Apunevych \& Novosyadlyj 2000]{an00} Apunevych, S. \& Novosyadlyj,
1665: B., 2000, Journal of Physical Studies, 4, 470
1666: \bibitem[Bahcall \& Fan 1998]{bah98} Bahcall, N.A. \& Fan, X., 1998, ApJ,
1667: 504, 1
1668: \bibitem[Balbi \etal~ 2000]{balbi} Balbi, A. \etal, 2000, ApJ, 545, L1
1669: \bibitem [Bardeen 1980]{bar80} Bardeen, M., 1980, Phys. Rev., D22, 1882
1670: \bibitem [Bennett \etal~1996]{ben96}
1671: Bennett, C.L. \etal, 1996, ApJ, 464, L1
1672: \bibitem [de Bernardis et al. 2000]{boom00} de Bernardis, P. \etal, 2000,
1673: Nature, 404, 495
1674: \bibitem [de Bernardis \etal~ 2001]{ber01} de Bernardis, P. \etal, 2001,
1675: astro-ph/0105296
1676: \bibitem[Burles \etal~ 2001]{burles}
1677: Burles, S. Nollet, K.M and Turner, M.S., 2001,
1678: Astrophys. J. Lett. {\bf 552}, L1--L6
1679: \bibitem [Bunn \& White 1997] {bw97} Bunn, E.F. \& White, M., 1997, ApJ, 480, 6
1680: \bibitem[Carroll \etal~ 1992]{car92}
1681: Carroll, S.M., Press, W.H. and Turner, E.L., 1992, ARA\&A, 30, 499
1682: \bibitem[Doran \& Lilley 2001] {dl01} Doran, M. \& Lilley, M., 2002, \mnras,
1683: 330, 965
1684: \bibitem[Douspis \& Ferreira 2001] {df01} Douspis, M. \& Ferreira, P.G.,
1685: 2001, astro-ph/0111400
1686: \bibitem[Durrer 1990]{d90} Durrer, R., 1990, Phys. Rev. D42, 2533
1687: \bibitem[Durrer 2001]{dur01} Durrer, R., 2001, astro-ph/0109522; 2001,
1688: Journal of Physical Studies, 5, No 2, 177
1689: \bibitem[Durrer \& Novosyadlyj 2001]{dn01} Durrer, R. \& Novosyadlyj, B.,
1690: 2001, \mnras, 324, 560
1691: \bibitem[Durrer \& Straumann 1999]{dur99} Durrer, R. \& Straumann, N., 1999,
1692: {\em New methods for the determination of cosmological parameters.}
1693: Troisi\`eme Cycle de la Physique en
1694: Suisse Romande, Universit\'e de Lausanne.
1695: \bibitem [Efstathiou \& Bond 1999]{efs99} Efstathiou, G. \& Bond, J.R., 1999,
1696: \mnras, 304, 75
1697: \bibitem [Eisenstein \& Hu 1998]{eh98} Eisenstein, D.J. \& Hu, W., 1998,
1698: ApJ, 496, 605
1699: \bibitem [Eisenstein \& Hu 1999]{eh99} Eisenstein, D.J. \& Hu, W., 1999,
1700: ApJ, 511, 5
1701: \bibitem [Gorski \etal~1996]{cobe96} Gorski, K. M. \etal, 1996, ApJ, 464, L11
1702: \bibitem [Halverson \etal~ 2001]{dasi} Halverson, N.W. \etal, 2001,
1703: astro-ph/0104489
1704: \bibitem [Hanany \etal~2000] {maxima00} Hanany, S. \etal, 2000, ApJ, 545, L5
1705: \bibitem [Hu \etal~2001]{hu01} Hu, W., Fukugita, M., Zaldarriaga, M. and
1706: Tegmark, M., 2001, ApJ, 549, 669
1707: \bibitem [Hu \& Sugiyama 1995] {hs95} Hu,W. \& Sugiyama, N., 1995, \apj,
1708: 444, 489
1709: \bibitem [Hu \& Sugiyama 1996] {hs96} Hu,W. \& Sugiyama, N., 1996, \apj,
1710: 471, 542
1711: \bibitem [Hu \& White 1996] {hw96} Hu,W. \& White, M., 1996, \apj,
1712: 471, 30
1713: \bibitem [Kofman \& Starobinsky 1985] {ks85} Kofman, L. \& Starobinsky, A.A.,
1714: 1985, Sov. Astron. Lett., 11(5), 271
1715: \bibitem [Knox \& Page 2000] {kp00} Knox, L. \& Page, L., 2000, Phys.Rev.
1716: Lett., 85, 1366
1717: \bibitem [Lahav \etal~1991] {la} Lahav, O., Lilje, P.B., Primack, J. and
1718: Rees, M.J., 1991, \mnras, 251, 128
1719: \bibitem [Lange \etal~ 2001]{lange} Lange, A.E. \etal, 2001, Phys. Rev. D63,
1720: 042001
1721: \bibitem [Lee \etal~ 2001] {maxima01} Lee, A.T. \etal, 2001, astro-ph/0104459
1722: \bibitem [Miller \& Nichol 2001] {mn01} Miller, C.J. \& Nichol, R.C., 2001,
1723: astro-ph/0112049
1724: \bibitem [Netterfield \etal~ 2001]{boom01} Netterfield, C.B. \etal, 2001,
1725: astro-ph/0104460
1726: \bibitem[Novosyadlyj, Durrer \& Lukash 1999]{ndl99}
1727: Novosyadlyj, B., Durrer, R. and Lukash, V.N., 1999, A\&A, 347, 799
1728: \bibitem[Novosyadlyj \etal~ 2000]{nd00}
1729: Novosyadlyj, B., Durrer, R., Gottloeber, S., Lukash, V.N. and Apunevych, S.,
1730: 2000, A\&A, 356, 418
1731: \bibitem[Perlmutter \etal~1999]{per99} Perlmutter, S. \etal, 1999, \apj
1732: 517, 565
1733: \bibitem[Pryke \etal~2001]{pry01} Pryke, C. \etal, 2001, astro-ph/0104490
1734: \bibitem [Sachs \& Wolfe 1967] {sw} Sachs, R.K. \& Wolfe, A.M., 1967,
1735: \apj, 147, 73
1736: \bibitem [SDSS Collaboration 2001]{sdss01} SDSS Collaboration: Szalay, A.S.
1737: \etal, 2001, astro-ph/0107419
1738: \bibitem[Seljak \& Zaldarriaga 1996]{cmbfast}Seljak, U. \& Zaldarriaga M.,
1739: 1996, ApJ 469, 437
1740: \bibitem[Smoot \etal 1992]{cobe92}Smoot, G.F \etal 1992, ApJ {\bf 396}, L1
1741: \bibitem[Tegmark \etal~ 2001]{teg01} Tegmark, M., Zaldarriaga, M. \&
1742: Hamilton, A.J.S., 2001, Phys. Rev. D63, 43007
1743: \bibitem[Wang \etal~2001]{wtz01} Wang, X., Tegmark, M. \& Zaldarriaga, M.,
1744: 2001, astro-ph/0105091
1745: \end{thebibliography}
1746:
1747:
1748: \end{document}
1749:
1750:
1751:
1752:
1753:
1754: