1: %\documentclass[preprint]{aastex}
2: \documentclass[useAMS,usenatbib]{mn2e}
3: %\documentclass[preprint,useAMS,usenatbib]{mn2e}
4: \bibliographystyle{mn2e}
5: \usepackage{epsfig}
6: \usepackage{amsmath}
7: %\usepackage{mn}
8: %\usepackage{mn-nat}
9: \newcommand{\be}{\begin{equation}}
10: \newcommand{\beq}{\begin{equation}}
11: \newcommand{\ba}{\begin{eqnarray}}
12: \newcommand{\ee}{\end{equation}}
13: \newcommand{\eeq}{\end{equation}}
14: \newcommand{\ea}{\end{eqnarray}}
15: \newcommand{\msun}{$M_{\odot}\hspace{1mm}$}
16: \newcommand{\wmap}{{\it WMAP }}
17: \newcommand{\kel}{$^{\circ} K\hspace{1mm}$}
18: \newcommand{\hs}{\hspace{1mm}}
19: \newcommand{\vs}{\vspace{2mm}}
20: \newcommand{\lya}{Ly$\alpha \hspace{1mm}$}
21: \newcommand{\ha}{H$\alpha \hspace{1mm}$}
22: \newcommand{\kms}{km s$^{-1}\hspace{1mm}$}
23: \newcommand{\cm}{cm$^{2}\hspace{1mm}$}
24: \newcommand{\GHz}{{\rm GHz}}
25: \newcommand{\Mpc}{{\rm Mpc}}
26: \newcommand{\kpc}{{\rm kpc}}
27: \newcommand{\dPsidL}{$\partial \log \Psi/\partial \log L_B$}
28: \newcommand{\apj}{ApJ}
29: \newcommand{\aap}{A\&A}
30: \newcommand{\apjl}{ApJL}
31: \newcommand{\mnras}{MNRAS}
32: \newcommand{\aj}{AJ}
33: \newcommand{\apjs}{ApJS}
34: \newcommand{\procspie}{Proc. SPIE}
35: \newcommand{\nat}{{\it Nature}}
36: \newcommand{\pasj}{PASJ}
37: \newcommand{\araa}{ARA\&A}
38: % definition to produce a "less than or similar to" symbol
39: \def\lsim{~\rlap{$<$}{\lower 1.0ex\hbox{$\sim$}}}
40: % definition to produce a "greater than or similar to" symbol
41: \def\gsim{~\rlap{$>$}{\lower 1.0ex\hbox{$\sim$}}}
42: %
43: \begin{document}
44:
45: \title[CMB Anisotropies From LBGs]{CMB Anisotropies from Outflows in Lyman Break Galaxies}
46:
47: \author[Daniel Babich and Abraham Loeb]
48: {Daniel Babich$^{1,2}$\thanks{E-mail:dbabich@cfa.harvard.edu} and Abraham
49: Loeb$^{1}$\thanks{E-mail:aloeb@cfa.harvard.edu}\\ $^1$
50: Astronomy Department, Harvard University, 60 Garden Street, Cambridge, MA
51: 02138, USA\\
52: $^2$ Current address: California Institute of Technology, Theoretical Astrophysics, MC 130-33
53: Pasadena, CA 91125, USA}
54:
55: %
56: \date{\today}
57: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2006}
58: %
59: \maketitle
60: %
61: \label{firstpage}
62: %
63: \begin{abstract}
64: Thomson scattering of the Cosmic Microwave Background (CMB) on moving
65: electrons in the outflows of Lyman Break Galaxies (LBGs) at redshifts 2--8
66: contributes to the small-scale CMB anisotropies. The net effect produced by
67: each outflow depends on its level of deviation from spherical symmetry,
68: caused either by an anisotropic energy injection from the nuclear starburst
69: or quasar activity, or by an inhomogeneous intergalactic environment. We
70: find that for plausible outflow parameters consistent with spectroscopic
71: observations of LBGs, the induced CMB anisotropies on arcminute scales
72: reach up to $\sim 1 \mu$K, comparable to the level produced during the
73: epoch of reionization.
74: \end{abstract}
75:
76: \begin{keywords}
77: cosmology -- cosmic microwave background -- galaxies:high-redshift
78: \end{keywords}
79:
80: \section{Introduction}
81: Several experiments to observe the Cosmic Microwave Background (CMB)
82: anisotropies on arcminute scales are currently, or will soon be, underway
83: \citep{Kosowsky03,Ruhl04, Lo05}. These experiments plan to measure the CMB
84: power-spectrum for a spherical harmonic multipole index of $10^3 \le \ell
85: \le 10^4$ at several frequencies centred around the thermal
86: Sunyaev-Zel'dovich null of 217 \GHz. Photon diffusion damps the CMB
87: anisotropies on these small scales during cosmological recombination at
88: redshift $z\sim 10^3$ \citep{Silk68}, and so any observed signal must
89: originate at much lower redshifts. Indeed, the above experiments plan to
90: constrain the epoch of reionization and the growth of structure in the low
91: redshift universe \citep{Zahn05}.
92:
93: The primary physical mechanism which is responsible for the small scale CMB
94: anisotropies is Thomson scattering of CMB photons off moving electrons.
95: Any peculiar velocity induces a Doppler anisotropy of the scattered
96: radiation along the direction of motion. This accounts for the CMB
97: anisotropies produced by the peculiar velocities of clusters (the so-called
98: {\it kinetic Sunyaev Zel'dovich effect}) \citep{Zeldovich80}, by peculiar
99: velocities of linear overdensities in the intergalactic medium (the
100: so-called {\it Ostriker-Vishniac effect}) \citep{Ostriker86, Vishniac87}
101: and by the peculiar velocities of the fluctuations in the ionization
102: fraction during patchy reionization \citep{Gruzinov98}.
103:
104: In this {\it Letter} we examine the contribution of outflows in Lyman Break
105: Galaxies (LBGs) to the small-scale CMB anisotropies. LBGs are believed to
106: be the ancestors of present-day luminous elliptical galaxies. They are
107: observed to produce gas outflows with velocities of several hundred \kms
108: (see \cite{Giavalisco02} for a comprehensive review). In contrast with the
109: traditional kinetic Sunyaev-Zel'dovich effect where the bulk velocity of
110: the virialized gas is responsible for the induced CMB anisotropy, we focus
111: here on the Doppler effect of the outflowing gas and ignore any bulk motion
112: of the LBG as a whole (which produced a smaller effect at the redshifts of
113: interest). This bulk effect is included in standard calculations of the
114: non-linear generalization of the Ostiker-Vishniac effect \citep{Hu00}.
115:
116: The contribution from a single LBG to the fractional temperature
117: fluctuation of the CMB can be expressed as \be \label{los_int} \frac{\Delta
118: T}{T} = - \int dl \sigma_T n_e\frac{\vec{n} \cdot \vec{v}}{c}, \ee where
119: $\sigma_T = 6.65 \times 10^{-25}$ \cm is the Thomson cross section, $n_e$
120: is the electron number denisty, $\vec{v}$ is the electron peculiar
121: velocity, $c$ is the speed of light and $\hat{n}$ is the observer's
122: line-of-sight toward the LBG \citep{Zeldovich80}. The integration traces
123: the photon's path through the LBG outflow.
124:
125: The radial extent of the outflow is found by solving the corresponding
126: hydrodynamics equations. These coupled non-linear partial differential
127: equations can be reduced to a single ordinary differential equation
128: \citep{Tegmark93, Furlanetto03} under the assumption that the gas swept-up
129: by the outgoing blast wave lies in a thin shell behind the propagating
130: shock front (the so-called {\it thin shell approximation}). The validity of
131: this approximation is illustrated by the self-similar Sedov-Taylor-von
132: Neumann solution for a point explosion in which $90\%$ of the swept-up mass
133: resides in a shell of thickness $10\%$ of the outflow's radius (see
134: Ostriker \& McKee 1988 and Ikeuchi et al. 1983 for additional discussion on
135: this approximation). The thin shell approximation allows us to treat the
136: radiative transfer of CMB photons through the shock front in a plane
137: parallel geometry. When the thickness of the shock front is small compared
138: with the shock front's radius of curvature, the path length through the
139: shock front can be expressed as \be \delta l \approx \frac{\delta R}{|
140: \hat{n}\cdot \hat{v} |}, \ee where $\delta R$ is the thickness of the shock
141: front. Using conservation of mass and the density compression ratio for a
142: strong adiabatic shock, one gets $\delta R/R = (\gamma -1)/(3\gamma + 3)$,
143: where $R$ is the radius of the outflow and $\gamma$ is the adiabatic index
144: of the gas \citep{Ostriker88}.
145:
146: Within the thin shell approximation the line-of-sight integration in
147: Eq. (\ref{los_int}) is simplified to \be \left(\frac{\Delta T}{T}
148: \right)_i = - \frac{\sigma_T~\delta R}{c}~[n_e(\vec{r}_1) v(\vec{r}_1) -
149: n_e(\vec{r}_2) v(\vec{r}_2)], \ee where $\vec{r}_1$ and $\vec{r}_2$ are the
150: location, in a coordinate system centred on the LBG, on the shock front
151: where the line-of-sight respectively enters and exits the shock front.
152: The subscript $i$ labels the contribution from a particular LBG.
153:
154: If the outflow is spherically symmetric then there is an exact cancellation
155: of the anisotropies produced at the entry and exit points of the
156: line-of-sight through the LBG shock front\footnote{Note that even if the
157: outflow had a perfect spherical shape, the finite light crossing time
158: through the outflow would produce a net non-zero signal because the flow
159: parameters are time-dependent. This effect would produce a signal
160: $\sigma_T n_e R (\Delta v)/c \sim \sigma_T n_e (dv/dt) (R/c)^2$ that is
161: extremely small and is ignored here.}. However, observations of
162: low-redshift starburst galaxies, which serve as analogs of the higher
163: redshift LBGs, show evidence for highly non-spherical outflow geometries
164: \citep{Martin99}. It is therefore reasonable to expect that the early
165: stages of LBG outflows, whether they are driven by starburst or quasar
166: activity\footnote{In this {\it Letter} we will only include the feedback
167: driven by supernovae. Energy input from quasars will lead to enhancements
168: in the bubble size and outflow velocity [see \cite{Furlanetto01} for a
169: description of quasar outflows].}, would produce CMB anisotropies.
170:
171: For simplicity, we will assume that the outflow is axisymmetric about some
172: axis $\hat{z}$ (possibly the rotation axis of the galactic disc or possibly
173: the jet axis of a central quasar) and expand both the outflow velocity and
174: shock front density in Legrendre polynomials, $ \vec{v}(\vec{r}) =
175: \hat{r}~\sum_{\ell} v_{\ell}(r) P_{\ell}(\hat{z} \cdot \hat{r})$ and
176: $n_e(\vec{r}) = \sum_{\ell} n_{\ell}(r) P_{\ell}(\hat{z} \cdot \hat{r})$,
177: where we have assumed that the velocity field is radial ($\hat{v} \equiv
178: \hat{r}$). Even if the outflow began highly collimated, it will eventually
179: isotropize as it propagates into the intergalactic medium (IGM; an
180: analogous tendency exists in relativistic flows, see Ayal \& Piran
181: 2001). On longer timescales, IGM inhomogeneities, such as filaments and
182: voids, will once again make the outflows non-spherical. Numerical
183: simulations are needed in order to properly model these effects.
184: Here we parameterize the level of asphericity in the outflow by a
185: coefficient $\epsilon$ (see Eq. \ref{eq:dT2} below).
186:
187: The outline of this {\it Letter} is as follows. In \S 2 we analyze the
188: CMB anisotropy induced by a single LBG with an arbitrarily non-spherical
189: outflow. In \S 3 we calculate the resulting CMB power spectrum. In \S 4
190: we present numerical results and in \S 5 we summarize our conclusions.
191: Throughout our discussion, we will assume the WMAP3 cosmological model
192: \citep{Spergel06}.
193:
194: \section{Single LBG Signal}
195: The signal along a given line-of-sight includes contributions from LBGs
196: with different masses, outflow ages, orientations of the outflow symmetry
197: axis and line-of-sight impact parameters. For simplicity, we begin by
198: analyzing the expectation value for LBGs formed at a certain redshift and
199: of certain mass and age. Since the net effect from a single LBG can be
200: either positive or negative we will find that the mean of the signal along
201: a given direction is always zero, but a non-zero variance will be produced
202: due to Poisson fluctuations in a manner equivalent to a random walk. The
203: analysis will be done separately for the distinct cases where either one or
204: two lines of sight intersect the same LBG outflow.
205:
206: \subsection{One Sightline}
207: First we consider the case where one line-of-sight intersects a single LBG
208: outflow. Averaging over symmetry axis orientation and impact parameter,
209: the expectation value of the fractional temperature perturbation is \be
210: \label{eq:sing_los} \left \langle \left( \frac{\Delta T}{T} \right)_i
211: \right \rangle = \int d^2\hat{z} P(\hat{z}) \int d^2 \vec{b}~P(b)
212: \left(\frac{\Delta T}{T} \right)_i. \ee Here $\hat{z}$ is the symmetry
213: axis of the outflow and $\vec{b}$ is the impact parameter at which the
214: line-of-sight enters the LBG (see Fig. \ref{fig:outflow} for definitions
215: of the variables we use in describing the LBG outflow).
216: \begin{figure}
217: \includegraphics[width=8cm]{diag_tmp.ps}
218: \caption[]{Schematic illustration of the geometry for a single LBG
219: outflow. The black circle represents the shock front, the orange arrow
220: labelled $\hat{n}$ is the line-of-sight from the observer, the red arrows
221: labelled $\vec{r}_1, \vec{r}_2$ are the entry and exit points of the
222: line-of-sight on the outflow, the blue arrow labelled $\hat{z}$ is the
223: symmetry axis of the outflow and the green arrow labelled $\vec{b}$ is the
224: impact parameter of the line-of-sight.}
225: \label{fig:outflow}
226: \end{figure}
227: We assume uniform probability distributions for the symmetry axis
228: orientation, $P(\hat{z}) = 1/4\pi$, and for the impact parameter,
229: $P(\vec{b}) = 1/\pi R^2.$
230:
231: Now we change the integration variable from the impact parameter $\vec{b}$
232: to the location of entrance point of the line-of-sight into the LBG
233: outflow $\hat{r}$. Thus, the expectation value defined in Eq. (\ref{eq:sing_los})
234: can be written as
235: \be \left \langle \left( \frac{\Delta T}{T} \right)_i \right \rangle =
236: \int_{4\pi} \frac{d^2\hat{z}}{4\pi} \int_{2\pi} \frac{d^2\hat{r}}{\pi}
237: \left(\frac{\Delta T}{T} \right)_i, \ee
238: where the integration over the
239: location of the impact parameter is restricted to a single hemisphere.
240: Performing this integration we find that the signal vanishes on average, as
241: expected from the fact that the net signal for a given LBG is just as
242: likely to be negative as it is to be positive.
243:
244: There will be a non-zero contribution from a given LBG since the induced
245: anisotropies will be different at the entry and exit points of the
246: line-of-sight through the shock front. Because the LBG symmetry axes is
247: randomly oriented with respect to the direction of the observer, the signal
248: vanishes once the average over this direction is done. This implies that
249: the mean signal and therefore the one-point function vanishes. We are
250: ultimately interested in the two-point correlation function and the related
251: power spectrum. There are two contributions to the power spectrum
252: \citep{Cooray02}. The first is a clustering (two halo) term, originating
253: from the correlated perturbations in the cold dark matter density produced
254: during inflation. The second is a Poisson (one halo) term, originating
255: from Poisson fluctuations in the number density of halos. The one-point
256: function vanishes because net temperature anisotropy produced by a given
257: LBG is uncorrelated with the signal from other LBGs along the
258: line-of-sight. The clustering term only implies that the number density of
259: LBGs nearby another LBG is greater than average, not that the symmetry axes
260: are somehow correlated\footnote{On small scales, the asphericity of the
261: outflows may be correlated because they propagate into the same
262: inhomgeneous IGM, or because tidal gravitational forces produced
263: correlations in the shapes of nearby galaxies \citep{Mackey02}.}. Since
264: there is no correlation in the signals between the two distinct
265: lines-of-sight, there will be no contribution to the resulting CMB power
266: spectrum from a clustering term.
267:
268: \subsection{Two Sightlines}
269: For a single sightline through each outflow we found that the two-point
270: correlation function vanishes, as the symmetry axes of LBGs are randomly
271: oriented. When both lines-of-sight intersect the same LBG this cancellation
272: does not take place. The average value of the two-point temperature
273: anisotropy when both lines of sight intersect the same LBG is \be \left
274: \langle \left( \frac{\Delta T}{T} \right)_i^2 \right \rangle = \int
275: d^2\hat{z} P(\hat{z}) \int d^2 \vec{b}_1~P(\vec{b}_1) \int
276: d^2\vec{b}_2~P(\vec{b}_2) \left(\frac{\Delta T}{T} \right)^2_i \ee
277: Performing the relevant integrations we find a non-zero answer when the
278: product $n_e(\vec{r})~v(\vec{r})$ has odd parity. The induced fluctuations
279: produced when two lines of sight intersect the same LBG are due to Poisson
280: fluctuations. As mentioned above, we parameterize the deviation from
281: sphericity with a fudge factor $\epsilon$. Then the expectation value for
282: two lines-of-sight intersecting the same LBG outflow region is \be
283: \label{eq:dT2} \left \langle \left( \frac{\Delta T}{T} \right)_i^2 \right
284: \rangle = \sigma^2_T n^2_e \delta R^2 \frac{\epsilon^2 v^2}{c^2}. \ee
285:
286: \section{Poisson Fluctuations}
287: We have seen that there is an exact cancellation when the two
288: lines-of-sight intersect two different LBG outflows and that a non-zero
289: signal arises when the two lines-of-sight intersect a single LBG
290: outflow. Correlations in the one LBG terms
291: are produced by Poisson fluctuations in the number of intercepted LBGs. We
292: will analyze this effect in two stages: first, we will consider the effect
293: along a given line-of-sight (pencil-beam survey) and then generalize to the
294: case of a finite beam size.
295:
296: \subsection{Pencil Beam Survey}
297: The temperature anisotropies induced by LBGs of halo mass between $M$ and
298: $M+dM$, formed between redshifts $z_f$ and $z_f + dz_f$ and scattering the
299: CMB between redshifts $z$ to $z+dz$ is \be \frac{\Delta T}{T} =
300: \sum_{n=1}^{\infty} P_{dN}(n_{\rm LBG}) \sum_{i=1}^{n_{\rm LBG}}
301: \left(\frac{\Delta T}{T}\right)_i, \ee where $P_N(n_{\rm LBG})$ is the
302: Poisson probability that $n_{\rm LBG}$ LBGs are observed and $dN$ is the
303: mean number of LBGs in the redshift interval between $z$ and $z+dz$, \be dN
304: = \pi R^2 \frac{-c~dz}{(1+z)H(z)} \frac{d^2n}{dM dz_f} dM dz_f, \ee where
305: $R$ is the radius of an outflow at redshift $z$ produced by an LBG of mass $M$
306: formed at redshift $z_f$. The total signal is found by integrating over
307: $dM$, $dz_f$ and $dz$. The expectation value toward a given line of sight
308: vanishes since the expectation value for the temperature anisotropy
309: produced by a single LBG vanishes. Nevertheless the variance of the signal
310: does not vanish due to Poisson fluctuations. For a single sightline through
311: an LBG outflow region the contribution to the temperature anisotropies can
312: be either positive or negative with equal probability, however for two
313: sightlines the contribution is always non-negative. The resultant
314: variance, which is the case of two sightlines at a separation less than the
315: characteristic angular size of LBG outflows, is \be \left\langle
316: \left(\frac{\Delta T}{T}\right)^2\right\rangle = \int dN \left\langle
317: \left(\frac{\Delta T}{T}\right)_i^2\right\rangle. \ee The {\it rms} value
318: of the anisotropy can be evaluated in terms of the unknowns and the time
319: changing Legendre coefficients $n_{\ell}$ and $v_{\ell}$. For simplicity,
320: we adopt the spherically symmetric solution for the shock radius and
321: velocity, and parameterize the degree of asymmetry in the outflow by the
322: fudge factor $\epsilon$.
323:
324: \subsection{Window Function Effects}
325: The average angular size of an LBG outflow $\bar{\theta}_{\rm LBG} \approx
326: 10\arcsec$ is below the resolution of the upcoming generation of
327: experiments, and so we must properly account for the beam's window
328: function. The observed temperature anisotropy will be an average over a
329: window function $W(\hat{n})$ \be \frac{\Delta \tilde{T}}{T}(\hat{n}) = \int
330: d^2 \hat{n}'~W(\hat{n}-\hat{n}')~\frac{\Delta T}{T}(\hat{n}'). \ee For
331: simplicity, we will take the window function shape to be a top hat of
332: angular size $\theta$, namely $W(\hat{n}) = 1/\theta^2 \mbox{ if }
333: |\hat{n}| \le \theta$, and $W(\hat{n})= 0 \mbox{ if } |\hat{n}| > \theta$.
334:
335: The variance in an angular aperature defined by the window function, \be \label{eq:cmb_var}
336: \left \langle \left( \frac{\Delta \tilde{T}}{T}\right)^2
337: \right\rangle_{\theta} = \int d^2\hat{n}'_1
338: d^2\hat{n}'_2~W(\hat{n}'_1)~W(\hat{n}'_2)~\left\langle\frac{\Delta
339: T}{T}(\hat{n}'_1)~\frac{\Delta T}{T}(\hat{n}'_2) \right\rangle, \ee is
340: related to the power spectrum as \be \frac{\ell(\ell+1)C_{\ell}}{2\pi}
341: \approx \left \langle\left( \frac{\Delta \tilde{T}}{T}\right)^2
342: \right\rangle_{\theta = 2\pi/\ell}. \ee
343:
344: In order for Poisson fluctuations to give a nonzero result, the two lines
345: of sight $\hat{n}'_1$ and $\hat{n}'_2$ must intersect the same
346: LBG. Therefore they must be separated by less than $\bar{\theta}_{\rm
347: LBG}$. This requirement allows us to evaluate Eq. (\ref{eq:cmb_var}) as \be
348: \left \langle \left( \frac{\Delta \tilde{T}}{T}\right)^2
349: \right\rangle_{\theta} = \frac1{\theta^2} \int dN \left( \frac{R}{D_{\rm
350: A}(z)} \right)^2 \left \langle \left( \frac{\Delta T}{T}\right)_i^2
351: \right\rangle. \ee Note that the Fourier multipole corresponding to
352: $\bar{\theta}_{\rm LBG}$ is $\ell = 2\pi/\bar{\theta}_{\rm LBG}$. Here
353: $D_{\rm A}(z)$ is the angular diameter distance to redshift $z$.
354:
355: A broad window function allows for a larger number of LBGs within the beam.
356: The window function is normalized such that it integrates to unity, and so
357: one is observing the fractional fluctuations in the signal as an average
358: over $\theta^2/\theta^2_{\rm LBG}$ independent coherence patches in the
359: beam. This is equivalent to a fractional fluctuation of $1/\sqrt{N}$ as
360: expected from Poisson fluctuations.
361:
362: \section{Results}
363: We numerically solve for the evolution of the shock front in the thin shell
364: approximation \citep{Tegmark93, Furlanetto03}, taking into account the
365: effects of the halo gravity and the self gravity of the mass shell, the
366: internal pressure of shocked IGM and the acceleration due to the
367: cosmological constant. When calculating the internal pressure we include
368: Compton cooling, adiabatic cooling, and the addition of shock heated
369: gas. Initially we assume that a fraction $f_{*} = 0.1$ of the baryons
370: assembled into the central LBG form stars with a Scalo initial mass
371: function. In this case there is one supernova per $126$ \msun of star
372: formation \citep{Furlanetto03}. Supernovae characteristically produce $10^{51}$
373: ergs of energy, but only a
374: small fraction of this energy, $f_{\rm SN}$, couples to the outflow, with
375: the rest being radiated away. We adopt a value of $f_{\rm SN}=0.01$; see
376: \cite{Furlanetto03} and reference within for a more detailed description
377: of our outflow model.
378:
379: Assuming a Sheth-Tormen mass function \citep{Sheth99} we include the
380: effects from all possible halos above the minimum galaxy mass. The lowest
381: galaxy mass is determined by the maximum between the Jeans filtering mass
382: and the cooling mass (dictated by the halo's ability to cool through atomic
383: hydrogen line emission)\footnote{Note that we include haloes with masses
384: below the observational sensitivity for LBGs, as there is no fundamental
385: reason to exclude these haloes.}. We allow the LBGs to form
386: between\footnote{We include this lower redshift limit because the
387: phenomenon of downsizing, as well as the evolution of AGN luminosity
388: function, implies that massive galaxies finished forming stars around that
389: redshift.} $z=2$ and $z=8$. We continue to allow the outflow to evolve
390: until its velocity equals to Hubble flow at its radius from the LBG or
391: until $z=0$. Increasing the upper redshift has little effect on our results
392: because of the higher minimum LBG mass, as well as, the higher velocity of
393: the Hubble flow which causes the LBG outflows to merge with the IGM at an
394: relatively earlier time. Decreasing the minimum LBG mass could have a
395: significant effect on our results because the ratio of the initial outflow
396: velocity to the halo escape velocity at the initial radius scales as
397: $v_{\rm init}/v_{esc} \propto M^{-2/9}$. However, in low mass haloes the
398: star formation timescale increases and supernova feedback becomes capable
399: of decreasing $f_{*}$. Some starburst activity continues to lower
400: redshifts, albeit with a reduced intensity compared to high redshift star
401: formation, changing the lower redshift of from $z_F = 2$ to $z_F = 1$ would
402: change our results by a factor of $2$. Since the observed star formation
403: efficiency decreases with decreasing $z_F$, our model, which assumes that
404: star formation and the subsequent supernova feedback only depends on the
405: LBG mass, overestimates this change.
406:
407: In Fig. \ref{fig:power_spec} we show the CMB power spectrum produced by
408: several secondary mechanisms. The three blue dot-dashed curves denote the LBG outflow
409: signal calculated in this {\it Letter} for the values of $\epsilon = 1,
410: 0.5, 0.25$ from top to bottom. The green dotted curve delineates the Ostriker-Vishniac
411: effect and the dashed cyan curve describes the patchy reionization effect as
412: calculated by \cite{Mcquinn06}\footnote{Note that \cite{Mcquinn06} used a
413: different cosmological model with a higher value of $\sigma_8 = 0.9$. This
414: artificially raises the amplitude of the results compared to value of $\sigma_8 = 0.73$
415: adopted in this work in line with WMAP3.}
416: Also, for comparison, we show the primary CMB anisotropies (solid, black curve) and
417: the small scale Cosmic Background Imager (CBI) data points (red triangles)
418: \citep{Readhead04}.
419:
420: \begin{figure}
421: \includegraphics[width=8cm]{cl_3yr.ps}
422: \caption[]{Power spectra produced by several secondary mechanisms. The
423: contribution from the LBG outflow is shown as the three blue dot-dashed curves for
424: $\epsilon = 1, 0.5, 0.25$ from top to bottom. The primary CMB anisotropies are shown in the
425: solid-black curve; the Ostriker-Vishniac and patchy reionization
426: contributions as calculated by \cite{Mcquinn06} are dotted-green and
427: dashed-light-blue respectively. The observed CBI data points are plotted as
428: red triangles.}
429: \label{fig:power_spec}
430: \end{figure}
431:
432: To characterize our results, let us mention some typical quantities for a
433: common LBG halo of mass $5 \times10^{9} M_{\odot}$ formed at $z_F = 2$. In
434: this case the outflow reaches a maximum comoving radius of 150 \kpc ~ at
435: $z= 0.9$ before it merges with the Hubble flow. At late times the outflow
436: velocity with respect to the local Hubble flow is $(v-HR)/c \sim 7 \times
437: 10^{-4}$ and the Thomson scattering optical depth through the shock front
438: is $\tau \approx 10^{-5}$.
439:
440: These small characteristic values for the Thomson scattering optical depth
441: and outflow velocity imply that the signal should not be notably
442: polarized. Thomson scattering of a radiation field containing a quadrupole
443: moment will produce polarized radiation. There are two standard ways in
444: which scattering by a halo can produce polarization -- (i) photons will
445: double scatter in the halo; (ii) the peculiar velocity of the scatterer
446: will induce a quadrupole moment in the radiation field
447: \citep{Zeldovich80}. In the first case, the radiation can scatter in the
448: halo producing an anisotropic radiation field; a second scattering of that
449: radiation field can produce polarization at the level $\mathcal{O}(\tau^2
450: v/c)$. In the second case, the peculiar velocity of the scatterer
451: perpendicular to its line-of-sight with respect to the observer will induce
452: a quadrupole in the incident radiation field at the order
453: $\mathcal{O}(v^2/c^2)$. A fraction $\tau$ of the radiation field will
454: scatter and become polarized at the level $\mathcal{O}(\tau v^2/c^2)$.
455: Since values of $\tau$ and $v/c$ are so small, we conclude that the
456: polarization power spectrum (which is sixth order in the small parameters
457: of $\tau$ and $v/c$) is negligible.
458:
459: \section{Discussion}
460:
461: The contribution of outflows from Lyman-break galaxies to the CMB
462: power-spectrum on arcminute scales is proportional to the square of their
463: characteristic level of deviation from sphericity, $\epsilon^2$. Future CMB
464: experiments could therefore calibrate the intricate feedback process of
465: galactic outflows on the IGM. Most tools used to study these feedback
466: processes focus on the inner few \kpc ~of the LBGs, even though the shock
467: front is typically located at several of tens or hundreds of \kpc. The
468: secondary CMB anistropies calculated in this work provide a unique probe of
469: these extended perturbed regions around starburst galaxies at high
470: redshifts.
471:
472: Even though the amplitude of the power spectrum produced by this effect is
473: small, the signal has distinctive spectral and spatial characteristics. The
474: power spectrum has the same scaling with Fourier multipole ($\propto
475: \ell^2$) as radio or IR point sources. However, the frequency dependence of
476: the anisotropies produced by our effect has the standard blackbody
477: spectrum, whereas the radio point sources have a power law frequency
478: spectrum. This distinct feature of our effect may allow future small scale
479: experiments, many of which have excellent frequency coverage, to separate
480: the anisotropies produced by LBGs from radio point sources.
481:
482: \bigskip
483:
484: {\bf Acknowledgements}
485:
486: This work was supported in part by Harvard university grants. D.B. thanks
487: the Harvard Institute for Theory and Computation for its hospitality when
488: this work was completed, and acknowledges helpful conversations with Re'em
489: Sari and Niayesh Afshordi. We also thank Matt McQuinn for making available
490: the numerical results from his work.
491:
492: \begin{thebibliography}{19}
493:
494: \bibitem[Ayal \& Piran(2001)]{Ayal01}
495: Ayal, S., \& Piran, T.\ 2001, \apj, 555, 23
496:
497: \bibitem[Cooray \& Sheth(2002)]{Cooray02}
498: Cooray, A., \& Sheth, R.\ 2002, Phys. Repts., 372, 1
499:
500: \bibitem[Furlanetto \& Loeb(2001)]{Furlanetto01}
501: Furlanetto, S.~R., \& Loeb, A.\ 2001, \apj, 556, 619
502:
503: \bibitem[Furlanetto \& Loeb(2003)]{Furlanetto03}
504: Furlanetto, S.~R., \& Loeb, A.\ 2003, \apj, 588, 18
505:
506: \bibitem[Giavalisco(2002)]{Giavalisco02}
507: Giavalisco, M.\ 2002, \araa, 40, 579
508:
509: \bibitem[Gruzinov \& Hu(1998)]{Gruzinov98}
510: Gruzinov, A., \& Hu, W.\ 1998, \apj, 508, 435
511:
512: \bibitem[Hu(2000)]{Hu00}
513: Hu, W.\ 2000, \apj, 529, 12
514:
515: \bibitem[Ikeuchi et al.(1983)]{Ikeuchi83}
516: Ikeuchi, S., Tomisaka, K., \& Ostriker, J.~P.\ 1983, \apj, 265, 583
517:
518: \bibitem[Kosowsky(2003)]{Kosowsky03} Kosowsky, A.\ 2003, New
519: Astronomy Review, 47, 939
520:
521: \bibitem[Lo et al.(2005)]{Lo05}
522: Lo, K.~Y., et al.\ 2005, IAU Symposium, 201, 306
523:
524: \bibitem[Mackey et al.(2002)]{Mackey02}
525: Mackey, J., White, M., \& Kamionkowski, M.\ 2002, \mnras, 332, 788
526:
527: \bibitem[Martin(1999)]{Martin99} Martin, C.~L.\ 1999, \apj, 513, 156
528:
529: \bibitem[McQuinn et al.(2006)]{Mcquinn06}
530: McQuinn, M., et al.\ 2006, New Astronomy Review, 50, 84
531:
532: \bibitem[Ostriker \& Vishniac(1986)]{Ostriker86}
533: Ostriker, J.~P., \& Vishniac, E.~T.\ 1986, \apjl, 306, L51
534:
535: \bibitem[Ostriker \& McKee(1988)]{Ostriker88}
536: Ostriker, J.~P., \& McKee, C.~F.\ 1988, Reviews of Modern Physics, 60, 1
537:
538: \bibitem[Readhead et al.(2004)]{Readhead04}
539: Readhead, A.~C.~S., et al. \ 2004, \apj, 609, 498
540:
541: \bibitem[Ruhl et al.(2004)]{Ruhl04} Ruhl, J., et al.\ 2004,
542: \procspie, 5498, 11
543:
544: \bibitem[Sheth \& Tormen(1999)]{Sheth99}
545: Sheth, R.~K., \& Tormen, G.\ 1999, \mnras, 308, 119
546:
547: \bibitem[Silk(1968)]{Silk68} Silk, J.\ 1968, \apj, 151, 459
548:
549: \bibitem[Spergel et al.(2006)]{Spergel06}
550: Spergel, D.N. et al.\ 2006, submitted to \apj, astro-ph 0603449
551:
552: \bibitem[Tegmark et al.(1993)]{Tegmark93}
553: Tegmark, M., Silk, J., \& Evrard, A.\ 1993, \apj, 417, 54
554:
555: \bibitem[Vishniac(1987)]{Vishniac87}
556: Vishniac, E.~T.\ 1987, \apj, 322, 597
557:
558: \bibitem[Zahn et al.(2005)]{Zahn05}
559: Zahn, O., et al. \ 2005, \apj, 630, 657
560:
561: \bibitem[Zel'dovich \& Sunyaev(1980)]{Zeldovich80}
562: Zel'dovich, Y.~B., \& Sunyaev, R.~A.\ 1980, Pis ma Astronomicheskii Zhurnal, 6, 545
563:
564: \end{thebibliography}
565:
566: \end{document}
567:
568:
569:
570:
571:
572:
573:
574:
575:
576:
577:
578:
579:
580: