astro-ph0005244/ms.tex
1: %\documentclass[]{article}
2: %\documentstyle[12pt,aasms4]{article}
3: \documentstyle[11pt,aaspp4]{article}
4: %\usepackage{emulateapj}
5: \newcommand\beq{\begin{equation}}
6: \newcommand\eeq{\end{equation}}
7: 
8: 
9: \begin{document}
10: 
11: \title{SYNCHROTRON EMISSION FROM HOT ACCRETION FLOWS AND 
12: THE COSMIC MICROWAVE BACKGROUND ANISOTROPY}
13: 
14:  \author{Rosalba Perna\altaffilmark{1} and Tiziana Di Matteo\altaffilmark{2}}
15:  \affil{Harvard-Smithsonian Center for Astrophysics,
16: 60 Garden St., Cambridge, MA 02138;\\
17: rperna, tdimatteo@cfa.harvard.edu}
18: 
19: \altaffiltext{1}{Harvard Junior Fellow}
20: \altaffiltext{2}{{\em Chandra\/} Fellow}
21: 
22: 
23: \begin{abstract} 
24: 
25: Current estimates of number counts of radio sources in the frequency
26: range where the most sensitive Cosmic Microwave Background (CMB)
27: experiments are carried out significantly under-represent sources with
28: strongly inverted spectra.  Hot accretion flows around supermassive
29: black holes in the nuclei of nearby galaxies are expected to produce
30: inverted radio spectra by thermal synchrotron emission.  We calculate
31: the temperature fluctuations and power spectra of these sources in the
32: Planck Surveyor 30 GHz energy channel, where their emission is
33: expected to peak.  We find that their potential contribution is
34: generally comparable to the instrumental noise, and approaches the CMB
35: anisotropy level at small angular scales.  Forthcoming CMB missions, which
36: will provide a large statistical sample of inverted-spectra sources,
37: will be crucial for determining the distribution of hot accretion flows
38: in nearby quiescent galactic nuclei. Detection of these sources in
39: different frequency channels will help constrain their spectral
40: characteristics, hence their physical properties.
41: 
42: \end{abstract}
43: 
44: \keywords{accretion, accretion disks --- black hole
45: physics --- cosmic microwave background: anisotropies}
46: 
47: \section{INTRODUCTION}
48: 
49: The upcoming cosmic microwave background (CMB) experiments, e.g.~MAP
50: and the Planck Surveyor, will be able determine the primordial
51: anisotropies to an unprecedented level of accuracy. Because of its high
52: sensitivity, excellent angular resolution and wide range of
53: frequencies, Planck in particular, will be extremely sensitive to
54: extragalactic foreground point sources, which provide the major source
55: of uncertainty in the measurement of the intrinsic fluctuations.
56: 
57: Several studies have therefore been carried out to calculate the
58: contribution of point sources to the CMB anisotropies.  Much of this
59: work (see Toffolatti et al. 1999a,b; De Zotti et al. 1999; Gawiser \&
60: Smoot 1997; Sokasian, Gawiser \& Smoot 1998) has dealt with the
61: contribution from radio sources, the number counts of which are
62: determined down to $\mu$Jy but only up to frequencies $\la$ 8 GHz.
63: These counts are usually extrapolated to the higher frequencies
64: relevant for the CMB experiments.  This implies that the available
65: counts are sensitive enough to include the most significant
66: contribution from the ``steep'' and ``flat'' spectrum sources (with
67: $F_\nu\propto \nu^{-\alpha}$, and $\alpha\ge 0$, such as compact radio
68: galaxies and radio loud quasars), but are missing, or are strongly
69: under-representing, an important contribution from a class of sources
70: with inverted spectra ($\alpha\la 0$; e.g. De Zotti et al. 1999). This
71: is further emphasized by recent observations at 28.5 GHz, which find
72: up to a factor of 7 more sources than predicted from low-frequency
73: surveys (Cooray et al.~1998). Inverted-spectrum sources, such as those
74: discussed here, may peak in the frequency range of a few tens to a few
75: hundreds GHz, and could therefore provide a considerable contribution
76: in the region where the most sensitive CMB experiments are carried
77: out.
78: 
79: GHz Peaked Spectrum (GPS) sources (see O'Dea et al.~1998, Guerra,
80: Haarsma \& Partridge 1998) have been recognized to be an important
81: class of inverted-spectrum sources.  Their emission is attributed to
82: synchrotron radiation from compact and high density regions often
83: associated with the early stages of the formation of more classical
84: double radio sources (the so called ``young source'' scenario; Philips
85: \& Mutel 1982). However, as pointed out by Toffolatti et al. (1999),
86: there may be another, distinct, class of strongly inverted spectra due
87: to thermal synchrotron emission in hot or advection dominated
88: accretion flows (ADAFs). Unlike the relatively rare and bright GPS
89: sources (peak fluxes of $\sim 1-10$ Jy), usually associated with
90: bright active galaxies or quasars at high redshifts, ADAF sources
91: should be common in nearby galaxies and provide the most significant
92: contribution to the emission in the high radio frequencies of the
93: faint ($\sim$ a few mJy) radio cores observed in such galaxies. 
94: 
95: The reason why we consider hot accretion flows to be common in nearby
96: galactic nuclei is that, in recent years, it has become apparent
97: (e.g. Fabian \& Rees 1995; Narayan \& Yi 1995; Di Matteo et al. 2000
98: and references therein) that the nuclei of such galaxies, which host
99: the largest black holes known with masses of $10^8-10^{10} M_{\odot}$
100: (e.g., Magorrian et al. 1998), are remarkably underluminous for the
101: typically expected accretion rates (determined from measuraments of
102: densities and sound speeds of their hot interstellar medium). In
103: particular, it has been shown (e.g. Di Matteo et al. 2000) that the
104: relative quiescence and spectral characteristics of the early-type
105: galactic nuclei can be well-explained if the central black holes
106: accrete via low radiative-efficiency accretion flows or ADAFs (Rees et
107: al. 1982; for a review see, e.g., Narayan, Mahadevan \& Quataert
108: 1998).  Moreover, it has been proposed (Di Matteo and Allen 1999) that
109: such flows, which also produce significant emission in the $X$-ray
110: band, could provide a significant contribution to the cosmic $X$-ray
111: background (XRB). Within the context of these models, a significant
112: fraction of the hard number counts in the X-ray energies should arise
113: from sources at low redshift ($z \la 1$).  This picture is supported
114: by recent deep {\em Chandra} observations, which have resolved about
115: 40 per cent of the hard XRB in point sources in bright early-type
116: galaxies (Mushotzky et al. 2000).
117: 
118: The potential contribution of GPS sources to fluctuations in the CMB
119: anisotropy has been discussed by De Zotti et al. (1999).  In this {\em
120: Paper}, we examine the specific contribution of inverted spectra ADAF
121: sources in the nuclei of early-type galaxies to the CMB anisotropy. We
122: evaluate their foreground contribution to the small-scale cosmic
123: microwave fluctuations in the low-energy channels foreseen for the
124: Plank surveyor mission.  These sources, if indeed common in elliptical
125: galaxies, should be much more numerous albeit fainter than the GPS
126: population, and may therefore provide a stronger noise contribution at
127: the small angular scales.
128: 
129: While it is important to assess the potential contribution of
130: advection-dominated sources to the CMB fluctuations, the forthcoming
131: CMB experiments themselves will, for the first time, provide a large
132: statistical sample of objects with inverted radio spectra.  Because
133: most of the ADAF emission occurs in the high radio and in the $X$-ray
134: band, Planck observations will possibly provide the most powerful test
135: for the presence of ADAFs around supermassive black holes.  In
136: particular, such studies will provide strong constraints on the
137: spectral properties of this class of objects, and will help determine
138: how common they are in the nearby Universe. Confirming the presence of
139: these sources would also support the conjecture that they provide a
140: significant contribution to the hard XRB.
141: 
142: \section{SYNCHROTRON EMISSION FROM ACCRETION FLOWS IN EARLY-TYPE GALAXIES}
143: 
144: Radio continuum surveys (at $\nu \la $ 8 GHz) of elliptical and S0
145: galaxies have shown that the sources in radio--quiet galaxies tend to
146: be extended but with a compact component with relatively flat or
147: slowly rising radio spectra (with typical spectral indexes of
148: 0.3--0.4). Recent VLA studies at high radio frequencies (up to 43 GHz),
149: although carried out only on a limited sample of objects, have shown that
150: all of the observed compact cores have spectra rising up to 
151: $\sim 20-30$ GHz (e.g., Di Matteo et al.~1999).
152: 
153: Although the low-frequency radio emission in these galaxies might
154: still have a significant contribution from the scaled-down radio jets
155: also present in these systems, it has been proposed that the
156: high-frequency emission can be easily accounted for if the supermassive black
157: holes in elliptical galaxies are accreting via ADAFs (Fabian \& Rees
158: 1995; Mahadevan 1997; Di Matteo et al.~1999). In an ADAF around a
159: supermassive black hole, the majority of the observable emission is in
160: the high radio and X--ray bands.  In the high-frequency radio band, the
161: emission results from synchrotron radiation due to the strong magnetic
162: field in the inner parts of the accretion flow. The X-ray emission is
163: due either to bremsstrahlung or inverse Compton scattering.  In the
164: thermal plasma of an ADAF, synchrotron emission rises steeply with
165: decreasing frequency. Under most circumstances the emission becomes
166: self--absorbed and gives rise to a black--body spectrum (in the
167: Rayleigh-Jeans limit) below a turnover frequency $\nu_{\rm c}$. Above
168: this frequency it decays exponentially as expected from a thermal
169: plasma, due to the superposition of cyclotron harmonics.
170: 
171: The spectral models and self-consistent temperature profile
172: calculations for the ADAF flows in the elliptical galaxy cores are
173: described in detail by Di Matteo et al. (1999; 2000b). Recent
174: studies have also shown that large outflows may be important in such
175: low--radiative efficiency accretion flows (Igumenshchev \& Abramowicz
176: 1999; Blandford \& Begelman 1999; Stone, Pringle \& Begelman 1999). This
177: should lead to a suppression of the synchrotron component with
178: respect to the standard ADAF model with no outflow. Flatter density
179: profiles, as expected from strong mass loss, are usually required to explain the 
180: high-resolution VLA observations at high-radio frequencies of a number of
181: ellipticals. As in our previous work, therefore, we model the accretion
182: flows by adopting a density profile which satisfies $\rho \propto
183: R^{-3/2 +p}$, where $R$ is the radius of the flow and $0 \le p
184: \le 1$. Figure 1 shows the synchrotron emission expected from an ADAF
185: around a supermassive black hole with $M_{BH} \sim 10^9 M_\odot$, for
186: different values of $p$ (effectively for radially dependent mass
187: accretion rates, $\dot{M} \propto R^p$). The uppermost curve is the
188: standard ADAF model calculated with the accretion rates in the flows
189: determined from Bondi accretion theory, $\dot{M} \sim \dot{M}_{\rm
190: Bondi} \propto M_{\rm BH}^2 n$(ISM)$/c_{\rm S}$(ISM); where the ISM
191: density, $n$(ISM), and sound speed, $c_{\rm S}$(ISM), can be
192: determined from X-ray deprojection analysis, and $M_{\rm BH}$ is given
193: in recent studies by Magorrian et al. (1998).
194: 
195: So far, only a limited sample of sources has been observed at
196: high--resolution radio frequencies (up to 43 GHz), such that both the
197: flux and the position of the synchrotron peak could be determined. The
198: shaded region in Figure 1 identifies the range of models that have
199: been shown to best fit the observed fluxes. However, as the sample is
200: not statistically significant, we will also consider various
201: populations of accretion sources which may have different luminosities
202: due to their different density profiles.  Note also, from
203: Figure 1, that because of their sharp spectral cut-offs, these sources
204: are expected to contribute only to the Planck low-frequency channels
205: ($\nu \la 100$ GHz), and in particular to the 30 GHz one.
206: 
207: It is worth pointing out that the synchrotron emission, and the
208: position of its peak in an ADAF, is a strong function of many model
209: variables: the emission at the self-absorbed synchrotron peak arises
210: from the inner regions of an accretion flow and scales as $L_{\nu}
211: \propto \nu_{c}^2 T$, where $\nu_{c} \propto T^{2} B \propto T^2
212: \dot{M}^{1/2} M_{\rm BH}^{1/4} R^{-5/4}$ (Narayan \& Yi 1995); $T$ is the electron
213: temperature, and $B$ the magnetic field strength.  Because of the
214: strong dependences on a number of parameters, we will estimate the
215: contribution to the CMB anisotropy by making the most conservative
216: assumptions for the model source parameters.  We stress that our
217: analysis is not intended to explore the full range of parameter space
218: available for the accretion flow models. At present, the theoretical
219: and observational uncertainties involved in such calculations are too
220: large to merit such work. However, although schematic, these models
221: may provide a useful guide for the prediction of the confusion noise
222: due to these sources.
223: 
224: \section{CONTRIBUTION TO CMB FLUCTUATIONS}
225: 
226: The contribution to CMB fluctuations from randomly distributed sources
227: has been extensively discussed in the literature (Scheuer 1957, 1974;
228: Condon 1974; Cavaliere \& Setti 1976; Franceschini et al. 1989;
229: Tegmark \& Efstathiou 1996).
230: 
231: We let $x=Sf(\theta)$  be the telescope response to a source of flux $S$
232: located at a distance $\theta$ from the beam axis, and let
233: $R(x)$ be the mean number of source responses of intensity $x$. 
234: The fluctuation level generated by randomly distributed sources with
235: a Poisson distribution is given by the second moment $\sigma$ of the
236: $R(x)$ distribution. If the angular power pattern of the detector,
237: $f(\theta)$, is taken to be a Gaussian with full width half maximum 
238: $\theta_0$, one obtains:
239: \beq
240: \sigma^2=\int_0^{x_c}x^2 R(x) dx=\pi\theta_0^2 I(x_c)\;
241: \label{eq:sigma}
242: \eeq 
243: where
244: \beq
245: I(x_c)=\int_0^{x_c}dx x^2\int_0^{\infty}d\psi N\left(\frac{x}{f(\psi)} \right)
246: \exp(4\psi\, {\rm ln}2)\;. 
247: \label{eq:Ic}
248: \eeq Here $\psi\equiv(\theta/\theta_0)$, $N(S_\nu)$ are the differential
249: source counts per steradian at a given frequency $\nu$, and $x_c=q\sigma$
250: is the threshold flux above which sources are considered to be
251: individually detected.  In our calculations we adopt the standard value of $q=5$.
252: 
253: The rms brightness temperature fluctuation $(\Delta T/T)_{\rm rms}\equiv
254: \left< (\Delta T/T)^2\right>^{1/2}$ at the  frequency $\nu$ is related to 
255: the confusion standard deviation $\sigma$ by
256: \beq
257: \left(\frac{\Delta T}{T}\right)_{\rm rms}=\frac{\sigma}{T\omega_b}
258: \left(\frac{\partial B_\nu}{\partial T}\right)^{-1}\;,
259: \label{eq:delT}
260: \eeq
261: where $\omega_b=\pi\int_0^\infty d\theta^2 f(\theta)=\pi(\theta_0/2)^2\log 2$
262: is the effective beam area, and 
263: \beq
264: \frac{\partial B_\nu}{\partial T} = \frac{2k}{c^2}\left(\frac{k T}{h}\right)^2
265: \frac{x^4 e^x}{(e^x-1)^2} 
266: %=\left(\frac{99270 mJy sr^{-1}}{10^{-6} K}\right)\frac{x^4 e^x}{(e^x-1)^2}
267: \label{eq:conv}
268: \eeq 
269: is the conversion factor from temperature to flux (per steradian).  
270: Here $B_\nu (T)$ is the Planck function, $x\equiv h\nu/kT$,  
271: and $T=2.725$ K (Mather et al. 1999) is the CMB temperature. 
272: 
273: We take the number density of our sources to be that of ellipticals, and 
274: use the fit\footnote{The fit was derived from   
275: observations of over 1700 galaxies in various magnitude
276: limited samples (the Autofib Redshift Survey).} 
277: provided by Heyl et al. (1997) for galaxies with $L\approx L_*$,
278: \beq
279: n(z) = n_0 (1+z)^{[\gamma_\phi-\gamma_L(\alpha_0+\gamma_\alpha z)]}
280: \exp[-1/(1+z)^{\gamma_L}]\;,
281: \label{eq:nz}
282: \eeq where the set of parameters \{$n_0,\gamma_\phi,\gamma_L,\alpha_0,
283: \gamma_\alpha $\} is given by Heyl et al. for red and blue
284: ellipticals. These are \{$1.62\times 10^{-3}$ Mpc$^{-3}$, -6.15,
285: -1.77, -1.05, 1.3\} and {$1.87\times 10^{-3}$ Mpc$^{-3}$, 1.56, -0.35,
286: -1.06, 1.23\} for red and blue respectively. We consider the total
287: $n(z)$ and add both components. We take a cutoff at $z\sim 1$ for
288: their contribution. This is consistent with the normalization of the
289: ADAF sources in ellipticals required to explain the hard XRB (Di
290: Matteo \& Allen 2000). Notice, however, that our results are rather
291: insensitive to the precise value of this cut-off, as most of the
292: contribution is produced by sources at very low
293: redshift.
294: 
295: Given a population of low-redshift sources with luminosity per unit
296: frequency $L_\nu$, the differential number counts are given by
297: $N(S_\nu)\approx [4\pi(c/H_0)n(z)d_L^2(z)dz/dS_\nu]_{z=z(S_\nu)}$,
298: where $z(S_\nu)$ is derived by inverting the relation $S_\nu=L_{\nu}
299: (1 + z)/4\pi d_L^2(z)$, and $d_L$ is the luminosity distance. We adopt
300: a flat cosmology with $\Omega_0=1$, and take $H_0=3.2\times 10^{-17}
301: {\rm s}^{-1} h$, with $h=0.65$. The particular choice of cosmology is
302: not relevant to our calculations, again because the bulk of the
303: contribution from these relatively faint objects comes from nearby
304: sources.
305: 
306: For a Poisson distribution of point sources, it is 
307: straightforward to compute their angular power spectrum $C_l$.
308: If only the contribution  from sources with $S_\nu\le S_{\rm lim}$
309: is included, this is given by (Tegmark \& Efstathiou 1996; Scott \& White 1999)
310: \beq
311: C_l(\nu)= \frac{1}{T^2}\left(\frac{\partial B_\nu}{\partial T}\right)^{-2}
312: \int_0^{\rm S_{\rm lim}} S_\nu^2\; N(S_\nu) dS_\nu\;. 
313: \label{eq:cl}
314: \eeq
315: ~~~
316: ~~~
317: 
318: 
319: \section{RESULTS}
320: 
321: Figure 2 shows the expected temperature fluctuations (Eq.~3) due to
322: synchrotron emission from accretion flows in the nuclei of ellipticals
323: as a function of the angular scale $\theta_0$, at $\nu = 30$ GHz. The
324: solid line corresponds to the case where the luminosities of all
325: sources fall in the region marked by the shaded region of Figure
326: 1. These are ADAF models with a significant amount of outflow and
327: relatively low luminosities. The other lines show the contribution to
328: the fluctuations due to a mixed population of sources containing a
329: fraction $f$ of accretion flows with higher synchrotron
330: luminosities\footnote{ We cannot attempt to model a proper luminosity
331: function for these sources, as the observed sample in Di Matteo et
332: al. (2000b) is too small to allow such modelling.}  (e.g. as expected
333: from ADAFs with no outflows corresponding to the uppermost curve of
334: Figure 1). Note that even a small fraction of high-luminosity sources
335: can have large effects on the level of fluctuations. This is due to
336: the fact that the source number counts roughly scale as $N(S)\propto f
337: L^{1.5}$, and therefore the dependence on the luminosity is stronger
338: than the dependence on $f$. Also note that our derived number counts are 
339: comparable (if $f=0$) to the number counts extrapolated from low-frequency surveys
340: by Toffolatti et al. (1999a), but can be up to a factor of 10 higher 
341: if a considerable fraction of higher-luminosity ($f\ga 0.5$)
342: sources is considered.
343: 
344: Figure 3 shows, in the case of a fraction $f=0.5$ of high-luminosity
345: sources in the sample, the expected level of fluctuations if all
346: sources above a flux $S_{\rm lim}$ were individually identified and
347: substracted out from the sample. This is shown for different choices
348: of $S_{\rm lim}$. Identification and removal of individual sources can
349: in principle be done by means of independent surveys.
350: 
351: Using the same parameters as in Figure 3, Figure 4 shows a comparison
352: between the poissonian power spectrum (Eq.~6) produced by the
353: accretion flows in ellipticals and the predicted power spectrum for a
354: standard CDM model. The heavy solid line shows the contribution to the
355: power spectrum from noise in the 30 GHz channel of the Planck LFI.
356: Note that the ``flat'' ($C_l$ = constant) angular power spectrum of
357: the fluctuations due to ADAF sources differs substantially from the
358: power spectrum of primordial fluctuations. We find that the source signal 
359: is generally well below that of the intrinsic fluctuations,
360: and it only becomes comparable to these on the small angular
361: scales, where also the instrumental noise increases to roughly the
362: same level.  Removal of source signal should be possible even in cases
363: where it gives a strong contribution. Even if the Poisson component of
364: the sources and the noise due to the instrument have similar power
365: spectra, they are indeed different in their nature (as emphasized by Scott \&
366: White 1999). In fact, while the sources on the sky contribute to the
367: flux in every observation of a given pixel, the noise, on the other
368: hand, differs from observation to observation, and, by assumption, it
369: is uncorrelated with the signal in that pixel. Therefore, if a given
370: direction in the sky is observed multiple times (as expected for
371: Planck), the instrumental noise component can be separated from the
372: sky signal.
373: 
374: \section{DISCUSSION}
375: 
376: We have computed the temperature fluctuations and power spectrum
377: produced by inverted radio spectra from hot accretion flows in the
378: nuclei of nearby elliptical galaxies in the Planck 30 GHz channel,
379: where their emission is expected to peak. We have shown that the
380: contribution from this class of sources approaches the intrinsic CMB
381: fluctuation level only at small angular scales.  However, because of
382: the different nature of its power spectrum, the source contribution
383: should not affect the most important goal of the Planck mission, that
384: is the accurate measurement of the primary CMB anisotropy.
385: 
386: On the other hand, Planck will provide a large statistical sample of
387: sources characterized by inverted spectra. Therefore, it should be
388: possible to use this study to determine how common this mode of
389: accretion is in the nearby supermassive black holes.  In particular,
390: as most of the contribution from this population is expected to peak
391: at high radio frequencies, Planck should allow us to study their
392: spectral characteristics.  In turn, because different spectral
393: distributions and luminosities reflect the shape of the density
394: profiles, CMB experiments could allow us to gain important information
395: on the physical conditions in these accretion flows.  As already noted
396: by Toffolatti et al. (1999b), the implications of such a study could,
397: more generally, be significant as a way of testing the physical
398: processes in the medium surrounding massive black holes, and the
399: evolution of the interstellar medium in galaxies up to moderate
400: redshifts. Even more, it would provide a test for current ideas
401: according to which a significant fraction of the $X$-ray background
402: may due to accretion in this regime in early-type galaxies in the
403: local universe (Di Matteo \& Allen 1999). Note that such a significant
404: statistical study would be more difficult to carry out with surveys at
405: other wavelengths because of the rapid decline of the ADAF flux, which
406: makes the emission from this type of accretion flows extremely weak in
407: the far infrared and optical bands.
408: 
409: We need to stress that, in principle, the contribution from ADAF
410: sources should be easily disentangled not only from that due to
411: sources with a flat and steep spectrum, but also from that due to GPS
412: sources which also have strongly inverted spectra. GPS sources are
413: typically much brighter (with fluxes typically ranging from a few to
414: 10 Jy) but rarer (usually associated with QSOs) than the expected ADAF
415: sources. The number of GPS sources rapidly decreases with decreasing
416: flux, whereas ADAFs are expected to be much more numerous at faint
417: flux levels.  As a result, GPS are only minor contributors to the
418: fluctuations at small angular scales, whereas ADAFs would be mostly
419: significant at these scales.  Therefore it should be possible to study
420: these two populations independently.
421: 
422: Note that we have shown the expected temperature fluctuations due to
423: ADAF sources only for the lowest energy channel of Planck.  If most of
424: the sources are indeed in the range of luminosities consistent with
425: those observed so far, then this channel is expected to have the
426: largest (possibly major) contribution, due to the high-frequency
427: cutoff in the spectrum of these sources. However, if a substantial
428: population of high-luminosity sources is present, then some
429: contribution should also be present in the other channels of the
430: Planck LFI.  The availability of multifrequency data should allow an
431: efficient identification of pixels contaminated by discrete
432: sources. In order to carry out a substraction of the contaminating
433: flux one should therefore take into account that strongly inverted
434: spectra such as those considered here may not be present in most
435: frequency channels but give rise to a strong contamination up to a
436: certain frequency, and then abruptly drop. It should also be pointed
437: out that, contrary to some of the GPS sources for which variability
438: has been observed (e.g. Stanghellini et al. 1998), the radio sources in
439: the hot accretion flows are usually not very variable. A lack of
440: variability is particularly important for a proper removal of sources
441: from the spectral fitting.
442: 
443: We note that ADAFs around massive black holes could also be found in
444: spiral galaxies such as the Galactic nucleus Sgr A$^*$. However, even
445: if ADAFs were indeed common in spiral nuclei, their potential
446: contribution to the CMB anisotropy would still be dominated by that
447: from ellipticals. Inferred black hole masses are found to be
448: proportional to the mass of the bulge component of their host galaxy,
449: implying $M_{BH} \sim 10^6-10^7 M_\odot$ for spiral galaxies. As a
450: consequence, the contribution from spirals should be much lower, as
451: the radio flux scales as $M_{\rm BH}^{2.5-3}$ (Franceschini et
452: al. 1998), and (see \S3) their spectrum would peak at frequencies
453: higher than those of elliptical cores and affect higher energy
454: (e.g. sub-mm, mm) channels of CMB experiments. Because of this, given
455: enough sensitivity, the relevance of ADAFs in quiescent spiral nuclei
456: may also be assesed separately by the forthcoming experiments.
457: 
458: Finally, we note that in our analysis we do not take into account the
459: effects of source clustering. Clustering decreases the effective
460: number of objects in randomly placed cells and, consequently, enhances
461: the cell to cell fluctuations.  There is indeed evidence that the
462: positions in the sky of a wide variety of extragalactic sources are
463: correlated (Shaver 1988).  However, the specific correlation function
464: of our radio--submm sources in early-type galaxies is not
465: well-constrained. The analyses of Toffolatti et al. (1998) have
466: shown that the contribution due to clustering (using the two-point
467: correlation function from sources selected at 5 GHz; Loan, Wall
468: \& Lahav 1997) is generally small in comparison with the Poisson
469: term; however, the relative importance of clustering increases if
470: sources are substracted out from the Planck maps down to faint flux
471: levels.
472: 
473: \acknowledgements 
474: We thank Ramesh Narayan for motivating this work,
475: and Martin White for useful discussions.  We also thank the anonimous
476: referee whose comments greatly improved the presentation of this
477: paper.  T.\,D.\,M.\ acknowledges support provided by NASA through
478: Chandra Postdoctoral Fellowship grant number PF8-10005 awarded by the
479: Chandra Science Center, which is operated by the Smithsonian
480: Astrophysical Observatory for NASA under contract NAS8-39073.
481: 
482: 
483: \begin{references}
484: 
485: \reference{} Allen S.W., Di Matteo T., Fabian A.C., 2000, MNRAS, 311, 493
486: 
487: \reference{} Blandford, R. D. \& Begelman, M. C. 1999, \mnras, 303, L1
488: 
489: \reference{} Cavaliere, A. \& Setti, G. 1976, A\&A, 46, 81
490: 
491: \reference{} Condon, J. J. 1974, ApJ, 188, 279
492: 
493: \reference{} Cooray, A. R., Grego, L., Holzappel, W. L., Marshall, J.,
494: \& Carlstrom, J. E. 1998, ApJ, 115, 1388
495: 
496: \reference{} De Zotti, G., Granato, G. L., Silva, L., Maino, D., \& Danese, L.
497: astro-ph/9912282
498: 
499: \reference{} Di Matteo T., Allen S.W., 1999, ApJ, 527, L21
500: 
501: \reference{DM99b} Di Matteo, T., Fabian, A.\,C., Rees, M.\,J., Carilli, C.\,L.,
502: Ivison, R.\,J.\, 1999, MNRAS, 305, 492
503: 
504: \reference{DM99a} Di Matteo, T., Quataert, E., Allen, S.\,W., Narayan, R., Fabian
505: A.\,C., 2000a, MNRAS, 311, 507
506: 
507: \reference{} Di Matteo, T.,  Carilli, C.L., Fabian A.C., 2000b, ApJ, submitted
508: 
509: \reference{} Fabian A.~C., Rees M.~J., 1995, MNRAS, 277, L55
510: 
511: \reference{} Franceschini, A., Toffolatti, L., Danese, L. \& De Zotti. G.
512: 1989, ApJ, 344, 35
513: 
514: \reference{} Franceschini, A., Vercellone, S., \& Fabian, A. C. 1998, MNRAS, 297, 817
515: 
516: \reference{} Gawiser, E., \& Smoot, G. F. 1997, ApJ, 480, L1
517: 
518: \reference{} Guerra, E. J., Haarsma, D. B., Partridge, R. B. 1998, AAS, 193, 4003
519: 
520: \reference{} Igumenshchev, I. V.  Abramowicz, M. A. 1999, \mnras, 303, 309 
521: 
522: \reference{} Kogut, A., et al. 1996, ApJ, 464, L5
523: 
524: \reference{} Loan, A. J., Wall, J. V. \& Lahav, O. 1997, MNRAS, 286, 994
525: 
526: \reference{} Mahadevan, R. 1997, ApJ, 477, 585
527: 
528: \reference{} Magorrian, J. et al. 1998, AJ, 115, 2285
529: 
530: \reference{} Mather, J. C., Fixsen, D. J., Shafer, R. A., Mosier, C., \&
531: Wilkinson, D. T. 1999, ApJ, 512, 511
532: 
533: \reference{} Mushotzky, R.F., Cowie, L.L., Barger, A.J., Arnaud A., 2000, Nature, in press
534: 
535: \reference{} Narayan, R. \& Yi, I. 1995, ApJ, 444, 231 
536: 
537: \reference{N99} Narayan, R., Mahadevan, R., Quataert, E.\ 1998,
538: Theory of Black Hole Accretion Disks, edited by Marek A. Abramowicz,
539: Gunnlaugur Bjornsson, and James E. Pringle. Cambridge University Press, p.148
540: 
541: \reference{} Philips, R. B. \& Mutel, R. L., 1982, A\&A, 106, 21
542: 
543: \reference{} Rees M.~J., Begelman M.~C., Blandford R.~D., Phinney E.~S.,
544: 1982, Nature, 295, 17 
545: 
546: \reference{} Scheuer, P. A. G. 1957, {\em Proc. Cambridge Phil. Soc.}, 53, 764
547: 
548: \reference{} Scheuer, P. A. G. 1974, MNRAS, 166, 329
549: 
550: \reference{} Scott, D. \& White, M. 1999, A\&A, 346, 1
551: 
552: \reference{} Shaver, P. A. 1988, in 'IAU Symposium 130, Evolution of large Scale Structures
553: in the Universe', ed. J. Audouze, M.-C. pelletan, and A. Szalay
554: (Dordrecht: Reidel) 
555: 
556: \reference{} Slee O.B., Sadler E.M., Reynolds J.E., Ekers R.D., 1994, MNRAS, 269, 92
557: 
558: \reference{} Smoot, G. F., et al. 1992, ApJ, 396, L1
559: 
560: \reference{} Sokasian, A., Gawiser, E., \& Smoot, G. F. 1998, ApJ submitted, 
561: astro-ph/9811311 
562: 
563: \reference{} Stanghellini, C., O'Dea, C. P., Dallacasa, D., Baum, S. A.,
564: Fanti, R., \& Fanti, C. 1998, A\&AS, 131, 303
565: 
566: \reference{} Stone, J. M., Pringle, J. E.  Begelman, M. C. 1999, \mnras,
567: 310, 1002
568: 
569: \reference{} Tegmark, M. \& Efstathiou, G. 1996, MNRAS, 281, 1297
570: 
571: \reference{} Toffolatti, L., Gomez, F. A., De Zotti, G., Mazzei, P.,
572: Franceschini, A., Danese, L. \& Burigana, C. 1999a, MNRAS, 297, 117
573: 
574: \reference{} Toffolatti, L., De Zotti, G., Argueso, F., \& Burigana, C.
575: astro-ph/9902343
576: 
577: \reference{} Wrobel J.~M., 1991, AJ, 101, 12
578: 
579: \reference{} Yi, I., \& Boughn, S. P. 1998, APJ, 499, 198
580: 
581: \end{references}
582: 
583: \newpage
584: 
585: 
586: \begin{figure}[t]
587: \centerline{\epsfysize=5.7in\epsffile{fig1.ps}}
588: \caption{Synchrotron emission from low-radiative efficiency accretion
589: flows.  The various lines correspond to various amounts of mass loss
590: in the flows ($p$ = 0, 0.2, 0.4, 0.6 from top to bottom).  The
591: uppermost curve is the standard ADAF model with no outflows.  The
592: shaded region indicates the range of luminosities which best fit the
593: data in the cores of ellipticals observed so far (Di Matteo et
594: al. 2000b).}
595: \label{fig:1}
596: \end{figure}
597: 
598: \begin{figure}[t]
599: \centerline{\epsfysize=5.7in\epsffile{fig2.ps}}
600: \caption{Potential contribution to the CMB anisotropy from accretion
601: flows in the cores of ellipticals as a function of the beamsize $\theta_0$
602: at $\nu = 30$ GHz.  The solid line shows the expected fluctuation
603: level if the luminosities of all sources were in the range
604: corresponding to the shaded region of Figure~1.  The other lines show
605: the fluctuations expected if a fraction $f$ of sources with a higher
606: luminosity (corresponding to the uppermost curve of Figure 1) were
607: mixed with these sources. The noise level indicated for Planck is the
608: average $\Delta T/T$ per pixel for 1$\sigma$ detection after two full
609: sky coverages.}
610: \label{fig:2}
611: \end{figure}
612: 
613: \begin{figure}[t]
614: \centerline{\epsfysize=5.7in\epsffile{fig3.ps}}
615: \caption{Fluctuations in the CMB anisotropy due to accretion flows  
616: in the cores of ellipticals, after removal of all sources above
617: a flux $S_{\rm lim}$. Here a fraction 
618: $f=0.5$ of high-luminosity sources has been assumed in the sample.}
619: \label{fig:3}
620: \end{figure}
621: 
622: \begin{figure}[t]
623: \centerline{\epsfysize=5.7in\epsffile{fig4.ps}}
624: \caption{Poisson component of the angular power spectrum of our sources
625: for a range of flux cuts: $S_{\rm lim}=100$ mJy (dotted line), 
626: $S_{\rm lim}=10$ mJy (dashed line), $S_{\rm lim}=1$ mJy (dotted-dashed line).
627: This is shown for the case where a fraction $f=0.5$ of high-luminosity sources
628: is present in the sample. To compare with the level of primary anisotropy
629: expected, the prediction for a standard CDM spectrum is also shown, normalized
630: to COBE. The thick solid line is the expected contribution to the power spectrum 
631: from noise in the 30 GHz channel of the Planck LFI.}
632: \label{fig:4}
633: \end{figure}
634: 
635: 
636: \end{document}
637: 
638: 
639: 
640: 
641: 
642: 
643: 
644: 
645: 
646: 
647: 
648: 
649: 
650: 
651: 
652: 
653: 
654: 
655: 
656: 
657: 
658: 
659: 
660: 
661: 
662: 
663: 
664: 
665: 
666: