From rohan@kailash.harvard.edu Thu Jun 12 10:25:56 1997 Date: Thu, 12 Jun 1997 10:25:01 -0400 From: rohan@kailash.harvard.edu (Rohan Mahadevan) To: gcnews@sgrastar.astro.umd.edu Subject: Sgr A* \documentstyle[12pt,aaspp4,flushrt]{article} %\documentstyle[aaspp4]{article} \input psfig \catcode`\@=11 % This allows us to modify PLAIN macros. \def\lsim{\mathrel{\mathpalette\@versim<}} \def\gsim{\mathrel{\mathpalette\@versim>}} \def\@versim#1#2{\vcenter{\offinterlineskip \ialign{$\m@th#1\hfil##\hfil$\crcr#2\crcr\sim\crcr } }} \newcommand{\Ref}{\hangindent=20pt \hangafter=1 \noindent} \newcommand{\StartRef}{\hyphenpenalty=10000 \raggedright} \newcommand{\rb}[1]{\raisebox{1.5ex}[0pt]{#1}} \spaceskip=0.4em plus 0.15em minus 0.15em \xspaceskip=0.5em \hsize=17 true cm \hoffset=0 true cm \vsize=22 true cm \def\be{\begin{equation}} \def\ee{\end{equation}} \def\msun{M_\odot} \def\sgr{Sgr A$^*$ } \def\mdot{\dot m} \def\ergs{\rm erg\,s^{-1}} \def\sles{\lower2pt\hbox{$\buildrel {\scriptstyle <} \over {\scriptstyle\sim}$}} \def\sgreat{\lower2pt\hbox{$\buildrel {\scriptstyle >} \over {\scriptstyle\sim}$}} \def\hangin#1{\hangindent=1cm \hangafter=1 \noindent #1} \def\arcsec{^{\prime \prime}} \def\arcmin{^{\prime}} \def\degree{^{\circ}} \begin{document} {\centerline {\bf Submitted to {\em The Astrophysical Journal} on June 11, 1997.}} \title{Advection-Dominated Accretion Model of Sagittarius A$^*$: Evidence for a Black Hole at the Galactic Center} \authoremail{authors@cfa.harvard.edu} \author{Ramesh~Narayan, Rohan~Mahadevan, Jonathan~E.~Grindlay, Robert~G.~Popham, Charles~Gammie} \affil{Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138} \begin{abstract} Sgr A$^*$ at the Galactic Center is a puzzling source. It has a mass $M=(2.5\pm0.4)\times10^6M_\odot$ which makes it an excellent black hole candidate. Observations of stellar winds and other gas flows in its vicinity suggest a mass accretion rate $\dot M\gsim {\rm few}\times10^{-6}\msun\,{\rm yr^{-1}}$. However, such an accretion rate would imply a luminosity $> 10^{40}~{\rm erg\,s^{-1}}$ if the radiative efficiency is the usual 10\%, whereas observations indicate a bolometric luminosity $<10^{37}~{\rm erg\,s^{-1}}$. The spectrum of Sgr A$^*$ is unusual, with emission extending over many decades of wavelength. We present a model of Sgr A$^*$ which is based on a two-temperature optically-thin advection-dominated accretion flow. The model is consistent with the estimated $M$ and $\dot M$, and it fits the observed fluxes in the radio and X-ray bands as well as measured upper limits in the sub-millimeter and infrared bands. The model explains the very low luminosity of Sgr A$^*$ by invoking advection. Most of the viscously dissipated energy is carried into the central star by the accreting gas, and therefore the radiative efficiency is extremely low, $\sim5\times10^{-6}$. A critical element of the model is the presence of an event horizon at the center. The success of the model may thus be viewed as confirmation that Sgr A$^*$ is a black hole. \end{abstract} \keywords{accretion, accretion disks --- black holes --- galaxies: nuclei --- Galaxy: center --- radiation mechanisms: bremsstrahlung, inverse Compton, synchrotron --- radio sources: Sgr A$^*$} \section{Introduction} The enigmatic radio source, Sagittarius A$^*$ (Sgr A$^*$), has for many years been a puzzle (Genzel \& Townes 1987; Genzel, Hollenbach \& Townes 1994; Mezger, Duschl \& Zylka 1996). The source is located at the dynamical center of the Galaxy and is presumed to be associated with a supermassive black hole. However, observations provide conflicting indications both for and against the black hole hypothesis and there is currently no model that explains all the observations. The paradoxical clues may be summarized briefly as follows: \noindent 1. Dynamical measurements indicate a dark mass of $\sim(2.5\pm0.4) \times10^6 M_\odot$ within the central 0.1 pc of the Galactic Center (Eckart \& Genzel 1997). This presumably represents the mass $M$ of the putative supermassive black hole in Sgr A$^*$. \noindent 2. Observations of stellar winds and gas flows near Sgr A$^*$, coupled with the above estimate of the black hole mass, suggest that \sgr must accrete gas from its surrounding at a mass accretion rate $\dot M \gsim{\rm few}\times 10^{-6}M_\odot{\rm yr^{-1}}$ (Genzel et al. 1994). \noindent 3. This $\dot M$ implies an accretion luminosity $L\sim0.1\dot Mc^2> 10^{40}~\ergs$, assuming a nominal radiative efficiency of 10\%. However, \sgr is an unusually dim source with a total luminosity from radio to $\gamma$-rays under $10^{37}~\ergs$. The extremely low luminosity has been used to argue against \sgr being an accreting black hole (Goldwurm et al. 1994). \noindent 4. For the above $M$ and $\dot M$ the peak emission from a standard thin accretion disk will be in the near infrared (cf. Frank, King \& Raine 1992). In fact, this is a generic prediction of any model that involves an optically thick flow radiating as a blackbody. However, Menten et al. (1997) have obtained a strong upper limit on the 2.2 micron flux of Sgr A$^*$ that effectively rules out such models. \noindent 5. \sgr is brightest in the radio/mm band (see \S2 and Fig. 1), it is weakly detected in X-rays (Predehl \& Trumper 1994), and it may have been detected between 100 MeV and 2 GeV (Merck et al. 1996). These observations, combined with the Menten et al. infrared upper limit, imply a spectral distribution completely unlike a blackbody. The observations suggest that \sgr is optically thin, and that many different radiation processes are in operation. Over the years a number of models have been proposed for Sgr A$^*$. Some of these have been phenomenological approaches which aim to explain the radio and infrared spectrum without including any detailed dynamics (e.g. Falcke 1996, Duschl \& Lesch 1994). These models generally require mass accretion rates lower than those indicated by the observations. Other models (Melia 1992, 1994; Mastichiadis \& Ozernoy 1994) have attempted to incorporate dynamics, but with drastic simplifying assumptions such as ignoring the angular momentum of the accreting gas. One model of \sgr that has attempted a self-consistent treatment of both viscous hydrodynamics and radiation processes is the advection-dominated accretion flow (ADAF) model of Narayan, Yi \& Mahadevan (1995). (See Rees 1982 for some early ideas along these lines.) The present paper is based on this model. An advection-dominated accretion flow is one in which most of the energy released by viscous dissipation is stored in the gas and advected to the center, and only a small fraction of the energy is radiated (Narayan \& Yi 1994, 1995a, 1995b; Abramowicz et al. 1995; Chen et al. 1995; see Narayan 1997 for a recent review). Most current work on ADAFs has been concerned with a branch of low $\dot M$ solutions which is present for mass accretion rates below a few percent of the Eddington rate (Narayan 1997; Esin, McClintock \& Narayan 1997). These low-$\dot M$ solutions are based on a two-temperature plasma (Shapiro, Lightman \& Eardley 1976; Rees et al. 1982); by assumption, the bulk of the viscous energy in these flows is deposited in the ions, and energy then flows from the ions to the electrons through Coulomb collisions. The ions achieve a nearly virial temperature, $T_i\sim10^{12}~{\rm K}/r$, where $r$ is the radius in Schwarzschild units, while the electron temperature saturates at around $10^9-10^{10}$ K for $r\ \sles\ 100$. The gas is optically thin and the flow does not suffer from any serious thermal or viscous instabilities (Narayan \& Yi 1995b; Abramowicz et al. 1995; Kato, Abramowicz \& Chen 1996). The low-$\dot M$ branch of ADAFs has several appealing features for explaining the observations of \sgr and other low-luminosity accretion systems. First, since most of the energy is advected with the gas and lost into the black hole, the model naturally explains the low luminosity of Sgr A$^*$. Second, being optically thin, the spectrum is quite different from blackbody. Indeed, the high electron and ion temperatures allow a variety of radiation processes to operate: synchrotron, bremsstrahlung and inverse Compton from the electrons (Narayan \& Yi 1995b), and $\gamma$-ray emission from the ions via pion production (Mahadevan, Narayan \& Krolik 1997). Therefore, the gas is likely to radiate over a wide range of wavelengths. The Narayan et al. (1995) ADAF model of \sgr gave a reasonable fit to the observations available at that time and provided a convincing explanation of the low luminosity of the source. However, the model required a black hole mass of $7\times10^5M_\odot$ which differs from the dynamical mass estimate of $2.5\times10^6M_\odot$ (Eckart \& Genzel 1997). This problem prompted us to take another look at the model. In addition, there have been important developments on the observational and theoretical front which further motivate the present study. The recent Menten et al. (1997) limit on the infrared flux of \sgr is well below previous ``detections'' (Rosa et al. 1991; Eckart et al. 1992) which were used by Narayan et al. (1995) to fit the ADAF model. Similarly, ASCA observations by Koyama et al. (1996) have shown that the X-ray emission from the Galactic Center is dominated by an X-ray burster which lies within 1.3 arcminutes of \sgr. This implies that most previous measurements of the X-ray flux of \sgr (Pavlinskii, Grebenev, \& Sunyaev 1992, 1994; Skinner et al. 1987) are suspect since they were obtained with inadequate spatial resolution. Only the ROSAT PSPC detection by Predehl \& Trumper (1994) survives as a reliable X-ray detection, although the corresponding luminosity of the source in the ROSAT band is uncertain due to uncertainties in the absorbing hydrogen column (cf. \S2). The modeling techniques have also advanced significantly during the intervening two years. The Narayan et al. (1995) model was a local one in the sense that the dynamics were calculated using a local self-similar solution (Narayan \& Yi 1994) and the radiation was computed with a local approximation for Compton scattering (Dermer, Liang \& Canfield 1991; Narayan \& Yi 1995b). Consistent global dynamical solutions have since been calculated, initially with a pseudo-Newtonian potential (Narayan, Kato \& Honma 1997; Chen, Abramowicz \& Lasota 1997) and more recently for the full relativistic Kerr problem (Abramowicz et al. 1996; Peitz \& Appl 1997; Gammie \& Popham 1997). A global scheme for Comptonization has also been developed (Narayan, Barret \& McClintock 1997), based on techniques due to Poutanen \& Svensson (1996). Finally, Nakamura et al. (1997) have shown that in addition to energy advection by ions, which had been considered in previous work, advection by electrons can be important in some circumstances. This effect is now included in the calculations (Appendix A). The present models are thus superior to those used by Narayan et al. (1995). In this paper we present improved ADAF models of Sgr A$^*$. We make use of the spectral data described in \S2 and employ the modeling techniques outlined in \S3. We present detailed results in \S4 and conclude with a discussion in \S5. \section{Spectral Data} The radio to near infrared (NIR) spectrum of Sgr A$^*$ has been of constant interest since the discovery of the source by Balick \& Brown (1974), and observations have been carried out from 400 MHz (Davies et al. 1976) to $\sim 10^{14}$ Hz (Menten et al. 1997). Table 1 is a compilation of the radio to NIR observations. The data are also plotted in Fig. 1. The two general features in the table and figures are 1) we have given only the maximum and minimum fluxes for frequencies at which variability has been observed, and 2) open circles correspond to low resolution measurements, which we treat as upper limits because of possible contamination, while filled circles correspond to high resolution measurements which most likely probe the accretion flow in Sgr A$^*$. This is discussed in detail below. In determining the spectrum of the ADAF from Sgr A$^*$, special attention has to made to the angular resolution of the observations, which must be comparable to the size of the ADAF. The angular size of an ADAF at the Galactic Center is $\lsim 0.06\arcsec$ ($\sim 500$ AU), where we have taken $M = 2.5\times10^{6} M_{\odot}$, and assumed a linear size of $\sim 10^4$ Schwarzschild radii. For frequencies $\lsim 40$ GHz, electron scattering in the ISM between Sgr A$^*$ and the observer leads to source broadening. The apparent size of the source is larger than the intrinsic size, and is proportional to the square of the observed wavelength (e.g. Davies et al. 1976; Backer 1982). Since the intrinsic source size ($\lsim 0.06{\arcsec}$) is smaller than the scattering size ($\sim 1{\arcsec}$ at low frequencies), the observed flux could be contaminated by structures (e.g. outflows or jets) which exist on scales smaller than the scattering size. Therefore, for frequencies $\lsim 40$ GHz, we label the points in Fig. 1 with open circles, indicating that the flux measurements should be considered as upper bounds rather than as firm detections of the central engine. For frequencies $\gsim 40$ GHz, the scattering size falls below the intrinsic size. High resolution observations can therefore determine directly the emission from the central engine. However, only a few high resolution ($\lsim 1$ mas) observations have been made at these wavelengths using VLBI or the VLBA. These radio observations are indicated by a $\star$ in Table 1, and are represented as filled circles in Fig. 1. We believe that these points best represent the emission from the ADAF \footnote{However since the scattering size approaches the intrinsic size at $\nu \sim 43$ GHz, this VLBI point might be slightly contaminated.} and we take the point at 86 GHz to be the best determined radio flux of \sgr for the purposes of modeling the source. Apart from these VLBI/VLBA points, all other radio points $\gsim 40$ GHz have been represented by open circles since these observations were done with poor resolution ($\gsim 4{\arcsec}$). The sub--millimeter to NIR observations are all upper limits, and are represented by open circles due to the low angular resolution of the measurements. However, the recent upper limit of Menten et al. (1997) at 2.2$\mu$m, is a high resolution measurement since the authors were able to determine accurately the relative position of the Galactic Center in their speckle images by comparing with a radio map of nearby stellar SiO masers. The effective resolution of the speckle image is $0.15\arcsec$, which is comparable to the size of the ADAF. The flux measured by Menten et al. is therefore an excellent upper limit for the emission of \sgr at NIR wavelengths, and we plot this point in Fig. 1 as a filled circle. X--ray and hard X--ray as well as gamma-ray observations of Sgr A* have long been limited by the relatively poor angular resolution available at these energies and yet very crowded field of the Galactic Center region. Even at hard X--ray energies (e.g. 20--100 keV) where the density of sources is relatively low, the Sgr A$^*$ source region contains more sources ($\sim$10) in the central 5$\degree$ of the Galaxy than any other region (Goldwurm et al. 1994). At soft X--ray (e.g. 0.5--4 keV) and low-medium X--ray energies (e.g. 2--10 keV), this same region is even more crowded, with complex diffuse emission and at least 10 sources within the central 1$\degree$ of Sgr A* (cf. Watson et al 1991 and Koyama et al 1996). Thus the flux and spectral distribution of the Galactic Center in the $\sim$1--100 keV band, where the ADAF bremsstrahlung component peaks (cf. Fig. 1), is particularly critical and angular resolution is paramount. At the soft X--ray end of this band, the relatively uncertain absorbing column density, $N_H$, towards the Sgr A$^*$ source is especially important since the value chosen (which can in principle be determined by future observations with AXAF) greatly affects the derived source flux and luminosity. We have included in the X--ray/hard X--ray region of the spectrum plotted in Figure 1 only 4 points or upper limits which all are the highest angular resolution available in their respective energy bands. These data are listed in Table 2. The one detection of a point-like source with position fully consistent with Sgr A$^*$ is the ROSAT detection (PSPC) with $\sim 20\arcsec$ resolution in the 0.8--2.5 keV band reported by Predehl and Trumper (1994). This detection is plotted for an assumed $N_H$ = 6 $\times$ 10$^{22}$ cm$^{-2}$, which is the usual best estimate (cf. Watson et al 1981) for the interstellar column density and which corresponds to the usually quoted visual extinction of A$_V \sim$ 25--30 mag. The total integrated luminosity in the 0.8--2.5 keV band is $L_X = 1.5 \times 10^{34}$ erg s$^{-1}$. \footnote{The $L_X$ and $\nu L_{\nu}$ calculations were done for the ROSAT data using the PIMMS program, supplied by the HEASARC at GSFC.} The $N_H$ used here differs from the much higher value (1.5 $\times$ 10$^{23}$ cm$^{-2}$) assumed by Predehl and Trumper. These authors chose a higher $N_H$ in order to make the soft X--ray luminosity more compatible with the variable compact source within $\sim 1\arcmin$ of Sgr A$^*$ reported by Skinner et al (1987) and Pavlinskii, Grebenev and Sunyaev (1992) from 2--20 keV coded aperture imaging observations with modest (few arcmin) resolution. However, recent ASCA observations of the Galactic center in the 2--10 keV band, with $\sim 1\arcmin$ angular resolution but much higher spectral resolution, have been reported by Koyama et al. (1996) and Maeda et al (1996). They show the Sgr A$^*$ region to be complex, with diffuse emission over a $\sim$2 $\times 3 \arcmin$ region and a point source at $1.3 \arcmin $ from Sgr A$^*$ which is an X--ray burster and therefore a neutron star in a (8.4h) binary system. This burster has very likely dominated the coded aperture imaging flux measurements. Koyama et al. therefore quote an upper limit of 10$^{36}$ erg/s as the 2--10 keV luminosity for the entire Sgr A$^*$ complex, with a possible actual value of $\sim$10$^{35}$ erg/s for the point source alone. We have plotted this ASCA upper limit in Fig. 1. The ROSAT PSPC detection, although only with angular resolution of $\sim 20\arcsec$, is a likely detection of Sgr A$^*$ with minimal contamination from the surrounding diffuse source since it appears in the ROSAT image to be consistent with a point source. However, higher spatial resolution (e.g. AXAF) X--ray imaging, and/or evidence for time variability, are needed to confirm this identification. In this sense, even the ROSAT detection might be regarded as an upper limit, although it is plotted as a solid symbol. The flux level plotted in Fig. 1 and listed in Table 2 corresponds to the likely $N_H$ value of $6\times10^{22}~{\rm cm^{-2}}$ and was obtained by considering two extreme values of a power law spectral index: photon index 1.0 and 2.0. These indices bracket the corresponding index ($\sim$1.4) for a bremsstrahlung spectrum as predicted by our model. The detected flux in the soft (0.8--2.5 keV) band is, however, still highly sensitive to the assumed $N_H$ (as also pointed out by Predehl and Trumper): for $N_H$ = 8 $\times$ 10$^{22}$ the flux increases by a factor of 2.8 over that shown; whereas for $N_H$ = 5 $\times$ 10$^{22}$, it decreases by a factor of 0.6. Thus the fits to ADAF models described in the next section, which are otherwise greatly constrained by this ROSAT detection, should be regarded as uncertain in normalization by a factor perhaps as large as 2 if uncertainties in the interstellar $N_H$ are allowed for completely and perhaps as large as 3 if internal self-absorption in the vicinity of the AGN is allowed for (as suggested by Predehl and Trumper). The box around the ROSAT point in Fig. 1 includes both the uncertainty due to spectral index as well as an additional factor of 2 uncertainty in normalizaton. The uncertain value of $N_H$ can, in principle, be directly measured or at least greatly constrained by observations of Sgr A$^*$ with the ACIS instrument on AXAF. This could observe the interstellar absorption edge of Oxygen (at 0.8 keV) at the high angular resolution (1$\arcsec$) needed to completely isolate the surrounding diffuse emission. At still higher energies, 20--100 keV, we plot the two upper limits derived from the deep SIGMA observations of the Galactic Center complex reported by Goldwurm et al. (1994). These observations are not able to resolve the burster, and are likely affected by it (or, rather, constrain the hard X--ray flux from the burster). More sensitive observations in the hard X--ray band at energies above 60--100 keV, where the hard X--ray spectral component of neutron stars in bursters usually is not detectable (e.g. Barret \& Grindlay 1995) could provide a more stringent test of the ADAF spectrum as well as the predicted spectral break. The $\gamma$--ray spectrum shown in Fig. 1 corresponds to emission from the EGRET source 2EG 1746--2852 which is coincident with the Galactic Center (Merck et al. 1996). The source appears to be point--like to within the resolution of the instrument ($ \sim 1\degree$), and is a significant excess ($\sim$10$\sigma$) above the local diffuse emission. The source has a very hard spectrum with a spectral slope of 1.7 which differs significantly from other unidentified Galactic EGRET sources. The source might therefore represent emission from Sgr A*, or it might be unresolved emission from a giant molecular cloud in the vicinity of the Galactic Center. If the source is indeed Sgr A$^*$, the flux and spectrum may be explained as $\gamma$-ray emission via pion production and decay in the ADAF (Mahadevan et al. 1997). If the source is instead associated with a dense cloud of molecular hydrogen situated at the Galactic Center, it will again show a pion decay spectrum but with a spectral index (at energies above 100 MeV) given by the local cosmic ray proton spectrum (i.e. a power law with energy index $\sim-1.6$). Due to the low resolution of EGRET and uncertainty in the nature of the source, we plot the $\gamma$-ray data in Fig. 1 with open circles. The EGRET observations (as reported by Mayer-Hasselwander, 4th Compton Symp.) indicate possible variability, which would of course suggest that the source is indeed Sgr A$^*$, but the same EGRET data also indicate the source may be slightly resolved, which would argue that the radiation is significantly contaminated by surrounding diffuse emission. Future observations with higher sensitivity and better resolution would be invaluable in resolving this issue. \section{Modeling Techniques} We consider a black hole of mass $M$ accreting through an ADAF at a rate $\dot{m} \dot{M}_{\rm Edd}\msun\,{\rm yr^{-1}}$, where $\dot{M}_{\rm Edd} = 10 L_{\rm Edd}/c^2 = 1.39 \times 10^{18}(M/\msun) ~{\rm g\,s^{-1}} = 2.21 \times 10^{-8}(M/\msun) ~\msun{\rm yr^{-1}}$. The ADAF is in some sense dynamically intermediate between a thin disk and a spherical accretion flow: it is hot ($H/R \sim 1$), with rapid radial inflow, yet centrifugal support still plays a significant role. In describing the ADAF we shall generally refer to the scaled radius $r \equiv R/R_s$, where $R_s = 2 G M/c^2$ is the Schwarzschild radius. We assume that the ADAF extends from the black hole horizon, $r=1$, to an outer radius $r_{\rm out} = 10^5$. We take the mean angular momentum vector of the ADAF to be inclined at an angle $i$ to the line of sight. To find the dynamical structure of the ADAF, we use the fully relativistic, self-consistent, steady-state global solutions in the Kerr metric developed by Popham \& Gammie (1997). Their model uses a nearly standard viscosity prescription which is parameterized by the efficiency of angular momentum transport, $\alpha$ (assumed constant, independent of radius). The viscosity prescription has been modified, however, so as to preserve causality and to include relativistic effects. We also use a quasi-spherical prescription for the vertical structure based on Abramowicz, Lanza, and Percival (1997). The model is uniquely specified by four parameters: $\alpha$, the viscosity parameter; $\gamma$, the adiabatic index of the fluid, which is assumed to be a mixture of gas and magnetic fields; $f$, the advection parameter which gives the ratio of advected energy to the viscous heat input; and $a$, the rotation parameter of the black hole. The models considered here are extremely advection-dominated and so $f$ is very close to unity. Also, we consider only the case a = 0, i.e. a Schwarzschild black hole. Comparison of relativistic and nonrelativistic global models, and the self-similar solution of Narayan \& Yi (1994) shows that they are rather similar outside about $r = 5$. Inside this radius the density in our solution is very close to self-similar, but the temperature is slightly lower. The thermodynamic state of the flow is described by the ion temperature $T_i$, the electron temperature $T_e$, and the magnetic pressure $p_{\rm mag} \equiv (1 - \beta) p_{\rm tot}$. The total pressure is fixed by the dynamical model. We assume that $\beta$ is a constant, independent of radius; that implies $\gamma = (8 - 3\beta)/ (6 - 3\beta)$ (see Appendix A and Esin 1997). Given $\beta$ we know the run of gas pressure with radius. Since the electrons can cool efficiently and coupling between the ions and electrons is assumed to be weak (provided solely by Coulomb collisions), we have $T_i \gg T_e$. The run of gas pressure then gives $T_i$ to a good approximation. The heart of the problem is now calculating the electron temperature, which determines the spectral properties of the flow. A detailed description of our procedure is given in Esin et al. (1997). Basically, the electron temperature is determined by solving an energy balance equation: \be Q^{\rm e,adv} = Q^{\rm ie} + \delta\,Q^{\rm vis} - Q^{\rm rad}. \ee Here $Q^{\rm e,adv}$ is the rate at which energy in the electrons is advected inward by the flow. This term is discussed in greater detail in Appendix A. Our earlier work (with the exception of Esin et al. 1997) ignored the advective term, but recently Nakamura et al. (1997) (see also Mahadevan \& Quataert 1997) have shown that it can be important, particularly at low mass accretion rates. At low $\mdot$ the electrons are nearly adiabatic and the compressive heating as the gas flows in becomes large. The other terms in equation (1) are as follows: $Q^{\rm ie}$ is the rate at which the electrons are heated by Coulomb collisions, $Q^{\rm vis}$ is the total rate of viscous dissipation, and $Q^{\rm rad}$ is the radiative cooling. The parameter $\delta$ describes the fraction of the viscous heating that goes into the electrons. In general we set $\delta = 10^{-3} \sim m_e/m_p$. Because of Compton scattering, the radiative cooling term couples the flow at different radii. It is calculated using the iterative scattering method described in detail in Narayan, Barret \& McClintock (1997). The ADAF is represented by a logarithmically spaced grid of nested shells. Within each shell all flow variables are assumed uniform for $\pi/2 - \theta_H < \theta < \pi/2 + \theta_H$, where $\theta_H$ is an effective angular scale height, calculated as in Narayan, Barret \& McClintock (1997) Appendix A. We then guess a value of the electron temperature and compute the synchrotron and bremsstrahlung emission from each shell. Then, using the probability matrix elements $P_{ij}$, which give the probability that a photon emitted in shell $i$ of the ADAF is scattered by an electron in shell $j$, the rate of cooling through Compton scattering is calculated (see Narayan, Barret \& McClintock (1997) for details). This procedure is iterated until convergence is achieved. The iteration method ensures that multiple scattering within the ADAF is properly taken into account. Doppler shifts and ray deflections are ignored. Once the cooling term is calculated, we must get the photons out to the observer. We use essentially Newtonian photon transport, although we have taken a step toward relativistic photon transport by including gravitational redshift (this is also included in the Compton scattering calculation). We do not include Doppler shifts, which to lowest order in $v/c$ broaden the spectrum. Since all features in our spectra have $\delta\nu/\nu$ of order unity, this should not greatly alter the gross properties of the spectrum, except where it is very steep. At next higher order in $v/c$ gravitational redshift, higher order Doppler, and various geometric effects all enter. In our model we have included a correction for the volume of the emitting region ($1/\sqrt{1 - 1/r} \times$ the Euclidean value) and for gravitational redshift. The sum of these effects is that $L_\nu(\nu)$ observed at large radius is now $L_\nu(\nu \sqrt{1 - 1/r})$, that is, $L_\nu$ is shifted redward by the gravitational redshift factor. Jaroszynski \& Kurpiewski (1997) have recently considered the full effects of photon transport near a Kerr black hole on the spectrum of an ADAF. The above discussion is concerned with radiation from electrons. We also compute pion production by the hot protons in the ADAF and the resulting $\gamma$-ray emission through pion decay. The procedure we follow for this calculation is described in Mahadevan et al. (1997) and is based on earlier work by Dermer (1986). The present calculations differ in three respects from those in Mahadevan et al.: 1) the density profile of the protons is now given by the relativistic global solution of Popham \& Gammie (1997) instead of the pseudo-Newtonian solution of Narayan, Kato \& Honma (1997), 2) we allow for the non--spherical geometry of the flow via the angle $\theta_H$ mentioned above, and 3) we include the effect of gravitational redshift. The EGRET spectrum of the Galactic Center is consistent with a power--law distribution of proton energies (Mahadevan et al. 1997). In the present calculations we assume that the protons are nonthermal with a power-law index $p=2.3$. Note that it is consistent for us to assume a nonthermal distribution of proton energies, since Mahadevan \& Quataert (1997) have shown that the protons in an ADAF do not have time to thermalize. \section{Results} In this section we combine the modeling techniques outlined in \S3 with the observational constraints described in \S2 to come up with ADAF models of Sgr A$^*$. Our standard parameters are as follows. We assume that the ADAF extends from $r=1$ to 10$^{5}$. The outer radius is large enough that the results are insensitive to its value. For the black hole mass, we take the estimate of Eckart \& Genzel (1997): $M=2.5\times10^6M_\odot$. We assume exact equipartition between gas and magnetic pressure in the accreting gas: $\beta=0.5$. Following Hawley, Gammie \& Balbus (1996) we take the viscosity parameter to be $\alpha\sim0.6(1-\beta)=0.3$. We assume that electrons receive only a small fraction $\delta\sim m_e/m_p$ of the viscous dissipation as direct viscous heating: $\delta=0.001$. We assume a Schwarzschild black hole. Since we do not know the orientation of the angular momentum vector of the accreting gas, we set the inclination of the system to a generic value: $i=60^o$. The models considered here are extremely optically thin and so the results are virtually independent of $i$. The only parameter that we consider fully adjustable is the mass accretion rate $\mdot$. In each model described below, we have adjusted $\mdot$ so as to fit the ROSAT X-ray flux (but see the discussion in \S2 for uncertainties in the X-ray flux due to the hydrogen column). Although we treat $\mdot$ as a free parameter, there are in fact some constraints. The infrared star IRS 16 appears to be the primary supplier of gas to Sgr A$^*$. Assuming that \sgr captures a fraction of the wind of IRS 16 by Bondi accretion, Melia (1992) estimated $\dot M\approx 2\times10^{-4}~M_\odot{\rm yr^{-1}}$ for a wind velocity of $600~{\rm km\,s^{-1}}$ while Genzel et al. (1994) estimated a mass accretion rate of $\dot M\approx 6\times10^{-6}~M_\odot{\rm yr^{-1}}$ for a wind velocity of $1000~{\rm km\,s^{-1}}$ . We take these two estimates to be low and high extremes. Converting to Eddington units, $\mdot$ is thus likely to be in the range $10^{-4}<\mdot<3\times10^{-3}$. The solid line in Fig. 1 shows our baseline model where $M$, $\alpha$, $\beta$, and $\delta$ have been set to their standard values ($2.5\times10^6M_{\odot}, ~0.3, ~0.5, ~0.001$ respectively) and $\mdot$ has been adjusted to fit the X-ray flux. This model has a mass accretion rate of $\mdot=1.3\times10^{-4}$, which lies within the acceptable range of $\mdot$ discussed above. The computed spectrum has four well-defined peaks. From the left, these correspond to self-absorbed thermal synchrotron emission, singly Compton-scattered synchrotron radiation, bremsstrahlung emission, and $\gamma$-rays from neutral pion decay. The calculated spectrum fits the radio and infrared constraints quite well. The model spectrum passes through the VLBI radio flux measurement at 86 GHz (which we have identified as the most reliable of the radio observations, cf. \S2) and turns over sharply in the sub-millimeter band exactly as required by the sub-millimeter and infrared upper limits. The model also satisfies the X-ray upper limits and detection. Considering that only one parameter, $\mdot$, has been adjusted, we consider the agreement with the data impressive. Note that the infrared and X-ray fluxes have come down significantly compared to the data shown in Narayan et al. (1995). The present model is compatible with both the new measurements. Interestingly, we cannot find any model that fits either the new infrared upper limit with the old X-ray data, or the new X-ray flux with the old infrared data. The model does have a discrepancy in the $\gamma$-ray band; the predicted flux is lower than the observations by approximately an order of magnitude. The pion decay model requires an accretion rate of $\mdot=4.5\times10^{-4}$ to fit the observed flux whereas the fit to the rest of the spectrum gives $\mdot=1.3\times10^{-4}$. There is thus a discrepancy of a factor $\sim3-4$ between the two values of $\mdot$. We have been unable to come up with a reasonable resolution of this discrepancy; perhaps the EGRET source does not correspond to Sgr A$^*$, but rather to a central concentration of molecular gas in the Galaxy and resulting diffuse emission which appears unresolved to the $1^o$ beam of EGRET. The model also does not agree with the radio flux at frequencies below about 50 GHz. We discuss this discrepancy at the end of the section. Two features of the ADAF model in Fig. 1 should be highlighted. First, the model fits the data using a reasonable mass for the black hole, $M=2.5\times10^6M_\odot$. This is an improvement over the model described in Narayan et al. (1995) where the data could be fitted only with a mass $M=7\times10^5M_\odot$. The primary reason for the improvement is the inclusion of electron energy advection (Nakamura et al. 1997), or compressive heating of the electrons, as we explain below. Second, the model is extremely advection-dominated. The bolometric luminosity $L_{\rm bol}$ integrated over all frequencies is only $2.1\times10^{36}~\ergs$, which corresponds to a radiative efficiency, $\epsilon=L_{\rm bol}/\dot Mc^2=5\times10^{-6}$. It is this extraordinarily low radiative efficiency that allows the model to fit the observations with such a large mass accretion rate. In contrast, a standard thin accretion disk model of \sgr is ruled out quite comprehensively by the data. The dotted line in Fig. 1 shows a thin disk model with $\mdot=10^{-4}$, the lowest mass accretion rate we consider reasonable. The spectrum was calculated assuming that the emission is blackbody at each radius. The model predicts an infrared flux which is many orders of magnitude above the measured upper limits. The short dashed line in Fig. 1 shows another thin disk model where $\mdot$ has been reduced to $10^{-9}$. This model does satisfy the IR upper limits but does not fit any of the observations, either in the radio or X-ray bands. Further, the mass accretion rate is unreasonably small. A number of improvements have taken place in the modeling techniques since the publication of our first model of \sgr (Narayan et al. 1995). It is interesting to investigate what effect each improvement has had on the calculated spectrum. Figure 2 shows a sequence of models in which we start with the simplest version of the model and progressively add features one by one. The dotted line in Fig. 2 corresponds to the most primitive version of the ADAF model, in which the flow is assumed to have a self-similar form (Narayan \& Yi 1994) and neither compressive heating of electrons nor gravitational redshift is included. The optimized accretion rate is $\mdot=6\times10^{-5}$. The general shape of the spectrum is similar to that of our standard model (the solid line), but this model differs in three ways. First, the emission is stronger in nearly all bands compared to our standard model; in fact, the model is inconsistent with the infrared limit. Second, the mass accretion rate is lower than in our standard model by a factor of 2. This, combined with the higher luminosity, means that this model is not as advection-dominated as our standard model; we calculate a radiative efficiency of $\epsilon=2.4\times10^{-4}$. Third, the synchrotron peak is well above the 86 GHz VLBI flux measurement. The discrepancy in the radio flux is serious since the flux predicted by the model is more than an order of magnitude larger than that observed. Within the framework of the self-similar flow assumption, the only way to eliminate this problem is by changing one or more of the model parameters. This was, in fact, the primary reason why the Narayan et al. (1995) model required a black hole mass of $7\times10^5M_\odot$ instead of the measured mass of $2.5\times10^6M_\odot$ (see Fig. 3 below which shows how the spectrum is modified when the black hole mass is reduced). The short-dashed line in Fig. 2a shows the effect of replacing the self-similar flow by a global flow based on a pseudo-Newtonian potential (Narayan, Kato \& Honma 1997; Chen, Abramowicz \& Lasota 1997). The changes are minimal. The long-dashed line in Fig. 2a next shows the effect of including compressive heating (Appendix A and Nakamura et al. 1997). This model, which has $\mdot=1.1\times10^{-4}$, has a very different spectrum compared to the previous two models. The overall emission is significantly reduced, notably in the synchrotron peak, and the flow is substantially more advection-dominated: $\epsilon=2.1\times10^{-6}$. This model fits the 86 GHz data point very well without requiring any adjustment to the black hole mass. We thus confirm the result of Nakamura et al. (1997) that compressive heating is an important effect. Figure 2b shows the electron temperature profiles of the various models. The model with compressive heating has a significantly different temperature structure than the models without. At very low $\mdot$, as in our model of Sgr A$^*$, the dominant terms in the electron energy equation are the two pieces of $Q^{\rm e,adv}$ written in the Appendix, viz. the terms proportional to $dT_e/dR$ and $d\rho/dR$. These two dominant terms balance each other, while the rest of the terms in equation (1) are small. In other words, the electrons are essentially adiabatic. The adiabatic condition gives a lower electron temperature than in the previous two models and this accounts for the change in the shape of the synchrotron peak. The dash-dot line in Fig. 2a shows next the effect of including a fully relativistic global solution ($\mdot=1.3\times10^{-4}$), taken from Popham \& Gammie (1997), instead of the pseudo-Newtonian global solution employed in the previous two models. Most of the features are similar, but the overall emission is increased. The primary reason for this is that the relativistic solution has a lower radial velocity close to the black hole compared to the pseudo-Newtonian solution. (This is because $v$ is constrained to be less than $c$ whereas the pseudo-Newtonian model gives $v>c$ close to the black hole.) Consequently, the density is higher and this leads to increased emission. Finally, the solid line in Fig. 2a shows our standard model, which includes gravitational redshift. As expected, this model has a lower luminosity than the previous model, but is otherwise quite similar. Figures 3--5 show the effect of varying the other parameters in the model. Figure 3a shows the effect of varying the mass. In addition to the baseline model, two other models are shown with $M = 1/3$ and $3$ times $M$ of the baseline model. As Mahadevan (1997) has shown, with increasing $M$, the magnetic field strength decreases as $M^{-1/2}$ and this causes the synchrotron peak to move to lower frequencies $\propto M^{-1/2}$. Of the three models shown in Fig. 3a, the one with $M=2.5\times10^6M_\odot$ gives by far the best fit to the VLBI radio data. Assuming our choices of $\alpha$ and $\beta$ are correct, this independently confirms the black hole mass measured by Eckart \& Genzel (1997). Figure 3b shows the effect of varying the mass accretion rate. In addition to the baseline model, four other models are shown, with $\mdot=1/2, ~1/\sqrt{2}, ~\sqrt{2}, $ and $~2$ times $\mdot$ of the baseline model. The radiative efficiency of ADAFs varies rapidly with $\mdot$: $\epsilon\propto\mdot$ (Narayan \& Yi 1995b). Therefore, the luminosity varies as $\mdot^2$. This can be seen in both the synchrotron and bremsstrahlung peaks. The Compton peak in the infrared shows an even stronger dependence on $\mdot$. This is because the Comptonized flux is proportional to the product of the synchrotron emission (which is $\propto \mdot^2$) and the optical depth (which is $\propto\mdot$), so that the amplitude of this peak varies approximately as $\mdot^3$. The three intermediate models in Fig. 3b are consistent with all the data, but the two extreme models lie outside the X-ray error box and one of them also violates the sub-millimeter and infrared limits. Figure 4a shows the effect of varying $\alpha$. For each $\alpha$, we have adjusted the accretion rate to fit the X-ray flux. The spectrum is not very sensitive to $\alpha$ (except in the optical/IR), though the fitted mass accretion rates show modest variations: $\mdot=6.3\times10^{-4}, ~1.0\times10^{-3}, ~1.3\times10^{-3}, ~1.5\times10^{-3}$ for $\alpha=0.1,~0.2, ~0.3,~0.4$ respectively. Figure 4b shows the effect of varying $\beta$. Again, for each $\beta$, we have adjusted the accretion rate to fit the X-ray flux. In this case, we see quite substantial changes in the predicted spectrum. An increase in $\beta$ leads to a decrease in the magnetic field strength and one may be tempted to think that this would cause the synchrotron peak to reduce in amplitude. In fact, the opposite behavior is seen. The reason can be traced to compressive heating. Since the electrons are effectively adiabatic, their temperature profile is determined by the adiabatic index $\gamma_e$. As Appendix A shows, because the gas is a mixture of particles and magnetic field, the effective $\gamma_e$ depends on the parameter $\beta$. As $\beta$ goes up, the magnetic pressure goes down and $\gamma_e$ increases. A larger $\gamma_e$ leads to hotter electrons, see Fig. 4c, which causes more synchrotron emission. Unfortunately, this means that the results are sensitive to the details of how we model energy advection in the electrons. In this context we note that our equation of state for the electrons differs somewhat from the one used by Nakamura et al. (1997). We use a relativistic equation of state which causes $\gamma_e$ to change as the electron temperature approaches and crosses $T_e\sim m_ec^2/k$. In addition, we modify the adiabatic index of the electrons to allow for the equipartition magnetic field which is present in the gas and is coupled to the electrons. Nakamura et al. (1997) treat the particles as a separate component and ignore the field in their equation of state. Figure 4d shows the effect of varying $\delta$. For all $\delta\ \sles\ 0.01$ the spectrum is generally unaffected. This result is different from that found in Narayan, Barret \& McClintock (1997). Once again, the reason is the inclusion of compressive heating. Because the electrons are now essentially adiabatic, their temperature profile is not affected by modest changes in the heating or cooling. Only when $\delta$ is large, e.g. $\delta=0.03162$ (the dot-dashed line in Fig. 4d), do the electrons experience significant additional heating and only then does the spectrum show a noticeable change. We turn now to the low frequency radio data, where the model deviates substantially from the measurements. In our model, different parts of the synchrotron peak are produced at different radii in the flow, the emission at $10^{12}$ Hz coming from close to the black hole and the emission at lower frequencies coming from farther out. Thus, one simple way of improving the fit to the radio data is to modify the electron temperature beyond a few tens of Schwarzschild radii. Figure 5 is {\em only} an illustration. The solid line in Fig. 5a corresponds to our baseline model, while the dashed line is another model which is identical in all respects except that we have arbitrarily set $T_e=2\times10^9$ K over the radius range $r=20-1000$ (Fig. 5b). This model fits the data quite well. Is there any reason to think that the electrons might have the profile shown in Fig. 5b? Yes, there are several radial transport mechanisms which could drive the electrons to a nearly isothermal state. First, the long mean free path of the electrons can lead to fairly strong radial heat conduction (parallel to field lines). Second, synchrotron radiation, which has a thermalizing effect on the electrons (Mahadevan \& Quataert 1997), can also cause significant energy diffusion. Finally, the tangled magnetic fields which we assume to be present in the flow may move outwards as a result of buoyancy and may dissipate their energy and heat the electrons at larger radii. Some of these effects could well be episodic; it is interesting in this connection that the radio flux of \sgr is known to be quite variable (Zhao et al. 1989). These effects need further study. \section{Discussion} The main result of this paper is that the ADAF model (Narayan et al. 1995) provides a viable explanation of the spectral properties of Sgr A$^*$. Our basic model, shown by the solid line in Fig. 1, fits all the high resolution measurements from the radio to the X-ray band, including the stringent infrared limit of Menten et al. (1997) and the significantly reduced X-ray flux discussed in \S2. In contrast to the models presented in Narayan et al. (1995), where a black hole mass of $7\times 10^5M_\odot$ was required to accommodate the radio data, here we find that the model quite naturally fits the data with the correct mass of $2.5\times10^6M_\odot$ (Eckart \& Genzel 1997). This removes one of the main problems with the previous work. The model also predicts a mass accretion rate $\mdot\sim 1.3\times10^{-4}$ (in Eddington units) which is within the range considered likely by a direct estimate of the accretion rate (Genzel et al. 1994). We should emphasize that the model is fully consistent with a detailed treatment of hydrodynamics, radiation processes, and thermal balance of ions and electrons in the two-temperature plasma. Most other models in the literature are more primitive and/or ad hoc. In addition, the model is fairly robust to changes in the parameters (Fig. 4). The predicted spectrum is insensitive to large variations in the viscosity parameter $\alpha$ and the electron heating parameter $\delta$. The results do, however, vary with the equipartition parameter $\beta$ (Fig. 4b, 4c). There are two relatively minor problems with the present model. First, we under--predict the $\gamma$-ray flux in the EGRET band by an order of magnitude. This might indicate a residual uncertainty in the relative normalization of the protons and electrons, but we have not been able to come up with any specific proposal to eliminate the discrepancy. Perhaps the $\gamma$-ray source detected by EGRET with its one degree beam is not Sgr A$^*$, but surrounding diffuse emission. Second, the model under--predicts the radio flux below about 50 GHz. We can explain the observed emission by modifying the electron temperature profile in the accretion flow (see Fig. 5), but this is rather ad hoc. Another possibility is that the radio emission arises from a different component of the source, such as a jet, which is outside the scope of the model. (We note that Krichbaum et al. 1993 claim to have seen evidence for a jet in their 43 GHz VLBI image, though Backer et al. 1993 could not verify this.) In any case, there is no region of the spectrum where the model {\it over-predicts} the flux and this is important because such a discrepancy would be much harder to explain away. A more serious problem is the somewhat low electron temperatures predicted by the model. As Fig. 4c shows, we obtain temperatures below $10^{10}$ K at all radii, whereas VLBI observations at 43 GHz and 86 GHz indicate brightness temperatures in excess of $10^{10}$ K (Backer et al. 1993, Rogers et al. 1994). Thus, although the model fits the 86 GHz flux well, it seems to predict a larger angular size for the source, by a factor of $\sim 1.5$, than observations indicate. For comparison with future observations, we list here linear source radii at various radio frequencies according to the baseline model shown in Fig. 1: $1.3\times10^{12}$ cm at $10^{12.5}$ Hz, $4.1\times10^{12}$ cm at $10^{12}$ Hz, $1.0\times10^{13}$ cm at $10^{11.5}$ Hz, $2.1\times10^{13}$ cm at $10^{11}$ Hz, $4.1\times10^{13}$ cm at $10^{10.5}$ Hz, $8.3\times10^{13}$ cm at $10^{10}$ Hz. Note that the electron temperature depends on the value of the parameter $\beta$ (Fig. 4c). Thus, one might be able to fit the observed brightness temperatures by adjusting this parameter. An outstanding feature of the ADAF model presented here is its extraordinarily low radiative efficiency: $\epsilon=L_{\rm bol}/ \dot Mc^2=5\times10^{-6}$. The low efficiency permits the model to fit the very low luminosity of \sgr with a fairly large $\mdot$. Figure 1 shows two thin disk models with the standard radiative efficiency of $\epsilon\sim0.1$. A thin disk with the ``correct'' $\mdot$ (dotted line) over--predicts the infrared flux by four or five orders of magnitude. A model with $\mdot$ reduced by a factor of $10^5$ accommodates the IR limits but does not fit any of the other data and has a mass accretion rate which is wildly discrepant with independent estimates of $\mdot$. Some models in the literature which are based on a thin disk attempt to solve the luminosity problem by hiding a large fraction of the emission in the optical band where there is substantial extinction by dust (e.g. Falcke et al. 1993a, Melia 1994). However, the models invariably predict emission in the infrared at a level well above the Menten et al. (1997) limit. Other models in the literature which are not based on a thin disk also have problems fitting the low luminosity (e.g. Duschl \& Lesch 1994; Mastichiadis \& Ozernoy 1994). These models generally require values of $\mdot$ significantly below the lower limit of $10^{-4}$. Can we save the thin disk model by assuming that the disk has no viscosity at all, so that the gas in the disk does not accrete at the present time? Falcke \& Melia (1997) have recently shown that even such an extreme model is inconsistent because the mere presence of the disk, even though it does not accrete, will still lead to fairly strong infrared emission. The argument is that if there is inflow of material towards the Galactic Center via a Bondi-Hoyle-like flow, when the inflowing gas hits the disk and circularizes it will produce a substantial amount of thermal radiation. For the mass accretion rate of $\mdot\ \sgreat\ 10^{-4}$ estimated in the case of \sgr (cf. \S4) the predicted infrared flux is well above the Menten et al. (1997) limit in nearly all the cases considered by Falcke \& Melia (1997). The ADAF model circumvents this argument relatively easily. Since the accreting gas in the ADAF has a quasi-spherical shape (Narayan \& Yi 1995a), the outer Bondi-Hoyle flow will shock with the ADAF at a large radius $\sim10^5R_s$. The luminosity associated with this impact is quite low. Inside the impact radius, the accretion becomes advection-dominated and the gas again does not radiate very much. Several recent papers have made the point that the successful application of the ADAF model to any observed system is direct evidence that the accreting star is a black hole (Narayan, McClintock \& Yi 1996; Narayan, Yi \& Mahadevan 1996; Narayan, Barret \& McClintock 1997; Narayan, Garcia \& McClintock 1997). The argument is that if the object has an event horizon the advected energy in an ADAF will disappear from sight, whereas if the central object has a surface, then the stellar surface will be heated by the hot inflow from the ADAF and the advected thermal energy will be emitted as thermal radiation from the star. Thus, for a standard star with a surface, the radiative efficiency will be the canonical 10\% even if the accretion occurs via an ADAF. Only if the central star is a black hole can the radiative efficiency be truly low. \sgr is perhaps one of the best objects for this argument. This source has two strong observational constraints: (1) The mass accretion rate is estimated to be at least a few$\times10^{-6} M_\odot\,{\rm yr^{-1}}$, and (2) the bolometric luminosity is no greater than $10^{37}~\ergs$. In our opinion the only plausible resolution of these two conflicting pieces of evidence is to postulate (1) that the accretion in \sgr occurs via an ADAF and (2) that the central object is a supermassive black hole. This ``proof'' of the black hole nature of \sgr is qualitatively different from usual proofs which rely on a measurement of the mass. The usual argument is that if an object is too massive to be a neutron star it must be a black hole; it is a proof by elimination. Our ``proof'' is somewhat more direct and cuts to the essence of what constitutes a black hole, namely the presence of an event horizon. We argue that \sgr has an enormous luminosity deficit for which the only reasonable explanation is that the object has an event horizon; therefore, \sgr must be a black hole. One of the results to come out of this work is our confirmation that compressive heating of electrons cannot be ignored and must be included consistently in computations of ADAF spectra of low luminosity systems like \sgr (see Fig. 2). This point was made recently in an important paper by Nakamura et al. (1997). In view of this result it would be useful to revisit other low-$\mdot$ systems to which the ADAF model has been applied (e.g. soft X-ray transients, Narayan, Barret \& McClintock 1997; Hameury et al. 1997; NGC 4258, Lasota et al. 1996; low-luminosity nuclei of giant ellipticals, Fabian \& Rees 1995; Mahadevan 1997; Reynolds et al. 1996) and redo the analysis with the inclusion of compressive heating. Interestingly, Esin et al. (1997) have shown that compressive heating has a much weaker effect on ADAFs at higher values of $\mdot$. The reason is that with increasing $\mdot$ the other terms in the electron energy equation (1), notably the Coulomb collision term $Q^{\rm ie}$ and the radiative cooling term $Q^{\rm rad}$, become more important, and compressive heating no longer dominates. In the model presented in this paper the emission in the infrared and the radio (above about 50 GHz) comes from fairly small radii $\sim10$ Schwarzschild radii. Since the gas in the ADAF is nearly in free-fall, the characteristic time scales of the flow are quite short. We may thus expect rapid variability in \sgr at these wavelengths. The shortest likely time scale is the dynamical time, $t_{\rm dyn}\sim(GM/R^3)^{-1/2}$, which gives $t_{\rm dyn}\sim5000$ s at $R=10R_s$. The viscous time is longer than this by a factor of $1/\alpha \sim3$. The longer wavelength radio emission is from larger radii; for instance, the emission below 1 GHz comes from $R\gsim 10^3R_s$. Variability at these wavelengths will be correspondingly slower: $t_{\rm dyn}(10^3R_s)\sim5\times10^6$ s. The bremsstrahlung emission in X-rays is from a broad range of radii, but is dominated by large radii. Therefore, the X-ray flux should show much slower variations (timescale $\gsim$ 1 year) than the radio, millimeter or infrared fluxes. By assumption, the electrons in our model are fully thermal. Mahadevan \& Quataert (1997) have, however, shown that at low mass accretion rates similar to that in \sgr, electrons may not be thermalized at larger radii. The electron energy distribution at these radii will then be truncated even more sharply than in a Maxwellian, which will cause a reduction in the synchrotron emission. This will act to increase the discrepancy between the model and the data at low radio frequencies. This is an area for further work. Note, however, that a power-law distribution of electrons extending over any reasonable range of energies is ruled out by the observations. For instance, since optically thin nonthermal synchrotron emission usually has a spectral form $L_\nu\sim\nu^{-0.7}$, this means that in a $\nu L_\nu$ plot the optically thin emission would continue to rise as $\nu^{+0.3}$. Such a rise beyond $\sim 10^{12} - 10^{13}$Hz is strongly ruled out by the observed upper limits in the infrared. In this connection, we note that both Falcke (1996) and Duschl \& Lesch (1994) have proposed a nonthermal model for \sgr in which most of the electrons have Lorentz factors of around a few hundred. Such a mono-energetic electron distribution is required in order to reproduce the sharp cutoff observed in the sub--millimeter band. Falcke assumes a mono-energetic distribution in the ``nozzle'' of his jet-disk model, while Duschl \& Lesch assume a homogeneous sphere of mono-energetic electrons. Such mono-energetic models are not ruled out by the argument of the previous paragraph and indeed Falcke and Duschl \& Lesch are able to fit the radio spectrum of \sgr reasonably well. We, however, find the idea of a nonthermal but mono-energetic distribution somewhat artificial. Our ADAF model fits the data in a much more natural way by making use of a thermal distribution of electrons. Further, because of thermal synchrotron self-absorption, the cutoff of the spectrum in the sub-millimeter band is very sharp without requiring any fine--tuning. Further high spatial resolution observations of \sgr are highly desirable. The ADAF model can be tested and constrained with better observations of the spectrum and variability, especially in the millimeter, sub--millimeter, infrared and X-ray bands. Better $\gamma$-ray observations might also resolve the issue of whether the source detected by EGRET is Sgr A$^*$. \acknowledgments This work was supported in part by NSF grant AST 9423209 and NASA grant NAG 5-2837. We thank Shoji Kato for sending a preprint of Nakamura et al. (1997) prior to publication. RM thanks Jun-Hui Zhao and Mark Reid for useful discussions on the radio observations of the Galactic Center. \vfill\eject \begin{appendix} \section{Energy Advection by Electrons} The advection term in equation (1) can be written per unit volume as \be Q^{\rm e,adv}=\rho T_e v \frac{d s_e}{d R}, \ee where $s_e$ is the entropy of the electrons per unit total gas mass. This term was ignored in much of the earlier work under the assumption that the temperature of the electrons is determined primarily by a balance between the Coulomb transfer term $Q^{\rm ie}$ and the radiative cooling term $Q^{\rm rad}$ in equation (1). However, Nakamura et al. (1997) have shown that energy advection by electrons can be very important under some circumstances and may play a dominant role in determining the electron temperature. This is the case especially when the mass accretion rate is low, as in Sgr A$^*$. We begin with the relation \be T_e d s_e = d u_e + P_e d \left(\frac{1}{\rho}\right), \ee where $u_e$ is the internal energy of the electrons per unit mass and $P_e$ is the electron pressure. We consider a mixture of gas and magnetic fields. If gas pressure contributes a constant fraction $\beta$ to the total pressure $P_{tot}$, then \be P_{tot} = \frac{\rho k T_i}{\mu_i m_u} + \frac{\rho k T_e}{\mu_e m_u} + \frac{B^2}{24 \pi} = \frac{\rho k T_i}{\beta \mu_i m_u} + \frac{\rho k T_e}{\beta \mu_e m_u}. \ee It seems natural to denote the two terms on the right as the effective ion and electron pressure, each including an appropriate fraction of the magnetic pressure. Therefore, we have $P_e = \rho k T_e/(\beta \mu_e m_u)$. The internal energy of the gas is a sum of the ion, electron and magnetic field internal energies: \be \label{inten} u = \frac{3}{2} \frac{k T_i}{\mu_i m_u} + a(T_e) \frac{k T_e}{\mu_e m_u} + \frac{B^2}{8 \pi \rho} = \frac{6-3\beta}{2 \beta} \frac{k T_i}{\mu_i m_u} + \left[\frac{3 (1-\beta)}{\beta} + a(T_e)\right] \frac{k T_e}{\mu_e m_u}, \ee where the coefficient $a(T_e)$ varies from $3/2$ in the case of a non-relativistic electron gas, to $3$ for fully relativistic electrons. The general expression for $a$ as a function of the dimensionless electron temperature $\theta_e = k T_e/m_e c^2$ was derived by Chandrasekhar (1939, Chapter X, eq.[236]): \be a(\theta_e) = \frac{1}{\theta_e} \left(\frac{3 K_3 (1/\theta_e) + K_1 (1/\theta_e)}{4 K_2 (1/\theta_e)} - 1\right). \ee Note that the ions never become relativistic in these flows, so that the corresponding coefficient for the ions is always $\sim 3/2$. As we have done for the pressure, the right hand side of Eq. (\ref{inten}) may be naturally divided into two terms, the ion and electron internal energies: \be u_i = \frac{6-3\beta}{2 \beta} \frac{k T_i}{\mu_i m_u},\ \ \ {\rm and}\ \ \ u_e= \left[\frac{3 (1-\beta)}{\beta}+ a(T_e)\right] \frac{k T_e}{\mu_e m_u}. \ee In this interpretation, $u_i$ and $u_e$ are again ``effective'' quantities, which include contributions from the particles as well as the associated magnetic field. Note that the contribution of the magnetic field to the internal energy of each particle species is proportional to the contribution of these particles to the total pressure, a natural choice in our model where the ratio of the magnetic to gas pressure is fixed. Having defined $P_e$ and $u_e$, we can now write the energy advection term for the electrons as \be Q^{\rm e,adv}= \frac{\rho v k}{\mu_e m_u} \left[\frac{3 (1-\beta)}{\beta} + a(T_e) + T_e \frac{d a}{d T_e}\right] \frac{d T_e}{d R} - \frac{v k T_e}{\beta \mu_e m_u} \frac{d \rho}{d R}. \ee From the pressure and internal energy of the electrons, we calculate the effective adiabatic index of the electrons via the relation $\gamma_e-1=P_e/u_e$. This gives \be \gamma_e =\frac{4-3\beta+a\beta}{3-3\beta+a\beta}. \ee If the particles are relativistic, then $a=3$ and $\gamma_e=4/3$ regardless of $\beta$. This is because both the particles and the tangled field behave like radiation. If the particles are non-relativistic, however, $a=3/2$ and in this case \be \gamma_e=\frac{8-3\beta}{6-3\beta}, \ee as shown by Esin (1997). We see that $\gamma_e$ varies from 4/3 when $\beta=0$ to 5/3 when $\beta=1$. Thus, the adiabatic index depends fairly sensitively on $\beta$ so long as the electrons are non-relativistic. If the electrons in the accreting gas behave adiabatically, then their temperature varies as $T_e\propto\rho^{\gamma_e-1}$ (technically there is also the term proportional to $da/dT_e$ in eq A7 which becomes important when the electrons are quasi-relativistic, but we neglect this for simplicity). Thus, for a given amount of compression, the electrons are hotter when $\beta$ is large and cooler when $\beta$ is small. This explains the trend seen in Figs. 4b and 4c. \end{appendix} \newpage \begin{references} \reference{A} Abramowicz, M.A., Chen, X., Granath, M., \& Lasota, J.-P. 1996, ApJ, 471, 762 \reference{A} Abramowicz, M.A., Chen, X., Kato, S., Lasota, J.-P., \& Regev, O. 1995, ApJ, 438, L37 \reference{A} Abramowicz, M.A., Lanza, A., \& Percival, M.J. 1997, ApJ 479, 179 \reference{A} Alberdi, A., et al., 1993, A\&A, 277, L1 \reference{A} Backer, D., C, 1982, in Proc. IAU Symposium, eds. D.S. Heeschen \& C. M. Wade, 97, 389 No. 97 \reference{A} Balick, B., \& Brown, R., 1974, ApJ, 194,265 \reference{A} Barret, D., \& Grindlay, J. E., 1995, ApJ, 440, 841 \reference{A} Beckert, T., Duschl, W. J., Mezger, P. G., \& Zylka, R., 1996, A\&A, 307, 450 \reference{A} Brown, R. L, \& Lo, K. Y., 1982, ApJ, 253, 10 \reference{A} Chandrasekhar, S., 1939, in Introduction to the Study of Stellar Structure, (New York: Dover Publications) \reference{A} Chen, X., Abramowicz, M. A., \& Lasota, J.-P. 1997, ApJ, 476, 61 \reference{A} Chen, X., Abramowicz, M. A., Lasota, J.-P., Narayan, R., \& Yi, I. 1995, ApJ, 443, L61 \reference{A} Davies, R. D., Walsh, D., \& Booth, R. S., 1976, MNRAS, 177, 319 \reference{A} Dent, W. R. F., Matthews, H. E., Wade, R., \& Duncan, W. D., 1993, ApJ, 410, 650 \reference{A} Dermer, C. D. 1986, A\&A, 157, 223 \reference{A} Dermer, C. D., Liang, E. P., \& Canfield, E. 1991, ApJ, 369, 410 \reference{A} Duschl, W. J., \& Lesch, H. 1994, A\&A, 286, 431 \reference{A} Eckart, A., et al. 1992, Nature, 355, 526 \reference{A} Eckart, A., \& Genzel, R. 1997, MNRAS, 284, 576 \reference{A} Esin, A. A. 1997, ApJ, 482, in press (astro-ph/9701039) \reference{A} Esin, A. A., McClintock, J. E., \& Narayan, R. 1997, ApJ, submitted (astro-ph/9705237) \reference{A} Fabian, A. C., \& Rees, M. J., 1995, MNRAS, 277, L55 \reference{A} Falcke, H., 1996, in IAU Symp. 169, Unsolved Problems of the Milky Way, eds. L. Blitz \& P. J. Teuben (Dordrecht: Kluwer), 163 \reference{A} Falcke, H., Biermann, P. L., Duschl, W. J., Mezger, P. G., 1993a, A\&A, 270, 102 \reference{A} Falcke, H., Mannheim, K., \& Biermann, P. L. 1993b, A\&A, 278, L1 \reference{A} Falcke, H., \& Melia, F., 1997, ApJ, 479, 740 \reference{A} Frank, J., King, A., \& Raine, D. 1992, Accretion Power in Astrophysics, (Cambridge: Cambridge Univ. Press) \reference{A} Gammie, C., \& Popham, R. 1997, ApJ, submitted \reference{A} Genzel, R., \& Townes, C. H. 1987, ARAA, 25, 377 \reference{A} Genzel, R., Hollenbach, D., \& Townes, C. H. 1994, Rep. Prog. Phys., 57, 417 \reference{A} Gezari, D., Ozernoy, L., Varosi, F., McCreight, D., \& Joyce, R., 1994, in The Nuclei of Normal Galaxies, eds. R. Genzel \& A. I. Harris (Dordrecht: Kluwer), 427 \reference{A} Goldwurm, A. et al. 1994, Nature, 371, 589 \reference{A} Hameury, J. M., Lasota, J. P., McClintock, J. E., \& Narayan, R., 1997, ApJ, in press (astro-ph/9703095) \reference{A} Hawley, J. F., Gammie, C. F., \& Balbus, S. A., 1996, ApJ, 464, 690 \reference{A} Jaroszynski, M., \& Kurpiewski, A. 1997 A\&A, submitted (astro-ph 9705044). \reference{A} Lasota, J. P., Abramowicz, M. A., Chen, X., Krolik, J., Narayan, R., \& Yi, I., 1996, ApJ, 462, 142 \reference{A} Kato, S., Abramowicz, M. A., \& Chen, X. 1996, PASJ, 48, 67 \reference{A} Krichbaum, T. P., et al., 1993, A\&A, 274, L37 \reference{A} Koyama, K. et al. 1996, PASJ, 48, 249 \reference{A} Maeda, Y. et al., 1996, PASJ, 48, 417 \reference{A} Mahadevan, R. 1997, ApJ, 477, 585 \reference{A} Mahadevan, R., \& Quataert, 1997, ApJ, in press (astro-ph/9705067) \reference{A} Mahadevan, R, Narayan, R., \& Krolik, J. 1997, ApJ, 486, in press (astro-ph/9704018) \reference{A} Mastichiadis, A., \& Ozernoy, L. M., 1994, ApJ, 426, 599--603 \reference{A} Melia, F. 1992, ApJL, 387, L25 \reference{A} Melia, F. 1994, ApJ, 426, 577 \reference{A} Menten, K. M., Reid, M. J., Eckart, A., \& Genzel, R. 1997, ApJL, 475, L111 \reference{A} Merck, M., et al. 1996, A\&A Sup., 120, 465 \reference{A} Mezger, P. G., 1994, in The Nuclei of Normal Galaxies, eds. R. Genzel \& A. I. Harris (Dordrecht: Kluwer), 415 \reference{A} Mezger, P. G, Duschl, W. J., \& Zylka, R., 1996, ARA\&A, 7, 289 \reference{A} Nakamura, K. E., Kusunose, M., Matsumoto, R., \& Kato, S. 1997, PASJ, in press \reference{A} Narayan, R. 1997, Proc. IAU Colloq. 163, Accretion Phenomena and Related Outflows, A.S.P. Conf. Series, eds. D. T. Wickramasinghe, L. Ferrario, G. V. Bicknell, in press (astro-ph/9611113) \reference{A} Narayan, R., Barret, D., \& McClintock, J. E. 1997, ApJ, 482, 448 \reference{A} Narayan, R., Garcia, M. R., \& McClintock, J. E. 1997, ApJ, 478, L79 \reference{A} Narayan, R., Kato, S., \& Honma, F. 1997, ApJ, 476, 49 \reference{A} Narayan, R., McClintock, J. E., \& Yi, I. 1996, ApJ, 457, 821 \reference{A} Narayan, R., \& Yi, I. 1994, ApJ, 428, L13 \reference{A} Narayan, R., \& Yi, I. 1995a, ApJ, 444, 231 \reference{A} Narayan, R., \& Yi, I. 1995b, ApJ, 452, 710 \reference{A} Narayan, R., Yi, I., \& Mahadevan, R. 1995, Nature, 374, 623 \reference{A} Narayan, R., Yi, I., \& Mahadevan, R. 1996, A\&A Sup., 120, 287 \reference{A} Pavlinsky, M. N., Grebenev, S. A., \& Sunyaev, R. A., 1992, Soviet Astr. Lett., 18, 116 \reference{A} Pavlinsky, M. N., Grebenev, S. A., \& Sunyaev, R. A., 1994, ApJ, 425, 110 \reference{A} Peitz, J., \& Appl, S., 1997, MNRAS, 286, 681 \reference{A} Popham, R., \& Gammie, C. 1997, ApJ, submitted \reference{A} Poutanen, J., \& Svensson, R. 1996, ApJ, 470, 249 \reference{A} Predehl, P., \& Trumper, J. 1994, A\&A, 290, L29 \reference{A} Rogers, A. E. E., et al., 1994, ApJ, 434, L59 \reference{A} Rosa, M. R., Zinnecker, H., Moneti, A., \& Melnick, J., A\&A, 257, 515 \reference{A} Rees, M. J. 1982, Nature, 333, 523 \reference{A} Rees, M. J., Begelman, M. C., Blandford, R. D., \& Phinney, E. S. 1982, Nature, 295, 17 \reference{A} Reynolds, C. S., Di Matteo, T., Fabian, A. C., Hwang, U., \& Canizares, C. R., 1996, MNRAS, 283, L111 \reference{A} Serabyn, E., Carlstrom, J. E., \& Scoville, N. Z., 1992, ApJ, 401, L87 \reference{A} Shapiro, S.L., Lightman, A.P., \& Eardley, D.M. 1976, ApJ, 204, 187 \reference{A} Skinner, G. K. et al., 1987, Nature, 330, 544 \reference{A} Telesco, C. M, Davidson, J. A., \& Werner, M. W., 1996, ApJ, 456, 541 \reference{A} Watson, M. G., Willingale, R., Grindlay, J. E., Hertz, P., 1981, ApJ, 250, 142 \reference{A} Wright, M. C. H., Genzel, R., G\"usten, R., Jaffe, T., 1987, in The Galactic Center AIP conf. Proc, ed. D. C. Backer, 133 \reference{A} Yusef--Zadeh, F., Morris, M., \& Ekers, R. D., 1990, Nature, 348, 45 \reference{A} Zhao, J., Ekers, R. D., Goss, W. M., Lo, K. Y., \& Narayan, R., 1989, in The Center of the Galaxy, ed. M. Morris, 535 \reference{A} Zhao, J., Goss, W. M., Lo, K. Y., \& Ekers, R. D., 1992, in Relationships Between Active Galactic Nuclei and Starburst Galaxies, ed. A. Filippenko, 31 \reference{A} Zylka, R., \& Mezger, 1988, A\&A, 190, L25 \reference{A} Zylka, R., Mezger, P. G., \& Lesch, H., 1992, A\&A, 261, 119 \reference{A} Zylka, R., Mezger, Ward--Thompson, D., Duschl, W. J., \& Lesch, H., 1995, A\&A, 297, 83 \end{references} \newpage \noindent{\bf Figure Captions} \bigskip \noindent Figure~1. The open and filled circles represent various flux measurements and upper limits of Sgr A$^*$. We consider the filled circles to be more important as model constraints. The box at 1 keV represents the uncertainty in the X-ray flux. The solid line is our baseline ADAF model with the following parameters: $M=2.5\times10^6M_\odot$, $\alpha=0.3$, $\beta=0.5$, $\delta=0.001$. The Eddington-scaled mass accretion rate $\mdot$ has been adjusted to fit the X-ray flux, giving $\mdot=1.3\times10^{-4}$. The peak at the left is due to synchrotron radiation, the next peak is due to Compton scattering, the peak between 10--100 keV is due to bremsstrahlung and the peak above 100 MeV is due to pion production. The long--dashed line is a model in which the pion peak has been artificially raised by about an order of magnitude to fit the data. The dotted line is the spectrum corresponding to a standard thin accretion disk with $\mdot=10^{-4}$ while the short--dashed line is a thin disk with $\mdot=10^{-9}$. Neither of these models is satisfactory. \noindent Figure~2. (a) Spectra corresponding to five models, showing the effect of various approximations; see the text for details. (b) Electron temperature profiles corresponding to the same five models. The models represented with long-dashed and solid lines have nearly identical temperatures. \noindent Figure ~3. (a) The solid line is the baseline model shown in Fig. 1, with a black hole mass of $M=2.5\times10^6M_\odot$. The dotted line corresponds to a model with $M=8.3\times10^5M_\odot$ and the dashed line corresponds to $M=7.5\times10^6M_\odot$. (b) The solid line is the baseline model with a mass accretion rate $\mdot=1.3\times10^{-4}$. The dotted, short-dashed, long-dashed and dash-dotted lines corresponds to models with $\mdot=6.4\times10^{-5}, ~9.0\times10^{-5}, ~1.8\times10^{-4}, ~\mdot=2.5\times10^{-4}$, respectively. \noindent Figure ~4. (a) The solid line is the baseline model shown in Fig. 1, with $\alpha=0.3$. The dotted, short-dashed and long-dashed lines correspond to models with $\alpha=0.1, ~0.2, ~0.4$ respectively. (b) The solid line is the baseline model shown in Fig. 1, with $\beta=0.5$. The dotted, short-dashed, long-dashed lines and dot-dashed lines correspond to models with $\beta=0.3, ~0.4, ~0.6, ~0.7$ respectively. (c) Electron temperature profiles corresponding to the models shown in (b). (d) The solid line is the baseline model shown in Fig. 1, with $\delta=0.001$. The dotted, short-dashed, long-dashed and dot-dashed lines correspond to models with $\delta=0, ~0.003162, ~0.01, ~0.03162$ respectively. \noindent Figure ~5. (a) The solid line is the baseline model shown in Fig. 1. The dotted line corresponds to a model in which the temperature of the electrons has been fixed at $T_e=2\times10^9$ K over the radius range $r=20-1000$. This model fits the low frequency radio data better. (b) Electron temperature profiles corresponding to the two models shown in (a). \newpage \begin{table} \caption[tab 1] {Sgr A$^*$: Radio and NIR Observations (Distance = 8.5 kpc)} \vskip-3.0cm \begin{center} \begin{tabular}{lccccl}\hline & & & & & \\ \rb{$\nu$} & \rb{$\lambda$}& \rb{$\theta$} & \rb{$S_{\nu}$} & \rb{$\nu L_{\nu}$} & \rb {Ref.} \\ \rb{Hz} &\rb{$\mu$m}& \rb{$ \prime \prime$} &\rb{Jy}& \rb{ergs s$^{-1}$} & \\[-.5ex] \hline \hline 4.08$\times 10^{8}$ & $735294$ & $4.3$ & $\le 0.05$& $\le 1.76 \times 10^{30}$& Davies et al. 76 \\ 9.6$\times 10^{8}$ & $312500$ & $10$ & $0.29$& $ 2.41 \times 10^{31}$& Davies et al. 76 \\ 9.6$\times 10^{8}$ & $312500$ & $10$ & $0.27$& $ 2.24 \times 10^{31}$& Davies et al. 76 \\ 9.6$\times 10^{8}$ & $312500$ & $10$ & $0.30$& $ 2.49 \times 10^{31}$& Davies et al. 76 \\ 1.66$\times 10^{9}$ & $180722$ & $2.5$ & 0.56& $8.04\times 10^{31}$& Davies et al. 76 \\ 1.5 $\times 10^{9}$ & $200000$ & $\sim 1.2$ & 0.8& 1.04$\times 10^{32}$& Backer 82 \\ 1.5 $\times 10^{9}$ & $200000$ & $\sim 1.2$ & 0.3& 3.89$\times 10^{31}$& Zhao et al. 89 \\ 2.7 $\times 10^{9}$ & $110000$ & $\sim 0.65$ & .73& $ 1.7 \times 10^{32}$& Backer 82 \\ 2.7 $\times 10^{9}$ & $110000$ & $\sim 0.65$ & .42& $ 9.78 \times 10^{31}$& Brown \& Lo 82\\ 5.0 $\times 10^{9}$ & 60000 & $\sim 0.35$ & 1.13& 1.45$\times 10^{32}$& Zhao et al. 89 \\ 5.0 $\times 10^{9}$ & 60000 & $\sim$ 0.35 & 0.55& 7.13$\times 10^{31}$& Zhao et al. 89 \\ 8.1 $\times 10^{9}$ & 37000 & $\sim 0.22$ & 0.9& 6.23$\times 10^{32}$& Backer 82 \\ 8.1 $\times 10^{9}$ & 37000 & $\sim 0.22$ & 0.58& $4.01 \times 10^{32}$& Brown \& Lo 82 \\ 8.4 $\times 10^{9}$ & 35714 & $\sim 0.21$ & 1.07& $ 7.77 \times 10^{32}$& Zhao et al. 92 \\ 8.4 $\times 10^{9}$ & 35714 & $\sim 0.21$ & 0.55& $3.99 \times 10^{32}$& Zhao et al. 92 \\ 1.5 $\times 10^{10}$ & 20000 & $\sim 0.12$& 1.64& 2.1$\times 10^{33}$& Zhao et al. 92 \\ 1.5 $\times 10^{10}$ & 20000 & $\sim 0.12$& .68& 8.7$\times 10^{32}$& Zhao et al. 92 \\ 1.5 $\times 10^{10}$ & 20000 & 0.12$\times$0.24 & 1.15+0.01& (1.49+0.06)$\times 10^{33}$& Yusef--Zadeh et al. 90 \\ 2.2 $\times 10^{10}$ & 13600 & $\sim 0.08$& 2.1& $3.99\times 10^{33}$& Zhao et al. 92 \\ 2.2 $\times 10^{10}$ & 13600 & $\sim 0.08$& 0.8& 1.52$\times 10^{33}$& Zhao et al. 92 \\ \end{tabular} \end{center} \end{table} \newpage \addtocounter{table}{-1} \begin{table} \caption[] { Sgr A$^*$: Observations (continued).} \vskip-3.0cm \begin{center} \begin{tabular}{lccccl}\hline & & & & & \\ \rb{$\nu$} & \rb{$\lambda$}& \rb{$\theta$} & \rb{$S_{\nu}$} & \rb{$\nu L_{\nu}$} & \rb {Ref.} \\ \rb{Hz} &\rb{$\mu$m}& \rb{$ \prime \prime$} &\rb{Jy}& \rb{ergs s$^{-1}$} & \\[-.5ex] \hline \hline \hspace{-.29cm}$\!\star$ 4.3 $\times 10^{10}$ & 7000 & 0.75$\times 10^{-3}$ & 1.4$\pm 0.1$& $(5.2\pm 0.4)\times 10^{33}$& Krichbaum et al. 94 \\ 8.6 $\times 10^{10}$ & 3488 & 0.02 & 1.3& 9.67$\times 10^{33}$& Backer 82 \\ 8.6 $\times 10^{10}$ & 3488 & 4$\times8$ & 1.05& 7.8$\times 10^{33}$& Wright et al. 87 \\ \hspace{-.29cm}$\!\star$ 8.6 $\times 10^{10}$ & 3488 & $0.16 \times 10^{-3}$ & 1.4$\pm0.2$& $(1.04\pm 0.15)\times 10^{34}$& Rogers et al. 94 \\ 2.2$\times 10^{11}$ &1350& 1.9$\times 4.3$& $2.4\pm 0.5$ &$(4.57\pm .95)\times 10^{34}$ & Serabyn et al. 92 \\ \hspace{-.29cm}$\!\star$ 2.2$\times 10^{11}$ &1350& 2.6$\times 10^{-3}$& $1.07\pm 0.15$ &$(2.04\pm 0.29)\times 10^{34}$ & Alberdi et al. 93 \\ 2.3$\times 10^{11}$ &1300& 11& $2.5$ &$4.97\times 10^{34}$ & Zylka \& Mezger 88 \\ 2.3$\times 10^{11}$ &1300& 11& $2.6\pm0.6$ &$(5.2\pm 1.2)\times 10^{34}$ & Zylka et al. 92 \\ 3.5$\times 10^{11}$ &870& 8 & $4.8\pm 1.2$ &$(1.45\pm 0.36)\times 10^{35}$ & Zylka et al. 92 \\ 3.75$\times 10^{11}$ &800& 13 & $3.5\pm 0.5$ &$(1.13\pm 0.16)\times 10^{35}$ & Zylka et al. 95\\ 5.0$\times 10^{11}$ &600& 10 & $4.0\pm 1.2$ &$(5.19\pm 0.52)\times 10^{35}$ & Zylka et al. 95\\ 6.7$\times 10^{11}$ &450& 7 & $\le 1.5$ &$ \le 8.69 \times 10^{34}$ & Dent et al. 93\\ 6.7$\times 10^{11}$ &450& 8 & $3.0\pm 1.0$ &$(1.74\pm 0.58)\times 10^{35}$ & Zylka et al. 95\\ 8.6$\times 10^{11}$ &350& 11 & $\le 10$ &$\le 7.4\times 10^{35}$ & Mezger 1994\\ 8.6$\times 10^{11}$ &350& 30& $18.5\pm 9$ &$(1.38\pm 0.67)\times 10^{36}$ & Zylka et al. 92 \\ 1.0$\times 10^{13}$ &30& 8 & $\le 120$ &$\le 1.04 \times 10^{38}$ & Zylka et al. 92 \\ 1.0$\times 10^{13}$ &30& 4 & $\le 20$ &$\le 1.7 \times 10^{37}$ & Telesco et al. 96 \\ 1.5$\times 10^{13}$ &20& 1.6 & $\le 1$ &$ \le 1.3 \times 10^{36}$ & Gezari et al. 94\\ 1.56$\times 10^{13}$ &19.2& 4 & $\le 1.4$ &$ \le 1.9 \times 10^{36}$ & Telesco et al. 96\\ 1.7$\times 10^{13}$ &18& 2.3 $\times 1.3$ & $\le 0.3$ &$\le 4.4\times 10^{35}$ & Zylka et al. 92 \\ 2.3--3.6$\times 10^{13}$ &13--8& 2.3 $\times 1.3$ & $\le 0.1$ &$\le 2.6\times 10^{35}$ & Zylka et al. 92 \\ \hspace{-.29cm}$\!\star$ 1.4$\times 10^{14}$ &2.2& 0.15 &$\le 9\times 10^{-3}$ & $\le 1.1\times 10^{35}$ & Menten et al. 97 \\ \end{tabular} \end{center} \end{table} \begin{table} \caption[] { Sgr A$^*$: X--Ray \& $\gamma$--Ray Observations (Distance = 8.5 kpc).} %\vskip-3.0cm \begin{center} \begin{tabular}{lccccl}\hline & & & & & \\ \rb{Energy} & \rb{Telescope/}& \rb{$\theta$}& \rb{$L_{\rm EB}$ \footnotemark[1]} & \rb{$\nu L_{\nu}$} & \rb{Ref.} \\ \rb{Band (EB)}&\rb{Instrument} & &\rb{erg s$^{-1}$} & \rb{erg s$^{-1}$} & \\[-.5ex] \hline \hline 0.8 - 2.5 keV& ROSAT& $\sim 20 \arcsec$ & $ 1.55 \times 10^{34}$ & 1.6 $\times 10^{34}$ & Predehl \& Tr\"umper 94 \footnotemark[2] \\ 2 - 10 keV& ASCA & $\sim 1\arcmin$ &$\le 6.4 \times 10^{35}$ & $\le 4.8 \times 10^{35}$ & Koyama et al. 96 \\ 35 - 75 keV &SIGMA &$\sim 15 \arcmin$& $\le 3.5\times 10^{35}$ & $\le 4.8 \times 10^{35}$ & Goldwrum et al. 94 \\ 75 - 150 keV &SIGMA & $\sim 15 \arcmin$& $ \le 2.4 \times 10^{35}$& $\le 3.6 \times 10^{35}$& Goldwrum et al. 94 \\ 30 - 50 MeV& EGRET & $\sim 1\degree$& & $\le 1.8 \times 10^{36}$ &Merck et al. 96\\ 50 - 70 MeV& EGRET & $\sim 1\degree$& & $\le 1.1 \times 10^{36}$ &Merck et al. 96\\ 70 - 100 MeV& EGRET & $\sim 1\degree$& & $\le 8.3 \times 10^{35}$ &Merck et al. 96\\ 100 - 150 MeV& EGRET & $\sim 1\degree$& & $\le 6.2 \times 10^{35}$ &Merck et al. 96\\ 150 - 300 MeV& EGRET & $\sim 1\degree$& & $(4.9^{+2.0}_{-2.1}) \times 10^{35}$&Merck et al. 96\\ 300 - 500 MeV& EGRET & $\sim 1\degree$& & $(1.2^{+0.24}_{-0.26}) \times 10^{36}$ &Merck et al. 96\\ 500 - 1000 MeV& EGRET & $\sim 1\degree$& & $(1.4^{+0.28}_{-0.28}) \times 10^{36}$ &Merck et al. 96\\ 1 - 2 GeV& EGRET & $\sim 1\degree$& & $(1.7^{+0.28}_{0.28}) \times 10^{36}$ &Merck et al. 96\\ 2 - 4 GeV& EGRET & $\sim 1\degree$& & $(2.2^{+0.82}_{-0.62}) \times 10^{36}$ &Merck et al. 96\\ 4 - 10 GeV& EGRET & $\sim 1\degree$& & $(8.3^{+4.2}_{-4.7}) \times 10^{35}$ &Merck et al. 96\\ & &&&& \\ \multicolumn{6}{l}{ \footnotesize{$^{1}L_{\rm EB}$ is the total luminosity integrated over the band.}} \\ \multicolumn{6}{l}{\footnotesize{$^2$ This flux is obtained using $N_H = 6\times 10^{22} $cm$^{-2}$ as opposed to the much higher }} \\ \multicolumn{6}{l}{\footnotesize{ column density used by Predehl \& Trumper (1994). This is discussed in \S 2.}} \end{tabular} \end{center} \end{table} \newpage \pagestyle{empty} \begin{figure} \epsffile{adsgr1.ps} \end{figure} \begin{figure} \epsffile{adsgr2.ps} \end{figure} \begin{figure} \epsffile{adsgr3.ps} \end{figure} \begin{figure} \epsffile{adsgr4ab.ps} \end{figure} \begin{figure} \epsffile{adsgr4cd.ps} \end{figure} \begin{figure} \epsffile{adsgr5.ps} \end{figure} \end{document} ----- End Included Message ----- ----- End Included Message -----