ptsrc_ApJL.tex accepted by ApJL \documentstyle[12pt,aasms4]{article} \def\msun{{\,M_\odot}} \def\simlt{\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}} \def\simgt{\lower.5ex\hbox{$\; \buildrel > \over \sim \;$}} \def\um{{\,\mu\rm m}} \def\cm{{\rm\,cm}} \def\km{{\rm\,km}} \def\au{{\rm\,AU}} \def\pc{{\rm\,pc}} \def\kpc{{\rm\,kpc}} \def\mpc{{\rm\,Mpc}} \def\sec{{\rm\,s}} \def\yr{{\rm\,yr}} \def\gm{{\rm\,g}} \def\kms{{\rm\,km\,s^{-1}}} \def\mdot{{\rm\,\msun\,yr^{-1}}} \def\gms{{\rm\,g\,s^{-1}}} \def\gcm3{{\rm\,g\,cm^{-3}}} \def\ncm3{{\rm\,cm^{-3}}} \def\kelvin{{\rm\,K}} \def\erg{{\rm\,erg}} \def\ev{{\rm\,eV}} \def\hz{{\rm\,Hz}} \def\>{$>$} \def\<{$<$} \def\bsl{$\backslash$} \def\refbook#1{\refindent#1} \def\refindent{\par\noindent\hangindent=3pc\hangafter=1 } \def\aa#1#2#3{\refindent#1, A\&A, {\bf#2}, #3.} \def\aalett#1#2#3{\refindent#1, A\&A {\it (Letters)}, {\bf#2}, #3.} \def\aasup#1#2#3{\refindent#1, A\&AS, #2, #3} \def\aj#1#2#3{\refindent#1, AJ, #2, #3} \def\apj#1#2#3{\refindent#1, {\it ApJ}, {\bf#2}, #3.} \def\apjlett#1#2#3{\refindent#1, {\it ApJ (Letters)}, {\bf #2}, #3.} \def\apjsup#1#2#3{\refindent#1, ApJS, #2, #3} \def\araa#1#2#3{\refindent#1, ARA\&A, #2, #3} \def\baas#1#2#3{\refindent#1, BAAS, #2, #3} \def\icarus#1#2#3{\refindent#1, Icarus, #2, #3} \def\mnras#1#2#3{\refindent#1, {\it MNRAS}, {\bf#2}, #3.} \def\nature#1#2#3{\refindent#1, {\it Nature}, {\bf #2}, #3.} \def\pasj#1#2#3{\refindent#1, PASJ, #2, #3} \def\pasp#1#2#3{\refindent#1, PASP, #2, #3} \def\qjras#1#2#3{\refindent#1, QJRAS, #2, #3} \def\science#1#2#3{\refindent#1, Science, #2, #3} \def\sov#1#2#3{\refindent#1, Soviet Astr., #2, #3} \def\sovlett#1#2#3{\refindent#1, Soviet Astr.\ Lett., #2, #3} \def\refpaper#1#2#3#4{\refindent#1, #2, #3, #4} \lefthead{Coker \& Melia} \righthead{Bondi-Hoyle Accretion onto Sgr A*} \begin{document} \title{Hydrodynamical Accretion Onto Sgr A* From\\ Distributed Point Sources} \author{Robert F. Coker\altaffilmark{1}$^*$ and Fulvio Melia\altaffilmark{2}$^{*\dag}$} \affil{$^*$Physics Department, The University of Arizona, Tucson, AZ 85721} \affil{$^{\dag}$Steward Observatory, The University of Arizona, Tucson, AZ 85721} \altaffiltext{1}{NASA GSRP Fellow.} \altaffiltext{2}{Presidential Young Investigator.} \begin{abstract} Spectral and kinematic studies suggest that the nonthermal radio source Sgr A*, located at the center of the Milky Way, is a supermassive compact object with a mass $\sim 2-3\times{10}^6\msun$. Winds from nearby stars, located $\approx 0.06$ pc to the east of Sgr A*, should, in the absence of any outflow from the putative black hole itself, be accreting onto this object. We report the results of the first 3D Bondi-Hoyle hydrodynamical numerical simulations of this process under the assumption that the Galactic center wind is generated by several different point sources (here assumed to be 10 pseudo-randomly placed stars). Our results show that the accretion rate onto the central object can be higher than in the case of a uniform flow since wind-wind shocks dissipate some of the bulk kinetic energy and lead to a higher capture rate for the gas. However, even for this highly non-uniform medium, most of the accreting gas carries with it a relatively low level of specific angular momentum, though large transient fluctuations can occur. Additionally, the post-bow-shock focusing of the gas can be substantially different than that for a uniform flow, but it depends strongly on the stellar spatial distribution. We discuss how this affects the morphology of the gas in the inner 0.15 pc of the Galaxy and the consequences for accretion disk models of Sgr A*. \end{abstract} \keywords{black hole physics---hydrodynamics---Galaxy: center---galaxies: nuclei---ISM: jets and outflows---stars: mass-loss} \section{Introduction} Sgr A* may be a massive ($\sim 2-3\times 10^6\;M_\odot$) point-like object dominating the gravitational potential in the inner $\la 0.5$ pc region of the Galaxy. This inference is based on the large proper motion of nearby stars (\cite{HA95}; \cite{EG97}; \cite{GEOE97}), the spectrum of Sgr A* (\cite{MJN92}), its low proper motion ($\simlt 20 \kms$; \cite{B96}), and its unique location (\cite{LAS91}). The gaseous flows in this region are themselves rather complex, and key constituents appear to be the cluster of mass-losing, blue, luminous stars comprising the IRS 16 assemblage, which is located within several arc seconds ($1^{\prime\prime} \approx$ 0.04 pc in projection at the distance to the Galactic center) from Sgr A*. Measurements of high outflow velocities associated with IR sources in Sgr A West (\cite{K91}) and in IRS 16 (\cite{G91}), the $H_2$ emission in the circumnuclear disk (CND) from molecular gas being shocked by a nuclear mass outflow (\cite{G86}), broad Br$\alpha$, Br$\gamma$ and He I emission lines (\cite{HKS82}; \cite{AHH90}; \cite{G91}), and radio continuum observations of IRS 7 (\cite{YM91}), provide clear evidence of a hypersonic wind, with velocity $v_w \sim500-1000\; \kms$, a number density $n_w\sim10^{3-4}\;\ncm3$, and a total mass loss rate $\dot M_w\sim3-4\times10^{-3}\mdot$, pervading the inner parsec of the Galaxy. Many of Sgr A*'s radiative characteristics may be due to its accretion of the IRS 16 wind. In the classical Bondi-Hoyle (BH) scenario (\cite{BH44}), the mass accretion rate for a uniform hypersonic flow is $\dot M_{BH} = \pi {R_A}^2 m_H n_w v_w$, in terms of the accretion radius $R_A \equiv 2 G M / {v_w}^2$. At the Galactic center, with $n_w \sim 5.5 \times 10^3 \ncm3$ and $v_w \sim700 \kms$, we would therefore expect an accretion rate $\dot M_{BH} \sim 10^{22} \gms$ onto the black hole, with a capture radius $R_A \sim .02 \pc$. Since this accretion rate is sub-Eddington for a one million solar mass object, the accreting gas is mostly unimpeded by the escaping radiation field and is thus essentially in hydrodynamic free-fall starting at $R_A$. Our initial numerical simulations of this process, assuming a highly simplistic uniform flow past a point mass (\cite{RM94}; \cite{CM96}) have verified these expectations. On the other hand, the nature of Bondi-Hoyle accretion onto a point like object also presents somewhat of a challenge in understanding what happens to the gas as it settles down into a planar configuration close to the event horizon. Fluctuations beyond the bow shock (located at $\sim R_A$) produce a transient accretion of net angular momentum that ought to result in the formation of a temporary (albeit small) disk. The circularization radius is $r_c \approx 2 \lambda^2 r_g$, where $r_g = 2 G M / c^2$ is the Schwarzschild radius and $\lambda$ is the specific angular momentum in units of $c r_g$. Our simulations of the BH accretion from a uniform flow suggest that in this configuration $\langle\lambda\rangle\sim 3-20$. More realistically, the inflow itself carries angular momentum, so that the formation of a disk-like structure at small radii (i.e., $r\approx 10^{2-3}\;r_g$) may be difficult to avoid (\cite{M94}). However, the observations do not appear to favor the presence of a {\it standard} $\alpha$-disk at small radii (\cite{M96}). The current upper limits on the infrared flux from Sgr A* (\cite{M97}) suggest that either (1) the circularized flow does not form an $\alpha$-disk, but rather advects most of its dissipated energy through the event horizon (e.g., \cite{NYM95}), (2) the Bondi-Hoyle flow merges into a massive, fossilized disk, storing most of the deposited matter at large radii (\cite{FM97}), or (3) Sgr A* is not a point-like object (\cite{HA95}). Added to this is the fact that in reality the flow past Sgr A* is not likely to be uniform. For example, one might expect many shocks to form as a result of wind-wind collisions within the IRS 16 comples, even before the plasma reaches $R_A$. With this consequent loss of bulk kinetic energy, it would not be surprising to see the black hole accrete at an even larger rate than in the uniform case. The implications for the spectral characteristics of Sgr A*, and thus its nature, are significant. We have therefore undertaken the task of simulating the BH accretion from the spherical winds of a distribution of 10 individual point sources located at an average distance of a few $R_A$ from the central object. As we shall see below, the accretion rate depends not only on the distance of the mass-losing star cluster from the accretor but also on the relative spatial distribution of the sources. As suggested by related work involving linear gradients (\cite{RA95}), the average value of $\lambda$ is also larger than that of the uniform case, exhibiting large temporal fluctuations, but still not as large as one might expect. In this {\sl Letter} we present the results of the first 3D hydrodynamical calculations of multiple-source stellar winds accreted by a point mass. \section{The Calculational Algorithm and Physical Setup} Although we here attempt to model the wind source more realistically than in previous work, we shall still focus on the hydrodynamical aspects of the flow and thus we do not include heating or cooling (see, e.g., \cite{M94}). Nor do we include any relativistic effects. Instead, we assume the medium to be an unmagnetized, polytropic gas with pressure $P = (\gamma - 1)\rho e$ where $\rho$, and $e$ are the mass and internal energy densities, respectively, and $\gamma = 4/3$ is the adiabatic index. This value of $\gamma$ was chosen so that the sonic point $d_s = R_A \times (5-3\gamma)/4$ occurs outside the accretor, which typically is modeled with a radius $\sim 0.1R_A$. Some earlier work (e.g., \cite{P89}) found that varying $\gamma$ affects the flow pattern significantly, particularly at radii much smaller than $R_A$, while others (e.g., \cite{R96}) found that it does not, particularly at the larger radii considered here. We use the numerical algorithm ZEUS-3D, a general purpose code for MHD fluids developed at NCSA (\cite{No94}). The calculations are carried out within a cubical domain of solution with ${112}^3$ active zones geometrically scaled so that the central zones are 1/32 times the size of the outermost zones, closely mimicking the ``multiply nested grids'' arrangement used by other researchers (e.g., \cite{RM94}). This allows for maximal resolution of the accretor within the computer memory limits available while sufficiently resolving the wind sources and minimizing zone-to-zone boundary effects. The total volume is $(16 R_A)^3$ or $\sim(0.28 \pc)^3$ using a point mass $M = 1\times{10}^6\msun$ located at the origin and a Mach 10 Galactic center wind with $v_w = 700\;\kms$. The initial density is set to a small value and the velocity is set to zero. The internal energy density is chosen such that the temperature is $\sim10^4\;\kelvin$. Free outflow conditions are imposed on the outermost zones and each time step is determined by the Courant condition with a Courant number of 0.5. The 10 identical stellar wind sources are modeled by forcing the velocity in 10 subregions of $5^3$ zones to be constant (at $v_w$) while the densities in these subvolumes are set so that the total mass flow into the volume of solution, $\dot M_w$, is 3$\times10^{-3}\mdot$. Recent observations (\cite{N94}) suggest that the IRS 16 wind is somewhat colder than the temperature assumed here, with a Mach number of $\sim$ 30, and that it may originate from more than 15 sources of varying strength rather than 10 sources of equal strength. This level of detail will be addressed in future work. To gauge the dependence of our results on the source configuration, we here simulate the BH process using two different stellar distributions. For numerical reasons, the sources need to be placed a minimum of 8 zones ($\simlt .2 R_A$) apart so no attempt is made here to imitate exactly the distribution within the IRS 16 cluster. The {\it average} location for the sources in run 1 is $\sim 4 R_A \hat{z}$ (with the accretor located at the origin) and $\sim 3 R_A \hat{z}$ for run 2. For run 1, the sources are distributed fairly randomly with x, y, and z being allowed to vary by $\pm3 R_A$ of the average location. For run 2, x and y are allowed to be as much as 5 $R_A$ from the average location while z can only vary by $\pm1 R_A$. These distributions are chosen to represent the extremes of a spherical stellar distribution versus a more or less planar one. For comparison, IRS 16 NW and IRS 16 NE, two bright members of the IRS 16 cluster, are $\sim3$ and $\sim7 R_A$ away (in projection) from Sgr A* (\cite{M97}). Each simulation is evolved for 2000 years. The wind crossing time is $\sim$ 400 years and equilibrium, the point at which the original gas is swept out of the volume of solution and the mass accretion rate stabilizes, appears to be reached within 2 crossing times, or $\sim$ 800 years. \section{Comparison of the Flow Patterns} Figure 1 shows a logarithmic negative grey scale image of the density profile for a slice running through the center of the accretor, for run 1 taken 2000 years after the winds are ``turned on''. The image is 6 $R_A$ on a side with the $0.1 R_A$ radius accretor located at the center. One stellar source can be clearly seen to the upper right of the accretor; the other 9 sources are either not in the plane of the slice or are off to the right. The image spans 4 magnitudes of density with white being $\simlt0.01 n_w$ and black being $\simgt10 n_w$. Figure 2 is a similar image for run 2. Note the large density fluctuations across the overall region. Also, the flow pattern is clearly different for the two configurations. However, once the stellar winds have cleared the region of the original low density gas, both simulations point to an overall average density ($\sim10^3\ncm3$) in agreement with observations. A substantial difference between these point source simulations and previous uniform calculations is that although there are transient fragmentary shocks extending from the accretor boundary (see Figure 1), there is no large-scale bow shock structure. There are, however, bow shocks around some of the stellar sources as their wind impacts the stronger accumulated wind of the other sources (see the source in Figure 1), and, for run 2, a semi-stable partial bow shock appears in front of the accretor (see Figure 2). The wind-wind collisions convert kinetic energy into thermal energy with the result that a larger fraction of the gas is captured by the central engine. Although this effect is likely to be sensitive to the actual stellar distribution, a larger density of point sources should produce more wind-wind collisions and perhaps further raise the accretion rate. \section{The Accretion of Mass and Angular Momentum onto the Central Object} For a wind originating from a single point source located at a distance $R$ (\> $R_A$) from the accretor, $\dot M$ is less than $\dot M_{BH}$ due to the divergent flow. Specifically, for $R= 4 R_A$, $\dot M \sim 0.15\; \dot M_{BH}$. However, wind-wind collisions from multiple sources reduce the effective $v_w$, thereby increasing the effective $R_A$ so that all of the shocked gas will then accrete and more than compensate for the geometric losses. In Figure 3 we present the mass accretion rate, $\dot M$, and the accreted specific angular momentum, $\lambda$, versus time for run 1, starting 2 crossing times ($\sim$ 800 years) after the winds are ``turned on''. Figure 4 presents the results for run 2. The average values for the mass accretion rate once the systems have reached equilibrium are $\dot M = 2.1 \pm 0.3 \dot M_{BH}$ for run 1 and $1.1 \pm 0.2 \dot M_{BH}$ for run 2. The mass accretion rate shows high frequency temporal fluctuations (with a period of $\simlt 0.25 \yr$) due to the finite numerical resolution of the simulations. The low frequency aperiodic variations (on the order of $20\%$ in amplitude) reflect the time dependent nature of the flow. Thus, the mass accretion rate onto the central object, and consequently the emission arising from within the accretor boundary, is expected to vary by $\simlt 20-40\%$ (since in some models $L$ may vary by as much as $\propto {\dot M}^2$) over the corresponding time scale of $<100$ years, even though the mass flux from the stellar sources remains constant. Similarly, for run 1, the accreted $\lambda$ can vary by $50\%$ over $\simlt$ 200 years with an average equilibrium value of $37 \pm 10$. For run 2, $\langle\lambda\rangle = 62 \pm 5$ with smaller amplitude long term variability. Since previous uniform simulations resulted in $\langle\lambda\rangle \sim 3-20$, it appears that even with a large amount of angular momentum present in the wind, relatively little specific angular momentum is accreted. This is understandable since clumps of gas with a high specific angular momentum do not penetrate within 1 $R_A$. The variability in the sign of the components of $\lambda$ suggests that if an accretion disk forms at all, it dissolves, and reforms (perhaps) with an opposite sense of spin on a time scale of $\sim 100$ years. \section{Discussion, Conclusions, and Future Work} A variety of accretion scenarios for Sgr A* have been proposed over the years (\cite{M94}; \cite{FO94}; \cite{NYM95}; \cite{Be96}), each with its own restrictions on $\dot M$ and $\lambda$. While the accreted specific angular momentum determined in the present simulations is an order of magnitude too small to support the fossil disk scenario (since then the energy liberated as the wind impacts the fossil disk should be visible; \cite{FM97}), it is still large enough that any standard $\alpha$-disk would be easily detectable (\cite{M94}). The advection dominated disk scenario (\cite{NYM95}), while permitting a large range of values for $\lambda$, requires an accretion rate of $\simlt 10^{-5}\mdot$ or roughly $0.1 \dot M_{BH}$, which is 10-20 times smaller than the value derived here. In addition, our $\dot M$ is more than $10$ times larger than that permitted by the ``mono-energetic'' electron model of Sgr A* (\cite{Be96}). It appears that additional work is needed to reconcile disk models with the fact that the observed multiple wind sources result in a large mass accretion rate onto the central engine, if its mass is $\sim {10}^6\msun$. In view of this, it may not be unreasonable to conjecture that in fact no flattened disk actually forms in Sgr A*, but rather that the excess angular momentum is dissipated in a quasi-spherical configuration and that the thermalized energy is then advected inwards through the event horizon before the gas settles onto a plane (\cite{Me92}). These simulations suggest that the $\sim 0.1 \pc$ region of the Galaxy, centered on the wind sources, is swept clear of gas, leaving a hot, low density gas filling the central cavity. This is consistent with observations of the region within the CND (\cite{YW93}), and may be acting in concert with other mechanisms to produce the sharp inner edge of the CND (e.g., the abrupt change in gravitational potential; \cite{D89}). Additionally, a tongue of hot, dense gas has been observed that connects members of the IRS 16 cluster to Sgr A* (\cite{G91}). It is worthwhile noting that the images in Figures 1 and 2 show ridges of dense gas connecting the sources in the figures to the accretor. \section{Acknowledgments} This work was partially supported by NASA under grants NAGW-2518 and NGT-51637, and utilized the Origin 2000 computer system in Friendly User mode at the National Center for Super computing Applications, University of Illinois at Urbana-Champaign. \begin{thebibliography}{} \bibitem[Allen, Hyland \& Hillier 1990] {AHH90} \mnras{Allen, D., Hyland, A., \& Hillier, D. 1990}{244}{706} \bibitem[Backer 1996] {B96} \refbook{Backer, D. 1996, in Proc. of IAU Symp. No. 169, ed. XXX (Dordrecht: Kluwer), XXX.} \bibitem[Beckert, et al. 1996] {Be96} \aa{Beckert, T., Duschl, W., Mezger, P., \& Zylka R. 1996}{307}{450} \bibitem[Bondi \& Hoyle 1944] {BH44} \mnras{Bondi, H, \& Hoyle, F. 1944} {104}{273} \bibitem[Coker \& Melia 1996] {CM96} \refbook{Coker, R., \& Melia, F. 1996, in The Galactic Center (ASP Conf. Vol. 102), ed. R. Gredel (San Francisco: ASP), 403.} \bibitem[Duschl 1989] {D89} \mnras{Duschl, W. 1989}{240}{219} \bibitem[Eckart \& Genzel 1997] {EG97} \mnras{Eckart, A. \& Genzel, R. 1997}{284}{576} \bibitem[Falcke \& Melia 1997] {FM97} \apj{Falcke, H. \& Melia, F. 1997}{479}{740} \bibitem[Falcke \& Heinrich 1994] {FO94} \aa{Falcke, H. \& Heinrich, O. 1994}{292}{430} \bibitem[Gatley, I., et al. 1986] {G86} \mnras{Gatley, I., Jones, T., Hyland, A., Wade, R., Geballe, T., \& Krisciunas, K. 1986}{222}{299} \bibitem[Geballe, et al. 1991] {G91} \apjlett{Geballe, T., Krisciunas, K., Bailey, J., \& Wade, R. 1991}{370}{L73} \bibitem[Genzel, Eckart, Ott \& Eisenhauer 1997] {GEOE97} Genzel, R., Eckart, A., Ott, T. \& Eisenhauer, F. 1997, {\it M.N.R.A.S.}, in press. \bibitem[Hall, Kleinmann \& Scoville 1982] {HKS82} \apjlett{Hall, D., Kleinmann, S., \& Scoville, N. 1982}{260}{L53} \bibitem[Haller, et al. 1995] {HA95} \apj{Haller, J.W., Rieke, M.J., Rieke, G.H., Tamblyn, P., Close, L. \& Melia, F. 1995}{456}{194} \bibitem[Krabbe, et al. 1991] {K91} \apjlett{Krabbe, A., Genzel, R., Drapatz, S., \& Rotaciuc, V. 1991}{382}{L19} \bibitem[Lacy, Achtermann, \& Serabyn 1991] {LAS91} \apjlett{Lacy, J., Achtermann, J., \& Serabyn, E. 1991}{380}{L71} \bibitem[Melia 1992] {Me92} \apjlett{Melia, F. 1992}{387}{L25} \bibitem[Melia 1994] {M94} \apj{Melia, F. 1994}{426}{577} \bibitem[Melia 1996] {M96} \refbook{Melia, F. 1996, in The Galactic Center (ASP Conf. Vol. 102), ed. R. Gredel (San Francisco: ASP), 395.} \bibitem[Melia, Coker \& Yusef-Zadeh 1996] {MCY96} \apjlett{Melia, F., Coker, R. F. \& Yusef-Zadeh, F. 1996}{460}{L33} \bibitem[Melia, Jokipii, \& Narayanan 1992] {MJN92} \apjlett{Melia, F., Jokipii, R., \& Narayanan, A. 1991}{395}{L87} \bibitem[Menten, et al. 1997] {M97} \apjlett{Menten, K., Reid, M., Eckart, A., \& Genzel, R. 1997}{475}{L111} \bibitem[Najarro, et al. 1994] {N94} \aa{Najarro, F., Hillier, D., Kudritzki, R., Krabbe, A., Genzel, R., Lutz, D., Drapatz, S., \& Geballe, T. 1994}{285}{573} \bibitem[Narayan, Yi, \& Mahadevan 1995] {NYM95} \nature{Narayan, R., Yi, I., \& Mahadevan, R. 1995}{374}{623} \bibitem[Norman 1994] {No94} \baas{Norman, M. 1994}{184}{50.01} \bibitem[Petrich, et al. 1989] {P89} \apj{Petrich, L., Shapiro, S., Stark, R. \& Teukolsky, S. 1989}{336}{313} \bibitem[Ruffert 1996] {R96} \aa{Ruffert, M. 1996}{311}{817} \bibitem[Ruffert \& Anzer 1995] {RA95} \aa{Ruffert, M, \& Anzer, U. 1995} {295}{108} \bibitem[Ruffert \& Melia 1994] {RM94} \aalett{Ruffert, M., \& Melia, F. 1994}{288}{L29} \bibitem[Yusef-Zadeh \& Melia 1991] {YM91} \apjlett{Yusef-Zadeh, F. \& Melia, F. 1991}{385}{L41} \bibitem[Yusef-Zadeh \& Wardle 1993] {YW93} \apj{Yusef-Zadeh, F. \& Wardle, M. 1993}{405}{584} \end{thebibliography}{} \clearpage \figcaption[]{A logarithmic negative grey scale snapshot, 6 $R_A$ on a side, of the gas density for run 1 taken 2000 years after the winds are ``turned on''. It is a slice in the $\hat x-\hat z$ plane (with $\hat z$ to the right) through the accretor (marked with a filled white circle). White corresponds to a number density $\simlt 0.01 n_w$ and black corresponds to a number density $\simgt 10 n_w$.\label{fig1}} \figcaption[]{Same as Figure 1 but for run 2 at 2000 years.\label{fig2}} \figcaption[]{Accretion results for run 1. The upper solid curve is accreted specific angular momentum $\lambda$ (in units of $cr_g$). The scale for $\lambda$ is on the left side. The lower dotted curve is the mass accretion rate $\dot M$ (${10}^{-4}\mdot$) versus time. The scale for $\dot M$ is shown on the right side.\label{fig3}} \figcaption[]{Accretion results for run 2.\label{fig4}} \clearpage \plotone{fig1.ps} \clearpage \plotone{fig2.ps} \clearpage \plotone{fig3.ps} \clearpage \plotone{fig4.ps} \end{document}