0705.2283/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: \newcommand{\myemail}{lixue@xmu.edu.cn}
3: 
4: \begin{document}
5: 
6: \title{Studies of Thermally Unstable Accretion Disks around Black
7: Holes with Adaptive Pseudo-Spectral Domain Decomposition Method\\
8: I. Limit-Cycle Behavior in the Case of Moderate Viscosity}
9: 
10: \author{Shuang-Liang Li, Li Xue\altaffilmark{*} and Ju-Fu Lu}
11: \affil{Department of Physics and Institute of Theoretical Physics
12: and Astrophysics, Xiamen University, Xiamen, Fujian 361005, China}
13: 
14: \altaffiltext{*}{\myemail}
15: 
16: \begin{abstract}
17: We present a numerical method for spatially 1.5-dimensional and
18: time-dependent studies of accretion disks around black holes, that
19: is originated from a combination of the standard pseudo-spectral
20: method and the adaptive domain decomposition method existing in
21: the literature, but with a number of improvements in both the
22: numerical and physical senses. In particular, we introduce a new
23: treatment for the connection at the interfaces of decomposed
24: subdomains, construct an adaptive function for the mapping between
25: the Chebyshev-Gauss-Lobatto collocation points and the physical
26: collocation points in each subdomain, and modify the
27: over-simplified 1-dimensional basic equations of accretion flows
28: to account for the effects of viscous stresses in both the
29: azimuthal and radial directions. Our method is verified by
30: reproducing the best results obtained previously by Szuszkiewicz
31: \& Miller on the limit-cycle behavior of thermally unstable
32: accretion disks with moderate viscosity. A new finding is that,
33: according to our computations, the Bernoulli function of the
34: matter in such disks is always and everywhere negative, so that
35: outflows are unlikely to originate from these disks. We are
36: encouraged to study the more difficult case of thermally unstable
37: accretion disks with strong viscosity, and wish to report our
38: results in a subsequent paper.
39: \end{abstract}
40: 
41: \keywords{accretion, accretion disks --- black hole physics ---
42: hydrodynamics --- instabilities}
43: 
44: \section{Introduction}
45: The radiation pressure-supported inner region of geometrically
46: thin, optically thick Shakura-Sunyaev accretion disks (SSD) around
47: black holes \citep{SS1973} is known to be thermally unstable
48: \citep[e.g.][Chap. 4]{Kato98}, but the occurrence of an
49: instability does not necessarily mean that the disk will be
50: disrupted after the characteristic growth-time. A possible fate of
51: the thermally unstable inner region of SSDs is the so-called
52: limit-cycle behavior, i.e., the nonlinear oscillation between two
53: stable states. Similar to the case of dwarf novae, the limit-cycle
54: behavior was realized from a local and steady analysis
55: \citep[e.g.][Chap.5]{Kato98}, i.e., from an S-shaped sequence of
56: steady state solutions at a certain radius in the $\dot{M}-\Sigma$
57: (mass accretion rate vs. surface density) plane, with the lower
58: and middle branches of the S-shaped sequence corresponding to
59: stable gas pressure-supported SSD solutions and unstable radiation
60: pressure-supported SSD solutions, respectively, and the upper
61: branch corresponding to stable slim disk solutions constructed by
62: \citet{Abr88}; and has been justified by a number of works
63: performing global and time-dependent numerical computations
64: \citep{Honma91,Szuszkiewicz1997,Szuszkiewicz1998,Szuszkiewicz2001,Teresi04a,Teresi04b,Mayer06}.
65: Unlike the case of dwarf novae, however, only one astrophysical
66: object, the Galactic microquasar GRS 1915+105, has been known to
67: show the theoretically predicted limit-cyclic luminosity
68: variations
69: \citep{Nayakshin00,Janiuk02,Watarai03,Ohsuga06,Kawata06}.
70: 
71: We select the paper of \citet[][hereafter SM01]{Szuszkiewicz2001}
72: as the representative of existing theoretical works on the
73: limit-cycle behavior of black hole accretion disks with the
74: following two reasons. First, SM01 adopted a diffusion-type
75: prescription for viscosity, i.e., the $r\phi$ component of the
76: viscous stress tensor is expressed as
77: \begin{equation}\label{standar}
78: \tau_{r\phi}=\alpha H c_s\rho r \frac{\partial \Omega}{\partial
79: r},
80: \end{equation}
81: where $\rho$ is the density, $\Omega$ is the angular velocity,
82: $c_s$ is the sound speed, $H$ is the half-thickness of the disk,
83: and $\alpha$ is a dimensionless constant parameter; whereas all
84: other relevant works used a simple prescription
85: \begin{equation}\label{alphap}
86: \tau_{r\phi}=-\alpha p,
87: \end{equation}
88: where $p$ is the pressure, and $\alpha$ is also a dimensionless
89: constant but has been rescaled (denoting $\alpha$ in expressions
90: [\ref{standar}] and [\ref{alphap}] as $\alpha_1$ and $\alpha_2$,
91: respectively, then $\alpha_2=[3\sqrt{6}/2]\alpha_1$). It is known
92: that the direct integration of the differential equations
93: describing transonic accretion disks with the diffusive form of
94: viscosity is extremely difficult, while that with the $\alpha p$
95: viscosity prescription becomes much easier (see the discussion in
96: SM01). It should be noted, however, that expression (\ref{alphap})
97: is only an approximation of expression (\ref{standar}) under a
98: number of conditions \citep[including assuming that the disk is
99: stationary, geometrically thin, Newtonian Keplerian rotating, and
100: in vertical hydrostatic equilibrium, e.g.][Chap. 3]{Kato98}. More
101: seriously, as shown recently by \citet{Becker05}, expression
102: (\ref{standar}) is the only one proposed so far that is physically
103: consistent close to the black hole event horizon because of its
104: diffusive nature, whereas expression (\ref{alphap}) as well as
105: some other viscosity prescriptions would imply an unphysical
106: structure in the inner region of black hole accretion disks.
107: Second, SM01 did complete very nice numerical computations, all
108: the curves in their figures showing the evolution of disk
109: structure are perfectly continuous and well-resolved on the grid;
110: while some fluctuations appear on the curves in the figures of
111: other relevant works, which might make one to worry whether there
112: had been some hidden numerical instabilities in the code.
113: 
114: As evidenced by SM01, thermally unstable accretion disks undergo
115: limit-cycles when viscosity is moderate, i.e., the viscosity
116: parameter $\alpha\sim0.1$  (hereafter all the numerical values of
117: $\alpha$ are for $\alpha_2$ unless otherwise specified); and the
118: instability seems to be catastrophic when viscosity is weak, i.e.,
119: $\alpha\sim0.001$. On the other hand, in the case of very strong
120: viscosity, i.e., $\alpha\sim1$, \citet{Chen95} found that the
121: S-shaped sequence of steady state solutions in the
122: $\dot{M}-\Sigma$ plane does not form, instead, slim disk solutions
123: and optically thin advection-dominated accretion flow (ADAF)
124: solutions \citep{NY94,Abr95} are combined into a single straight
125: line. Accordingly, \citet{TM98} performed time-evolutionary
126: computations using the $\alpha p$ viscosity prescription with
127: $\alpha=1$ and proposed another possible fate of thermally
128: unstable accretion disks: the very inner region of the disk
129: finally becomes to be an ADAF-like feature, while the outer region
130: keeps being the SSD state, forming a persistent two-phased
131: structure. While this result is really interesting since a
132: phenomenological SSD+ADAF model has been quite successfully
133: applied to black hole X-ray binaries and galactic nuclei
134: \citep[e.g.,][]{Narayan98}, SM01 stated that they could not make
135: computations for $\alpha=1$ because of difficulties in keeping
136: their code numerically stable, and pointed out that it is worth
137: checking whether the persistent ADAF feature obtained in
138: \citet{TM98} would survive changing the viscosity prescription to
139: the diffusive form.
140: 
141: We purpose to study thermally unstable accretion disks if they are
142: not disrupted by instabilities, that is, we wish to check whether
143: the limit-cycle behavior is the only possible fate of these disks
144: provided viscosity is not too weak, or a transition from the SSD
145: state to the ADAF state is the alternative. As in SM01, we adopt
146: the diffusive viscosity prescription of equation (\ref{standar})
147: and make spatially 1.5-dimensional, time-dependent computations.
148: But we choose a numerical method that is different from either of
149: SM01 or of \citet{TM98}, and that is the adaptive pseudo-spectral
150: domain decomposition method. With this method, we hope to be able
151: to perform computations for various values of $\alpha$ ranging
152: from $\sim0.1$ to $\sim 1$, and to obtain numerical results at the
153: quality level of SM01. In this paper, we describe our numerical
154: algorithm and techniques in details and present computational
155: results for $\alpha=0.1$ as a test of our algorithm. We wish to
156: report our results for larger values of $\alpha$ in a subsequent
157: paper.
158: 
159: \section{Numerical Algorithm}
160: As the main intention of this paper, in this section we present a
161: numerical algorithm to solve a partial differential equation (or
162: equations) in the general form
163: \begin{equation}\label{pde}
164:     \frac{\partial u(r,t)}{\partial t}=L(u(r,t)),\quad
165:     r\in[r_{min},r_{max}],
166: \end{equation}
167: where $u(r,t)$ is a physical quantity that is a function of the
168: spatial independent variable $r$ (e.g., the radius in the
169: cylindrical coordinate system) and the time $t$, and $L$ is a
170: partial differential operator of $r$ and can be linear or
171: nonlinear.
172: 
173: \subsection{Scheme of Spacial Discretization}\label{psedo}
174: We first describe the standard Chebyshev pseudo-spectral method
175: that is used to discretize the spatial differential operator $L$.
176: This method has been explained in several textbooks
177: \citep{gottlieb1983,canuto1988,boyd2000,peyret2002}. Recently,
178: \citet{Chan2005,Chan2006} applied it to studies of astrophysical
179: accretion flows and discussed its advantages.
180: 
181: Concretely, a series with finite terms is used to approximate a
182: physical quantity $u(r)$ as
183: \begin{equation}\label{appseries}
184:     u(r_k)=u[g(\bar{r}_k)]=\sum\limits_{n=0}^N\hat{u}_nT_n(\bar{r}_k)=\sum\limits_{n=0}^N\hat{u}_n\cos\left(\frac{nk\pi}{N}\right),
185: \end{equation}
186: where $T_n(\bar{r}_k)$ is the $n$-th order Chebyshev polynomial;
187: $\bar{r}_k$ ($k=0, 1, 2, ... N$) is the Chebyshev-Gauss-Lobatto
188: collocation points and is defined as $\bar{r}_k\equiv
189: \cos(k\pi/N)$, with $N$ being the number of collocation points;
190: $r_k=g(\bar{r}_k)$ is the mapping from the Chebyshev-Gauss-Lobatto
191: collocation points $\bar{r}_k\in[-1,1]$ to the physical
192: collocation points $r_k\in[r_{min},r_{max}]$ that is a strictly
193: increasing function and satisfies both $g(-1)=r_{min}$ and
194: $g(1)=r_{max}$; $\hat{u}_n$ is the spectral coefficients and can
195: be calculated from the physical values $u(r_k)$ by a fast discrete
196: cosine transform \citep[hereafter FDCT,][Chap. 12]{Press1992};
197: contrarily, if one has $\hat{u}_n$, then $u(r_k)$ can be obtained
198: immediately by a inverted FDCT.
199: 
200: The radial derivative $\partial u(r)/\partial r$ is also a
201: function of $r$ and in principle can also be approximated by a
202: series that is obtained by using the chain rule
203: \begin{equation}\label{chain}
204:    \frac{\partial u(r_k)}{\partial
205:    r}=\frac{1}{dg/d\bar{r}}\frac{\partial u[g(\bar{r}_k)]}{\partial
206:    \bar{r}}=\frac{1}{dg/d\bar{r}}\sum\limits_{n=0}^N\hat{u}_n^{'}T_n(\bar{r}_k).
207: \end{equation}
208: The spectral coefficients $\hat{u}_n^{'}$ can be calculated from
209: $\hat{u}_n$ by a three-term recursive relation
210: \begin{eqnarray}\label{three-term}
211:   \nonumber \hat{u}_N^{'} &=& 0, \\
212:   \nonumber \hat{u}_{N-1}^{'} &=& 2N\hat{u}_N, \\
213:   c_n \hat{u}_n^{'} &=& \hat{u}_{n+2}^{'}+2(n+1)\hat{u}_{n+1},
214: \end{eqnarray}
215: where $c_0=2$, and $c_n=1$ for $n=1,2,...,N$. Subsequently,
216: $\partial u[g(\bar{r}_k)]/\partial\bar{r}$ is calculated from
217: $\hat{u}_n^{'}$ by a inverted FDCT, and then substituted into
218: equation (\ref{chain}) to obtain discrete spatial derivatives
219: $\partial u(r_k)/\partial r$.
220: 
221: To summarize, we define a discretized differential operator $D$
222: for the continuous differential operator $\partial/\partial r$.
223: The operator $D$ carries out the following works: (1) using FDCT
224: to calculate $\hat{u}_n$ from $u(r_k)$; (2) using the three-term
225: recursive relation equation (\ref{three-term}) to obtain
226: $\hat{u}_n^{'}$ from $\hat{u}_n$; (3) using a inverted FDCT and
227: equation (\ref{chain}) to obtain $\partial u(r_k)/\partial r$.
228: Finally, we use $D$ to construct a discretized operator
229: $\tilde{L}$ to approximate the operator $L$ in equation
230: (\ref{pde}). For example, if
231: $L\equiv\partial_r(u_1\partial_ru_2)$, where $\partial_r$ denotes
232: $\partial/\partial r$, then $\tilde{L}$ can be constructed as
233: $\tilde{L}=D[u_1D(u_2)]$.
234: 
235: \subsection{Scheme of Time-Discretization}\label{time-discre}
236: We adopt two schemes to perform the time-integration, that is, we
237: use a third order total variation diminishing (TVD) Runge-Kutta
238: scheme \citep{shu1988} to integrate the first two time-steps, and
239: then change to a low CPU-consumption scheme, the so-called third
240: order backward-differentiation explicit scheme
241: \citep[][pp.130-133]{peyret2002}, to carry out the rest
242: computations.
243: 
244: The third order TVD Runge-Kutta scheme is expressed as
245: \begin{eqnarray}\label{RK3}
246:   \nonumber u^{(1)}&=& u^n+\Delta t \tilde{L}(u^n),\\
247:   \nonumber u^{(2)}&=& \frac{3}{4}u^n+\frac{1}{4}u^{(1)}+\frac{1}{4}\Delta t\tilde{L}(u^{(1)}),\\
248:   u^{n+1}&=& \frac{1}{3}u^n+\frac{2}{3}u^{(2)}+\frac{2}{3}\Delta
249:   t\tilde{L}(u^{(2)}),
250: \end{eqnarray}
251: where $\Delta t$ is the time-step; $u^n$ and $u^{n+1}$ are the
252: values of the physical quantity $u$ at the $n$-th and $(n + 1)$-th
253: time-levels, respectively; and $u^{(1)}$ and $u^{(2)}$ are two
254: temporary variables.
255: 
256: The third order backward-differentiation explicit scheme can be
257: written as
258: \begin{equation}\label{BDE3-1}
259:     \frac{1}{\Delta t}\sum\limits_{j=0}^{3} a_j
260:     u^{n+1-j}=\sum\limits_{j=0}^{2} b_j \tilde{L}(u^{n-j}),
261: \end{equation}
262: where
263: \begin{eqnarray}\label{BDE3-2}
264:   \nonumber a_0&\equiv& 1+\frac{1}{1+k_n}+\frac{1}{1+k_n+k_{n-1}},\\
265:   \nonumber a_1&\equiv& -\frac{(1+k_n)(1+k_n+k_{n-1})}{k_n(k_n+k_{n-1})},\\
266:   \nonumber a_2&\equiv& \frac{1+k_n+k_{n-1}}{k_nk_{n-1}(1+k_n)},\\
267:  a_3&\equiv&
268:  -\frac{1+k_n}{k_{n-1}(k_n+k_{n-1})(1+k_n+k_{n-1})};\\
269:  \nonumber\\
270:   \nonumber b_0&\equiv& \frac{(1+k_n)(1+k_n+k_{n-1})}{k_n(k_n+k_{n-1})},\\
271:   \nonumber b_1&\equiv& -\frac{1+k_n+k_{n-1}}{k_nk_{n-1}},\\
272:   b_2&\equiv& \frac{1+k_n}{k_{n-1}(k_n+k_{n-1})};
273: \end{eqnarray}
274: and
275: \begin{eqnarray}\label{BDE3-3}
276:   \nonumber k_n&\equiv& \frac{t^n-t^{n-1}}{\Delta t},\\
277:   k_{n-1}&\equiv& \frac{t^{n-1}-t^{n-2}}{\Delta t};
278: \end{eqnarray}
279: with $t^n$, $t^{n-1}$, and $t^{n-2}$ being the times of the
280: $n$-th, $(n-1)$-th, and $(n-2)$-th time-levels, respectively.
281: 
282: Of these two time-integration schemes, the former spends three
283: times of the latter's CPU-time per time-step, but the latter is
284: not able to start the time-integration by itself while the former
285: is able to do. Therefore, we combine these two schemes in order to
286: achieve a sufficient high order accuracy with minimal CPU-time
287: consumption.
288: 
289: Hereto, we have fully discretized equation (\ref{pde}). In order
290: to obtain a physically sound and numerically stable solution in a
291: finite domain, it is additionally necessary to impose appropriate
292: boundary conditions and to apply some filtering techniques to
293: overcome the inevitable spurious nonlinear numerical instabilities
294: in the code. We leave the details of these in the Appendix.
295: 
296: \subsection{Domain Decomposition}\label{domain-decom}
297: The numerical algorithm described in the above two subsections and
298: the Appendix has been a useful implement for solving partial
299: differential equations and is essentially what was adopted in
300: \citet{Chan2005}. However, it turns out that, as we have
301: experienced in our computations, the above algorithm is
302: insufficient for resolving the so-called stiff problem. This
303: problem is a one whose solution is characterized by two or more
304: space-scales and/or time-scales of different orders of magnitude.
305: In the spatial case, common stiff problems in fluid mechanics are
306: boundary layer, shear layer, viscous shock, interface, flame
307: front, etc. In all these problems there exists a region (or exist
308: regions) of small extent (with respect to the global
309: characteristic length) in which the solution exhibits a very large
310: variation \citep[][p. 298]{peyret2002}. When the Chebyshev
311: pseudo-spectral method described in \S\ref{psedo} is applied to a
312: stiff problem, the accuracy of the method can be significantly
313: degraded and there may appear spurious oscillations which can lead
314: to nonlinear numerical instabilities or spurious predictions of
315: solution behavior \citep[the so-called Gibbs
316: phenomenon,][]{gottlieb1997}. The spectral filtering technique
317: described in the Appendix is not able to completely remove these
318: spurious oscillations, so that the solution is still not well
319: resolved, and sometimes the computation can even be destroyed by
320: the growing spurious oscillations. A special method that has been
321: developed to overcome these difficulties is the domain
322: decomposition \citep{bayliss1995,peyret2002}. Here we mainly
323: follow \citet{bayliss1995} to use this method, but with a
324: different technique to connect the decomposed subdomains.
325: 
326: The basic idea of domain decomposition is to divide a wide
327: computational domain into a set of subdomains, so that each
328: subdomain contains at most only one single region of rapid
329: variation (i.e., with a stiff problem), and more grid points are
330: collocated into this region by a special mapping function to
331: enhance the resolution while the total consumption of CPU-time is
332: not substantially increased.
333: 
334: In each subdomain the solution is obtained by taking into account
335: some connection conditions at the interface between the two
336: conjoint subdomains. In general, appropriate connection conditions
337: are the continuities of both the solution and its spatial
338: derivative normal to the interface \citep{bayliss1995,peyret2002}.
339: The continuity of the solution is satisfied naturally, but the
340: continuity of its derivative cannot be achieved directly with the
341: pseudo-spectral method because of the use of FDCT. To see this,
342: let us divide the entire computational domain
343: $r\in[r_{min},r_{max}]$ into $M$ subdomains,
344: \begin{equation}
345:   S_i\equiv [r_{min}^{(i)},r_{max}^{(i)}],\ \ i=1,2,...,M,
346: \end{equation}
347: where $r_{min}^{(1)}=r_{min},\ r_{max}^{(1)}=r_{min}^{(2)},\
348: r_{max}^{(2)}=r_{min}^{(3)},\ ...$ and $r_{max}^{(M)}=r_{max}$ are
349: the locations of the interfaces between the subdomains. Because
350: FDCT is used to calculate the numerical derivative in each
351: subdomain $S_i$, one obtains two values of the derivative at each
352: interface. Let $\partial^{-}u$ and $\partial^{+}u$ denote the left
353: and right numerical derivatives of the physical quantity $u$ at a
354: certain interface, respectively, a seemingly rational choice for
355: keeping the continuity of derivative is to set the numerical
356: derivative at the interface to be the mean of $\partial^{-}u$ and
357: $\partial^{+}u$, i.e.,
358: \begin{equation}\label{connect}
359:    \left(\frac{\partial u}{\partial r}\right)_{interface}=\frac{\partial^{-}u+\partial^{+}u}{2}.
360: \end{equation}
361: Unfortunately, in practice the connection technique of equation
362: (\ref{connect}) will often cause a numerical instability at the
363: interfaces.
364: 
365: We find that the connection between two certain subdomains $S_i$
366: and $S_{i+1}$ can be numerically stable when their discretizations
367: satisfy an additional practical condition. Let
368: $r_{int}$($=r_{max}^{(i)}=r_{min}^{(i+1)}$) denotes the location
369: of interface between $S_i$ and $S_{i+1}$; $r_{N-1}^{(i)}$ and
370: $r_1^{(i+1)}$ ($r_{N-1}^{(i)}\in S_i$, $r_1^{(i+1)}\in S_{i+1}$
371: and $r_{N-1}^{(i)}<r_{int}<r_1^{(i+1)}$) denote the locations of
372: the two nearest points to the interface, respectively; our
373: computations show that if the condition
374: \begin{equation}\label{connect2}
375:    \left|r_{int}-r_{N-1}^{(i)}\right|=\left|r_{int}-r_1^{(i+1)}\right|
376: \end{equation}
377: is satisfied, then the connection of derivative represented by
378: equation (\ref{connect}) will be numerically stable.
379: 
380: If the stiff problem always appeared in a fixed spatial region,
381: then the domain decomposition would be kept unchanged. However, in
382: general this is not the case. Instead, the location of region in
383: which the stiff problem appears changes with time
384: \citep{bayliss1995}. Therefore, the domain decomposition must be
385: adjusted adaptively. To ensure the connection condition equation
386: (\ref{connect2}) at the interfaces of newly divided subdomains, an
387: adjustable mapping between the physical collocation points $r_k$
388: in each new subdomain $S_i$ and the Chebyshev-Gauss-Lobatto
389: collocation points $\bar{r}_k$ is needed. We adopt such a mapping
390: in the form (see eq.[\ref{appseries}])
391: \begin{equation}\label{adjmapp}
392:    r^{(i)}=g(\bar{r})=r_{max}^{(i)}+\frac{2}{\pi}\left(r_{max}^{(i)}-r_{min}^{(i)}\right)\arctan\left[a
393:    \tan\frac{\pi}{4}(\bar{r}-1)\right],\ \ \bar{r}\in [-1,1]\ \mathrm{and}\ r^{(i)}\in
394:    S_i,
395: \end{equation}
396: in the subdomain $S_i$, which is a combination of two mapping
397: functions,
398: \begin{equation}\label{trivialmap}
399:    r^{(i)}=\frac{r_{max}^{(i)}}{2}(\tilde{r}+1)-\frac{r_{min}^{(i)}}{2}(\tilde{r}-1)
400: \end{equation}
401: and
402: \begin{equation}\label{adjmapp2}
403:    \tilde{r}=\frac{4}{\pi}\arctan\left[a
404:    \tan\frac{\pi}{4}(\bar{r}-1)\right].
405: \end{equation}
406: Equation (\ref{trivialmap}) is a trivial linear mapping
407: \citep{Chan2005}, and equation (\ref{adjmapp2}) is the mapping
408: proposed by \citet{bayliss1995}. The parameter $a$ in equation
409: (\ref{adjmapp2}) is an adjustable parameter, but equation
410: (\ref{adjmapp2}) is only a mapping from $\bar{r}\in[-1,1]$ to
411: $\tilde{r}\in[-1,1]$. Therefore, we add equation
412: (\ref{trivialmap}) in order to make a complete mapping from
413: $\bar{r}\in[-1,1]$ to $r^{(i)}\in S_i$. The combined mapping
414: equation (\ref{adjmapp}) will concentrate the discrete grid points
415: toward $r_{min}^{(i)}$ when $a>1$ and toward $r_{max}^{(i)}$ when
416: $a<1$, and will be reduced to equation (\ref{trivialmap}) when
417: $a=1$.
418: 
419: The adjustability of mapping equation (\ref{adjmapp}) is crucially
420: important for achieving a numerically stable connection at the
421: interfaces of subdomains. By substituting equation (\ref{adjmapp})
422: into equation (\ref{connect2}), we obtain
423: \begin{equation}\label{para-a}
424:    a^{(i+1)}=\frac{\cot\left\{\omega\arctan\left[a^{(i)}\tan\frac{\pi}{4}(\bar{r}_l-1)\right]\right\}}{\tan\frac{\pi}{4}(\bar{r}_r-1)},
425: \end{equation}
426: with
427: \begin{eqnarray}
428:   \nonumber
429:   \omega&\equiv&\frac{r_{max}^{(i)}-r_{min}^{(i)}}{r_{max}^{(i+1)}-r_{min}^{(i+1)}},\\
430:   \nonumber \bar{r}_l&\equiv&\cos\left(\frac{\pi}{N}\right),\\
431:   \nonumber \bar{r}_r&\equiv&\cos\left[\frac{(N-1)\pi}{N}\right],
432: \end{eqnarray}
433: where $a^{(i)}$ and $a^{(i+1)}$ are the mapping parameters for
434: subdomains $S_i$ and $S_{i+1}$, respectively. Equation
435: (\ref{para-a}) can be used to determine the mapping parameters of
436: every subdomain after giving a decomposition of computational
437: domain $\{S_i\}$ and the mapping parameter $a^{(1)}$ of the
438: innermost subdomain $S_1$ ($=[r_{min},r_{max}^{(1)}]$). As a
439: result, we obtain a particular collocation of discrete grid-points
440: within the whole computational domain $[r_{min},r_{max}]$. This
441: collocation ensures a stable connection of the derivatives between
442: any two conjoint subdomains (eq.[\ref{connect}]), and thus ensures
443: a correct implementation of the pseudo-spectral method in each
444: subdomain. The combination of the standard pseudo-spectral method
445: and the adaptive domain decomposition method finally names our
446: numerical algorithm as that in the title of this paper.
447: 
448: \section{Limit-cycle Solutions}
449: We now verify our numerical algorithm by applying it to studies of
450: thermally unstable black hole accretion disks with moderate
451: viscosity and comparing our results with that of the
452: representative work SM01.
453: 
454: \subsection{Basic Equations}
455: We write basic equations for viscous accretion flows around black
456: holes in the Eulerian form rather than Lagrangian form as in SM01
457: because partial differential equations in the Eulerian description
458: take the general form of equation (\ref{pde}), to which our
459: numerical algorithm is suited. The basic equations to be solved
460: are
461: \begin{eqnarray}
462:   \frac{\partial \Sigma}{\partial t}&=&-v_r\frac{\partial \Sigma}{\partial
463:    r}-\frac{\Sigma}{r}\frac{\partial}{\partial
464:    r}(rv_r),\label{eq-mass}\\
465:   \frac{\partial v_r}{\partial t}&=&-v_r\frac{\partial v_r}{\partial r}-\frac{1}{\rho}
466:     \frac{\partial p}{\partial r}+\frac{l^2-l_K^2}{r^3},\label{eq-radial}\\
467:   \frac{\partial l}{\partial t}&=&-v_r\frac{\partial l}{\partial r}+\frac{\alpha}{r\Sigma}
468:     \frac{\partial}{\partial r}\left(r^3c_sH\Sigma\frac{\partial\Omega}{\partial r}\right),\label{eq-angular}\\
469:   \frac{\partial H}{\partial t}&=&-v_r\frac{\partial H}{\partial r}+V_z,\label{eq-height}\\
470:   \frac{\partial V_z}{\partial t}&=&-v_r\frac{\partial V_z}{\partial r}+6\frac{p}{\Sigma}-\frac{GM_{BH}}{(r-r_g)^2}\left(\frac{H}{r}\right),\label{eq-vertical}\\
471:   \nonumber\frac{\partial T}{\partial t}&=&-v_r\frac{\partial T}{\partial
472:   r}\\
473:    & &+\frac{T}{12-10.5\beta}\left\{\frac{\alpha\Sigma
474: c_sH(r\partial\Omega/\partial
475:   r)^2-F^{-}}{0.67pH}-(4-3\beta)\left[\frac{V_z}{H}+\frac{1}{r}\frac{\partial}{\partial
476:   r}(rv_r)\right]\right\}.\label{eq-energy}
477: \end{eqnarray}
478: 
479: Equations (\ref{eq-mass}), (\ref{eq-radial}), (\ref{eq-angular}),
480: and (\ref{eq-energy}) are conservations of mass, radial momentum,
481: angular momentum, and energy, respectively. As in SM01, we adopt
482: the diffusive form of viscosity (eq. [\ref{standar}]) in equations
483: (\ref{eq-angular}) and (\ref{eq-energy}); and abandon vertical
484: hydrostatic equilibrium assumed in 1-dimensional studies, and
485: instead introduce two new dynamical equations for the vertical
486: acceleration (eq. [\ref{eq-vertical}]) and the evolution of the
487: disk's thickness (eq. [\ref{eq-height}]), thus our studies can be
488: regarded as 1.5-dimensional. In these equations $v_r$, $l$, $l_K$,
489: $V_z$, $M_{BH}$, $r_g$, $T$, $\beta$, and $F^{-}$ are the radial
490: velocity, specific angular momentum, Keplerian specific angular
491: momentum, vertical velocity at the surface of the disk, black hole
492: mass, gravitational radius ($\equiv 2GM_{BH}/c^2$), temperature,
493: ratio of gas pressure to total pressure, and radiative flux per
494: unit area away from the disk in the vertical direction,
495: respectively. We use the 'one-zone' approximation of the
496: vertically-averaged disk as in SM01, so that $v_r$, $\Omega$, $l$,
497: $l_K$, $\rho$, $p$, $c_s$, and $T$ are all equatorial plane
498: quantities, while $V_z$ and $F^{-}$ are quantities at the disk's
499: surface. Additional definitions and relations of these quantities
500: are
501: \begin{eqnarray}
502:    \rho&=&\frac{\Sigma}{H},\\
503:    l_K&=&\sqrt{\frac{GM_{BH}r^3}{(r-r_g)^2}},\\
504:    c_s&=&\sqrt{\frac{p}{\rho}},\\
505:    \Omega&=&\frac{l}{r^2},\\
506:    F^{-}&=&\frac{24\sigma T^4}{3\tau_R/2+\sqrt{3}+1/\tau_P},\label{eq-bridge}\\
507:    p&=&k\rho T+p_{rad},\\
508:    p_{rad}&=&\frac{F^{-}}{12c}\left(\tau_R+\frac{2}{\sqrt{3}}\right),\\
509:    \tau_R&=&0.34\Sigma(1+6\times10^{24}\rho T^{-3.5}),\\
510:    \tau_P&=&\frac{1.24\times10^{21}\Sigma\rho T^{-3.5}}{4\sigma},\\
511:    \beta&=&\frac{k\rho T}{p},
512: \end{eqnarray}
513: where equation (\ref{eq-bridge}) is a bridging formula that is
514: valid for both optically thick and thin regimes, $p_{rad}$ is the
515: radiation pressure, and $\tau_R$ and $\tau_P$ are the Rosseland
516: and Planck mean optical depths.
517: 
518: \subsection{Specific Techniques}
519: As in SM01, in our code the inner edge of the grid is set at
520: $r\simeq2.5r_g$, close enough to the central black hole so that
521: the transonic point can be included in the solution; and the outer
522: boundary is set at $r\simeq 10^4r_g$, far enough away so that no
523: perturbation from the inner regions could reach. A stationary
524: transonic disk solution calculated with the $\alpha p$ viscosity
525: prescription is used as the initial condition for the evolutionary
526: computations. The $\alpha p$ viscosity prescription may seem
527: inconsistent with the evolutionary computations adopting the
528: diffusive viscosity prescription, but this does not matter. In
529: fact, the initial condition affects only the first cycle of the
530: obtained limit-cycle solutions, and all following cycles are
531: nicely regular and repetitive. The time-step $\Delta t$ is
532: adjusted also in the same way as in SM01 to maintain numerical
533: stability. We emphasize some techniques specific to our numerical
534: algorithm below.
535: 
536: The solution to be obtained covers so wide a range of the whole
537: computational domain, and in particular, the thermal instability
538: causes a violent variation of the solution (the stiff problem) in
539: the inner region (inside $200r_g$) of the domain. In such
540: circumstances the standard one-domain spectral method is certainly
541: insufficient. Accordingly, we divide the whole computational
542: domain into $6$ subdomains and let each subdomain contain $65$
543: grid points, so the total number of grid points is
544: $65\times6-5=385$ (there are $5$ overlapping point at the
545: interfaces of subdomains). At each time-level we apply the
546: one-domain spectral method described in \S\ref{psedo} into each
547: subdomain. In doing this, various techniques are used to remove or
548: restrain spurious non-linear oscillations and to treat properly
549: the boundary conditions, in order to have a numerically stable
550: solution in each subdomain as well as a stable connection at each
551: interface. Then we use the scheme described in \S\ref{time-discre}
552: to carry out the time-integration over the time-step $\Delta t$ to
553: reach the next time-level.
554: 
555: In general, spurious oscillations are caused by three factors: the
556: aliasing error, the Gibbs phenomenon, and the absence of viscous
557: stress tensor components in the basic equations. The aliasing
558: error is a numerical error specific to the spectral method when it
559: is used to solve differential equations that contain non-linear
560: terms, and can be resolved by the spectral filtering technique
561: described in Appendix (see \citealt{peyret2002} for a detailed
562: explanation and \citealt{Chan2005} for a quite successful
563: application of this technique).
564: 
565: However, the spectral filtering itself cannot resolve the Gibbs
566: phenomenon characteristic to the stiff problem, and the adaptive
567: domain decomposition method described in \S\ref{domain-decom}
568: becomes crucially important. In our computations, we set the
569: spatial location where a large variation appears (e.g., the peak
570: of the surface density $\Sigma$) as the interface of two certain
571: subdomains and use the mapping equation (\ref{adjmapp}) along with
572: the connection conditions (\ref{connect}) and (\ref{connect2}), so
573: that more grid points are concentrated on the two sides of the
574: interface to enhance the resolution, and a stable connection
575: between the two subdomains is realized. As the location of large
576: variation is not fixed and instead shifts during time-evolution,
577: we follow this location, redivide the computational domain and
578: recalculate the mapping parameter for each new subdomain with
579: equation (\ref{para-a}). In practice, the stiff problem appears
580: and its location shifts during the first about $30$ seconds of
581: each cycle whose whole length is $\sim700$ seconds. In these $30$
582: seconds we have to redivide the domain and reset the grid after
583: every interval of $0.1$ seconds (or every $0.01$ seconds for the
584: first a few seconds), and the typical length of time-step $\Delta
585: t$ is about $10^{-6}-10^{-7}$ seconds. For the rest time of a
586: cycle (more than $600$ seconds), the stiff problem ceases, then
587: the grid reset is not needed and the time-step can be much longer.
588: 
589: In addition to the above two factors in the numerical sense, there
590: is a factor in the physical sense that can also cause spurious
591: oscillations. The viscous stress tensor has nine spatial
592: components, but in 1-dimensional studies usually only the $r\phi$
593: component is included in the basic equations, and omitting other
594: components can result in numerical instabilities. In particular,
595: \citet{Szuszkiewicz1997} noted that if the tensor components
596: $\tau_{rr}$ and $\tau_{\phi\phi}$ were neglected from the radial
597: momentum equation, some instability would develop and cause
598: termination of the computation because of the non-diffusive nature
599: of the equation; and that the instability was suppressed when a
600: low level of numerical diffusion was added artificially into the
601: equation. In our code we resolve the similar problem in a slightly
602: different way. We add into the radial momentum equation
603: (\ref{eq-radial}) two viscous forces $F_{rr}$ and $F_{\phi\phi}$
604: in the form as given by equation (5) of \citet{Szuszkiewicz1997},
605: and accordingly add into the energy equation (\ref{eq-energy}) a
606: heating term due to viscous friction in the radial direction as
607: given by equation (15) of \citet{Szuszkiewicz1997}. We find by
608: numerical experiments that when a very small viscosity in the
609: radial direction is introduced, i.e., with the radial viscosity
610: parameter $\alpha_r\simeq0.05\alpha$, where $\alpha$ is the
611: viscosity parameter in the $\alpha p$ viscosity prescription, the
612: spurious oscillations due to the absence of viscous stress tensor
613: components disappear and the solution keeps nicely stable.
614: 
615: Solving partial differential equations in a finite domain usually
616: requires the Dirichlet boundary condition, i.e., changing the
617: values of physical quantities at the boundary points to be their
618: boundary values supplied a priori, or/and the Neumann boundary
619: condition, i.e., adjusting the values of derivatives of physical
620: quantities at the boundary points to be their boundary values
621: supplied a priori. For spectral methods, \citet{Chan2005}
622: introduced a new treatment, namely the spatial filter as described
623: in the Appendix, in order to ensure that the Dirichlet or/and the
624: Neumann boundary condition can be imposed and numerical
625: instabilities due to boundary conditions are avoided. We note,
626: however, that their spatial filter treatment is applicable only
627: for physical quantities whose boundary derivatives are equal, or
628: approximately equal, to zero (e.g., the radial velocity $v_r$, its
629: spatial derivative at the outer boundary is nearly zero); while
630: for quantities whose boundary derivatives are not negligible
631: (e.g., the specific angular momentum $l$, its spatial derivative
632: at the outer boundary is not very small), the boundary treatment
633: in finite difference methods still works, i.e., to let directly
634: the values of those quantities at the boundary points be the
635: specified boundary values.
636: 
637: SM01 supplied boundary conditions at both the inner and outer
638: edges of the grid, i.e., at the inner edge
639: $Dl/Dt=(\partial/\partial r+v_r\partial/\partial r)l=0$ and
640: $\partial p/\partial r=0$, and at the outer edge both $v_r$ and
641: $l$ are constant in time and $l$ is slightly smaller than the
642: corresponding $l_K$. In our computations we find that it is not
643: necessary to supply a priori the two inner boundary conditions as
644: in SM01. With the two outer boundary conditions as in SM01 and
645: free inner boundary conditions instead, we are able to obtain
646: numerically stable solutions of the basic equations, and these
647: solutions automatically lead to an almost zero viscous torque,
648: i.e., $Dl/Dt\simeq0$, and a nearly vanishing pressure gradient,
649: i.e., $\partial p/\partial r\simeq0$, in the innermost zone. This
650: proves that the inner boundary conditions assessed in SM01 are
651: correct, but we think that our practice is probably more natural
652: and more physical. Once the state of an accretion flow at the
653: outer boundary is set, the structure and evolution of the flow
654: will be controlled by the background gravitational field of the
655: central black hole, and the flow should adjust itself in order to
656: be accreted successfully into the black hole. In particular, both
657: the viscous torque and the pressure gradient of the flow must
658: vanish on the black hole horizon, and in the innermost zone a
659: correct solution should asymptotically approach to such a state,
660: thus in the computations no inner boundary conditions are
661: repeatedly needed.
662: 
663: \subsection{Numerical Results}
664: We have performed computations for a model accretion disk with
665: black hole mass $M_{BH}=10M_{\odot}$, initial accretion rate
666: $\dot{m}\equiv\dot{M}/\dot{M}_{cr}=0.06$ ($\dot{M}=-2\pi r \Sigma
667: v_r$ is the accretion rate and $\dot{M}_{cr}$ is the critical
668: accretion rate corresponding to the Eddington luminosity), and
669: viscosity parameter $\alpha=0.1$ (the equivalent $\alpha$ in the
670: diffusive viscosity prescription is
671: $0.1\times(2/3\sqrt{6})\simeq0.0272$). It is known that the inner
672: region of a stationary accretion disk with such physical
673: parameters is radiation pressure-supported ($\beta<0.4$) and is
674: thermally unstable, and the disk is expected to exhibit the
675: limit-cycle behavior \citep[][Chaps. 4 and 5]{Kato98}. We have
676: continued computations for several complete limit-cycles, and a
677: representative cycle is illustrated in Figures \ref{fig1} -
678: \ref{fig6}, which are for the time evolution of the radial
679: distribution of the half-thickness of the disk $H$, temperature
680: $T$, surface density $\Sigma$, effective optical depth
681: $\tau_{eff}=(2/3)(3\tau_R/2+\sqrt{3}+1/\tau_P)$, ratio of gas
682: pressure to total pressure $\beta$, and accretion rate $\dot{m}$,
683: respectively. Note that negative values of $\dot{m}$ signify an
684: outflow in the radial direction, not in the vertical direction as
685: the word 'outflow' in the literature usually means.
686: 
687: The first panel of Figure \ref{fig1} and the solid lines in
688: Figures \ref{fig2} - \ref{fig6} show the disk just before the
689: start of the cycle ($t=0s$). The disk is essentially in the SSD
690: state, i.e., it is geometrically thin ($H/r\ll1$) and optically
691: thick ($\tau_{eff}\gg1$) everywhere, its temperature has a peak at
692: $r\simeq6r_g$, its accretion rate is nearly constant in space, and
693: its inner region (from $\sim5r_g$ to $\sim14r_g$) has $\beta<0.4$
694: and is thermally unstable. Note that this configuration is not a
695: stationary state and is with the diffusive viscosity prescription,
696: so it is very different from the initial condition at the
697: beginning of the computation, which is a stationary solution with
698: the $\alpha p$ viscosity prescription.
699: 
700: As the instability sets in ($t=2s$, the second panel of Fig.
701: \ref{fig1} and the thin dashed lines in Figs. \ref{fig2} -
702: \ref{fig6}), in the unstable region ($r<24r_g$) the temperature
703: rises rapidly, the disk expands in the vertical direction, a very
704: sharp spike appears in the surface density profile and accordingly
705: in the optical depth and accretion rate profiles (exactly the
706: stiff problem). The spikes move outwards with time, forming an
707: expansion wave, heating the inner material and pushing it into the
708: black hole, and perturbing the outer material to departure from
709: the SSD state. The expansion region is in fact essentially in the
710: state of slim disk, as it is geometrically thick ($H/r\lesssim1$),
711: optically thick, very hot, and radiation pressure-supported
712: ($\beta<0.4$); and the front of the expansion wave forms the
713: transition surface between the SSD state and the slim disk state.
714: At $t=12s$ (the third panel of Fig. \ref{fig1} and the thin
715: dot-dashed lines in Figs. \ref{fig2} - \ref{fig6}), in the
716: expansion region $H$ and $\dot{m}$ (negative, radial outflow)
717: reach their maximum values, and the local $\dot{m}$ (positive,
718: inflow) exceeds $3$ which is far above the initial value
719: $\dot{m}=0.06$ and is even well above the critical value
720: $\dot{m}=1$.
721: 
722: Once the wavefront has moved beyond the unstable region
723: ($r\lesssim120r_g$), the expansion starts to weaken, the
724: temperature drops in the innermost part of the disk and the
725: material there deflates ($t=23s$, the fourth panel of Fig.
726: \ref{fig1} and the thick dashed lines in Figs. \ref{fig2} -
727: \ref{fig6}). Subsequently, deflation spreads out through the disk,
728: and the disk consists of three different parts: the inner part is
729: geometrically thin, with the temperature and surface density being
730: lower than their values at $t=0s$; the middle part is what remains
731: of the slim disk state; and the outer part is still basically in
732: the SSD state ($t=27s$, the fifth panel of Fig. \ref{fig1} and the
733: thick dot-dashed lines in Figs. \ref{fig2} - \ref{fig6}).
734: 
735: The 'outburst' part of the cycle ends when it has proceeded on the
736: thermal time-scale ($t=32s$, the sixth panel of Fig. \ref{fig1}
737: and the dotted lines of Figs. \ref{fig2} - \ref{fig6}). What
738: follows is a much slower process (on the viscous time-scale) of
739: refilling and reheating of the inner part of the disk. Finally
740: ($t=722s$, the seventh panel of Fig. \ref{fig1} and again the
741: solid lines of Figs. \ref{fig2} - \ref{fig6}), the disk returns to
742: essentially the same state as that at the beginning of the cycle.
743: Then the thermal instability occurs again and a new cycle starts.
744: 
745: The bolometric luminosity of the disk, obtained by integrating the
746: radiated flux per unit area $F^{-}$ over the disk at successive
747: times, is drawn in Figure \ref{fig7} for three complete cycles.
748: The luminosity exhibits a burst with a duration of about $20$
749: seconds and a quiescent phase lasting for the remaining about
750: $700$ seconds of the cycle. The amplitude of the variation is
751: around two orders of magnitude, during the outburst a
752: super-Eddington luminosity is realized.
753: 
754: All these results obtained with our numerical method are similar
755: to that of SM01, not only in the sense that the limit-cycle
756: behavior of thermally unstable accretion disks is confirmed, but
757: also in the sense that the numerical solutions are of very good
758: quality. In our computations we have been able to suppress all
759: numerical instabilities and to remove all spurious oscillations,
760: so that in our figures all curves are perfectly continuous and
761: smooth and all spikes are well-resolved.
762: 
763: What is new, however, is that we have also computed the Bernoulli
764: function (i.e., the specific total energy) of the accreted matter
765: that is expressed as (cf. Eq. [11.33] of \citealt{Kato98})
766: \begin{equation}\label{Bernu}
767: B=\left[3(1-\beta)+\frac{\beta}{\gamma-1}\right]\frac{p}{\rho}+\frac{1}{2}\left(
768: v_r^2+V_z^2+\Omega^2r^2\right)-\frac{GM_{BH}}{\sqrt{r^2+H^2}-r_g}
769: \end{equation}
770: where $\gamma$ is the specific heat ratio and is taken to be
771: $5/3$. Figure \ref{fig8} shows the quantity $B$ obtained in the
772: whole computational domain ranging from $r\simeq2.5r_g$ to
773: $r\simeq10^4r_g$. It is clear that $B$ has negative values for the
774: whole spatial range (approaching to zero for very large $r$) and
775: during the whole time of the cycle (in the figure the thick
776: dot-dashed line for $t=27s$ and the dotted line for $t=32s$ are
777: coincided with the solid line for $t=0s$). Note that in equation
778: (\ref{Bernu}) the vertical kinetic energy $0.5V_z^2$ is included,
779: and the gravitational energy is that for the surface of the disk.
780: If the vertical kinetic energy is omitted and the gravitational
781: energy is taken to be its equatorial plane value as in
782: 1-dimensional models, then $B$ will have even larger negative
783: values. This result is in strong support of the analytical
784: demonstration of \citet{Abr00} that accretion flows with not very
785: strong viscosity ($\alpha\lesssim0.1$) have a negative Bernoulli
786: function; and implies that outflows are unlikely to originate from
787: thermally unstable accretion disks we consider here, because a
788: positive $B$ is a necessary, though not a sufficient, condition
789: for the outflow production.
790: 
791: \section{Summary and Discussion}
792: We have introduced a numerical method for studies of thermally
793: unstable accretion disks around black holes, which is essentially
794: a combination of the standard one-domain pseudo-spectral method
795: \citep{Chan2005} and the adaptive domain decomposition method
796: \citep{bayliss1995}. As a test of our method, for the case of
797: moderate viscosity we have reproduced the best numerical results
798: obtained previously by SM01. Despite these similarities, we have
799: made the following improvements over previous works in the
800: numerical algorithm and concrete techniques, which have been
801: proven to be effective in the practice of our computations.
802: 
803: 1. In applying the domain decomposition method to resolve the
804: stiff problem, we develop a simple and useful connection technique
805: to ensure a numerically stable continuity for the derivative of a
806: physical quantity across the interface of two conjoint subdomains,
807: i.e., equations (\ref{connect}) and (\ref{connect2}), instead of
808: the connection technique in \citet{bayliss1995} that is seemingly
809: complicated and was not explicitly explained.
810: 
811: 2. We construct a mapping function (eq. [\ref{adjmapp}]) by adding
812: the simple linear mapping function (eq. [\ref{trivialmap}]) into
813: the adaptive mapping function (eq. [\ref{adjmapp2}]) proposed by
814: \citet{bayliss1995}, so that the mapping between the
815: Chebyshev-Gauss-Lobatto collocation points $\bar{r}_k$ and the
816: physical collocation points $r_k^{(i)}$, not only the mapping
817: between two sets of collocation points $\bar{r}_k$ and
818: $\tilde{r}_k$, is completed; and the adjustability of equation
819: (\ref{adjmapp2}) is kept to enable us to follow adaptively the
820: region that is with the stiff problem and is shifting in space
821: during time-evolution.
822: 
823: 3. For the time-integration, we use two complementary schemes,
824: namely the third order TVD Runge-Kutta scheme and the third order
825: backward-differentiation explicit scheme. The former scheme is
826: popular in one-domain spectral methods and is essentially what was
827: used by \citet{Chan2005}, and the latter one can achieve the same
828: accuracy and has advantage of lower CPU-time consumption.
829: 
830: 4. For the treatment of boundary conditions, we notice that the
831: spatial filter technique developed by \citet{Chan2005} for
832: spectral methods is useful but is itself alone insufficient, and
833: the treatment traditionally used in finite difference methods is
834: still needed to complement. We also find that once reasonable
835: conditions are set at the outer boundary, our solutions behave
836: themselves physically consistent close to the black hole horizon,
837: and no inner boundary conditions are necessary as supplied by
838: SM01.
839: 
840: 5. We resolve the problem of spurious oscillations due to the
841: absence of viscous stress tensor components in the basic equations
842: in a way different from that of SM01. SM01 introduced an
843: artificial viscous term in the radial and vertical momentum
844: equations. We instead improve the basic equations by including two
845: viscous force terms $F_{rr}$ and $F_{\phi\phi}$ in the radial
846: momentum equation and a corresponding viscous heating term in the
847: energy equation, all these terms were already mentioned by the
848: same authors of SM01 in an earlier paper \citep{Szuszkiewicz1997}.
849: As for the vertical momentum equation, because of its crudeness in
850: our $1.5$-dimensional studies, we still adopt an artificial term
851: whose explicit form is kindly provided by Szuszkiewicz \& Miller
852: and is unpublished. We obtain solutions at the same quality level
853: as in SM01, but we think that our treatment is probably more
854: physical in some sense. In particular, any modification in the
855: momentum equation ought to require a corresponding modification in
856: the energy equation, otherwise the energy conservation would not
857: be correctly described.
858: 
859: Of these five improvements, we expect that the first two and the
860: last one will be particularly helpful for our subsequent studies
861: of the strong viscosity case ($\alpha\sim1$). In this case the
862: viscous heating becomes extremely huge, the 'outburst' of the disk
863: due to the thermal instability is predicted to be more violent,
864: and the Gibbs phenomenon related to the stiff problem can be even
865: more serious than in the case of moderate viscosity studied in
866: this paper. Our improved domain decomposition method is prepared
867: to front these difficulties. As for another nettlesome problem
868: that the absence of some viscous stress tensor components in $1$-
869: or $1.5$-dimensional equations can also cause serious spurious
870: oscillations, we think that, although in the moderate viscosity
871: case they are equivalently effective as what were made by SM01,
872: our modifications for both the radial momentum and energy
873: equations will show their advantages in the strong viscosity case.
874: In fact, the importance of the viscous forces $F_{rr}$ and
875: $F_{\phi\phi}$ has long since been pointed out
876: \citep[e.g.,][]{papaloizou86}. We think that the inclusion of a
877: heating term in the energy equation in accordance with these two
878: forces will be not only consistent in physics, but also hopefully
879: important in obtaining numerically stable solutions. With all
880: these preparations made in this paper, we wish to achieve the goal
881: to answer the question of the fate of thermally unstable black
882: hole accretion disks with very large values of $\alpha$: do these
883: disks finally form stable and persistent SSD+ADAF configurations
884: as suggested by \citet{TM98}, or they also undergo limit-cycles,
885: or something else? In view of the two facts that limit-cyclic
886: luminosity variations, even though with seemingly very reliable
887: theoretical warranties, are not usually observed for black hole
888: systems (GRS 1915+105 remains the only one known); and that
889: outflows are already observed in many high energy astrophysical
890: systems that are believed to be powered by black hole accretion,
891: but are unlikely to originate from thermally unstable accretion
892: disks we study here because of the negative Benoulli function of
893: the matter in these disks, it will be definitely interesting if
894: some behaviors other than the limit-cycle for non-stationary black
895: hole accretion disks and/or the outflow formation from these disks
896: can be demonstrated theoretically.
897: 
898: \acknowledgments
899: 
900: We are very grateful to Ewa Szuszkiewicz and John C. Miller for
901: many helpful instructions and providing an unpublished formula of
902: the numerical viscosity. We also thank Wei-Min Gu for beneficial
903: discussions. This work was supported by the National Science
904: Foundation of China under grant 10673009.
905: 
906: \appendix
907: 
908: \section{Spectral Filtering and Boundary Conditions}
909: When applied to solve non-linear partial differential equations, a
910: principle drawback of spectral methods is the aliasing error that
911: can cause  spurious oscillations at high frequencies. The spectral
912: filtering is a special technique developed to filter out the
913: high-frequency modes in each time-step to reduce the aliasing
914: error. As in \citet[][see~also
915: \citealt{gottlieb1997,peyret2002}]{Chan2005}, we use a exponential
916: filter in spectral space as
917: \begin{equation}
918:     \sigma_{\delta}\left(\frac{n}{N}\right)\equiv\exp\left(-|\ln\epsilon|\left|\frac{n}{N}\right|^{\delta}\right),
919: \end{equation}
920: where $\epsilon$ is the machine accuracy and $\delta$ is a
921: parameter that can be determined from the numerical practice. Then
922: instead of $u(r_k)$, the filtered collocation values of the
923: physical quantity $u(r)$ at a collocation point $r_k$ are given by
924: \begin{equation}
925:    \bar{u}(r_k)=\sum\limits_{n=0}^{N}\sigma_{\delta}\left(\frac{n}{N}\right)\hat{u}_n
926:     \cos\left(\frac{nk\pi}{N}\right).
927: \end{equation}
928: 
929: As for the boundary conditions, the Dirichlet condition,
930: $u|_{boundary}=u_0$, or/and the Neumann condition, $(\partial
931: u/\partial r)|_{boundary}=u_0^{'}$, are generally required. In
932: order to avoid the appearance of discontinuous step-functions at
933: the boundaries that would cause the Gibbs Phenomenon,
934: \citet{Chan2005} introduced another filtering technique, namely
935: the spatial filter. In our numerical algorithm, we either follow
936: \citet{Chan2005} to impose the boundary conditions by using the
937: spatial filter, or directly change the value of a certain physical
938: quantity to be its given value at the boundary point, depending on
939: whether the boundary derivative of the quantity is or is not very
940: small. The spatial filter is a monotonically decreasing filter
941: \begin{equation}
942: h(r)=\exp\left[-|\ln\epsilon|\left(\frac{r-r_{min}}{r_{max}-r_{min}}\right)^\delta\right]
943: \end{equation}
944: for the outer boundary, and a monotonically increasing filter
945: \begin{equation}
946: h(r)=\exp\left[-|\ln\epsilon|\left(\frac{r_{max}-r}{r_{max}-r_{min}}\right)^\delta\right]
947: \end{equation}
948: for the inner boundary. With this filter, the Dirichlet boundary
949: condition can be imposed in each time-step as
950: \begin{equation}
951: u_k^n \rightarrow (u_k^n-u_0)h(r_k)+u_0,
952: \end{equation}
953: and the Neumann boundary condition can be imposed as
954: \begin{equation}
955: \left(\frac{\partial u}{\partial r}\right)_k^n \rightarrow
956:    \left[\left(\frac{\partial u}{\partial
957:    r}\right)_k^n-u_0^{'}\right]h(r_k)+u_0^{'},
958: \end{equation}
959: where the superscript $n$ and subscript $k$ denote the relevant
960: values at the $n$-th time-level and the collocation point $r_k$,
961: respectively.
962: 
963: \begin{thebibliography}{}
964: \bibitem[Abramowicz et al.(2000)]{Abr00} Abramowicz, M. A.,
965: Lasota, J.-P., \& Igumenshchev, I. V. 2000, \mnras, 314, 775
966: 
967: \bibitem[Abramowicz et al.(1995)]{Abr95} Abramowicz, M. A., Chen, X., Kato, S., Lasota, J.-P.,
968: \& Regev, O. 1995, \apj, 438, L37
969: 
970: \bibitem[Abramowicz et al.(1988)]{Abr88} Abramowicz, M. A., Czerny, B., Lasota, J.-P., \&
971: Szuszkiewicz, E. 1988, \apj, 332, 646
972: 
973: \bibitem[Bayliss et al.(1995)]{bayliss1995} Bayliss, A., Garbey, M., \& Matkowsky, B. 1995, J. Comput.
974: Phys., 119, 132
975: 
976: \bibitem[Becker \& Subramanian(2005)]{Becker05} Becker, P. A., \& Subramanian, P. 2005, \apj, 622,
977: 520
978: 
979: \bibitem[Boyd(2000)]{boyd2000} Boyd, J.P. 2000, Chebyshev and Fourier Spectral
980: Methods, 2nd Edition (New York: Dover)
981: 
982: \bibitem[Chan et al.(2005)]{Chan2005} Chan, C.-K., Psaltis, D., \& \"{O}zel, F. 2005, \apj,
983: 628, 353
984: 
985: \bibitem[Chan et al.(2006)]{Chan2006} Chan, C.-K., Psaltis, D., \& \"{O}zel, F. 2006, \apj,
986: 645, 506
987: 
988: \bibitem[Chen et al.(1995)]{Chen95} Chen, X., Abramowicz, M. A., Lasota, J.-P., Narayan,
989: R., \& Yi, I. 1995, \apj, 443, L61
990: 
991: \bibitem[Canuto et al.(1988)]{canuto1988} Canuto, C., Hussaini, M.Y., Quarteroni, A., \& Zang,
992: T.A. 1988, Spectral Methods in Fluid Dynamics (New York: Springer)
993: 
994: \bibitem[Gottlieb \& Orszag(1983)]{gottlieb1983} Gottlieb, D., \& Orszag, S.A. 1983, Numerical
995: Analysis of Spectral Methods: Theory and Applications
996: (Philadelphia: SIAM)
997: 
998: \bibitem[Gottlieb \& Shu(1997)]{gottlieb1997} Gottlieb, D., \& Shu, C.-W. 1997, SIAM Rev., 39(4),
999: 644
1000: 
1001: \bibitem[Hawley et al.(2001)]{hawley2001} Hawley, J. F., Balbus, S.A., \& Stone, J.M. 2001,
1002: \apj, 554, L49
1003: 
1004: \bibitem[Honma et al.(1991)]{Honma91} Honma, F., Matsumoto, R., \& Kato, S. 1991, \pasj,
1005: 43, 147
1006: 
1007: \bibitem[Janiuk et al.(2002)]{Janiuk02} Janiuk, A., Czerny, B., \& Siemiginowska, A. 2002,
1008: \apj, 576, 908
1009: 
1010: \bibitem[Kato et al.(1998)]{Kato98} Kato, S., Fukue, J., \& Mineshige, S. 1998,
1011: Black-Hole Accretion Disks (Kyoto: Kyoto Univ. Press)
1012: 
1013: \bibitem[Kawata et al.(2006)]{Kawata06} Kawata, A., Watarai, K., \& Fukue, J. 2006, \pasj,
1014: 58, 447
1015: 
1016: %\bibitem[Lapidus(1967)]{lapidus1967} Lapidus, A. 1967, J. Comput. Phys., 2, 154
1017: 
1018: \bibitem[Mayer \& Pringle(2006)]{Mayer06} Mayer, M., \& Pringle, J. E. 2006, \mnras, 368, 396
1019: 
1020: \bibitem[Narayan et al.(1998)]{Narayan98} Narayan, R., Mahadevan, R., \& Quataert, E. 1998, in
1021: The Theory of Black Hole Accretion Disks, ed. M. A. Abramowicz, G.
1022: Bj\"{o}rnsson, \& J. E. Pringle (Cambridge: Cambridge Univ.
1023: Press), 148
1024: 
1025: \bibitem[Narayan \& Yi(1994)]{NY94} Narayan, R., \& Yi, I. 1994, \apj, 428, L13
1026: 
1027: \bibitem[Nayakshin et al.(2000)]{Nayakshin00} Nayakshin, S., Rappaport, S., \& Melia, F. 2000,
1028: \apj, 535, 798
1029: 
1030: \bibitem[Ohsuga(2006)]{Ohsuga06} Ohsuga, K. 2006, \apj, 640, 923
1031: 
1032: \bibitem[Papaloizou \& Stanley(1986)]{papaloizou86} Papaloizou J.
1033: C. B., \& Stanley G. Q. C. 1986, \mnras, 220, 593
1034: 
1035: \bibitem[Peyret(2002)]{peyret2002} Peyret, R. 2002, Spectral Methods for Incompressible
1036: Viscous Flow (New York: Springer)
1037: 
1038: \bibitem[Press et al.(1992)]{Press1992} Press, W.H., Teukolsky, S.A., Vetterling, W.T., \&
1039: Flannery, B.P. 1992, Numerical recipes in FORTRAN: the art of
1040: scientific computing, 2nd Edition (New York: Cambridge University
1041: Press)
1042: 
1043: \bibitem[Shakura \& Sunyaev(1973)]{SS1973} Shakura, N.I., \& Sunyaev, R.A. 1973, \aap, 24, 337
1044: 
1045: \bibitem[Shu \& Osher(1988)]{shu1988} Shu, C.W.,\& Osher, S. 1988, J. Comput. Phys., 77, 439
1046: 
1047: \bibitem[Szuszkiewicz \& Miller(1997)]{Szuszkiewicz1997} Szuszkiewicz, E., \& Miller, J.C. 1997, \mnras, 287,
1048: 165
1049: 
1050: \bibitem[Szuszkiewicz \& Miller(1998)]{Szuszkiewicz1998} Szuszkiewicz, E., \& Miller, J.C. 1998, \mnras, 298,
1051: 888
1052: 
1053: \bibitem[Szuszkiewicz \& Miller(2001)]{Szuszkiewicz2001} Szuszkiewicz, E., \& Miller, J.C. 2001, \mnras, 328,
1054: 36 (SM01)
1055: 
1056: \bibitem[Takeuchi \& Mineshige(1998)]{TM98} Takeuchi, M., \&
1057: Mineshige, S. 1998, \apj, 505, L19
1058: 
1059: \bibitem[Teresi et al.(2004a)]{Teresi04a} Teresi, V., Molteni, D., \& Toscano, E. 2004a,
1060: \mnras, 348, 361
1061: 
1062: \bibitem[Teresi et al.(2004b)]{Teresi04b} Teresi, V., Molteni, D., \& Toscano, E. 2004b,
1063: \mnras, 351, 297
1064: 
1065: %\bibitem[Toselli \& Widlund(2005)]{toselli2005} Toselli, A., \& Widlund, O. 2005, Domain Decomposition Methods - Algorithms and
1066: %Theory (New York: Springer)
1067: 
1068: %\bibitem[Weizs\"{a}cker(1948)]{weizsacker} von Weizs\"{a}cker C.F., 1948, Z. Naturforsch., 3a, 524
1069: 
1070: \bibitem[Watarai \& Mineshige(2003)]{Watarai03} Watarai K., \& Mineshige S. 2003, \apj, 596, 421
1071: 
1072: \end{thebibliography}
1073: 
1074: \clearpage
1075: 
1076: \begin{figure}
1077: \plotone{f1.eps} \caption{Evolution of the half-thickness of the
1078: disk during one full cycle.} \label{fig1}
1079: \end{figure}
1080: 
1081: \begin{figure}
1082: \plotone{f2.eps} \caption{Evolution of the temperature of the
1083: matter in the disk during one full cycle.} \label{fig2}
1084: \end{figure}
1085: 
1086: \begin{figure}
1087: \plotone{f3.eps} \caption{Evolution of the surface density of the
1088: disk.} \label{fig3}
1089: \end{figure}
1090: 
1091: \begin{figure}
1092: \plotone{f4.eps} \caption{Evolution of the effective optical depth
1093: of the disk.} \label{fig4}
1094: \end{figure}
1095: 
1096: \begin{figure}
1097: \plotone{f5.eps} \caption{Evolution of the ratio of gas pressure
1098: to total pressure in the disk.} \label{fig5}
1099: \end{figure}
1100: 
1101: \begin{figure}
1102: \plotone{f6.eps} \caption{Evolution of the local accretion rate in
1103: the disk.} \label{fig6}
1104: \end{figure}
1105: 
1106: \begin{figure}
1107: \plotone{f7.eps} \caption{Variation of the bolometric luminosity
1108: of the disk during three full cycles.} \label{fig7}
1109: \end{figure}
1110: 
1111: \begin{figure}
1112: \plotone{f8.eps} \caption{Bernoulli function of the matter of the
1113: disk. Note that the horizonal scale is very different from that in
1114: Figs. \ref{fig1} - \ref{fig6}.} \label{fig8}
1115: \end{figure}
1116: \end{document}
1117: