1: %\documentclass[useAMS,usenatbib,referee]{mn2e}
2: \documentclass[useAMS,usenatbib]{mn2e}
3: \usepackage{times}
4: \usepackage{graphics,epsfig}
5: \usepackage{graphicx}
6: \usepackage{amssymb}
7: %\documentclass[useAMS,twocolumn,usenatbib]{mn2e}
8:
9: \title[Spherical accretion in a fractal medium]
10: {Critical properties of spherically symmetric accretion in a
11: fractal medium}
12: \author[Roy and Ray]
13: {Nirupam Roy$^{1}$\thanks{nirupam@ncra.tifr.res.in}
14: and Arnab K. Ray$^{2}$\thanks{akr@iucaa.ernet.in}\\%\footnotemark[1]\\
15: $^{1}$National Centre for Radio Astrophysics, Tata Institute of
16: Fundamental Research, Post Bag 3, Ganeshkhind, Pune 411007, India\\
17: $^{2}$Inter--University Centre for Astronomy and Astrophysics, Post
18: Bag 4, Ganeshkhind, Pune 411007, India}
19:
20: \begin{document}
21:
22: %\date{}
23: %\pagerange{\pageref{}--\pageref{}} \pubyear{}
24:
25: \maketitle
26:
27: \label{firstpage}
28:
29: \begin{abstract}
30: Spherically symmetric transonic accretion of a fractal medium
31: has been studied in both the stationary and the dynamic regimes.
32: The stationary transonic solution is greatly sensitive to
33: infinitesimal deviations in the outer boundary condition, but
34: the flow becomes transonic and stable, when its evolution
35: is followed through time. The evolution towards transonicity
36: is more pronounced for a fractal medium than what is it
37: for a continuum. The dynamic approach also shows that there
38: is a remarkable closeness between an equation of motion for
39: a perturbation in the flow, and the metric of an analogue
40: acoustic black hole. The stationary inflow solutions of a
41: fractal medium are as much stable under the influence of
42: linearised perturbations, as they are for the fluid continuum.
43: \end{abstract}
44:
45: \begin{keywords}
46: accretion, accretion discs -- hydrodynamics -- ISM: structure
47: \end{keywords}
48:
49: \section{Introduction}
50: \label{sec1}
51:
52: Accretion processes involve the non-self-gravitating flow dynamics
53: of astrophysical matter under the external gravitational influence
54: of a massive astrophysical object, like an ordinary star or a compact
55: object~\citep{fkr02}.
56: A paradigmatic model of an astrophysical accreting system is that
57: of spherically symmetric infall on to a central accretor. Ever since
58: the seminal work published by~\citet{bon52}, which effectively
59: launched the subject in the form in which it is recognised today,
60: the problem of spherically symmetric flows has been revisited time
61: and again from various angles~\citep{par58,par66,an67,bal72,mich72,
62: mes75,blum76,ms77,beg78,cos78,sb78,gar79,bri80,monc80,pso80,vit84,
63: bona87a,bona92b,td92,km94,mar95,tuf95,tmk96,zmt96,tmk97,ke98,das99,
64: malec99,ttsrcl99,das00,ds01,rb02,ray03,das04,rb05,gai06,mrd07,roy07}.
65: This abiding appeal of the spherically symmetric model is explained
66: by the fact that almost always it lends itself to an exact mathematical
67: analysis, and in the process it allows a very clear insight to be had
68: into the underlying physical principles.
69:
70: Ease of mathematical manipulations, however, is not the only reason why
71: spherically symmetric flows are regularly invoked in accretion-related
72: literature. The details of the physics of many astrophysical flows
73: are, in fact, very faithfully described and understood with the help
74: of this relatively simple model. Accretion of the interstellar medium
75: (ISM) is a case in point.
76:
77: While formal fluid dynamical equations in the Newtonian construct of
78: space and time --- which would involve a momentum balance equation
79: (with gravity as an external driving force), the continuity equation
80: and an equation of state --- suffice to a great extent in
81: shedding light on the accretion of the ISM, it must
82: at the same time be recognised that the ISM is not entirely to be
83: seen as a fluid continuum. In fact, for many purposes essential to
84: grasping the underlying details, the ISM is believed to possess a
85: self-similar hierarchical structure over several orders of magnitude
86: in scale~\citep{lars,falg92,heith}.
87: Direct H~{\sc i} absorption observations and interstellar scintillation
88: measurements suggest that the structure extends down to a scale of
89: $10\, \mathrm{au}$~\citep{crov,lang,fais} and possibly even to
90: sub-$\mathrm{au}$ scales~\citep{hill}.
91: Numerous theories have attempted to explain the origin, evolution
92: and mass distribution of these clouds and it has been established,
93: from both observations~\citep{elme} and numerical
94: simulations~\citep{burk,kless,seme}, that the interstellar medium
95: has a clumpy hierarchical self-similar structure with a fractal
96: dimension in three-dimensional space. The main reason
97: for this is still not properly understood, but it can be the consequence
98: of an underlying fractal geometry that may arise due to turbulent
99: processes in the medium.
100:
101: A theoretical study of these astrophysical systems --- either a
102: fluid continuum or a fractal structure --- will necessitate the
103: application of the mathematical principles of nonlinear dynamics.
104: This is the principal objective of this work. The physical processes
105: in a fractal medium have been analysed by fractional
106: integration and differentiation~\citep[][and references therein]{zas}.
107: To do so, the fractal medium has had to be replaced
108: by a continuous medium and the integrals on the network of the fractal
109: medium has had to be approximated by fractional integrals~\citep{ren}.
110: The interpretation of fractional integration is connected with
111: fractional mass dimension~\citep{mand}. Fractional integrals can
112: be considered as integrals over fractional dimension space within
113: a numerical factor~\citep{tara}. This numerical factor has been
114: chosen suitably to get the right dimension of any physical parameter
115: and to derive the standard expression in the continuum limit. The
116: isotropic and homogeneous nature of dimensionality has also been
117: incorporated properly. All of these will give a self-consistent
118: description of the hydrodynamics in a fractal medium~\citep{roy07}.
119:
120: Once the hydrodynamic equivalence has been established, it has then
121: been a fairly easy exercise to model the steady fractal flow like a
122: simple first-order autonomous dynamical system~\citep{stro,js99}. This
123: has made it possible to gain an understanding of the critical aspects
124: of the stationary phase portrait of the fractal flow, especially the
125: behaviour of the transonic solution. The critical point in the phase
126: portrait has been shown to be a saddle point, and
127: the transonic solution that has to pass through this point has been
128: shown to be infinitely sensitive to the choice of a boundary condition.
129: While this bodes ill for the feasibility of transonicity itself within
130: the stationary framework, all steady global solutions
131: (transonic or otherwise) have been found to be stable under a
132: time-dependent linearised perturbation. An interesting fact that has
133: come to light is that the necessary mathematical conditions, which
134: include an equation of motion for the dynamic perturbation and its
135: relevant boundary
136: conditions, to argue for the stability of the steady fractal
137: flows, have been found to be entirely identical to what has been
138: reported earlier regarding the stability of continuous spherically
139: symmetric inflows~\citep{pso80,td92}. This similarity is fortunate
140: and armed with this knowledge, it can be safely claimed that fractal
141: flows are just as stable as continuous flows under the effects of
142: small linearised perturbations.
143:
144: Having said this, one will still have to confront the fact that
145: the time-dependent perturbative analysis has done nothing to indicate
146: the primacy of the transonic solution, and that the steady
147: transonic inflow solution would not be possible without an infinite
148: precision in prescribing a proper boundary condition. This obstacle
149: has, however, been overcome by taking into consideration explicit
150: dynamics in the
151: flow system, and then evolving a physical flow through time, after
152: having started with appropriate initial conditions. Transonicity
153: becomes evident very soon, and it has been argued with a simplified
154: analytical model in the ``pressure-free" limit, that the guiding physical
155: criterion to select the transonic solution is the one forwarded
156: by~\citet{bon52}, i.e. the transonic solution will be chosen because
157: it corresponds to a
158: minimum energy configuration. While the ``pressure-free" limit does
159: not involve the fractal properties directly, it has been demonstrated
160: through a numerical integration of the dynamic flow equations of the
161: fractal medium that transonicity is very much the favoured mode of
162: infall in this case too. And the most salient result to have emerged
163: from this numerical study has been that transonic features becomes
164: more pronounced, as the medium is more like a fractal.
165:
166: It has already been mentioned that the perturbative treatment has
167: been shown to give no clear-cut evidence to favour transonicity. Support
168: for transonicity, however, has come indirectly from the perturbative
169: angle too. The equation of motion for the propagation of the linearised
170: perturbation has been shown to have subtle similarities with an
171: equation implying the metric of an acoustic black hole~\citep{vis98}.
172: This hints at the fact that matter might cross the sonic horizon at
173: the greatest possible rate, i.e. transonically, just as matter has
174: to cross the event horizon of a black hole maximally.
175:
176: \section{The equations of the flow and its fixed points}
177: \label{sec2}
178:
179: Considering the existence of a medium that has a fractal structure of
180: mass dimensionality $D=3d$ (with $d<1$) embedded in a 3-dimensional space,
181: the mass enclosed in a sphere of radius $r$ can be written as~\citep{roy07}
182: \begin{equation}
183: M_D = kr^D \sim \rho l_{\mathrm c}^3
184: \left(\frac{r}{l_{\mathrm c}}\right)^{3d} ,
185: \label{md1}
186: \end{equation}
187: with $D$ referring to the dimension, $\rho$ to the constant density of
188: the medium, and $l_{\mathrm c}$ to a characteristic inner length of the
189: medium that can take an arbitrary value in the limit $d \longrightarrow 1$.
190: This is the scale below which the medium will be continuous. The fractional
191: integrals are computed as integrals over fractional dimension space
192: within a numerical factor. The fractional infinitesimal length for
193: a medium with isotropic mass dimension will, therefore, be
194: given by~\citep{roy07}
195: \begin{equation}
196: {\mathrm d}\overline{r}=\left(\frac{r}{l_{\mathrm c}}\right)^{d-1}
197: {\mathrm d}r ,
198: \label{dr}
199: \end{equation}
200: with the constant having been chosen to derive the standard expression in
201: the limit $d\longrightarrow 1$. It is to be noted that the infinitesimal
202: area and volume elements in this ``fractional continuous'' medium of
203: mass dimension $D=3d$ will be different, and hence the mass enclosed
204: in a sphere of radius $r$ for constant density $\rho$ will be~\citep{roy07}
205: \begin{equation}
206: M_D=\int_V\rho {\mathrm d}\overline{V}=\frac{4}{3}
207: \pi\rho \left(\frac{l_{\mathrm c}}{d}\right)^3
208: \left(\frac{r}{l_{\mathrm c}}\right)^{3d} \sim r^D .
209: \label{mdinteg}
210: \end{equation}
211:
212: In this medium the inviscid Euler equation, describing the dynamics of
213: the velocity field, $v$, can be expressed as~\citep{roy07}
214: \begin{equation}
215: \frac{\partial v}{\partial t}+v\frac{\partial v}{\partial r}
216: +\frac{1}{\rho}\frac{\partial p}{\partial r}+\phi^{\prime}(r)=0 ,
217: \label{euler}
218: \end{equation}
219: where $\phi(r)$ is the gravitational potential of the central
220: accretor that drives the flow (with the prime denoting the spatial
221: derivative of the potential). This is a local conservation law and,
222: as it is to be expected, this has exactly the same form as that of
223: the equation for the continuous medium. In the case of stellar accretion,
224: the flow is driven by the Newtonian potential, $\phi = -GM/r$. On the
225: other hand, frequently in studies of black hole accretion, it becomes
226: convenient to dispense
227: completely with the rigour of general relativity, and instead make
228: use of an ``effective'' pseudo-Newtonian potential that will imitate
229: general relativistic effects in the Newtonian construct of space and
230: time~\citep{pw80,nw91,abn96}. The choice of a particular form of the
231: potential will not affect the general arguments overmuch.
232:
233: The pressure, $p$, is related to the local density, $\rho$,
234: through a polytropic
235: equation of state $p=K \rho^\gamma$, in which $K$ is a constant, and
236: $\gamma$ is the polytropic exponent, whose admissible range is given
237: by $1< \gamma < 5/3$. This range is restricted by the isothermal
238: and the adiabatic limits, respectively~\citep{sc39}. The evolution
239: of $\rho$ is described by the equation of continuity~\citep{roy07},
240: \begin{equation}
241: \frac{\partial \rho}{\partial t}+\frac{1}{r^\alpha}
242: \frac{\partial}{\partial r}\left(\rho v r^\alpha \right)=0 ,
243: \label{con}
244: \end{equation}
245: in which $\alpha = 3d -1$.
246:
247: The flow system will, therefore, be specified by equations~(\ref{euler})
248: and~(\ref{con}), along with the polytropic equation of state. Of
249: particular interest are the steady state solutions for the case where
250: the fractal medium is at rest at a large distance from the accretor.
251: Since transonic flows are of primary concern here, it is required that
252: the static flow should evolve from $v \longrightarrow 0$ as
253: $r \longrightarrow {\infty}$ (the outer boundary condition) to $v> a(r)$
254: for small $r$, where $a$ is the speed of sound, given by
255: $a^2 = \partial p/ \partial \rho = \gamma K \rho ^{\gamma - 1}$. The
256: stationary state implies
257: $\partial v/ \partial t = \partial \rho /\partial t = 0$.
258: Consequently, equations~(\ref{euler}) and~(\ref{con}) will be reduced
259: to their steady state forms as
260: \begin{equation}
261: v\frac{{\mathrm d}v}{{\mathrm d}r}
262: +\frac{1}{\rho}\frac{{\mathrm d}p}{{\mathrm d}r}+\phi^{\prime}(r)=0
263: \label{stateuler}
264: \end{equation}
265: and
266: \begin{equation}
267: \frac{{\mathrm d}}{{\mathrm d}r}\left(\rho v r^\alpha \right) =0 ,
268: \label{statcon}
269: \end{equation}
270: respectively. It is easily seen that equations~(\ref{stateuler})
271: and~(\ref{statcon}) remain invariant under the transformation
272: $v \longrightarrow -v$, i.e. the mathematical problem for inflows
273: ($v<0$) and outflows ($v>0$) is identical in the stationary
274: state~\citep{arc}.
275:
276: \begin{figure}
277: \begin{center}
278: \includegraphics[scale=0.65, angle=0.0]{rrf1.eps}
279: \caption{\label{f1} \small{Integral solutions of the stationary
280: fractal flow, driven by the Newtonian potential, for $D=2.55$
281: and $\gamma=1.4$. The
282: continuous curves, $\mathrm A$ and $\mathrm W$ represent ``accretion"
283: and ``wind", respectively. The fixed point is at $r=r_{\mathrm c}$
284: and $\vert {\mathrm M} \vert =1$. This point is a saddle point.}}
285: \end{center}
286: \end{figure}
287:
288: It is in principle possible to eliminate either $v$ or $\rho$ and solve
289: for the other variable as a function of $r$. However, adopting a slightly
290: different approach, it is possible to recast equations~(\ref{stateuler})
291: and~(\ref{statcon}) in a combined form as
292: \begin{equation}
293: \frac{\mathrm d}{{\mathrm d}r} \left(v^2 \right) = \frac{2 v^2}{r}
294: \left[\frac{\alpha a^2 - r \phi^{\prime}(r)}{v^2 - a^2} \right] ,
295: \label{dvdr}
296: \end{equation}
297: whose integral solutions have been shown in Fig.~\ref{f1}, with
298: $\phi(r) = - GM/r$, and with the vertical axis scaled by the
299: absolute value of the Mach
300: number, $\mathrm{M} = v/a$. The two continuous curves labelled ``A" and
301: ``W" refer to accretion and wind, respectively. The critical points
302: in the flow will be derived from the standard requirement that the
303: flow solutions will have a finite gradient when they will cross the
304: sonic horizon (where the bulk flow velocity exactly matches the speed
305: of acoustic propagation), which will mean that both the numerator
306: and the denominator of equation~(\ref{dvdr}) will have to vanish
307: simultaneously and non-trivially~\citep{skc90,skc96}. This can only
308: happen when
309: \begin{equation}
310: v_{\mathrm c}^2 = a_{\mathrm{c}}^2 =
311: \frac{r_{\mathrm c}\phi^{\prime}(r_{\mathrm c})}{\alpha} ,
312: \label{critcon1}
313: \end{equation}
314: which gives the critical point (or the sonic point in this particular
315: case) conditions, with the subscript
316: ``$\mathrm c$" labelling the critical point values.
317:
318: It is not a difficult exercise to integrate equation~(\ref{stateuler})
319: and then transform the variable $\rho$ in it to $a$ with the help of
320: the equation of state. This, with the critical point conditions as
321: given by equation~(\ref{critcon1}), will give a relation for fixing
322: the critical point coordinates in terms of the flow parameters,
323: $\mathcal E$ (which is actually Bernoulli's constant), $\alpha$
324: and $\gamma$ as
325: \begin{equation}
326: \left(\frac{\gamma +1}{\gamma -1}\right) \frac{r_{\mathrm c}\phi^{\prime}
327: (r_{\mathrm c})}{2\alpha} + \phi(r_{\mathrm c}) = {\mathcal E} .
328: \label{eulerfix}
329: \end{equation}
330: The form of $\phi(r)$ will obviously determine the number of the
331: critical points, and for the Newtonian potential only one root of
332: $r_{\mathrm c}$ will be obtained from equation~(\ref{eulerfix}).
333: This root, for $\mathcal E$ fixed by the outer boundary condition
334: of the transonic inflow solution, will be given as
335: \begin{equation}
336: r_{\mathrm c} = \left[\frac{\left(\gamma+1\right)-2\alpha
337: \left(\gamma-1\right)}{2\alpha}\right] \frac{GM}{a_\infty^2} ,
338: \label{rcrit}
339: \end{equation}
340: with $a_\infty$ being the speed of sound at the outer boundary of the
341: flow~\citep{skc90,skc96}, where the influence of gravity is negligibly
342: weak.
343:
344: It should be worth mentioning here that although with the choice of
345: a pseudo-Newtonian potential multiple roots for $r_{\mathrm c}$
346: would result, practically speaking only one of these roots would
347: be a physically meaningful critical point, through which an integral
348: solution may pass. For spherically symmetric flows in the general
349: relativistic framework, this issue has been discussed by~\citet{mrd07}.
350:
351: \section{The flow as an autonomous dynamical system}
352: \label{sec3}
353:
354: So far the flow variables have been ascertained only at the critical
355: points. Since the flow equations are in general nonlinear differential
356: equations, short of carrying out a numerical integration, there is
357: no completely rigorous analytical prescription for solving these
358: differential equations to determine the global nature of the flow
359: variables.
360: Nevertheless, some analytical headway could be made after all by taking
361: advantage of the fact that equation~(\ref{dvdr}), which gives a
362: complete description of the $r$ --- $v^2$ phase portrait of the flow,
363: is an autonomous first-order differential equation, and as such, could
364: easily be recast into the mathematical form ${\dot{x}} = X(x,y)$ and
365: ${\dot{y}} = Y(x,y)$, which is that of the very familiar coupled
366: first-order dynamical system~\citep{stro,js99}.
367: Quite frequently for any nonlinear physical system, a linearised
368: analytical study of the properties of the fixed points of a
369: first-order dynamical system, affords a
370: robust platform for carrying out an investigation to understand
371: the global behaviour of integral solutions in the phase portrait.
372:
373: And so it is that to investigate the nature of the critical point,
374: equation~(\ref{dvdr}) will have to be decomposed in terms of a
375: mathematical parameter, $\tau$, to read as
376: \begin{eqnarray}
377: \frac{\mathrm d}{{\mathrm d}\tau} \left(v^2\right) &=&
378: 2 v^2 \left[\alpha a^2 - r \phi^{\prime}(r)\right]
379: \nonumber \\
380: \frac{{\mathrm d}r}{{\mathrm d}\tau} &=& r\left(v^2 - a^2\right) ,
381: \label{dynsys}
382: \end{eqnarray}
383: in both of which the parameter $\tau$ does not make an explicit
384: appearance in the right hand side, something of an especial advantage
385: that derives from working with autonomous systems. This kind of
386: parametrization is quite common in fluid dynamics~\citep{bdp93},
387: and in accretion studies especially, this approach has been made
388: before~\citep{rb02,ap03,crd06,mrd07,gkrd07}. Some earlier works in
389: accretion
390: had also made use of the general mathematical aspects of this
391: approach~\citep{mkfo84,mc86,ak89}. A further point that has to
392: be noted is that the function $a^2$ in the right hand side of
393: equation~(\ref{dynsys}) can be expressed entirely in terms of
394: $v^2$ and $r$, with the help of the equation of state and
395: equation~(\ref{statcon}). This will exactly satisfy the
396: criterion of a first-order autonomous dynamical system.
397:
398: The next task would be to make a linearised approximation about
399: the fixed point coordinates and extract a linear dynamical system
400: out of equations~(\ref{dynsys}). This will give a direct way to
401: establish the nature of the critical points (or fixed points).
402: Expanding about the fixed point values, a perturbation of the kind
403: $v^2=v_{\mathrm c}^2 +\delta v^2=v_{\mathrm c}^2(1 +\epsilon_1)$
404: and $r=r_{\mathrm c} +\delta r=r_{\mathrm c}(1 +\epsilon_2)$
405: can be applied. Using the continuity equation and the equation
406: of state, this perturbation scheme, when linearised, will also
407: give $\delta a/a^2_{\mathrm c}=-\left(\gamma -1\right)
408: \left(\epsilon_1+2\alpha\epsilon_2\right)/2$. Applying this
409: perturbative expansion on equation~(\ref{dynsys}), and linearising
410: in $\epsilon_1$ and $\epsilon_2$ will give,
411: \begin{eqnarray}
412: \frac{{\mathrm d}\epsilon_1}{{\mathrm d}\tau}&=&\alpha v_{\mathrm c}^2
413: \Bigg\{-\left(\gamma -1\right)\epsilon_1 \nonumber \\
414: & &-2 \left[\alpha\gamma -\alpha
415: +1 +\frac{\phi^{\prime\prime}(r_{\mathrm c}) r_{\mathrm c}}
416: {\phi^{\prime}(r_{\mathrm c})} \right]
417: \epsilon_2 \Bigg\} \nonumber \\
418: \frac{{\mathrm d}\epsilon_2}{{\mathrm d} \tau}&=&\alpha
419: v_{\mathrm c}^2 \left[\left(\frac{\gamma +1}{2\alpha} \right)
420: \epsilon_1 +\left(\gamma-1 \right)\epsilon_2 \right] .
421: \label{lindyn}
422: \end{eqnarray}
423: Using solutions of the type $\epsilon_1 \sim \exp(\Omega \tau)$
424: and $\epsilon_2 \sim \exp(\Omega \tau)$ in equations~(\ref{lindyn}),
425: the eigenvalues of the stability matrix associated with the critical
426: points will be derived as
427: \begin{equation}
428: \Omega^2 = \alpha a_{\mathrm c}^4 \left[\left(2\alpha-1\right)
429: -\gamma\left(2\alpha+1\right)-\left(\gamma+1\right)r_{\mathrm c}
430: \frac{\phi^{\prime\prime}(r_{\mathrm c})}{\phi^{\prime}(r_{\mathrm c})}
431: \right]
432: \label{eigen}
433: \end{equation}
434: with $a_{\mathrm c}$ itself being a function of $r_{\mathrm c}$, as
435: given by equation~(\ref{critcon1}).
436:
437: Once the position of a critical point, $r_{\mathrm c}$, has become
438: known from equation~(\ref{eulerfix}), it is then quite easy to determine
439: the nature of that critical point by using $r_{\mathrm c}$ in
440: equation~(\ref{eigen}). Since $r_{\mathrm c}$ is a function of
441: $\mathcal E$ and $\gamma$, it effectively implies that $\Omega^2$ can,
442: in principle, be regarded as a function of the flow parameters. From
443: the form of $\Omega^2$ in equation~(\ref{eigen}), a generic conclusion
444: that can be immediately drawn is that any critical point, as it may
445: be expected for a conservative system, will be either a saddle point
446: (for $\Omega^2 >0$) or a centre-type point (for $\Omega^2 <0$). For the
447: particular case of the Newtonian potential, $\phi=-GM/r$, the eigenvalues
448: of the stability matrix will be given by
449: \begin{equation}
450: \Omega_{\mathrm N}^2 =\alpha a_{\mathrm c}^4
451: \left[\left(2\alpha+1\right)-\gamma\left(2\alpha-1\right)\right] .
452: \label{neigen}
453: \end{equation}
454: For the values of $\gamma$ and $\alpha$ lying in the range of physical
455: interest, it can always be shown that $\Omega_{\mathrm N}^2 > 0$. Hence
456: in this case
457: the critical point is a saddle point, and the curves which have been
458: labelled ``accretion" and ``wind" in Fig.~\ref{f1} are in fact separatrices
459: of a dynamical system, rather than physical solutions.
460:
461: The understand the full import of this line of reasoning, what has to
462: be borne in mind is that saddle points are inherently unstable, and to
463: make a solution pass through such a point, after starting from an outer
464: boundary condition, will entail an infinitely precise fine-tuning of that
465: boundary condition~\citep{rb02}. This can be demonstrated through simple
466: arguments. Going back to equations~(\ref{lindyn}), the coupled set of
467: linear differential equations in $\epsilon_1$ and $\epsilon_2$ can be
468: set down as
469: \begin{equation}
470: \label{ratio}
471: \frac{{\mathrm d} \epsilon_1}{{\mathrm d} \epsilon_2}
472: = \frac{{\mathrm d} \epsilon_1/{\mathrm d} \tau}
473: {{\mathrm d} \epsilon_2/{\mathrm d} \tau} =
474: \frac{{\mathcal Q}_1 \epsilon_1 + {\mathcal Q}_2 \epsilon_2}
475: {{\mathcal Q}_3 \epsilon_1 + {\mathcal Q}_4 \epsilon_2} ,
476: \end{equation}
477: in which the constant coefficients ${\mathcal Q}_1$, ${\mathcal Q}_2$,
478: ${\mathcal Q}_3$ and ${\mathcal Q}_4$ are to be determined simply by
479: an inspection of equations~(\ref{lindyn}). It is also to be easily seen
480: that ${\mathcal Q}_1 = - {\mathcal Q}_4$. This makes the integration
481: of equation~(\ref{ratio}) a straightforward exercise, and it yields
482: \begin{equation}
483: \label{hyper}
484: {\mathcal Q}_2 \epsilon_2^2 +2{\mathcal Q}_1 \epsilon_1 \epsilon_2
485: - {\mathcal Q}_3 \epsilon_1^2 + {\mathcal C} = 0
486: \end{equation}
487: with ${\mathcal C}$ being an integration constant. Generally speaking
488: equation~(\ref{hyper}) is the equation of a conic section in the
489: $\epsilon_2$ --- $\epsilon_1$ plane. If the origin of this plane were
490: to be considered to have been shifted to the saddle point, then the
491: condition for solutions passing through the origin,
492: i.e. $\epsilon_1 = \epsilon_2 = 0$, would be ${\mathcal C} = 0$,
493: which would reduce equation~(\ref{hyper}) to a pair of straight
494: lines intersecting each other through the origin itself. All other
495: solutions in the vicinity of the origin will, therefore, be hyperbolic
496: in nature, a fact that is given by the condition
497: $({\mathcal Q}_1^2 + {\mathcal Q}_2 {\mathcal Q}_3) > 0$.
498: For the case of $\phi = -GM/r$, this contention can be verified
499: completely analytically, and this shows that even a very minute
500: deviation from a precise boundary condition for transonicity
501: (i.e. a boundary condition
502: that will generate solutions to pass {\em only} through the origin,
503: $\epsilon_1 = \epsilon_2 = 0$) will take the stationary solution far
504: away from a transonic state. This extreme sensitivity of transonic
505: solutions to boundary conditions is entirely in keeping with the
506: nature of saddle points. It may be imagined that in a proper
507: astrophysical system such precise fulfillment of a boundary
508: condition will make the transonic solution well-nigh physically
509: non-realisable. Indeed, this difficulty, for any kind of accreting
510: system, is readily appreciated by anyone trying to carry out a numerical
511: integration of equation~(\ref{dvdr}) to generate the transonic solutions,
512: which can only be obtained when the numerics is first biased in favour
513: of transonicity by using the saddle point condition itself as the
514: boundary condition for numerical integration.
515:
516: Apart from this, there is also a mathematical aspect of the physical
517: non-realisability of transonic solutions.
518: Using the condition ${\mathcal C} = 0$ will make it easy
519: to express $\epsilon_1$ in terms of $\epsilon_2$ and vice versa.
520: Going back to the set of linear equations given by equations~(\ref{lindyn})
521: and choosing the second one of the two equations (the choice of
522: the first would also have led to the same result), one gets
523: \begin{equation}
524: \label{aar}
525: \frac{{\mathrm d} \epsilon_2}{{\mathrm d} \tau}= \pm
526: \left(\sqrt{{\mathcal Q}_1^2 + {\mathcal Q}_2 {\mathcal Q}_3}\right)
527: \,\,\epsilon_2 ,
528: \end{equation}
529: which can be integrated for both the roots from an arbitrary
530: initial value of $\epsilon_2 = \epsilon_2^{\star}$ lying anywhere
531: on the transonic solution, to a point $\epsilon_2 = \Delta$, with
532: $\Delta$ being very close to the critical point given by
533: $\epsilon_2 =0$. Using the equivalence that
534: $\Omega^2 = {\mathcal Q}_1^2 + {\mathcal Q}_2 {\mathcal Q}_3$,
535: it can be shown that
536: \begin{equation}
537: \label{tau}
538: \tau = \pm \frac{1}{\Omega}\int_{\epsilon_2^{\star}}^{\Delta}
539: \frac{{\mathrm d} \epsilon_2}{\epsilon_2} %\nonumber \\
540: = \pm \frac{1}{\Omega}
541: \ln \Bigg{\vert} \frac{\Delta}{\epsilon_2^{\star}} \Bigg{\vert} ,
542: \end{equation}
543: from which it is easy to see that
544: $\vert\tau\vert\longrightarrow\infty$
545: for $\Delta \longrightarrow 0$. This implies
546: that the critical point may be reached along either of the
547: separatrices, only after $\vert \tau \vert$ has become
548: infinitely large. This divergence of the parameter $\tau$
549: indicates that in the stationary regime, solutions passing
550: through the saddle point are not actual solutions, but
551: separatrices of various classes of solutions~\citep{stro,js99}. This fact,
552: coupled with the sensitivity of the stationary transonic solutions
553: to the choice of an outer boundary condition, makes
554: their feasibility a seriously questionable matter.
555:
556: \section{A time-dependent perturbative approach}
557: \label{sec4}
558:
559: It has been demonstrated in the foregoing section that the steady
560: transonic accretion solution is unstable under infinitesimal deviations
561: from the precise outer boundary condition needed to generate the
562: solution. In
563: the astrophysical context, such precision is quite impossible,
564: and, therefore, the very feasibility of transonicity becomes a matter
565: of grave doubt. This difficulty can, however, be dispelled if one is
566: mindful of the fact that the real astrophysical flow is not static
567: but dynamic in character, i.e. it will have an explicit dependence
568: on time. This, of course, will mean that the time-dependent terms
569: involving both the velocity and the density fields in
570: equations~(\ref{euler}) and~(\ref{con}) will have to be retained.
571:
572: While full-fledged time-dependence of the flow variables will
573: undoubtedly reveal many new interesting mathematical facets (all of
574: them involving the mathematics of partial differential equations)
575: of the physical problem,
576: it would still be worthwhile to go back to studying the properties
577: of the background stationary flow under the influence of a linearised
578: perturbative effect. As a preliminary exercise in accounting for
579: explicit time-dependence, this will at least shed some light on the
580: global stability of the flow solutions.
581:
582: To that end, it will first be necessary to define, closely following a
583: perturbative procedure prescribed by~\citet{pso80} and~\citet{td92},
584: a new physical variable $f=\rho v r^\alpha$. It is quite obvious from
585: the form of equation~(\ref{con}) that the stationary value of $f$ will
586: be a global constant, $f_0$, which can be closely identified with the
587: matter flux rate. A perturbation prescription of the form
588: $v(r,t) = v_0(r) + v^{\prime}(r,t)$ and
589: $\rho (r,t) = \rho_0 (r) + \rho^{\prime}(r,t)$, will give,
590: on linearising in the primed quantities,
591: \begin{equation}
592: f^{\prime} = f_0 \left(\frac{\rho^{\prime}}{\rho_0}
593: + \frac{v^{\prime}}{v_0} \right) ,
594: \label{effprime}
595: \end{equation}
596: with the subscript ``$0$" denoting stationary values in all cases.
597: From equation~(\ref{con}), it then becomes possible to set down the
598: density fluctuations, $\rho^{\prime}$, in terms of $f^{\prime}$ as
599: \begin{equation}
600: \frac{\partial \rho^{\prime}}{\partial t} +\frac{v_0 \rho_0}{f_0}
601: \left(\frac{\partial f^{\prime}}{\partial r}\right)=0 .
602: \label{flucden}
603: \end{equation}
604: Combining equations~(\ref{effprime}) and~(\ref{flucden}) will then
605: render the velocity fluctuations as
606: \begin{equation}
607: \frac{\partial v^{\prime}}{\partial t}=
608: \frac{v_0}{f_0}\left(\frac{\partial f^{\prime}}{\partial t}
609: + v_0 \frac{\partial f^{\prime}}{\partial r}\right)
610: \label{flucvel}
611: \end{equation}
612: which, upon a further partial differentiation with respect to time,
613: will give
614: \begin{equation}
615: \frac{{\partial}^2 v^{\prime}}{\partial t^2}=\frac{\partial}{\partial t}
616: \left[\frac{v_0}{f_0} \left(\frac{\partial f^{\prime}}{\partial t}\right)
617: \right]+\frac{\partial}{\partial t} \left[ \frac{v_0^2}{f_0}
618: \left(\frac{\partial f^{\prime}}{\partial r}\right)\right] .
619: \label{flucvelder2}
620: \end{equation}
621:
622: From equation~(\ref{euler}) the linearised fluctuating part could be
623: extracted as
624: \begin{equation}
625: \frac{\partial v^{\prime}}{\partial t}+ \frac{\partial}{\partial r}
626: \left( v_0 v^{\prime} + a_0^2
627: \frac{\rho^{\prime}}{\rho_0}\right) =0
628: \label{fluceuler}
629: \end{equation}
630: with $a_0$ being the speed of sound in the steady state. Differentiating
631: equation~(\ref{fluceuler}) partially with respect to $t$, and making
632: use of equations~(\ref{flucden}),~(\ref{flucvel}) and~(\ref{flucvelder2})
633: to substitute for all the first and second-order derivatives of
634: $v^{\prime}$ and $\rho^{\prime}$, will deliver the result
635: \begin{displaymath}
636: \frac{\partial}{\partial t} \left[\frac{v_0}{f_0}
637: \left( \frac{\partial f^{\prime}}{\partial t}\right)\right]
638: + \frac{\partial}{\partial t} \left[\frac{v_0^2}{f_0}
639: \left( \frac{\partial f^{\prime}}{\partial r}\right)\right]
640: + \frac{\partial}{\partial r} \left[\frac{v_0^2}{f_0}
641: \left( \frac{\partial f^{\prime}}{\partial t}\right)\right]
642: \end{displaymath}
643: \begin{equation}
644: \qquad \qquad \qquad \qquad \qquad + \frac{\partial}{\partial r}
645: \left[\frac{v_0}{f_0}\left(v_0^2 - a_0^2 \right)
646: \frac{\partial f^{\prime}}{\partial r}\right] = 0 .
647: \label{interm}
648: \end{equation}
649: A little readjustment of terms in equation~(\ref{interm}) will finally
650: give an equation of motion for the perturbation as
651: \begin{equation}
652: \frac{{\partial}^2 f^{\prime}}{\partial t^2}+2\frac{\partial}{\partial r}
653: \left(v_0 \frac{\partial f^{\prime}}{\partial t}\right)+\frac{1}{v_0}
654: \frac{\partial}{\partial r}\left[ v_0 \left(v_0^2- a_0^2\right)
655: \frac{\partial f^{\prime}}{\partial r}\right] = 0 ,
656: \label{tpert}
657: \end{equation}
658: which is an expression that is exactly the same as what can be derived
659: upon perturbing the stationary solutions of spherically symmetric
660: inflows in a continuous medium~\citep{pso80,td92}. Another aspect of
661: equation~(\ref{tpert}) is that its form has no explicit dependence on
662: the potential --- Newtonian or pseudo-Newtonian --- that is driving the
663: flow. This is entirely to be expected, because the potential, being
664: independent of time, will only lend its direct presence to the stationary
665: background flow. Arguments regarding stability will, therefore, be more
666: dependent on the boundary conditions of the steady flow.
667: As the form of the equation of motion for the linearised perturbation
668: remains unchanged even for a flow in a fractal medium, and as the
669: physical boundary conditions are also not altered in this case, the
670: general conclusions reached by both~\citet{pso80} and~\citet{td92}
671: regarding flows in a continuous medium, will carry over here, and it
672: can be safely claimed that under all reasonable boundary conditions,
673: both the transonic and subsonic solutions will be stable.
674:
675: While this does nothing to cause any immediate worry, it also does
676: not reveal anything in particular either about the physical feasibility
677: of any solution
678: from a perturbative point of view. This is in keeping with the
679: conventional wisdom about spherically symmetric flows~\citep{gar79}
680: that the natural
681: preference of the system for any particular solution --- especially
682: the transonic solution --- cannot be justified by a linear stability
683: analysis, but by the more fundamental arguments forwarded
684: by~\citet{bon52}.
685:
686: For all that, some positive hint about the primary status of the
687: transonic solution can actually be derived if the whole issue of
688: a linear stability analysis is viewed from a different perspective.
689: It is known that there is a close one-to-one correspondence between
690: certain features of black hole physics and the physics of supersonic
691: acoustic flows \citep{vis98}. In this very context, a compact
692: rendering of equation~(\ref{interm}) can be obtained as
693: \begin{equation}
694: \partial_\mu \left({\mathrm{f}}^{\mu \nu}
695: \partial_\nu f^{\prime}\right) = 0 ,
696: \label{compact}
697: \end{equation}
698: in which the Greek indices are made to run from $0$ to $1$, with the
699: identification that $0$ stands for $t$, and $1$ stands for $r$. An
700: inspection of the terms in the left hand side of equation~(\ref{interm})
701: will then allow for constructing the symmetric matrix
702: \begin{equation}
703: {\mathrm{f}}^{\mu\nu}=\frac{v_0}{f_0}
704: \pmatrix
705: {1 & v_0 \cr
706: v_0 & v_0^2 - a_0^2} .
707: \label{matrix}
708: \end{equation}
709: Now in Lorentzian geometry the d'Alembertian for a scalar field in
710: curved space is given in
711: terms of the metric ${\mathrm{g}}_{\mu \nu}$ by~\citep{vis98}
712: \begin{equation}
713: \Delta \psi\equiv \frac{1}{\sqrt{-\mathrm{g}}}
714: \partial_\mu\left({\sqrt{-\mathrm{g}}}\,
715: {\mathrm{g}}^{\mu\nu}\partial_\nu\psi\right)
716: \label{alem}
717: \end{equation}
718: with $\mathrm{g}^{\mu\nu}$ being the inverse of the matrix implied
719: by ${\mathrm{g}}_{\mu\nu}$. Comparing equation~(\ref{compact})
720: with equation~(\ref{alem}),
721: it would be tempting to look for an exact equivalence between
722: ${\mathrm f}^{\mu\nu }$ and $\sqrt{-\mathrm g}\, {\mathrm g}^{\mu\nu}$.
723: This, however, cannot be done in a general sense. What can be
724: appreciated, nevertheless, is that equation~(\ref{compact}) gives
725: a relation for $f^{\prime}$ which is of the type given by
726: equation~(\ref{alem}). The metrical part of equation~(\ref{compact}),
727: as given by equation~(\ref{matrix}), may then be
728: extracted, and its inverse will incorporate the notion of a sonic
729: horizon of an acoustic black hole when $v_0^2 = a_0^2$.
730: This point of view does not make for a perfect acoustic analogue model,
731: but it has some similar features to the metric of a wave equation for a
732: scalar field in curved space-time, obtained through a
733: somewhat different approach, in which the velocity of an an irrotational,
734: inviscid and barotropic fluid flow is first represented as the gradient
735: of a scalar function $\psi$, i.e. ${\bf v}= -{\bf{\nabla}}\psi$, and
736: then a perturbation is imposed on this scalar function~\citep{vis98}.
737:
738: The foregoing discussion indicates that the physics of supersonic
739: acoustic flows closely corresponds to many features of black hole
740: physics. This closeness of form is very intriguing. For a black hole,
741: infalling matter crosses the event horizon maximally, i.e. at the
742: greatest possible speed. By analogy the same thing may be said of
743: matter crossing the sonic horizon of a spherically symmetric fluid
744: flow, falling on to a point sink. That this fact can be appreciated
745: for the spherically symmetric accretion problem, through a perturbative
746: result as given by equation~(\ref{interm}), is quite remarkable. This
747: is because it is universally recognised that that no insight into the
748: special status of any inflow solution may possibly be derived solely
749: through a perturbative technique~\citep{gar79}. It is the
750: transonic solution that crosses the sonic horizon at the greatest
751: possible rate~\citep{bon52},
752: and the similarity of form between equations~(\ref{interm})
753: and~(\ref{alem}) may very well be indicative of the primacy of the
754: transonic solution. If such an insight were truly to be had with the
755: help of the perturbation equation, then the perturbative linear
756: stability analysis might not have been carried out in vain after all.
757:
758: \section{Dynamic evolution towards transonicity}
759: \label{sec5}
760:
761: Much more direct and robust evidence in favour of transonicity could
762: be obtained if the
763: accreting system were to be made to evolve through time, as opposed
764: to making it suffer small linearised perturbations in time. Having
765: said this, it must also be stressed that equations~(\ref{euler})
766: and~(\ref{con}), which govern the temporal evolution of the flow,
767: are not amenable to a ready mathematical analysis; indeed, in the matter
768: of incorporating both the dynamic and the pressure effects in
769: the equations, short of a direct numerical treatment, the mathematical
770: problem, is very aptly described as ``insuperable"~\citep{bon52}.
771: Therefore, to have a preliminary
772: appreciation of the governing mechanism that underlies any possible
773: selection of a transonic flow, it should be necessary to adopt
774: some simplifications. This will pave the way for a more complete
775: physical understanding of the evolutionary properties of the flow.
776:
777: The evolutionary dynamics is, therefore, to be studied first in the
778: regime of what is understood to be the ``pressureless" motion of a fluid
779: in a gravitational field~\citep{shu}, as opposed to dropping the
780: dynamic effects to study a much simplified
781: stationary picture~\citep{bon52}. Simplification of
782: the mathematical equations, however, is not the only justification
783: for such a prescription. A greater justification lies in the fact
784: that the result delivered is in conformity with, what~\citet{gar79}
785: calls ``the more fundamental arguments" of~\citet{bon52}; that it
786: is the criterion of minimum total energy associated with a solution,
787: that will accord it a principal status over all the others.
788:
789: An immediate consequence of adopting dynamic equations is that
790: the invariance of the stationary solutions under the transformation
791: $v \longrightarrow -v$, is lost. As a result, one will have to
792: separately consider either the inflows $(v<0)$ or the outflows $(v>0)$,
793: a choice that has to be imposed upon the system at $t=0$. Euler's
794: equation, tailored according to the simplified requirements of a
795: ``pressureless" field, is rendered as
796: \begin{equation}
797: \frac{\partial v}{\partial t} + v \frac{\partial v}{\partial r}
798: + \frac{GM}{r^2} = 0
799: \label{presfree}
800: \end{equation}
801: which can be solved by the method of characteristics~\citep{ld97}.
802: The characteristic solutions are obtained from
803: \begin{equation}
804: \frac{{\mathrm d}t}{1}= \frac{{\mathrm d}r}{v}
805: = \frac{{\mathrm d}v}{-GM/r^2} .
806: \label{charcur}
807: \end{equation}
808: First solving the ${\mathrm d}v/{\mathrm d}r$ equation will give
809: \begin{equation}
810: \frac{v^2}{2}- \frac{GM}{r} = \frac{c^2}{2}
811: \label{dvdrchar}
812: \end{equation}
813: in which $c$ is an integration constant that derives from the spatial
814: part of the characteristic equation. This result is to be used to solve
815: the ${\mathrm d}r/{\mathrm d}t$ equation from equation~(\ref{charcur}),
816: which will finally lead to
817: \begin{equation}
818: \frac{2vr}{c r_{\mathrm s}} -\ln r -
819: \ln \left(\frac{v}{c}+1 \right)^2
820: - \frac{2ct}{r_{\mathrm s}} = {\tilde c}
821: \label{drdt}
822: \end{equation}
823: with $\tilde{c}$ being another integration constant, and $r_{\mathrm s}$
824: being a length scale in the system, defined as $r_{\mathrm s}=2GM/c^2$.
825:
826: A general solution of equation~(\ref{charcur}) is given by the
827: condition, $f({\tilde c})= c^2/2$, with $f$ being an arbitrary
828: function, whose form is to be determined from the initial condition.
829: The general solution can, therefore, be set down as
830: \begin{equation}
831: \frac{v^2}{2}-\frac{GM}{r}=f\left[\frac{2vr}{c{r_{\mathrm s}}} -\ln r
832: -\ln \left(\frac{v}{c}+1\right)^2-\frac{2ct}{r_{\mathrm s}}\right] ,
833: \label{gensol2}
834: \end{equation}
835: to determine whose particular form, the initial condition that will
836: have to be used is $v=0$ at $t=0$ for all $r$. This will lead to
837: \begin{equation}
838: \frac{v^2}{2}-\frac{GM}{r}= -\frac{GM}{r}\left(\frac{v}{c}+1
839: \right)^{-2}\exp\left(\frac{2vr}{cr_{\mathrm s}}
840: -\frac{2ct}{r_{\mathrm s}}\right)
841: \label{partform}
842: \end{equation}
843: from which it is easy to see that for $t \longrightarrow \infty$,
844: what is approached is the stationary solution,
845: \begin{equation}
846: \frac{v^2}{2} - \frac{GM}{r} = 0 .
847: \label{statsol}
848: \end{equation}
849: Corresponding to the given initial condition, this is evidently
850: the stationary solution associated with the lowest possible total
851: energy, and the temporal evolution selects this solution from among
852: all the others. The whole picture could be conceived of as one in
853: which a system with a uniform velocity distribution $v=0$ everywhere,
854: suddenly has a gravity mechanism switched on in its midst at $t=0$.
855: This induces a potential $-GM/r$ at all points in space. The system
856: then starts evolving to restore itself to another stationary state,
857: with the velocity increasing according to equation~(\ref{partform}),
858: so that for $t \longrightarrow \infty$, the total energy at all
859: points, $E=(v^2/2)-(GM/r)=0$, remains the same as at $t=0$.
860:
861: \begin{figure}
862: \begin{center}
863: \includegraphics[scale=0.65, angle=0.0]{rrf2.eps}
864: \caption{\label{f2} \small{For the temporal evolution of velocity,
865: the slope of this logarithmic plot shows that in the early stages
866: of the evolution, $-v$ varies linearly with $t$. Deviation from
867: linear growth sets in later. The horizontal
868: line on top shows the terminal value of the velocity field, whose
869: evolution is being followed at a fixed length scale.}}
870: \end{center}
871: \end{figure}
872:
873: \begin{figure}
874: \begin{center}
875: \includegraphics[scale=0.65, angle=0.0]{rrf3.eps}
876: \caption{\label{f3} \small{The $\log$-$\log$ plot of the Mach number
877: versus the radial distance, with the latter being scaled by the sonic
878: radius, $r_\mathrm{c}$, after $4000$ seconds of evolution.
879: Going from left to right each curve corresponds to $d=0.7$, $0.85$
880: and $1.0$, respectively. The slope of the curves indicates a power
881: law behaviour for the evolution on intermediate length scales. All
882: the curves approach a saturation velocity scale closer to the accretor.}}
883: \end{center}
884: \end{figure}
885:
886: \begin{figure}
887: \begin{center}
888: \includegraphics[scale=0.65, angle=0.0]{rrf4.eps}
889: \caption{\label{f4} \small{Velocity field, scaled as the Mach number,
890: $\mathrm M$, after $4000$ seconds, for various values of $d$. Moving
891: from the bottom to the top, successive solutions
892: have been shown for $d=1.0$, $0.85$ and $0.7$, respectively. The
893: radial distance has been normalised with respect to $r_\odot$, and
894: plotted logarithmically over a length scale of $r_\odot$ to
895: $4r_\odot$.}}
896: \end{center}
897: \end{figure}
898:
899: This contention has been borne out by a numerical integration of
900: equation~(\ref{presfree}) by the finite differencing technique. The
901: mass of the accretor has been chosen to be $M_\odot$, while its
902: radius is $r_\odot$. The evolution through time has been followed
903: at a fixed length scale of $51 r_\odot$. The result of the numerical
904: evolution of the velocity field, $-v$ (for inflows $v$ is actually
905: negative), through time, $t$, has been plotted in
906: Fig.~\ref{f2}. The limiting value of the velocity, as
907: the evolution progresses towards the long-time limit, is evidently
908: $\sqrt{2GM/r}$ (with $M= M_\odot$ and $r =51 r_\odot$),
909: as equation~(\ref{statsol}) would suggest.
910: This is what the plot in Fig.~\ref{f2} shows, as $-v$ approaches
911: its terminal value for $t \longrightarrow \infty$. The slope of
912: this logarithmic plot also indicates that in the early
913: stages of the evolution there is a linear growth of the velocity
914: field through time, but on later times, conspicuous deviation from
915: linearity sets in.
916:
917: The argument presented above, with the effects of pressure taken
918: into account, can now be extended to understand the
919: dynamic selection of the transonic solution. The inclusion of the
920: pressure term in the dynamic equation, fixes the total energy of
921: the system accordingly at $t=0$. A physically realistic initial
922: condition should be that $v=0$ at $t=0$, for all $r$, while $\rho$
923: has some uniform value. The temporal evolution of the accreting
924: system would then non-perturbatively select the transonic trajectory,
925: as it is this solution with which is associated the least possible
926: energy configuration. This argument is in conformity with the assertion
927: made by~\citet{bon52} that it is the criterion of minimum total energy
928: that should make a particular solution (the transonic solution in this
929: case) preferred to all the others. However, this selection mechanism
930: is effective only through the temporal evolution of the flow.
931:
932: To test this contention a numerical study has been carried out once
933: again using finite differencing, but this time using both the dynamic
934: equations for the velocity and the density fields, as given by
935: equations~(\ref{euler}) and (\ref{con}). The accretor has been chosen
936: to have a mass, $M_\odot$, and radius, $r_\odot$. The ``ambient"
937: conditions are $a_\infty = 10 \ {\rm{km}}~{\rm{s}}^{-1}$ and
938: $\rho_\infty = 10^{-21} \ {\rm{kg}}~{\rm{m}}^{-3}$. The polytropic
939: exponent, $\gamma$, has been set as $n \equiv (\gamma -1)^{-1}=1.61$.
940: For these values of the physical constants,
941: transonicity becomes apparent even at the very early stages of the
942: evolution. This is shown in Fig.~\ref{f3} in which the velocity field
943: (scaled as the Mach number) has been plotted after it has evolved for
944: $4000$ seconds. The horizontal distance has been scaled by the sonic
945: radius (which, going by equation~(\ref{rcrit}), effectively makes this
946: scaling dependent on the fractal dimension), and it shows a saturation
947: behaviour for the inflow velocity at smaller length scales. This
948: saturation scale for the velocity field is roughly the same for
949: any value of $d$, although at larger length scales, a curve placed
950: higher in the plot, corresponds to a higher value of $d$. Besides
951: this, all curves, as it is apparent from the slope of each of them,
952: display the same kind of power-law behaviour on larger length scales.
953: All this is somewhat reminiscent of the growth processes exhibited
954: by the ballistic deposition model, which is applied to generate
955: a nonequilibrium interface~\citep{barstan}.
956:
957: The comparative properties of the velocity field, at smaller length
958: scales, for various values of $d$, have been exhibited in
959: Fig.~\ref{f4}. Here the radial distance has been scaled in terms of
960: the radius of the accretor, which in this case is $r_\odot$. What is
961: interesting to note in this plot is that for all other conditions
962: remaining the same, on length scales close to the accretor, solutions
963: corresponding to lower values of the fractal dimension, $d$, grow
964: faster in time than solutions with higher values of $d$. It is
965: possible to argue that this is exactly how it should be.
966: The pressure of the infalling gas, in so far as it is connected to
967: the density through a polytropic equation of state, builds up
968: resistance against gravity, because of the growth of the density
969: field on small length scales. Transonicity can only be achieved when
970: gravity wins over pressure on length scales smaller than the sonic
971: radius. This will be all the more true near the accretor, where
972: the velocity field will evolve under free-fall conditions, and,
973: therefore, the more dilute the gas, the more efficient will be the
974: drive towards the transonic state. Now a fractal medium may be
975: viewed equivalently as a continuum with an effective lesser density.
976: In this situation a system with
977: a lesser value of $d$ will be more prone to losing against gravity
978: than a system with higher value of $d$, and so the race towards
979: transonicity will be more successful as $d$ decreases. It is
980: exactly this state of affairs that Fig.~\ref{f4} graphically
981: represents.
982:
983: \section{Concluding remarks}
984: \label{sec6}
985:
986: An earlier work reported by~\citet{roy07} was carried out under the
987: implicit assumption that the accretion process would take place
988: transonically. The present treatment bears out this assumption
989: self-consistently. It was also discussed by~\citet{roy07} that
990: the rate of accretion in a fractal medium can vary significantly
991: from the~\citet{bon52} rate for more massive accretors. This is
992: very much in conformity with the conclusions derived in this work,
993: through the numerical evolution of transonicity, and it would be
994: judicious to account for this fact, while studying the accretion
995: of a fractal medium on to a black hole.
996: Black hole accretion is necessarily transonic~\citep{skc90,skc96},
997: but even for accretion from a molecular cloud on to a star, in the
998: absence of any inner boundary condition being imposed on the flow,
999: the flow is expected to be transonic~\citep{pso80}. Therefore,
1000: whatever be the nature of the accretor, both transonicity and the
1001: quantitative modifications arising due to the fractal
1002: nature of the accreting medium, will be very much relevant for
1003: studies in spherically symmetric accretion.
1004:
1005: It has also been discussed here that transonic properties manifest
1006: themselves more noticeably with an increase in the fractal
1007: properties of the flow. The dynamic evolution has shown that
1008: the growth rate of the velocity field (as scaled against the
1009: speed of acoustic propagation) becomes significantly higher in
1010: this case. This, of course, will have a direct bearing on the
1011: mass accretion rate, and it is worth conjecturing that there
1012: might be some observational evidence for this kind of behaviour.
1013:
1014: \section*{Acknowledgements}
1015:
1016: This research has made use of NASA's Astrophysics Data System. The
1017: authors express gratitude to J. K. Bhattacharjee, J. N.
1018: Chengalur, T. Naskar and R. Nityananda for much encouragement and
1019: many helpful comments.
1020:
1021: \begin{thebibliography}{99}
1022:
1023: \bibitem[\protect\citeauthoryear{Abramowicz \& Kato}{1989}]{ak89}
1024: Abramowicz, M. A., Kato, S., 1989, ApJ, 336, 304
1025:
1026: \bibitem[\protect\citeauthoryear{Afshordi \& Paczy\'nski}{2003}]{ap03}
1027: Afshordi, N., Paczy\'nski, B., 2003, ApJ, 592, 354
1028:
1029: \bibitem[\protect\citeauthoryear{Artemova et al.}{1996}]{abn96}
1030: Artemova, I. V., Bj\"ornsson, G., Novikov, I. D., 1996, ApJ, 461, 565
1031:
1032: \bibitem[\protect\citeauthoryear{Axford \& Newman}{1967}]{an67}
1033: Axford, W. I., Newman, R. C., 1967, ApJ, 147, 230
1034:
1035: \bibitem[\protect\citeauthoryear{Balazs}{1972}]{bal72}
1036: Balazs, N. L., 1972, MNRAS, 160, 79
1037:
1038: \bibitem[\protect\citeauthoryear{Barab\'asi \& Stanley}{1995}]{barstan}
1039: Barab\'asi, A.-L., Stanley, H. E., 1995, Fractal Concepts in Surface
1040: Growth, Cambridge University Press, Cambridge
1041:
1042: \bibitem[\protect\citeauthoryear{Begelman}{1978}]{beg78}
1043: Begelman, M. C., 1978, A\&A, 70, 53
1044:
1045: \bibitem[\protect\citeauthoryear{Blumenthal \& Mathews}{1976}]{blum76}
1046: Blumenthal, G. R., Mathews, W. G., 1976, ApJ, 203, 714
1047:
1048: \bibitem[\protect\citeauthoryear{Bohr et al.}{1993}]{bdp93}
1049: Bohr, T., Dimon, P., Putkaradze, V., 1993, Journal of Fluid
1050: Mechanics, 254, 635
1051:
1052: \bibitem[\protect\citeauthoryear{Bonazzola et al.}{1987}]{bona87a}
1053: Bonazzola, S., Falgarone, E., Heyvaerts, J., P\' {e}rault, M.,
1054: Puget, J. L., 1987, A\&A, 172, 293
1055:
1056: \bibitem[\protect\citeauthoryear{Bonazzola et al.}{1992}]{bona92b}
1057: Bonazzola, S., P\' {e}rault, M., Puget, J. L., Heyvaerts, J.,
1058: Falgarone, E., Panis, J. F., 1992, Journal of Fluid Mechanics, 245, 1
1059:
1060: \bibitem[\protect\citeauthoryear{Bondi}{1952}]{bon52}
1061: Bondi, H., 1952, MNRAS, 112, 195
1062:
1063: \bibitem[\protect\citeauthoryear{Brinkmann}{1980}]{bri80}
1064: Brinkmann, W., 1980, A\&A, 85, 146
1065:
1066: \bibitem[\protect\citeauthoryear{Burkert et al.}{1997}]{burk}
1067: Burkert, A., Bate, M. R., Bodenheimer, P., 1997, MNRAS, 289, 497
1068:
1069: \bibitem[\protect\citeauthoryear{Chakrabarti}{1990}]{skc90}
1070: Chakrabarti, S. K., 1990, Theory of Transonic Astrophysical
1071: Flows, World Scientific, Singapore
1072:
1073: \bibitem[\protect\citeauthoryear{Chakrabarti}{1996}]{skc96}
1074: Chakrabarti, S. K., 1996, Physics Reports, 266, 229
1075:
1076: \bibitem[\protect\citeauthoryear{Chandrasekhar}{1939}]{sc39}
1077: Chandrasekhar, S., 1939, An Introduction to the Study of Stellar
1078: Structure, The University of Chicago Press, Chicago
1079:
1080: \bibitem[\protect\citeauthoryear{Chaudhury et al.}{2006}]{crd06}
1081: Chaudhury, S., Ray, A. K., Das, T. K., 2006, MNRAS, 373, 146
1082:
1083: \bibitem[\protect\citeauthoryear{Choudhuri}{1999}]{arc}
1084: Choudhuri, A. R., 1999, The Physics of Fluids and Plasmas: An
1085: Introduction for Astrophysicists, Cambridge University Press, Cambridge
1086:
1087: \bibitem[\protect\citeauthoryear{Cowie et al.}{1978}]{cos78}
1088: Cowie, L. L., Ostriker, J. P., Stark, A. A., 1978, ApJ, 226, 1041
1089:
1090: \bibitem[\protect\citeauthoryear{Crovisier et al.}{1985}]{crov}
1091: Crovisier, J., Dickey, J. M., Kaz\`{e}s, I., 1985, A\&A, 146, 223
1092:
1093: \bibitem[\protect\citeauthoryear{Das}{1999}]{das99}
1094: Das, T. K., 1999, MNRAS, 308, 201
1095:
1096: \bibitem[\protect\citeauthoryear{Das}{2000}]{das00}
1097: Das, T. K., 2000, MNRAS, 318, 294
1098:
1099: \bibitem[\protect\citeauthoryear{Das}{2004}]{das04}
1100: Das, T. K., 2004, Classical and Quantum Gravity, 21, 5253
1101:
1102: \bibitem[\protect\citeauthoryear{Das \& Sarkar}{2001}]{ds01}
1103: Das, T. K., Sarkar, A., 2001, A\&A, 374, 1150
1104:
1105: \bibitem[\protect\citeauthoryear{Debnath}{1997}]{ld97}
1106: Debnath, L., 1997, Nonlinear Partial Differential Equations for
1107: Scientists and Engineers, Birkh\"auser, Boston
1108:
1109: \bibitem[\protect\citeauthoryear{Elmegreen \& Falgarone}{1996}]{elme}
1110: Elmegreen, B. G., Falgarone, E., 1996, ApJ, 471, 816
1111:
1112: \bibitem[\protect\citeauthoryear{Faison et al.}{1998}]{fais}
1113: Faison, M. D., Goss, W. M., Diamond, P. J., Taylor, G. B., 1998,
1114: AJ, 116, 2916
1115:
1116: \bibitem[\protect\citeauthoryear{Falgarone et al.}{1992}]{falg92}
1117: Falgarone, E., Puget, J.-L., Perault, M., 1992, A\&A, 257, 715
1118:
1119: \bibitem[\protect\citeauthoryear{Frank et al.}{2002}]{fkr02}
1120: Frank, J., King, A., Raine, D., 2002, Accretion Power in
1121: Astrophysics, Cambridge University Press, Cambridge
1122:
1123: \bibitem[\protect\citeauthoryear{Gaite}{2006}]{gai06}
1124: Gaite, J., 2006, A\&A, 449, 861
1125:
1126: \bibitem[\protect\citeauthoryear{Garlick}{1979}]{gar79}
1127: Garlick, A. R., 1979, A\&A, 73, 171
1128:
1129: \bibitem[\protect\citeauthoryear{Goswami et al.}{2007}]{gkrd07}
1130: Goswami, S., Khan, S. N., Ray, A. K., Das, T. K., 2007, 378, 1400
1131:
1132: \bibitem[\protect\citeauthoryear{Heithausen et al.}{1998}]{heith}
1133: Heithausen, A., Bensch, F., Stutzki, J., Falgarone, E.,
1134: Panis, J. F., 1998, A\&A, 331, L65
1135:
1136: \bibitem[\protect\citeauthoryear{Hill et al.}{2005}]{hill}
1137: Hill, A. S., Stinebring, D. R., Asplund, C. T., Berkwick, D. E.,
1138: Everett, W. B., Hinkel, N. R., 2005, ApJ, 619, L171
1139:
1140: \bibitem[\protect\citeauthoryear{Jordan \& Smith}{1999}]{js99}
1141: Jordan, D. W., Smith, P., 1999, Nonlinear Ordinary Differential
1142: Equations, Oxford University Press, Oxford
1143:
1144: \bibitem[\protect\citeauthoryear{Kazhdan \& Murzina}{1994}]{km94}
1145: Kazhdan, Y. M., Murzina, M., 1994, MNRAS, 270, 351
1146:
1147: \bibitem[\protect\citeauthoryear{Klessen et al.}{1998}]{kless}
1148: Klessen, R. S., Burkert, A., Bate, M. R., 1998, ApJ, 501, L205
1149:
1150: \bibitem[\protect\citeauthoryear{Kovalenko \& Eremin}{1998}]{ke98}
1151: Kovalenko, I. G., Eremin, M. A., 1998, MNRAS, 298, 861
1152:
1153: \bibitem[\protect\citeauthoryear{Langer et al.}{1995}]{lang}
1154: Langer, W. D., Velusamy, T., Kuiper, T. B. H., Levin, S., Olsen, E.,
1155: Migenes, V., 1995, ApJ, 453, 293
1156:
1157: \bibitem[\protect\citeauthoryear{Larson}{1981}]{lars}
1158: Larson, R. B., 1981, MNRAS, 194, 809
1159:
1160: \bibitem[\protect\citeauthoryear{Malec}{1999}]{malec99}
1161: Malec, E., 1999, Phys. Rev. D, 60, 104043
1162:
1163: \bibitem[\protect\citeauthoryear{Mandal et al.}{2007}]{mrd07}
1164: Mandal, I., Ray, A. K., Das, T. K., 2007, 378, 1407
1165:
1166: \bibitem[\protect\citeauthoryear{Mandelbrot}{1983}]{mand}
1167: Mandelbrot, B., 1983, The Fractal Geometry of Nature, W. H. Freeman,
1168: New York
1169:
1170: \bibitem[\protect\citeauthoryear{Markovic}{1995}]{mar95}
1171: Markovic, D., 1995, MNRAS, 277, 11
1172:
1173: \bibitem[\protect\citeauthoryear{Matsumoto et al.}{1984}]{mkfo84}
1174: Matsumoto, R., Kato, S., Fukue, J., Okazaki, A. T., 1984, PASJ, 36, 71
1175:
1176: \bibitem[\protect\citeauthoryear{M\'esz\'aros}{1975}]{mes75}
1177: M\'esz\'aros, P., 1975, A\&A, 44, 59
1178:
1179: \bibitem[\protect\citeauthoryear{M\'esz\'aros \& Silk}{1977}]{ms77}
1180: M\'esz\'aros, P., Silk, J., 1977, A\&A, 55, 289
1181:
1182: \bibitem[\protect\citeauthoryear{Michel}{1972}]{mich72}
1183: Michel, F. C., 1972, Astrophys. Space Sci., 15, 153
1184:
1185: \bibitem[\protect\citeauthoryear{Moncrief}{1980}]{monc80}
1186: Moncrief, V., 1980, ApJ, 235, 1038
1187:
1188: \bibitem[\protect\citeauthoryear{Muchotrzeb-Czerny}{1986}]{mc86}
1189: Muchotrzeb-Czerny, B., 1986, Acta Astronomica, 36, 1
1190:
1191: \bibitem[\protect\citeauthoryear{Nowak \& Wagoner}{1991}]{nw91}
1192: Nowak, A. M., Wagoner, R. V., 1991, ApJ, 378, 656
1193:
1194: \bibitem[\protect\citeauthoryear{Paczy\'nski \& Wiita}{1980}]{pw80}
1195: Paczy\'nski, B., Wiita P. J., 1980, A\&A, 88, 23
1196:
1197: \bibitem[\protect\citeauthoryear{Parker}{1958}]{par58}
1198: Parker, E. N., 1958, ApJ, 123, 664
1199:
1200: \bibitem[\protect\citeauthoryear{Parker}{1966}]{par66}
1201: Parker, E. N., 1966, ApJ, 143, 32
1202:
1203: \bibitem[\protect\citeauthoryear{Petterson et al.}{1980}]{pso80}
1204: Petterson, J. A., Silk, J., Ostriker, J. P., 1980, MNRAS, 191, 571
1205:
1206: \bibitem[\protect\citeauthoryear{Ray}{2003}]{ray03}
1207: Ray, A. K., 2003, MNRAS, 344, 1085
1208:
1209: \bibitem[\protect\citeauthoryear{Ray \& Bhattacharjee}{2002}]{rb02}
1210: Ray, A. K., Bhattacharjee, J. K., 2002, Phys. Rev. E, 66, 066303
1211:
1212: \bibitem[\protect\citeauthoryear{Ray \& Bhattacharjee}{2005}]{rb05}
1213: Ray, A. K., Bhattacharjee, J. K., 2005, ApJ, 627, 368
1214:
1215: \bibitem[\protect\citeauthoryear{Ren et al.}{2003}]{ren}
1216: Ren, F-Y., Liang, J-R., Wang, X-T., Qiu, W-Y., 2003,
1217: Chaos, Solitons and Fractals, 16, 107
1218:
1219: \bibitem[\protect\citeauthoryear{Roy}{2007}]{roy07}
1220: Roy, N., 2007, (To appear in MNRAS Letters)
1221:
1222: \bibitem[\protect\citeauthoryear{Semelin \& Combes}{2000}]{seme}
1223: Semelin, B., Combes, F., 2000, A\&A, 360, 1096
1224:
1225: \bibitem[\protect\citeauthoryear{Shu}{1991}]{shu}
1226: Shu, F. K., 1991, The Physics of Astrophysics, Vol. II : Gas Dynamics,
1227: University Science Books, California
1228:
1229: \bibitem[\protect\citeauthoryear{Stellingwerf \& Buff}{1978}]{sb78}
1230: Stellingwerf, R. F., Buff, J., 1978, ApJ, 221, 661
1231:
1232: \bibitem[\protect\citeauthoryear{Strogatz}{1994}]{stro}
1233: Strogatz, S. H., 1994, Nonlinear Dynamics and Chaos, Addison-Wesley
1234: Publishing Company, Reading, MA
1235:
1236: \bibitem[\protect\citeauthoryear{Tarasov}{2004}]{tara}
1237: Tarasov, V. E., 2004, Chaos, 14, 123
1238:
1239: \bibitem[\protect\citeauthoryear{Theuns \& David}{1992}]{td92}
1240: Theuns, T., David, M., 1992, ApJ, 384, 587
1241:
1242: \bibitem[\protect\citeauthoryear{Titarchuk et al.}{1996}]{tmk96}
1243: Titarchuk, L., Mastichiadis, A., Kylafis, N. D., 1996, A\&A, 120, 171
1244:
1245: \bibitem[\protect\citeauthoryear{Titarchuk et al.}{1997}]{tmk97}
1246: Titarchuk, L., Mastichiadis, A., Kylafis, N. D., 1997, ApJ, 487, 834
1247:
1248: \bibitem[\protect\citeauthoryear{Toropin et al.}{1999}]{ttsrcl99}
1249: Toropin, Yu. M., Toropina, O. D., Savelyev, V. V., Romanova, M. M.,
1250: Chechetkin, V. M., Lovelace, R. V. E., 1999, ApJ, 517, 906
1251:
1252: \bibitem[\protect\citeauthoryear{Tsuribe et al.}{1995}]{tuf95}
1253: Tsuribe, T., Umemura, M., Fukue, J., 1995, PASJ, 47, 73
1254:
1255: \bibitem[\protect\citeauthoryear{Vitello}{1984}]{vit84}
1256: Vitello, P., 1984, ApJ, 284, 394
1257:
1258: \bibitem[\protect\citeauthoryear{Visser}{1998}]{vis98}
1259: Visser, M., 1998, Classical and Quantum Gravity, 15, 1767
1260:
1261: \bibitem[\protect\citeauthoryear{Zampieri et al.}{1996}]{zmt96}
1262: Zampieri, L., Miller, J. C., Turolla, R., 1996, MNRAS, 281, 1183
1263:
1264: \bibitem[\protect\citeauthoryear{Zaslavsky}{2002}]{zas}
1265: Zaslavsky, G. M., 2002, Physics Reports, 371, 461
1266:
1267: \end{thebibliography}
1268:
1269: \bsp
1270: \label{lastpage}
1271: \end{document}
1272: