astro-ph0210677/ms.tex
1: 
2: \documentclass[12pt,preprint]{aastex}
3: 
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: 
7: 
8: \shorttitle{Thermodynamics of partially relaxed stellar systems}
9: \shortauthors{Bertin and Trenti}
10: 
11: 
12: \begin{document}
13: 
14: 
15: \title{Thermodynamical description of a family of partially relaxed stellar systems}
16: 
17: 
18: \author{G. Bertin}
19: \affil{Dipartimento di Fisica, Universit\`{a} di Milano, via
20: Celoria 16, I-20133 Milano, Italy}
21: \email{Giuseppe.Bertin@unimi.it}
22: 
23: \and
24: 
25: \author{M. Trenti}
26: \affil{Scuola Normale Superiore, piazza dei Cavalieri 7, I-56126 Pisa, Italy}
27: \email{m.trenti@sns.it}
28: 
29: 
30: \begin{abstract}
31: We examine the thermodynamical
32: properties of a family of partially relaxed, anisotropic stellar
33: systems, derived earlier from the Boltzmann entropy under the
34: assumption that a third quantity $Q$ is conserved in addition to
35: the total energy and the total number of stars.  We now show that the family
36: of models conforms to the paradigm of the gravothermal
37: catastrophe, which is expected to occur (in the presence of
38: adequate energy transport mechanisms) when the one-parameter
39: equilibrium sequence attains sufficiently high values of the
40: concentration parameter; these are the values for which the models
41: are well fitted by the $R^{1/4}$ law. In the intermediate
42: concentration regime the models belonging to the sequence exhibit
43: significant deviations from the $R^{1/4}$ law. Curiously, in the
44: low-concentration regime, the global thermodynamical temperature
45: associated with the models becomes negative when the models become
46: too anisotropic so that they are unstable against the radial orbit
47: instability; this latter behavior, while offering a new clue to
48: the physical interpretation of the radial orbit instability, is at
49: variance with respect to the low-concentration limit of the
50: classical case of the isotropic, isothermal sphere investigated by
51: \citet{bon56}  and \citet{lyn68}.
52: \end{abstract}
53: 
54: 
55: \keywords{stellar dynamics --- galaxies: evolution --- galaxies: formation --- galaxies: kinematics and dynamics --- galaxies: structure}
56: 
57: 
58: \section{Introduction}
59: The possibility of providing a thermodynamical description of
60: self-gravitating stellar systems has motivated a number of
61: investigations in galactic dynamics, starting with the pioneering work
62: of Antonov and Lynden-Bell in the 60s. After the realization that
63: violent relaxation is likely to lead to partially relaxed
64: configurations in dynamical equilibrium \citep{lyn67}, a
65: re-examination of the problem of the isothermal sphere, studied
66: earlier by \citet{bon56} for a self-gravitating gas, led to the
67: interesting possibility that stellar systems may undergo the process
68: of gravothermal catastrophe (\citealt{lyn68}; see also
69: \citealt{ant62}).  However, a rigorous derivation of the onset of the
70: gravothermal catastrophe from a study of the Boltzmann entropy
71: appeared to be available only for the case of ideal systems confined
72: by a spherical reflecting wall. A number of convincing qualitative
73: arguments made it clear that also unbound stellar systems with finite
74: mass, such as those described by the King sequence (\citealt{kin66};
75: these spherical models have a finite radius, but do not require an
76: external wall), should probably fall into the same physical framework
77: and indeed the paradigm received a lot of attention, especially in the
78: context of the dynamics of globular clusters (see \citealt{spi87}),
79: which are known to possess, at least to some extent, the desired
80: internal collisionality (see also \citealt{lyn80}).  [An indirect
81: indication that the general physical picture of the gravothermal
82: catastrophe is likely to be robust comes also from the proof that the
83: behaviour of the classical gas case is basically independent of the
84: assumption of spherical geometry \citep{lom01}.]  Several
85: investigations have aimed at producing a rigorous derivation of the
86: gravothermal catastrophe for unbound stellar systems, focusing on the
87: underlying argument that refers to the Poincar\'{e} stability of
88: linear series of equilibria \citep{kat78,kat79,pad89}, but the proof
89: has always been centered on an unjustified {\it ansatz} in order to
90: connect the underlying entropy $S$ with the global temperature $T =
91: 1/(\partial{S}/\partial{E_{tot}})$ (see Appendix V in the article by
92: \citealt{lyn68}; \citealt{kat80}; \citealt{mag98}). Other
93: investigations have explored the possibility of setting the discussion
94: in the context of non-standard entropies (e.g., the Tsallis entropies;
95: see the study of the polytropic spheres by \citealt{cha02} and
96: \citealt{tar02}). Note that the concept of entropy for collisionless
97: systems is quite subtle (e.g., see \citealt{sti87} and references
98: therein). One might even argue whether it is actually compatible with
99: the long-range nature of gravity, given the fact that self-gravitating
100: systems lack additivity, a key ingredient in thermodynamics.
101: 
102: In the meantime, inspired by N-body simulations of collisionless
103: collapse \citep{van82}, which confirmed the general picture of
104: incomplete violent relaxation and showed that it can lead to systems
105: with realistic density profiles without ad hoc tuning of the initial
106: conditions, some families of models were constructed able to
107: reproduce, for quasi-spherical configurations, the characteristic
108: feature of the anisotropy profile with an inner isotropic core and an
109: outer radially biased envelope (\citealt{ber84}; see \citealt{ber93}
110: and references therein): these families turned out to exhibit the
111: characteristic $R^{1/4}$ projected density profile and indeed were
112: shown to match nicely the observed photometric and kinematical
113: characteristics of bright ellipticals. In an attempt at providing a
114: justification of these models (in particular, of the so-called
115: $f_{\infty}$ models, constructed initially only from dynamical
116: arguments) from statistical mechanics, two routes were pursued
117: \citep{sti87}.
118: 
119: The first combines an
120: explicit statement of partial relaxation, i.e. of a relaxation process
121: that is expected to be inefficient in the outer regions, and
122: the existence of a suitable weight, related to the orbital period,
123: for the cells that make the relevant partition of phase space; it follows
124: qualitative arguments proposed by \citet{lyn67} and is  physically appealing (see also \citealt{tre86a}).
125: It was indeed shown to lead naturally to the $f_{\infty}$ models.
126: However, this route is not fully satisfactory from the
127: mathematical point of view, especially since it involves an
128: explicit approximation for the orbital period that is applicable
129: only to the low binding energy limit of quasi-Keplerian orbits.
130: The second route is straightforward from the mathematical point of
131: view, being based on the classical Boltzmann entropy and on the
132: assumed explicit conservation of a third quantity $Q$, in
133: addition to the total mass $M$ and to the total energy $E_{tot}$.
134: It was shown to lead to an analytically different family of models
135: (that we may call the $f^{(\nu)}$ models; see definition in
136: Sect.~3 below), with qualitative properties very similar to those
137: of the $f_{\infty}$ models. Those models were not studied much
138: further and did not receive great attention, not only because the
139: relevant distribution function is not as simple as that of the
140: $f_{\infty}$ models, but especially because the conservation of
141: $Q$ could only be justified approximately by inspection of a
142: number of N-body simulations, without a clear-cut physical
143: justification (see \citealt{sti87}; in contrast, the
144: conservation of the additional ${\bf A} \cdot {\bf B}$ invariant
145: sometimes invoked in plasma physics is rather transparent; see
146: \citealt{cha58}).
147: 
148: In this paper we take advantage of the simple statistical
149: mechanics foundation of the $f^{(\nu)}$ family of unbound
150: partially relaxed stellar systems to explore the possibility of a
151: thermodynamical description of stellar-dynamical models that are
152: endowed with realistic properties.
153: 
154: \section{Comment on the evolution of elliptical galaxies}
155: 
156: Before proceeding to illustrate the results of this paper, we make a short digression in order to bring out the connections between the present analysis and the evolution of elliptical galaxies. We start by recalling that, formally, 
157: the sequence of \citet{kin66} models is one special family of solutions of
158: the collisionless Boltzmann equation. Yet, it is recognized to provide a
159: reasonable description of the current properties of globular clusters
160: (see \citealt{djo94}), within a framework where these stellar
161: systems continually evolve as a result of a variety of mechanisms (among
162: which star evaporation and disk shocking; see \citealt{ves97} and
163: references therein) and where the paradigm of the gravothermal
164: catastrophe can be applied (see \citealt{spi87}). Of course, it is well
165: known that the level of internal collisionality in globular clusters is
166: relatively high, so that the above approach is quite natural.
167: 
168: In contrast, one might at first think of dismissing the possibility
169: that the paradigm of the gravothermal catastrophe should be of
170: interest for the study of elliptical galaxies, because these large
171: stellar systems lack the desired level of collisionality, judging from
172: the estimate of the relevant star-star relaxation times. Here,
173: following the spirit of earlier investigations (starting with
174: \citealt{lyn68}), we note that real elliptical galaxies are actually
175: complex systems the evolution of which goes well beyond the idealized
176: framework of the collisionless Boltzmann equation. In other words,
177: splitting their description into past (formation) processes and
178: present (mostly collisionless equilibrium) conditions should be
179: considered only as an idealization introduced in order to assess the
180: properties that define their current basic state. 
181: 
182: In practice, elliptical galaxies are expected to be in a state of
183: continuous evolution, for which we can list several specific dynamical
184: causes: (1) Left-over granularity of the stellar system itself from
185: initial collapse.  Clumps of stars are likely to continue to form and
186: dissolve in phase space even after the system has reached an
187: approximate steady state.  This acts as internal collisionality thus
188: making some relaxation proceed even at current epochs. Indeed,
189: numerical simulations of violent (partial) relaxation show that some
190: evolution continues well after the initial collapse has taken
191: place. (2) Drag of a system of globular clusters or other heavier
192: objects towards the galaxy center. A globular cluster system or the
193: frequent capture of small satellites (mini-mergers) may provide an
194: internal heating mechanism associated with the process of dynamical
195: friction by the stars on the heavier objects \citep{ber02b}. (3)
196: Long-term action of tidal interactions of the galaxy with external
197: objects. (4) Presence of gas in various phases (cold, warm, and
198: hot). Significant cooling flows have been observed in bright
199: ellipticals. Traditionally, studies of processes of this kind focus on
200: the dynamics of the cooling gas and keep the background stellar system
201: as `frozen'. In reality, energy and mass exchanges take place between
202: the stellar system and the interstellar medium. (5) Interaction
203: between the galactic nucleus and the galaxy. A number of interesting
204: correlations have been found between the properties of galaxy nuclei
205: and global properties of the hosting galaxies (e.g., see
206: \citealt{pel99}). These correlations suggest that significant energy
207: exchanges are taking place between the galaxy and its
208: nucleus. Eventually, if a sufficiently concentrated nucleus is
209: generated, then star-star relaxation in the central regions may also
210: become a significant cause of dynamical evolution.
211: 
212: All of the above are specific mechanisms that are expected to make
213: elliptical galaxies evolve in spite of their very long typical
214: star-star relaxation time. Most of these processes are hard to model
215: and to calculate in detail. As for the evolution of other complex
216: many-body systems, it is hoped that thermodynamical arguments may help
217: us identify general trends characterizing such evolution. This is the
218: basic physical scenario in which the calculations presented in this
219: paper are expected to be of interest for real elliptical
220: galaxies.
221: 
222: 
223: 
224: \section{Partially relaxed, unbound, finite mass systems from the Boltzmann entropy}
225: 
226: Let us consider the standard Boltzmann entropy $S = - \int f \ln{f} d^3x d^3v$ and look for functions that extremize its value under the constraint that the total energy $E_{tot} = (1/3)\int E f d^3x d^3v$, the total mass $M = \int f d^3x d^3v$, and the additional quantity
227: 
228: \begin{equation}
229: Q = \int J^{\nu} |E|^{-3 \nu/4} f d^3x d^3v
230: \end{equation}
231: 
232: \noindent are taken to be constant. Here the functions $E$ and $J^2$ represent specific energy and specific angular momentum square of a single star subject to a spherically symmetric mean potential $\Phi(r)$. As shown elsewhere \citep{sti87}, this extremization process leads to the following family of distribution functions
233: 
234: \begin{equation}
235: f^{(\nu)} = A \exp {\left[- a E - d \left( \frac{J^2}{|E|^{3/2}}\right)^{\nu/2}\right]}~,
236: \end{equation}
237: 
238: \noindent where $a$, $A$, and $d$ are positive real constants. One
239: may think of these constants as providing two dimensional scales
240: (for example, $M$ and $Q$) and one dimensionless parameter; the
241: dimensionless parameter can be taken to be $\gamma = ad^{2/\nu}/(4
242: \pi GA)$. In principle, $\nu$ is any positive real number; in
243: practice, we will focus on values of $\nu \approx 1$. The
244: $f^{(\nu)}$ non-truncated models are constructed by taking this
245: form of the distribution function for $E \leq 0$, a vanishing
246: distribution function for $E>0$, and by integrating the relevant
247: Poisson equation under the condition that the potential $\Phi$ be
248: regular at the origin and behaves like $- G M/r$ at large radii.
249: This integration leads to an eigenvalue problem (see Appendix) for
250: which a value of $\gamma$ is determined by the choice of the
251: central dimensionless potential, $\gamma = \gamma (\Psi)$, with
252: $\Psi = - a \Phi (r=0)$.
253: 
254: The main point of the following analysis is the determination of
255: the Boltzmann entropy $S(M, Q, \Psi)$ and of the total energy
256: $E_{tot}(M, Q, \Psi)$ along the sequence of models, i.e. as a
257: function of the concentration parameter $\Psi$ defined above.
258: These functions, at constant $M$ and $Q$, are illustrated in Fig.~{\ref{fig1}}.
259: They have been obtained by noting that, from the definitions of $S$
260: and $f^{(\nu)}$,
261: 
262: \begin{equation}
263: S = - M \ln{A} + 3 a E_{tot} + d Q.
264: \end{equation}
265: 
266: \noindent From the definitions $Q = A a^{-9/4}d^{-1 - 3/\nu} \hat{Q}(\Psi)$
267: and $M = A a^{-9/4}d^{- 3/\nu} \hat{M}(\Psi)$ and the definition of $\gamma$,
268: we can express the variables $(A, a, d)$ in terms of the variables $(M, Q, \Psi)$
269: and thus find that the entropy per unit mass can be written as $S/M=S_0(M,Q)+\sigma(\Psi)$, where $S_0$ is constant when the values of $M$ and $Q$
270: are fixed, with
271: 
272: \begin{equation}
273: \sigma = -\ln{\left(\hat{M}^{\frac{4\nu - 6}{5 \nu}}\hat{Q}^{\frac{6}{5 \nu}}\gamma^{-\frac{9}{5}}\right)} + \frac{3 \hat{E}}{\hat{M}} + \frac{\hat{Q}}{\hat{M}}~.
274: \end{equation}
275: 
276: \noindent Here $\hat{E} = \hat{E}(\Psi)$ is the dimensionless total energy defined from $E_{tot} = A a^{-13/4} d^{-3/\nu}\hat{E}$. From the identity $a E_{tot}/M = \hat{E}/\hat{M}$ and the expression of $a = a(M, Q, \Psi)$ obtained previously, we find $E_{tot}/M = H(M,Q)\epsilon(\Psi)$, with:
277: 
278: \begin{equation}
279: \epsilon = \gamma^{\frac{4}{5}} \hat{M}^{-\frac{9\nu+4}{5\nu}} \hat{Q}^{\frac{4}{5\nu}} \hat{E}~.
280: \end{equation}
281: 
282: \noindent The factor $H(M,Q)$ is a constant when $M$ and $Q$ are taken to be constant. The quantities $\gamma (\Psi)$, $\hat{M}(\Psi)$, $\hat{Q}(\Psi)$, and $\hat{E}(\Psi)$ that enter the expression of $\sigma$ and $\epsilon$ depend only on $\Psi$ and are evaluated numerically on the equilibrium sequence.
283: 
284: This completes the derivation that allows us to draw the analogy with the classical paper of \citet{lyn68}. This step, straightforward for the $f^{(\nu)}$ models, is by itself interesting and new. In fact, other attempts at applying the paradigm of the gravothermal catastrophe to stellar dynamical equilibrium sequences were either based on an unjustified {\it ansatz} for the identification of the relevant temperature (e.g., see Appendix V in the article by \citealt{lyn68}; \citealt{kat80}; \citealt{mag98}) or on the use of non-standard entropies (for less realistic models; \citealt{cha02}).
285: 
286: 
287: \section{The high concentration regime: gravothermal catastrophe}
288: 
289: When the $f^{(\nu)}$ models were constructed \citep{sti87}, it was immediately realized that they have general
290: properties similar to those of the $f_{\infty}$ models \citep{ber84}; in particular, for values of $\nu \approx 1$,
291: sufficiently concentrated models along the sequence tend to settle
292: into a ``stable" overall structure, except for the development of
293: a more and more compact nucleus, as the value of $\Psi$ increases,
294: and are characterized by a projected density profile very well
295: fitted by the $R^{1/4}$ law characteristic of the surface
296: brightness profile of bright elliptical galaxies. This property is
297: illustrated in Fig.~{\ref{fig2}}.
298: 
299: Now, by inspection of Fig.~{\ref{fig1}} and by analogy with the study of the isothermal sphere \citep{lyn68}, we can identify the location at $\Psi \approx 9$ as the location for the {\it onset of the gravothermal catastrophe}. This sequence of models thus has the surprising result that the value of $\Psi$ that defines the onset of the gravothermal catastrophe is precisely that around which the models appear to become realistic representations of bright elliptical galaxies. We leave to other papers (see \citealt{ber93}) the detailed discussion of the issues that have to be addressed when comparison is made with the observations.
300: 
301: We note that in this regime of high concentration the general properties of the gravothermal catastrophe are reasonably well recovered by the use of the {\it ansatz} that the temperature parameter conjugate to the total energy is $a$, a quantity directly related to the velocity dispersion in the central regions. Basically, this was the {\it ansatz} made in the discussion of the possible occurrence of the gravothermal catastrophe for the King models or for other sequences of models (e.g., see \citealt{lyn68}, \citealt{kat80}, \citealt{mag98}). Here we have proved that the application of a rigorous derivation, which is available in our case, gives rise to relatively modest quantitative changes in the $(E_{tot},1/T)$ diagram for values of $\Psi$ close to and beyond the onset of the catastrophe (see Fig.~{\ref{fig3}}). However, in Sect.~6 we will draw the attention to an interesting, qualitatively new phenomenon missed in the previous derivations based on the use of the $a$-{\it ansatz}.
302: 
303: 
304: In passing, we note that in this regime of relatively high
305: concentrations, the $f^{(\nu)}$ models possess one intrinsic property
306: that makes them more appealing than the widely studied $f_{\infty}$
307: models. This is related to the way the models compare to the phase
308: space properties of the products of collisionless collapse, as
309: observed in N-body simulations \citep{van82}. In fact, one noted
310: unsatisfactory property of the concentrated $f_{\infty}$ models was
311: their excessive degree of isotropy with respect to the models produced
312: in the simulations.\footnote{ \citet{mer89} stressed this point and
313: thus argued that a better representation of N-body simulations would
314: be obtained by considering the $f_{\infty}$ family of models extended
315: to the case of negative values of the coefficient $a$ multiplying the
316: energy in the exponent. Unfortunately, their proposed solution, in
317: terms of models characterized by such a peculiar phase-space
318: structure, is unable to reproduce both the core velocity distribution
319: observed in numerical experiments and the modest amount of radial
320: anisotropy revealed by observed line profiles. In addition, their
321: proposed solution is not viable because for negative values of $a$ the
322: radial anisotropy level is so high that the models are violently
323: unstable, on an extremely short timescale, with the result that their
324: structural properties would be drastically changed by rapid evolution
325: \citep{sti91,ber94}.} Here we can easily check that the anisotropy
326: level of the concentrated $f^{(\nu)}$ models, while still within the
327: desired (radial orbit) stability boundary and still consistent with
328: the modest amount of radial anisotropy revealed by the observations,
329: seems much closer to that resulting from N-body simulations of
330: collisionless collapse; in particular, the anisotropy radius
331: $r_{\alpha}$, defined from the relation $\alpha (r_{\alpha}) = 1$,
332: with $\alpha = 2 - (\langle v^2_{\theta} \rangle + \langle v^2_{\phi}
333: \rangle )/ \langle v^2_r \rangle$, is close to the half-mass radius
334: $r_M$ (while for the $f_{\infty}$ models it is about three times as
335: large). This is illustrated in Fig.~{\ref{fig4}}. In any case, we
336: would like to emphasize that, while we have been clearly taking
337: inspiration from simulations of collisionless collapse, our main
338: interest is in comparing the structure of our models with that of
339: observed objects rather than in providing a detailed fit to the
340: results of N-body simulations.
341: 
342: \section{The intermediate concentration regime: the $R^{1/4}$ law and deviations from it}
343: 
344: The intermediate concentration regime (the precise point that
345: marks the low concentration regime will be identified in the next
346: Section) is a regime where the models appear to be stable, with
347: respect not only to the gravothermal catastrophe (following the
348: arguments provided earlier; but we should recall that the
349: catastrophe is expected to require a sufficiently high level of
350: effective collisionality in order to take off) but also to other
351: instabilities (see the discussion given by \citealt{ber94} and references therein). The relatively wide variations, between $\Psi = 3.5$ and
352: $\Psi = 9$, in all the representative quantities that characterize
353: the equilibrium models suggest that this part of the sequence
354: could be used to model the weak homology of bright elliptical
355: galaxies (see \citealt{ber02}), much like the sequence of King
356: models is able to capture observed systematic variations in the
357: structure of globular clusters (see \citealt{djo94}).
358: 
359: 
360: \section{The low concentration regime: negative global temperature and radial orbit instability}
361: 
362: The low concentration regime is marked by an unexpected and significant difference with respect to the low concentration limit of the classical isothermal sphere \citep{bon56,lyn68}. In fact, while the classical case reduces to the ideal non-gravitating gas, to which Boyle's law applies, for the $f^{(\nu)}$ models the system remains self-gravitating, although it develops a wide core in the density distribution. A clear-cut proof of this difference is given by inspecting the behavior of the global temperature $T$, identified from the thermodynamical definition $T = 1/(\partial{S}/\partial{E_{tot}})$. While the temperature defined by the {\it a-ansatz} remains obviously positive definite, by definition, if we look at Fig.~{\ref{fig1}} we see that the global temperature $T$ changes sign at $\Psi \approx 3.5$. This marks a drastic qualitative deviation from the classical studies.
363: 
364: Here we note a curious coincidence of this transition value of $\Psi$ with the value around which the sequence is bound to change its stability properties with respect to the radial orbit instability \citep{pol81}. Indeed, the location where the sequence is expected to become unstable in this regard is precisely that defined by $\Psi \approx 3.5$, as can be judged from inspection of Fig.~{\ref{fig4}}; around those values of $\Psi$ the level of radial anisotropy, as measured by $2 K_r/K_T$ reaches the threshold value of $1.8 - 2$, known to be sufficient for the excitation of the instability (the precise value of $2K_r/K_T$ corresponding to marginal stability is model dependent; for some sequences the reported value is below the range suggested by \citealt{pol81}).
365: 
366: These clues appear to be interesting and important, but more work is required before a final claim can be made that there is indeed a direct relation between the dynamical radial orbit instability and the fact that the system possesses a negative global temperature, as we found based on the simple work presented here.
367: 
368: \section{Conclusions}
369: 
370: A relatively straightforward and simple thermodynamical description of an equilibrium sequence constructed earlier and known to possess realistic characteristics with respect to bright elliptical galaxies shows that, on the one side, the paradigm of the gravothermal catastrophe may be adequate to explain the occurrence of realistic properties in models of collisionless stellar systems and, on the other side, a long-known dynamical instability might turn out to be interpreted in terms of a thermodynamical argument.
371: 
372: Probably the main open question regarding the models discussed in this paper, partially addressed in previous studies (see \citealt{sti87}), is to what extent the quantity $Q$ is actually reasonably well conserved during violent (partial) relaxation. It is likely that a thorough investigation of this issue may give indications that the quantity is best conserved only in certain ranges of $\nu$. Studies of this type, combined with other dynamical and thermodynamical considerations, may turn out to lead to the identification of a family of equilibrium models with optimal behavior with respect to statistical mechanics, with respect to what we know about collisionless collapse, and with respect to the problem of providing a realistic representation of bright elliptical galaxies.
373: 
374: 
375: \acknowledgments
376: 
377: We would like to thank L. Ciotti, M. Stiavelli, and T. van Albada for interesting discussions and suggestions. This work has been partially supported by MIUR of Italy.
378: 
379: \appendix
380: \section{Numerical integrations}
381: \paragraph{}
382: The $f^{(\nu)}$ models are constructed by solving the Poisson equation
383: \begin{equation} \label{eq:PoissonADIM}
384: \frac{1}{\hat{r}^2} \frac{d}{d \hat r} \hat{r} ^2 \frac{d}{d \hat r} \hat \Phi(\hat r) = \frac{1}{\gamma} \hat{\rho} (\hat{r},\hat{\Phi}),
385: \end{equation}
386: 
387: \noindent
388: in which $\gamma$ is considered as an eigenvalue to be determined by imposing the two natural boundary conditions
389: %%%%
390: $ \hat{\Phi}(0) = -\Psi $ and
391: $ \hat{\Phi}(\hat{r}) \sim - {\hat{M}(\hat{r})}/(4 \pi \gamma \hat{r}) $ as $\hat{r} \rightarrow \infty$.
392: %%%%
393: Here the symbol $\hat{}$  indicates that the quantity is suitably expressed in dimensionless form.
394: \paragraph{}
395: We have computed the two-dimensional integral for the density with an adaptive seven-point scheme \citep{bern91} in order to properly handle the presence of a peaked integrand for certain values of the pair $(\hat{r},\hat{\Phi})$. The Poisson equation has then been solved with a fourth order Runge-Kutta code by starting from $\hat{r}=0$ with a seed value for $\gamma$ and iterating the procedure until the boundary condition at large radii is matched within a certain accuracy. Finally, we have proceeded to calculate the global quantities $\hat{M}(\Psi)$, $\hat{Q}(\Psi)$, and $\hat{E}(\Psi)$ from their definitions.
396: \paragraph{}
397: In order to check the accuracy of the numerical integration we have performed the following tests:
398: (1) The virial theorem is satisfied with accuracy of the order $10^{-6}$ or better;
399: (2) The integrated mass (from its definition) and the mass derived from the asymptotic behaviour of the potential at large radii are the same with accuracy of the order $10^{-4}$;
400: (3) The expression for $\hat{\Phi}(\hat{r})$ at large radii to two significant orders in the relevant asymptotic expansion has been checked to be correct with an accuracy from $10^{-3}$ to $10^{-4}$;
401: (4) The asymptotic analysis allows us to estimate the contributions to $\hat{M}$, $\hat{Q}$, and $\hat{E}$  external to a sphere of large radius $R$; this has been checked to help improve the numerical determination of the relevant global quantities.
402: \paragraph{}
403: We estimate that the final relative error in the quantities along the equilibrium sequence is of the order of some parts times $10^{-4}$ for $\hat M$ and $\hat Q$ and some parts times $10^{-5}$ for $\hat E$. The total energy is less sensitive to the finite radius truncation error, due to its $1/R^{2}$ convergence. The propagation of these errors leads to the error bars plotted in Fig.~{\ref{fig1}}, where the entropy $\sigma$ is the more difficult to determine with good accuracy.
404: 
405: 
406: 
407: \begin{thebibliography}{}
408: 
409: \bibitem[Antonov(1962)]{ant62} Antonov, V.A. 1962, Vestnik Leningr. Univ., no. 19, 96 (Engl. Transl.: Structure and Dynamics of Elliptical Galaxies, T. de Zeeuw, Dordrecht: Reidel, 1986, 531)
410: 
411: \bibitem[Berntsen et al.(1991)]{bern91} Berntsen, J., Espelid T., and Genz, A. 1991, ACM Transactions on Mathematical Software, 17, 437
412: 
413: \bibitem[Bertin et al.(2002)]{ber02} Bertin, G., Ciotti, L., Del Principe, M. 2002, \aap, 386, 149
414: \bibitem[Bertin at al.(2002)]{ber02b} Bertin, G., Liseikina, T., and Pegoraro, F. 2002, paper submitted
415: 
416: \bibitem[Bertin et al.(1994)]{ber94} Bertin, G., Pegoraro, F., Rubini, F., Vesperini, E. 1994, \apj, 434, 94
417: \bibitem[Bertin and Stiavelli(1984)]{ber84} Bertin, G., and Stiavelli, M. 1984, \aap, 137, 26
418: \bibitem[Bertin and Stiavelli(1993)]{ber93} Bertin, G., and Stiavelli, M. 1993, Rep. Prog. Phys., 56, 493
419: \bibitem[Bonnor(1956)]{bon56} Bonnor, W.B. 1956, \mnras, 116, 351
420: \bibitem[Chandrasekhar and Woltjer(1958)]{cha58} Chandrasekhar, S., and Woltjer, L. 1958, Proc. Natl. Acad. Sci. USA, 44, 285
421: \bibitem[Chavanis(2002)]{cha02} Chavanis, P.H. 2002, \aap, 386, 732
422: \bibitem[Djorgovski and Meylan(1994)]{djo94} Djorgovski, S.,
423:     and Meylan, G. 1994, \aj, 108, 1292
424: \bibitem[Katz(1978)]{kat78} Katz, J. 1978, \mnras, 183, 765
425: 
426: \bibitem[Katz(1979)]{kat79} Katz, J. 1979, \mnras, 189, 817
427: 
428: \bibitem[Katz(1980)]{kat80} Katz, J. 1980, \mnras, 190, 497
429: \bibitem[King(1966)]{kin66} King, I. R.  1966, \aj, 71, 64
430: \bibitem[Lombardi and Bertin(2001)]{lom01} Lombardi, M., and Bertin, G. 2001,    \aap, 375, 1091
431: \bibitem[Lynden-Bell(1967)]{lyn67} Lynden-Bell, D. 1967, \mnras, 136, 101
432: \bibitem[Lynden-Bell and Eggleton(1980)]{lyn80} Lynden-Bell, D.,
433:     and Eggleton, P.P. 1980, \mnras, 191, 483
434: \bibitem[Lynden-Bell and Wood(1968)]{lyn68} Lynden-Bell, D.,
435:     and Wood, R. 1968, \mnras, 138, 495
436: \bibitem[Magliocchetti et al.(1998)]{mag98} Magliocchetti, M., Pucacco, G., and Vesperini, E. 1998, \mnras, 301, 25
437: \bibitem[Merritt et al.(1989)]{mer89} Merritt, D., Tremaine, S., and Johnstone, D. 1989, \mnras, 236, 829
438: \bibitem[Padmanabhan(1989)]{pad89} Padmanabhan, T. 1989, \apjs, 71, 751
439: 
440: \bibitem[Pellegrini(1999)]{pel99} Pellegrini, S. 1999, \aap, 351, 487
441: \bibitem[Polyachenko and Shukhman(1981)]{pol81} Polyachenko, V.L., and Shukhman, I.G. 1981, Sov. Astron., 25, 533
442: \bibitem[Spitzer(1987)]{spi87} Spitzer, L.  1987, Dynamical evolution of globular clusters, Princeton: Princeton University Press
443: \bibitem[Stiavelli and Bertin(1987)]{sti87} Stiavelli, M., and Bertin, G. 1987, \mnras, 229, 61
444: \bibitem[Stiavelli and Sparke(1991)]{sti91} Stiavelli, M., and Sparke, L.S. 1991, \apj, 382, 466
445: \bibitem[Taruya and Sakagami (2002)]{tar02} Taruya, A., and Sakagami, M. 2002, Physica A, 307, 185
446: \bibitem[Tremaine(1986)]{tre86a} Tremaine, S. in Structure and Dynamics of Elliptical Galaxies, T. de Zeeuw, Dordrecht: Reidel, 1986, 367
447: %\bibitem[Tremaine et al.(1986)]{tre86b} Tremaine, S., H\'{e}non, M., and Lynde%n-Bell, D. 1986, \mnras, 219, 285
448: \bibitem[van Albada(1982)]{van82} van Albada, T.S. 1982, \mnras, 201,
449: 939
450: \bibitem[Vesperini(1997)]{ves97} Vesperini, E. 1997, \mnras, 287, 915
451: \end{thebibliography}
452: 
453: \clearpage
454: 
455: \begin{figure}
456: \plotone{f1.eps}
457: \caption{Specific entropy and total energy along the equilibrium sequence of $f^{(\nu)}$ models with $\nu=1$ (as a function of the concentration parameter $\Psi$, at constant $M$ and $Q$, and thus expressed by means of the functions $\sigma(\Psi)$ and $\epsilon(\Psi)$ defined in the text). Note that for $\Psi \lesssim 3.5$ the models are characterized by a negative temperature, because the derivatives of $S$ and $E_{tot}$ have opposite signs.\label{fig1}}
458: \end{figure}
459: 
460: \clearpage
461: 
462: \begin{figure}
463: \plotone{f2.eps}
464: \caption{Residuals  $\mu^{(\nu)}-\mu^{1/4}$ obtained by fitting the $R^{1/4}$ law to the projected density profile of $f^{(\nu)} $ models for $\nu=1$ and some values of $\Psi$.\label{fig2}}
465: \end{figure}
466: 
467: \clearpage
468: 
469: \begin{figure}
470: \plotone{f3.eps}
471: \caption{Instability spiral of $f^{(\nu)}$ models with
472: $\nu=1$. The solid line refers to the results obtained with the
473: $a$-{\it ansatz} (with $\hat{a}= \gamma^{-4/5} \hat{M}^{(\nu+1)/(5
474: \nu)}\hat{Q}^{-4/(5 \nu)}$). Crosses represent the
475: global temperature from the definition $\partial S/
476: \partial E_{tot} $; other symbols indicate estimated points for which the adopted numerical differentiation is less reliable. The values of $\Psi$ and $\epsilon$ for points A and B with a vertical tangent remain unchanged.\label{fig3}}
477: \end{figure}
478: 
479: \clearpage
480: 
481: \begin{figure}
482: \plotone{f4.eps}
483: \caption{Anisotropy along the equilibrium sequence: anisotropy radius in units of the half mass radius $2 r_{\alpha}/r_M$ and the anisotropy parameter  $2K_r/K_T$ (ratio of total kinetic energy in the radial direction to that in the tangential directions) of $f^{(\nu)}$ models with $\nu=1$. \label{fig4}}
484: 
485: \end{figure}
486: 
487: 
488: \end{document}
489: 
490: