Skip to main content
Log in

Statistical mechanics of self-gravitating systems in general relativity: II. The classical Boltzmann gas

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

We study the statistical mechanics of classical self-gravitating systems confined within a box of radius R in general relativity. It has been found that the caloric curve \(T_{\infty }(E)\) has the form of a double spiral whose shape depends on the compactness parameter \(\nu =GNm/Rc^2\) (Roupas in Class Quantum Grav 32:135023, 2015; Alberti and Chavanis in arXiv:1908.10316). The double spiral shrinks as \(\nu \) increases and finally disappears when \(\nu _{{\mathrm{max}}}=0.1764\). Therefore, general relativistic effects render the system more unstable. On the other hand, the cold spiral and the hot spiral move away from each other as \(\nu \) decreases. Using a normalization \(\varLambda =-ER/GN^2m^2\) and \(\eta =GNm^2/R k_\mathrm{B} T_{\infty }\) appropriate to the nonrelativistic limit, and considering \(\nu \rightarrow 0\), the hot spiral goes to infinity and the caloric curve tends to a limit curve (determined by the Emden equation) exhibiting a single cold spiral, as found in former works. Using another normalization \({{{\mathcal {M}}}}=GM/Rc^2\) and \({{{\mathcal {B}}}}={Rc^4}/{GNk_\mathrm{B} T_{\infty }}\) appropriate to the ultrarelativistic limit, and considering \(\nu \rightarrow 0\), the cold spiral goes to infinity and the caloric curve tends to a limit curve (determined by the general relativistic Emden equation) exhibiting a single hot spiral. This result is new. We discuss the analogies and the differences between this asymptotic caloric curve and the caloric curve of the self-gravitating black-body radiation. Finally, we compare box-confined isothermal models with heavily truncated isothermal distributions in Newtonian gravity and general relativity.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. If the particles are fermions, the classical limit corresponds to the nondegenerate limit \(T\gg T_\mathrm{F}\) of the self-gravitating Fermi gas, where \(T_\mathrm{F}\) is the Fermi temperature. This is what we will assume here in order to make the connection with Paper I. However, our results are more general since they can also describe a gas of bosons in the classical limit \(T\gg T_\mathrm{c}\), where \(T_\mathrm{c}\) is the condensation temperature.

  2. A detailed history of the statistical mechanics of self-gravitating systems (classical and quantum) in Newtonian gravity and general relativity is given in Refs. [1, 29, 30].

  3. The Woolley [36] distribution, previously introduced by Eddington [38], corresponds to the isothermal distribution with positive energies removed. The King [37] distribution, previously introduced by Michie [39] in a more general form, is a lowered isothermal distribution such that the density of stars in phase space vanishes continuously above the escape energy.

  4. Truncated isothermal distributions of relativistic star clusters have been studied independently by Fackerell [59].

  5. The curve \(T_{\infty }(\rho _0)\) presents damped oscillations (see Fig. 2 of [58]).

  6. Ipser [63] (see also [62]) studied the dynamical stability with respect to the Vlasov–Einstein equations of heavily truncated isothermal distributions by using a variational principle based on the equation of pulsations derived by Ipser and Thorne [61]. By using a suitably chosen trial function, he numerically obtained an approximate expression of the square complex pulsation \(\omega _{\mathrm{app}}^2\) which, by construction, is always larger than the exact one \(\omega ^2\). He showed that \(\omega _{\mathrm{app}}^2>0\) for \(z<z_\mathrm{c}\sim 0.5\) and \(\omega _{\mathrm{app}}^2<0\) for \(z>z_\mathrm{c}\sim 0.5\). The transition (\(z_\mathrm{c}\sim 0.5\)) turns out to coincide with the turning point of fractional binding energy \(E/Nmc^2\). This proves that the system becomes unstable after the turning point of binding energy and suggests (but does not prove) that the system is stable before the turning point of binding energy. The dynamical stability of the system before the turning point of binding energy was proved later by Ipser [11]. He first showed that thermodynamical stability (in a very general sense) implies dynamical stability. Then, using the Poincaré [73] criterion, he showed (see also [10]) that the system is thermodynamically stable before the turning point of binding energy and thermodynamically unstable after the turning point of binding energy. This implies that the system is dynamically stable before the turning point of binding energy. On the other hand, since the system is dynamically unstable after the turning point of binding energy [63], Ipser [11] concluded that dynamical stability and thermodynamical stability coincide in general relativity. He conjectured that this equivalence between thermodynamical stability and dynamical stability remains valid for all isotropic star clusters, not only for those described by the heavily truncated Maxwell–Boltzmann distribution. This is in sharp contrast with the Newtonian case where it has been shown [52,53,54,55,56,57] that all isotropic stellar systems are dynamically stable with respect to the Vlasov–Poisson equations, even those that are thermodynamically unstable. To solve this apparent paradox, one expects that the growth rate \(\lambda \) of the dynamical instability decreases as relativity effects decrease and that it tends to zero in the nonrelativistic limit \(c\rightarrow +\infty \).

  7. Fackerell et al. [74] worked in terms of the curve \(E(z_0)\) which presents damped oscillations (see Fig. 2 of [63]). The caloric curve \(T_{\infty }(E)\) can be obtained from Table I of Ipser [63]. It is plotted in Fig. 5 of Appendix A and has the form of a spiral.

  8. The gravitational collapse of a supermassive star (fluid sphere) is a homologous radial infall of all the fluid [82, 83]. By contrast, the gravitational collapse of a collisionless relativistic star cluster is an inward spiralling of all the stars [78].

  9. Early works on relativistic star clusters were stimulated by the scenario of Hoyle and Fowler [84] that quasars could be supermassive star clusters. (They had previously considered the possibility that supermassive stars [85, 86] could be the energy sources for quasars and active galactic nuclei.) In their theory, the observed high redshifts of quasars (\(z > rsim 2\)) are explained by the fact that the clusters are very relativistic (gravitational redshift) instead of being far away (cosmological redshift). The fact that all relativistic star clusters studied by Ipser [62, 63] and Fackerell [64] were found to be unstable above \(z_\mathrm{c}\sim 0.5\) rapidly threw doubts on their scenario [87], privileging the scenario of Salpeter [88] and Zel’dovich [89] that supermassive black holes might be the objects responsible for the energetic activity of quasars and galactic nuclei. We now know that quasar redshifts have a cosmological origin. (They are far away.) We note, however, that there are examples of relativistic star clusters that are stable at any redshift. A first example was constructed by Bisnovatyi-Kogan and Zel’dovich [90] (see also Bisnovatyi-Kogan and Thorne [91]), but this cluster is singular, having infinite central density and infinite central redshift, so it is not very realistic. (In addition, the techniques for testing stability yielded inconclusive results.) Later, Rasio et al. [92] (see also Merafina and Ruffini [93]) reported a situation where the binding energy has no turning point so that there is no dynamical instability. Specifically, they followed the gravothermal catastrophe in the general relativistic regime and found that the binding energy decreases monotonically with the central redshift so that the clusters remain dynamically stable for all redshifts up to infinite central redshift \(z_0\rightarrow +\infty \). These clusters could represent the relativistic final states of initially Newtonian clusters undergoing the gravothermal catastrophe. Remark According to these results, it is not quite clear if the initial scenario proposed by Shapiro and Teukolsky [75,76,77,78,79,80,81] is correct. Indeed, in their early works [75,76,77,78,79,80,81] they assumed that the gravothermal catastrophe transforms a Newtonian star cluster into a relativistic one described by a truncated isothermal distribution and then showed that above a critical central redshift \(z_\mathrm{c}\sim 0.5\), this distribution undergoes a dynamical instability of general relativistic origin and collapses towards a black hole. However, in their later work (with Rasio) [92] they found that the relativistic generalization of the distribution function produced by the gravothermal catastrophe [48] differs from the truncated isothermal distribution and that it remains always dynamically stable up to infinite central redshift. This seems to preclude the formation of a black hole from the gravothermal catastrophe. This problem may be solved by the refined scenario developed by Balberg et al. [94] (see below).

  10. The analogies and the differences between box-confined isothermal models and heavily truncated isothermal distributions in Newtonian gravity and general relativity are discussed in Appendix B. We note that the presence of a box can substantially change the behaviour of the caloric curve and alter the physics of the problem. For nonrelativistic systems, the caloric curve of box-confined isothermal systems is relatively similar to the caloric curve of the truncated isothermal (King or Woolley) model. However, for general relativistic systems, they are very different from each other. The “box” may not be justified in that case.

  11. As recalled in the Introduction, the formalism developed in Paper I is valid for an arbitrary form of entropy.

  12. Here and in the following, (I-x) refers to Eq. (x) of Paper I.

  13. For a classical gas, the relation (24) is always valid locally with the local temperature T(r). The isentropic (adiabatic) condition \(s/n=\mathrm{cst}\) corresponds to \(n\lambda _T^3=\mathrm{cst}\), i.e. \(n(r)/T(r)^{3/2}=\mathrm{cst}\). For a gas at statistical equilibrium, as considered here, \(T(r)=T\).

  14. This result is valid for an arbitrary function E(p).

  15. We have not included the usual factor 2 in the effective Schwarzschild radius \(R_S^*=GNm/c^2\) because the relation (91) essentially results from dimensional analysis and does not have the same status as the fundamental Schwarzschild radius \(R_S=2GM/c^2\) of a black hole [137, 138].

  16. This quantity was denoted \(\alpha \) in Ref. [120]. Here, we use the notation a instead of \(\alpha \) to avoid confusion with the variable \(\alpha =\mu /k_\mathrm{B} T\) introduced in Sect. 2.

  17. It should not be confused with the chemical potential. (The notation \(\mu \) introduced in [126] is somewhat unfortunate.) The degeneracy parameter plays an important role for self-gravitating fermions as it determines the shape of their caloric curves [126]. For classical particles, or for fermions in the nondegenerate limit, it just appears as an additive constant in the Boltzmann entropy [see Eq. (119)] so it can be omitted in most applications. We note that the classical limit corresponds formally to \(\mu \rightarrow +\infty \). In this sense, the entropy \(S/Nk_\mathrm{B}\) diverges as \(\ln \mu \) (see the Remark at the beginning of Sect. 2).

  18. The connection between the sign of the specific heat and the instability in the canonical and microcanonical ensembles is explained in Appendix B of [30].

  19. As shown in Appendix A of Paper I, this result is valid for an arbitrary distribution function in the ultrarelativistic limit.

  20. This quantity was denoted \(\alpha \) in Ref. [17]. Here, we use the notation a instead of \(\alpha \) to avoid confusion with the variable \(\alpha =\mu /k_\mathrm{B} T\) introduced in Sect. 3.

  21. We note that globular clusters become rapidly canonically unstable, i.e. as soon as they are substantially different from a polytrope \(n=5/2\) (see the discussion in [32]). This is a clear sign of the fact that real globular clusters are described by the microcanonical ensemble in which they are stable longer. Indeed, if they were described by the canonical ensemble, most of the observed globular clusters would be unstable since they substantially differ from polytropes \(n=5/2\).

  22. This can be understood from different arguments: (i) under the effect of close encounters, stars leave the system with an energy positive or close to zero. Therefore, the energy of the cluster decreases or remains approximately constant. Since the number of stars in the cluster decreases, the cluster contracts (according to the virial theorem) and becomes more and more concentrated. (ii) For globular clusters described by the King model, one can show that the Boltzmann entropy \(S_\mathrm{B}\) is an increasing function of the concentration parameter k until a point \(k_{*}\) at which the Boltzmann entropy reaches a maximum before decreasing (see Table II of [40], Fig. 5 of [48], and Fig. 46 of [97]). Therefore, we can relate the temporal increase of the concentration parameter k(t) on the series of equilibria with the second principle of thermodynamics, i.e. the temporal increase of the Boltzmann entropy \({\dot{S}}_\mathrm{B}\ge 0\) (H-theorem). This adiabatic evolution continues (at most) until the point \(k_{*}\) at which the Boltzmann entropy is maximum since the Boltzmann entropy cannot decrease with time. At the instability point \(k_{\mathrm{MCE}}\), the system becomes unstable, undergoes the gravothermal catastrophe, and evolves away from the series of equilibria. The instability point \(k_{\mathrm{MCE}}\), corresponding to the first extremum of the King entropy S (defined in [97]) or equivalently to the first turning point of energy (since \(\delta S=\beta \delta E\)), occurs a bit sooner than the point \(k_*\) at which the Boltzmann entropy is maximum (see the discussion in [97]). It is a bit disturbing to note that the King entropy decreases as the concentration parameter increases up to \(k_\mathrm{MCE}\) along the series of equilibria (see Fig. 46 of [97]). There is, however, no paradox since the H theorem applies to the Boltzmann entropy not to the King entropy. On the other hand, for box-confined systems, the Boltzmann entropy is a decreasing function of the density contrast (see Fig 3 of [149]) but, in that case, the system does not evolve along the series of equilibria so there is no paradox either.

  23. They mentioned that the damped oscillations of \(T(\rho _0)\) are similar to the damped oscillations of \(M(\rho _0)\) for neutron stars discovered by Dmitriev and Kholin [150]. This is because, in the ultrarelativistic limit where the density (or the redshift) is large, the equation of state of a classical isothermal gas takes the form \(P=\epsilon /3=K n^{4/3}\), where \(\epsilon \) is the energy density and n the particle number, like the ultrarelativistic equation of state of a Fermi gas at \(T=0\) (or like the black-body radiation).

  24. Zel’dovich and Podurets [58] assumed a certain relation between the energy cut-off and the temperature. This relatively ad hoc choice was later criticized. This led to several generalizations of the problem by Katz et al. [6], Suffern and Fackerell [66], Fackerell and Suffern [67], Merafina and Ruffini [68,69,70], and Bisnovatyi-Kogan et al. [71, 72] that we do not review in detail here.

  25. The same is true in Newtonian gravity for the isotropic Wilson model which presents a similar series of equilibria (see Fig. 1 of [41]). This is confirmed by the fact that for \(k\rightarrow 0\), the Wilson model is equivalent to a polytrope of index \(n=7/2>3\) that is canonically unstable (the corresponding gaseous spheres are dynamically unstable with respect to the Euler-Poisson equations) while being microcanonically stable (see Ref. [139] for more details).

  26. Actually, this remains a conjecture because it has not been proven mathematically that isothermal spheres become unstable after the turning point of fractional binding energy. This is only a numerical result valid for heavily truncated isothermal distributions with a certain relation between the energy cut-off and the temperature. Furthermore, the main open question is to know whether this result is true for all isotropic distribution functions, i.e. if a collisionless relativistic star cluster always becomes dynamically unstable after the turning point of binding energy. This property has been observed numerically for all the distributions functions that have been considered [63, 76], but there is no rigorous proof of this result in general.

  27. This property immediately results from the equilibrium scalar virial theorem \(2K+W=0\) for an unbounded self-gravitating system implying \(E=K+W=(1/2)W=-K<0\).

  28. We recall that the isothermal model studied by Zel’dovich and Podurets [58] and Isper [63] is based on a certain ad hoc relation between the energy cut-off and the temperature (see footnote 24). This is why their series of equilibria is unique. If we relax this assumption, we get a family of caloric curves as investigated by [6, 66,67,68,69,70,71,72]. For box-confined isothermal models [30, 128], we also have a family of caloric curves parametrized by the compactness parameter \(\nu =GNm/Rc^2\). These considerations explain why the complete caloric curve of Fig. 6 depends on the parameter \(\nu \).

  29. It is interesting to note that the caloric curve of Fig. 6 for classical truncated isothermal star clusters resembles the caloric curve in Fig. 29 and 5 of [29, 151] for self-gravitating fermions confined within a box in general relativity.

  30. In general relativity, equilibrium states of star clusters may have a positive energy (as can be seen in Figs. 5 and 6), while this is not possible in Newtonian gravity. (The equilibrium virial theorem \(2K+W=0\) implies \(E=K+W=(1/2)W=-K<0\).) This observation was first made by Zel’dovich [152]. However, the equilibrium states with \(E>0\) are generally unstable.

References

  1. P.H. Chavanis, arXiv:1908.10806 (Paper I) (in press)

  2. R.C. Tolman, Phys. Rev. 35, 904 (1930)

    ADS  Google Scholar 

  3. W.J. Cocke, Ann. I.H.P. 4, 283 (1965)

    Google Scholar 

  4. G. Horwitz, J. Katz, Ann. Phys. USA 76, 301 (1973)

    ADS  Google Scholar 

  5. J. Katz, G. Horwitz, Astrophys. J. 194, 439 (1974)

    ADS  MathSciNet  Google Scholar 

  6. J. Katz, G. Horwitz, M. Klapisch, Astrophys. J. 199, 307 (1975)

    ADS  Google Scholar 

  7. J. Katz, Y. Manor, Phys. Rev. D 12, 956 (1975)

    ADS  MathSciNet  Google Scholar 

  8. J. Katz, G. Horwitz, Astrophys. J. 33, 251 (1977)

    ADS  Google Scholar 

  9. G.W. Gibbons, S.W. Hawking, Phys. Rev. D 15, 2752 (1977)

    ADS  Google Scholar 

  10. G. Horwitz, J. Katz, Astrophys. J. 223, 311 (1978)

    ADS  Google Scholar 

  11. J.R. Ipser, Astrophys. J. 238, 1101 (1980)

    ADS  MathSciNet  Google Scholar 

  12. R.D. Sorkin, R.M. Wald, Z.Z. Jiu, Gen. Relativ. Gravit. 13, 1127 (1981)

    ADS  Google Scholar 

  13. W.M. Suen, K. Young, Phys. Rev. A 35, 406 (1987)

    ADS  Google Scholar 

  14. W.M. Suen, K. Young, Phys. Rev. A 35, 411 (1987)

    ADS  Google Scholar 

  15. N. Bilic, R.D. Viollier, Gen. Relativ. Gravit. 31, 1105 (1999)

    ADS  Google Scholar 

  16. P.H. Chavanis, Astron. Astrophys. 381, 709 (2002)

    ADS  Google Scholar 

  17. P.H. Chavanis, Astron. Astrophys. 483, 673 (2008)

    ADS  Google Scholar 

  18. S. Gao, Phys. Rev. D 84, 104023 (2011)

    ADS  Google Scholar 

  19. S. Gao, Phys. Rev. D 85, 027503 (2012)

    ADS  Google Scholar 

  20. Z. Roupas, Class. Quantum Gravity 30, 115018 (2013)

    ADS  MathSciNet  Google Scholar 

  21. S.R. Green, J.S. Schiffrin, R.M. Wald, Class. Quantum Gravity 31, 035023 (2014)

    ADS  Google Scholar 

  22. X. Fang, S. Gao, Phys. Rev. D 90, 044013 (2014)

    ADS  Google Scholar 

  23. Z. Roupas, Class. Quantum Gravity 32, 119501 (2015)

    ADS  MathSciNet  Google Scholar 

  24. J.S. Schiffrin, Class. Quantum Gravity 32, 185011 (2015)

    ADS  MathSciNet  Google Scholar 

  25. K. Prabhu, J.S. Schiffrin, R.M. Wald, Class. Quantum Gravity 33, 185007 (2016)

    ADS  Google Scholar 

  26. X. Fang, X. He, J. Jing, Eur. Phys. J. C 77, 893 (2017)

    ADS  Google Scholar 

  27. J.R. Oppenheimer, G.M. Volkoff, Phys. Rev. 55, 374 (1939)

    ADS  Google Scholar 

  28. O. Klein, Rev. Mod. Phys. 21, 531 (1949)

    ADS  Google Scholar 

  29. G. Alberti, P.H. Chavanis, arXiv:1808.01007

  30. G. Alberti, P.H. Chavanis, arXiv:1908.10316

  31. P.-H. Chavanis, Astron. Astrophys. 556, A93 (2013)

    ADS  Google Scholar 

  32. P.H. Chavanis, M. Lemou, F. Méhats, Phys. Rev. D 91, 063531 (2015)

    ADS  Google Scholar 

  33. S. Chandrasekhar, Principles of Stellar Dynamics (University of Chicago Press, Chicago, 1942)

    MATH  Google Scholar 

  34. V.A. Ambartsumian, Ann. Leningr. State Univ. 22, 19 (1938)

    Google Scholar 

  35. L. Spitzer, Mon. Not. R. Astron. Soc. 100, 396 (1940)

    ADS  Google Scholar 

  36. R. Woolley, Mon. Not. R. Astron. Soc. 114, 191 (1954)

    ADS  Google Scholar 

  37. I. King, Astron. J. 71, 64 (1966)

    ADS  Google Scholar 

  38. A.S. Eddington, Mon. Not. R. Astron. Soc. 76, 572 (1916)

    ADS  Google Scholar 

  39. R.W. Michie, Mon. Not. R. Astron. Soc. 125, 127 (1963)

    ADS  MathSciNet  Google Scholar 

  40. D. Lynden-Bell, R. Wood, Mon. Not. R. Astron. Soc. 138, 495 (1968)

    ADS  Google Scholar 

  41. J. Katz, Mon. Not. R. Astron. Soc. 190, 497 (1980)

    ADS  Google Scholar 

  42. V.A. Antonov, Vest. Leningr. Gos. Univ. 7, 135 (1962)

    Google Scholar 

  43. R.B. Larson, Mon. Not. R. Astron. Soc. 147, 323 (1970)

    ADS  Google Scholar 

  44. R.B. Larson, Mon. Not. R. Astron. Soc. 150, 93 (1970)

    ADS  Google Scholar 

  45. M. Hénon, Astrophys. Space Sci. 13, 284 (1971)

    ADS  Google Scholar 

  46. I. Hachisu, Y. Nakada, K. Nomoto, D. Sugimoto, Prog. Theor. Phys. 60, 393 (1978)

    ADS  Google Scholar 

  47. D. Lynden-Bell, P.P. Eggleton, Mon. Not. R. Astron. Soc. 191, 483 (1980)

    ADS  Google Scholar 

  48. H. Cohn, Astrophys. J. 242, 765 (1980)

    ADS  Google Scholar 

  49. S. Inagaki, D. Lynden-Bell, Mon. Not. R. Astron. Soc. 205, 913 (1983)

    ADS  Google Scholar 

  50. D. Sugimoto, E. Bettwieser, Mon. Not. R. Astron. Soc. 204, 19 (1983)

    ADS  Google Scholar 

  51. D. Heggie, N. Ramamani, Mon. Not. R. Astron. Soc. 237, 757 (1989)

    ADS  Google Scholar 

  52. J.P. Doremus, M.R. Feix, G. Baumann, Phys. Rev. Lett. 26, 725 (1971)

    ADS  Google Scholar 

  53. J.P. Doremus, M.R. Feix, G. Baumann, Astron. Astrophys. 29, 401 (1973)

    ADS  Google Scholar 

  54. D. Gillon, M. Cantus, J.P. Doremus, G. Baumann, Astron. Astrophys. 50, 467 (1976)

    ADS  Google Scholar 

  55. J.F. Sygnet, G. Des Forets, M. Lachieze-Rey, R. Pellat, Astrophys. J. 276, 737 (1984)

  56. H. Kandrup, J.F. Sygnet, Astrophys. J. 298, 27 (1985)

    ADS  MathSciNet  Google Scholar 

  57. H. Kandrup, Astrophys. J. 370, 312 (1991)

    ADS  MathSciNet  Google Scholar 

  58. Y.B. Zel’dovich, M.A. Podurets, Sov. Astron. AJ 9, 742 (1966)

    ADS  Google Scholar 

  59. E.D. Fackerell, Ph.D. thesis, University of Sydney (1966)

  60. E. Fackerell, Astrophys. J. 153, 643 (1968)

    ADS  Google Scholar 

  61. J.R. Ipser, K.S. Thorne, Astrophys. J. 154, 251 (1968)

    ADS  MathSciNet  Google Scholar 

  62. J.R. Ipser, Astrophys. J. 156, 509 (1969)

    ADS  MathSciNet  Google Scholar 

  63. J.R. Ipser, Astrophys. J. 158, 17 (1969)

    ADS  MathSciNet  Google Scholar 

  64. E. Fackerell, Astrophys. J. 160, 859 (1970)

    ADS  MathSciNet  Google Scholar 

  65. A.W. Sudbury, Mon. Not. R. Astron. Soc. 147, 187 (1970)

    ADS  Google Scholar 

  66. K. Suffern, E. Fackerell, Astrophys. J. 203, 477 (1976)

    ADS  Google Scholar 

  67. E. Fackerell, K. Suffern, Aust. J. Phys. 29, 311 (1976)

    ADS  Google Scholar 

  68. M. Merafina, R. Ruffini, Astron. Astrophys. 221, 4 (1989)

    ADS  Google Scholar 

  69. M. Merafina, R. Ruffini, Europhys. Lett. 9, 621 (1989)

    ADS  Google Scholar 

  70. M. Merafina, R. Ruffini, Astron. Astrophys. 227, 415 (1990)

    ADS  Google Scholar 

  71. G.S. Bisnovatyi-Kogan, M. Merafina, R. Ruffini, E. Vesperini, Astrophys. J. 414, 187 (1993)

    ADS  Google Scholar 

  72. G.S. Bisnovatyi-Kogan, M. Merafina, R. Ruffini, E. Vesperini, Astrophys. J. 500, 217 (1998)

    ADS  Google Scholar 

  73. H. Poincaré, Acta Math. 7, 259 (1885)

    MathSciNet  Google Scholar 

  74. D. Fackerell, J. Ipser, K. Thorne, Comments on Astrophysics and Space Physics, vol. 1 (Gordon and Breach, New York, 1969), p. 134

    Google Scholar 

  75. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 298, 34 (1985)

    ADS  Google Scholar 

  76. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 298, 58 (1985)

    ADS  Google Scholar 

  77. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 292, L41 (1985)

    ADS  Google Scholar 

  78. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 307, 575 (1986)

    ADS  Google Scholar 

  79. C.S. Kochanek, S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 320, 73 (1987)

    ADS  Google Scholar 

  80. F. Rasio, S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 344, 146 (1989)

    ADS  Google Scholar 

  81. S.L. Shapiro, S.A. Teukolsky, Philos. Trans. R. Soc. Lond. A 340, 365 (1992)

    ADS  Google Scholar 

  82. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 234, L177 (1979)

    ADS  Google Scholar 

  83. S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 235, 199 (1980)

    ADS  Google Scholar 

  84. F. Hoyle, W.A. Fowler, Nature 213, 373 (1967)

    ADS  Google Scholar 

  85. F. Hoyle, W.A. Fowler, Mon. Not. R. Astron. Soc. 125, 169 (1963)

    ADS  Google Scholar 

  86. F. Hoyle, W.A. Fowler, Nature 197, 533 (1963)

    ADS  Google Scholar 

  87. H.S. Zapolsky, Astrophys. J. 153, L163 (1968)

    ADS  Google Scholar 

  88. E.E. Salpeter, Astrophys. J. 140, 796 (1964)

    ADS  Google Scholar 

  89. Ya. B. Zel’dovich, Sov. Phys. Dokl. 9, 195 (1964)

  90. G.S. Bisnovatyi-Kogan, Ya B. Zel’dovich, Astrofizika 5, 223 (1969)

    ADS  Google Scholar 

  91. G.S. Bisnovatyi-Kogan, K.S. Thorne, Astrophys. J. 160, 875 (1970)

    ADS  Google Scholar 

  92. F. Rasio, S.L. Shapiro, S.A. Teukolsky, Astrophys. J. 336, L63 (1989)

    ADS  Google Scholar 

  93. M. Merafina, R. Ruffini, Astrophys. J. 454, L89 (1995)

    ADS  Google Scholar 

  94. S. Balberg, S.L. Shapiro, S. Inagaki, Astrophys. J. 568, 475 (2002)

    ADS  Google Scholar 

  95. S. Balberg, S.L. Shapiro, Phys. Rev. Lett. 88, 101301 (2002)

    ADS  Google Scholar 

  96. J. Pollack, D. Spergel, P. Steinhardt, Astrophys. J. 804, 131 (2015)

    ADS  Google Scholar 

  97. P.H. Chavanis, M. Lemou, F. Méhats, Phys. Rev. D 92, 123527 (2015)

    ADS  Google Scholar 

  98. P.H. Chavanis, Phys. Rev. D 100, 083022 (2019)

    ADS  Google Scholar 

  99. J.R. Ipser, Astrophys. J. 193, 463 (1974)

    ADS  Google Scholar 

  100. L. Taff, H. van Horn, Astrophys. J. 197, L23 (1975)

    ADS  Google Scholar 

  101. G. Horwitz, J. Katz, Astrophys. J. 211, 226 (1977)

    ADS  MathSciNet  Google Scholar 

  102. Y. Nakada, Publ. Astron. Soc. Jpn. 30, 57 (1978)

    ADS  Google Scholar 

  103. I. Hachisu, D. Sugimoto, Progr. Theor. Phys. 60, 123 (1978)

    ADS  Google Scholar 

  104. G. Horwitz, J. Katz, Astrophys. J. 222, 941 (1978)

    ADS  Google Scholar 

  105. J. Katz, Mon. Not. R. Astron. Soc. 183, 765 (1978)

    ADS  Google Scholar 

  106. J.R. Ipser, G. Horwitz, Astrophys. J. 232, 863 (1979)

    ADS  MathSciNet  Google Scholar 

  107. S. Inagaki, Publ. Astron. Soc. Jpn. 32, 213 (1980)

    ADS  Google Scholar 

  108. M. Lecar, J. Katz, Astrophys. J. 243, 983 (1981)

    ADS  Google Scholar 

  109. J. Messer, H. Spohn, J. Stat. Phys. 29, 561 (1982)

    ADS  Google Scholar 

  110. J.F. Luciani, R. Pellat, Astrophys. J. 317, 241 (1987)

    ADS  MathSciNet  Google Scholar 

  111. M. Kiessling, J. Stat. Phys. 55, 203 (1989)

    ADS  Google Scholar 

  112. T. Padmanabhan, Astrophys. J. Suppl. 71, 651 (1989)

    ADS  Google Scholar 

  113. T. Padmanabhan, Phys. Rep. 188, 285 (1990)

    ADS  MathSciNet  Google Scholar 

  114. H.J. de Vega, N. Sanchez, F. Combes, Phys. Rev. D 54, 6008 (1996)

    ADS  Google Scholar 

  115. B. Semelin, H.J. de Vega, N. Sanchez, F. Combes, Phys. Rev. D 59, 125021 (1999)

    ADS  Google Scholar 

  116. J. Katz, I. Okamoto, Mon. Not. R. Astron. Soc. 317, 163 (2000)

    ADS  Google Scholar 

  117. B. Semelin, N. Sanchez, H.J. de Vega, Phys. Rev. D 63, 084005 (2001)

    ADS  Google Scholar 

  118. H.J. de Vega, N. Sanchez, Nucl. Phys. B 625, 409 (2002)

    ADS  Google Scholar 

  119. H.J. de Vega, N. Sanchez, Nucl. Phys. B 625, 460 (2002)

    ADS  Google Scholar 

  120. P.H. Chavanis, Astron. Astrophys. 381, 340 (2002)

    ADS  Google Scholar 

  121. P.H. Chavanis, C. Rosier, C. Sire, Phys. Rev. E 66, 036105 (2002)

    ADS  MathSciNet  Google Scholar 

  122. C. Sire, P.H. Chavanis, Phys. Rev. E 66, 046133 (2002)

    ADS  MathSciNet  Google Scholar 

  123. P.H. Chavanis, Astron. Astrophys. 401, 15 (2003)

    ADS  Google Scholar 

  124. J. Katz, Found. Phys. 33, 223 (2003)

    MathSciNet  Google Scholar 

  125. P.H. Chavanis, Astron. Astrophys. 432, 117 (2005)

    ADS  Google Scholar 

  126. P.H. Chavanis, Int. J. Mod. Phys. B 20, 3113 (2006)

    ADS  Google Scholar 

  127. M. Sormani, G. Bertin, Astron. Astrophys. 552, A37 (2013)

    ADS  Google Scholar 

  128. Z. Roupas, Class. Quantum Gravity 32, 135023 (2015)

    ADS  MathSciNet  Google Scholar 

  129. C. Sire, P.H. Chavanis, Phys. Rev. E 69, 066109 (2004)

    ADS  Google Scholar 

  130. J.W. Gibbs, Elementary Principles of Statistical Mechanics (Yale University Press, New Haven, 1902)

    MATH  Google Scholar 

  131. R. Balescu, Equilibrium and Nonequilibrium Statistical Mechanics (Wiley, Hoboken, 1975)

    MATH  Google Scholar 

  132. L. Boltzmann, Wien. Ber. 66, 275 (1872)

    Google Scholar 

  133. F. Juttner, Ann. Phys. 339, 856 (1911)

    Google Scholar 

  134. F. Juttner, Ann. Phys. 340, 145 (1911)

    Google Scholar 

  135. M. Planck, Ann. Phys. 26, 1 (1908)

    Google Scholar 

  136. S. Chandrasekhar, An Introduction to the Theory of Stellar Structure (University of Chicago Press, Chicago, 1939)

    MATH  Google Scholar 

  137. K. Schwarzschild, Berliner Sitzungsbesichte, 189 (1916)

  138. K. Schwarzschild, Berliner Sitzungsbesichte, 424 (1916)

  139. P.H. Chavanis, Astron. Astrophys. 451, 109 (2006)

    ADS  Google Scholar 

  140. A. Campa, P.H. Chavanis, J. Stat. Mech. 06, 06001 (2010)

    Google Scholar 

  141. S. Chandrasekhar, in General Relativity, Papers in Honour of J.L. Synge, ed. by L. O’Raifeartaigh (Oxford University Press, Oxford, 1972)

    Google Scholar 

  142. N. Riazi, M.R. Bordbar, Int. J. Theor. Phys. 45, 495 (2006)

    Google Scholar 

  143. F.M. Araujo, C.B.M.H. Chirenti, arXiv:1102.2393

  144. R. Singh, N. Das, J. Kumar, Eur. Phys. J. Plus 132, 251 (2017)

    Google Scholar 

  145. R.A. El-Nabulsi, J. Anal. 25, 301 (2017)

    MathSciNet  Google Scholar 

  146. R. Singh, Eur. Phys. J. Plus 133, 320 (2018)

    Google Scholar 

  147. M.I. Nouh, E.A.-B. Abdel-Salam, Eur. Phys. J. Plus 133, 149 (2018)

    Google Scholar 

  148. R. Singh, J. Shahni, H. Garg, A. Garg, Eur. Phys. J. Plus 134, 548 (2019)

    Google Scholar 

  149. P.H. Chavanis, Phys. Rev. E 65, 056123 (2002)

    ADS  Google Scholar 

  150. N.A. Dmitriev, S.A. Kholin, Voprosy Kosmogonii, vol. 9 (Izd-vo Akademii nauk SSSR, Moscow, 1963), p. 254

    Google Scholar 

  151. P.H. Chavanis, G. Alberti, Phys. Lett. B 801, 135155 (2020)

    MathSciNet  Google Scholar 

  152. Ya. B. Zel’dovich, Sov. Phys. JETP 15, 1158 (1962)

  153. Y. Pomeau, M. Le Berre, P.H. Chavanis, B. Denet, Eur. Phys. J. E 37, 26 (2014)

    Google Scholar 

  154. P.H. Chavanis, B. Denet, M. Le Berre, Y. Pomeau, Eur. Phys. J. B 92, 271 (2019)

    ADS  Google Scholar 

  155. P.H. Chavanis, B. Denet, M. Le Berre, Y. Pomeau, Europhys. Lett. 129, 30003 (2020)

    ADS  Google Scholar 

  156. Z. Roupas, Symmetry 11, 1435 (2019)

    Google Scholar 

  157. Z. Roupas, Universe 5, 12 (2019)

    ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre-Henri Chavanis.

Appendices

Appendix A: Series of equilibria of truncated isothermal distributions

In this Appendix, we discuss the series of equilibria of truncated isothermal distributions in Newtonian gravity and general relativity.

1.1 Nonrelativistic systems

The series of equilibria of globular clusters described by truncated isothermal distributions (Woolley [36] and King [37] models) have been determined by Lynden-Bell and Wood [40], Katz [41] and Chavanis et al. [32]. The caloric curve \(\beta (E)\) of the King model is reproduced in Fig. 4. It has the form of a spiral. It is parametrized by the concentration parameter k that increases monotonically along the series of equilibria. The curves \(\beta (k)\) and E(k) giving the inverse temperature and the energy as a function of the concentration parameter k display damped oscillations [32].

The thermodynamical stability of truncated isothermal distributions was analysed in Refs. [32, 40, 41] by using the Poincaré turning point criterion [73]. For \(E\rightarrow 0^-\) and \(\beta \rightarrow 0\), we know that the system is stable because it is equivalent to a polytrope of index \(n=5/2\) that is both canonically and microcanonically stable [139]. In the canonical ensemble (fixed temperature), the series of equilibria is stable up to the first turning point of temperature (corresponding to \(k_{{\mathrm{CE}}}=1.34\)) and becomes unstable afterwards. This is when the specific heat becomes infinite, passing from positive to negative values. In the microcanonical ensemble (fixed energy), the series of equilibria is stable up to the first turning point of energy (corresponding to \(k_{\mathrm{MCE}}=7.44\)) and becomes unstable afterwards. This is when the specific heat vanishes, passing from negative to positive values. The statistical ensembles are inequivalent in the region of negative specific heat (\(C<0\)), between the turning point of temperature CE and the turning point of energy MCE. From general arguments, it can be shown that canonical stability implies microcanonical stability [139, 140]. Basically, this is because the microcanonical ensemble is more constrained, hence more stable, than the canonical ensemble. In the present case, this manifests itself (in conjunction with the Poincaré theory) by the fact that the turning point of temperature occurs before the turning point of energy. For isolated stellar systems, only the microcanonical ensemble makes sense physically.Footnote 21

Let us now consider the dynamical stability of the system with respect to a collisionless evolution described by the Vlasov–Poisson equations. In Newtonian gravity, it has been shown that all isotropic stellar systems with a distribution function of the form \(f=f(\epsilon )\) with \(f'(\epsilon )<0\) are dynamically stable [52,53,54,55,56,57]. Therefore, the whole series of equilibria of isothermal stellar systems is dynamically stable, even the equilibrium states deep into the spiral that are thermodynamically unstable. We know from general arguments that thermodynamical stability implies dynamical stability [106, 140]. The results of [52,53,54,55,56,57] show that the converse is wrong in Newtonian gravity: before the first turning point of energy, the system is both thermodynamically stable (in the microcanonical ensemble) and dynamically stable; after the first turning point of energy, the system is thermodynamically unstable while it is still dynamically stable.

Fig. 4
figure 4

Series of equilibria (caloric curve) of the classical King model. The units for the curve \(\beta (E)\) are \((m v_0^{2})^{-1}\) for \(\beta \) and \(Mv_0^2\) for E, where \(v_0\) is a typical velocity defined in [32, 41]. The series of equilibria is parametrized by the dimensionless concentration parameter \(k=\beta [\varPhi (R)-\varPhi (0)]\). It has a snail-like (spiral) structure, but only the part of the curve up to CE (\(k_{\mathrm{CE}}=1.34\)) is stable in the canonical ensemble and only the part up to MCE (\(k_{\mathrm{MCE}}=7.44\)) is stable in the microcanonical ensemble. (The region between CE and MCE where the specific heat is negative corresponds to a region of ensembles inequivalence.) For \(k\rightarrow 0\), the King model is equivalent to a polytrope of index \(n=5/2\) [32, 41] that is both canonically and microcanonically stable [139]. (The same is true for the Wooley model which is equivalent, for \(k\rightarrow 0\), to a polytrope \(n=3/2\) [41].) The equilibrium states are all dynamically (Vlasov) stable

We also know from general arguments that the thermodynamical stability of a stellar system in the canonical ensemble is equivalent to the dynamical stability of the corresponding barotropic star with respect to the Euler–Poisson equations [139]. Therefore, barotropic stars with the equation of state corresponding to the truncated isothermal distribution function are dynamically stable before the first turning point of temperature and dynamically unstable after the first turning point of temperature.

The dynamical evolution of globular clusters is discussed in the introduction (see also [30, 31] and Appendix D of [32]). Because of collisions (more precisely close encounters) between stars and evaporation, the system evolves quasistatically along the series of equilibria. The natural evolution corresponds to an increase of the central density \(\rho _0\) that parametrizes the series of equilibria.Footnote 22 During the evolution, the energy decreases. On the other hand, the temperature decreases in the region of positive specific heat \(C=dE/dT>0\) and increases in the region of negative specific heat \(C=dE/dT<0\). At the first turning point of energy (minimum energy state), the system becomes unstable and undergoes the gravothermal catastrophe (core collapse). This is a thermodynamical instability taking place on a relaxation (secular) timescale. It ultimately leads to the formation of a binary star surrounded by a hot halo.

1.2 General relativistic systems

The series of equilibria of relativistic star clusters described by the heavily truncated Maxwell–Boltzmann distribution (relativistic Woolley model) was first determined by Zel’dovich and Podurets [58]. They plotted the temperature measured by an observer at infinity \(T_{\infty }\) as a function of the central density \(\rho _0\) and found that the curve \(T_{\infty }(\rho _0)\) displays damped oscillations.Footnote 23 As a result, there exists a maximum temperature \(k_\mathrm{B} T_{{\mathrm{max}}}/mc^2=0.273\) above which there is no equilibrium.Footnote 24 They argued (without rigorous justification) that the series of equilibria should become unstable at that point and that the system should experience a gravitational collapse that they called an “avalanche-type catastrophic contraction of the system”. Considering the same distribution function, Ipser [63] plotted the fractional binding energy \(E/Nmc^2\) as a function of the central redshift \(z_0\) and found that the curve \(E/Nmc^2(z_0)\) displays damped oscillations. Using a rigorous instability criterion based on the equation of pulsations derived by Ipser and Thorne [61], he found that the series of equilibria becomes dynamically (Vlasov) unstable above a critical redshift \(z_\mathrm{c}= 0.516\) and that this critical value happens to coincide with the turning point of fractional binding energy. At that point \([(M-Nm)/Nm]_{\mathrm{c}}=-0.0357\), \((Rc^2/2GNm)_{\mathrm{c}}=4.42\) and \((k_\mathrm{B} T_\infty /mc^2)_{\mathrm{c}}=0.23\). The turning point of binding energy found by Ipser [63] is different from the turning point of temperature found by Zel’dovich and Podurets [58] corresponding to \(z_0=1.08\), \((M-Nm)/Nm=-0.0133\), \(Rc^2/2GNm=3.92\), and \(k_\mathrm{B}(T_{\infty })_{\mathrm{max}}/mc^2=0.27\). In particular, the gravitational instability occurs sooner than predicted by Zel’dovich and Podurets [58]. Using the values of the fractional binding energy and of the temperature measured by an observer at infinity tabulated by Ipser [63] (see his Table 1), we have plotted in Fig. 5 the caloric curve of general relativistic heavily truncated Maxwell–Boltzmann distributions giving \(mc^2/k_\mathrm{B} T_{\infty }\) as a function of \(E/Nmc^2\). This curve has not been plotted before. It has the form of a spiral of which we only see the beginning.

Fig. 5
figure 5

Series of equilibria (caloric curve) of heavily truncated Maxwell–Boltzmann distributions in general relativity (according to the data of Ipser [63]). The series of equilibria is parametrized by the central redshift \(z_0\). It has a snail-like (spiral) structure of which we only see the beginning. The equilibrium states are all canonically unstable. In the microcanonical ensemble, only the part of the curve up to \(z_\mathrm{c}=0.516\) (corresponding to the turning point of binding energy) is stable. In this region, the specific heat is negative. The onset of thermodynamical instability coincides with the onset of dynamical instability

We can now investigate the dynamical and thermodynamical stability of the equilibrium states of relativistic star clusters in more detail. Let us first consider their thermodynamical stability. In a sense, we can say that Zel’dovich and Podurets [58] considered the canonical ensemble (fixed \(T_{\infty }\)), while Ipser [63] considered the microcanonical ensemble (fixed E). The fact that they obtained different results is related to the notion of ensembles inequivalence for systems with long-range interactions like self-gravitating systems.

Let us first consider the microcanonical ensemble which is the rigorous statistical ensemble that we should use in order to describe isolated star clusters. Let us assume that the system is stable at sufficiently high energies (i.e. close to \(0^{-}\)). Using the Poincaré criterion [73], we conclude that the system is microcanonically stable on the upper branch of Fig. 5 up to the first turning point of fractional binding energy and becomes unstable afterwards. We note that the system becomes microcanonically unstable when the specific heat becomes positive. This result was first stated by Horwitz and Katz [10].

Let us now turn to the canonical ensemble. If the system were stable at sufficiently low temperatures, using the Poincaré criterion [73], we would conclude that the system is canonically stable up to the first turning point of temperature and becomes canonically unstable afterwards. However, considering the topology of the series of equilibria in Fig. 5, this is not possible. Indeed, we know from general arguments that canonical stability implies microcanonical stability [139, 140]. Now, we note that the first turning point of temperature occurs after the first turning point of binding energy. Therefore, if the series of equilibria were canonically stable up to the first turning point of temperature it would also be microcanonically stable up to that point. But, we know that the series of equilibria becomes unstable after the first turning point of binding energy. Therefore, we arrive at a contradiction. Since a change of stability in the canonical ensemble can occur only at a turning point of temperature, we conclude that the whole series of equilibria is canonically unstable.Footnote 25 This is consistent with the fact that the beginning of the series of equilibria has a negative specific heat which is forbidden in the canonical ensemble.

Let us finally consider the dynamical stability of the system with respect to a collisionless evolution described by the Vlasov–Einstein equations. We know from general arguments that thermodynamical stability implies dynamical stability [11, 140]. From the previous results obtained in the microcanonical ensemble, we conclude that the system is both thermodynamically and dynamically stable up to the first turning point of fractional binding energy. This is a particular case (for isothermal systems) of the general binding energy criterion derived by Ipser [11]. After the first turning point of fractional binding energy, the system is thermodynamically unstable. On the other hand, Ipser [63] has shown numerically that the system is also dynamically unstable after that point. This lead Ipser [11] to the conclusionFootnote 26 that, in general relativity, dynamical stability and microcanonical thermodynamical stability coincide contrary to the case of Newtonian systems (see Appendix A 1).

From the previous results obtained in the canonical ensemble, we cannot conclude anything regarding the dynamical stability of collisionless isothermal star clusters since all equilibria are canonically unstable. However, we know that the canonical stability of a collisionless star cluster is equivalent to the dynamical stability of the corresponding barotropic star with respect to the Euler–Einstein equations [20, 21, 23, 26]. Therefore, our observation that all the equilibria of isothermal relativistic star clusters are canonically unstable is consistent with Ipser’s [63] finding that all the relativistic stars with the same equation of state as the isothermal star clusters are dynamically unstable.

The dynamical evolution of relativistic star clusters is reviewed in the Introduction (see also the introduction of [30]). Because of collisions between stars and evaporation, the system evolves quasistatically along the series of equilibria. The natural evolution corresponds to an increase of the central density \(\rho _0\) (or central redshift \(z_0\)) that parametrizes the series of equilibria. During the evolution, the energy decreases. On the other hand, the temperature increases since the specific heat is negative \(C=dE/dT_{\infty }<0\). At the turning point of energy (minimum energy state), the system becomes unstable and undergoes a gravitational collapse. This is a dynamical (and thermodynamical) instability of general relativistic origin taking place on a dynamical (short) timescale. It ultimately leads to the formation of a black hole surrounded by a halo of stars.

Appendix B: Analogies and differences between box models and truncated distributions

1.1 Newtonian gravity

In Newtonian gravity, the series of equilibria of truncated isothermal distributions (see Fig. 4) is qualitatively similar to the series of equilibria of box-confined isothermal systems (see Fig. 2 and the “cold spiral” of Fig. 1). The main difference stems from the fact that, for open clusters, equilibrium states necessarily have a negative energy \(E<0\).Footnote 27 Indeed, when the energy is positive (\(E>0\)) the stars are unbounded and disperse away. If the system is confined within a box, equilibrium states with a positive energy are possible because the stars bounce off the wall. Apart from this difference, the series of equilibria in Figs. 2 and 4 have a similar spiralling shape. They display a turning point of temperature before the turning point of energy. For the two systems, equilibrium states are canonically stable up to the turning point of temperature and microcanonically stable up to the turning point of energy. We note that the dimensionless temperature and the dimensionless energy in the two models are related by

$$\begin{aligned} \eta =\sigma \beta m v_0^2\quad \mathrm{and}\quad \varLambda =-\frac{1}{\sigma } \frac{E}{Mv_0^2}\quad \mathrm{where}\quad \sigma =\frac{GM}{Rv_0^2}. \end{aligned}$$
(B1)

In practice, \(\sigma \sim 1\) since \(v_0\) corresponds to the virial velocity of the cluster. We see that the orders of magnitude of the dimensionless energies and dimensionless temperatures in Figs. 2 and 4 are consistent with each other.

1.2 General relativity

We now turn to the general relativistic case. The values of E and \(T_{\infty }\) tabulated by Ipser [63] correspond to strongly relativistic stellar systems. Therefore, the caloric curve of Fig. 5 describes only the strongly relativistic part of the caloric curve (hot spiral). We need to complete this caloric curve with the nonrelativistic part from Fig. 4 (cold spiral). To that purpose, we first recall that, in the nonrelativistic regime, \(\beta m v_0^2\) and \({E}/{Mv_0^2}\) are of the same order of magnitude as \(\eta \) and \(\varLambda \) in the box model (see Appendix B 1). On the other hand, in the relativistic regime, we have the relations

$$\begin{aligned} \eta =\nu \frac{mc^2}{k_\mathrm{B} T_{\infty }}\quad \mathrm{and} \quad \varLambda =-\frac{1}{\nu } \frac{E}{Nmc^2}\quad \mathrm{where}\quad \nu =\frac{GNm}{Rc^2}. \end{aligned}$$
(B2)

As shown in Refs. [30, 128], the parameter \(\nu \) belongs to the interval [0, 0.1764]. In order to construct the complete caloric curve of truncated isothermal star clusters, we first determine the relativistic curve \(\eta (\varLambda )\) from Fig. 5 by multiplying \({mc^2}/{k_\mathrm{B} T_{\infty }}\) by \(\nu \) and dividing \({E}/{Nmc^2}\) by \(\nu \), and then, we add the nonrelativistic caloric curve of Fig. 4. We note that the result depends on the parameter \(\nu \) so we actually have a family of caloric curves. For illustration, we have taken \(\nu =0.015\). This leads to the complete caloric curve reported in Fig. 6.Footnote 28 Our procedure is of course very approximate, but it is sufficient to show the idea. It will be important in future works to improve this procedure and determine the family of caloric curves exactly.

From the complete caloric curve of Fig. 6, it is now easy to understand the thermodynamical stability/instability of isothermal clusters in both nonrelativistic and relativistic regimes. We again advocate the Poincaré turning point criterion [73].

In the microcanonical ensemble, the nonrelativistic branch is stable up to the first fundamental turning point of energy MCE (see Appendix A 1). At that point, the series of equilibria \(\beta (-E)\) rotates clockwise so that a mode of stability is lost. A new mode of stability is lost at each subsequent turning point of energy where the series of equilibria rotates clockwise. At some point, the series of equilibria unwinds and rotates anticlockwise. (The unwinding of the spiral is represented schematically by a dashed line in Fig. 6.) A mode of stability is gained at each turning point of energy where the series of equilibria rotates anticlockwise. Finally, the series of equilibria on the relativistic branch becomes stable again (after rotating anticlockwise an even number of times) and remains stable until the second fundamental turning point of energy MCE’ at which it becomes unstable again. After that point, the series of equilibria rotates only clockwise so that it remains unstable until the end. This is consistent with the results from Appendix A 2.

In the canonical ensemble, the nonrelativistic branch is stable up to the first turning point of temperature CE (see Appendix A 1). At that point, the series of equilibria \(\beta (-E)\) rotates clockwise so that a mode of stability is lost. The series of equilibria thus becomes unstable. It remains unstable until the end because it can never recover the first mode of stability that it has lost. (It rotates anticlockwise an odd number of times.) This is why we found in Appendix A 2 that the relativistic branch is always unstable in the canonical ensemble.

The caloric curve of Fig. 6 provides a nice illustration of the scenario discussed in the Introduction. Because of collisions (close encounters) and evaporation, a Newtonian stellar system evolves along the upper branch of the series of equilibria. When it reaches the first fundamental turning point of energy, it becomes thermodynamically unstable (while remaining dynamically stable) and undergoes a gravothermal catastrophe. It then evolves towards very hot and very dense configurations and becomes relativistic. It may then reach a stable relativistic isothermal equilibrium distribution. Again, because of collisions and evaporation, a relativistic star cluster evolves along the lower branch of the series of equilibria. When it reaches the second fundamental turning point of energy, it becomes thermodynamically and dynamically unstable and undergoes a catastrophic collapse towards a black hole.Footnote 29

Fig. 6
figure 6

Complete caloric curve of truncated isothermal star clusters showing the nonrelativistic cold spiral and the relativistic hot spiral. (For illustration, we have taken \(\nu =0.015\).) The units of \(\beta \) and E are the same as in Fig. 4. Nonrelativistic clusters undergo a gravothermal catastrophe at MCE and become relativistic. They can then undergo a catastrophic collapse at MCE’ leading to the formation of a black hole

We can now compare the caloric curve of truncated isothermal star clusters (see Fig. 6) with the caloric curve of box-confined isothermal systems (see Fig. 1). As we have seen in Appendix B 1, the nonrelativistic parts of the caloric curves are relatively similar. In particular, they present a “cold” spiral at negative energies. We now consider the relativistic parts of the caloric curves. In that case, a “hot” spiral arises because, in the ultrarelativistic limit (corresponding to very high energy densities), the equation of state of an isothermal gas is \(P=\epsilon /3\) like for the self-gravitating black-body radiation. However, in the relativistic domain, the series of equilibria of truncated isothermal distributions (see Figs. 5 and 6) and the series of equilibria of box-confined isothermal systems (see Figs. 1 and 3) are very different.

(i) The origin of the difference stems from the fact that, for open clusters, stable equilibrium states necessarily have a negative energy \(E<0\).Footnote 30 By contrast, when the system is confined within a box, stable equilibrium states with a positive energy are allowed because of the pressure exerted by the box. For box-confined systems, the “hot spiral” appears at very large positive energies. At such energies, the relativistic isothermal gas behaves as a form of radiation confined within a cavity like in the studies of Sorkin et al. [12] and Chavanis [17]. There are no such equilibrium states with very large positive energies in open star clusters. In open star clusters, the “hot spiral” appears at energies (positive and negative) around \(E=0\). In addition, the stable equilibrium states necessarily have a negative energy.

(ii) For box-confined isothermal systems, the cold and hot spirals are quite distinct. The cold spiral appears at negative energies, and the hot spiral appears at large positive energies. For open star clusters, the cold and hot spirals both appear at approximately the same range of (negative) energies. When general relativity is taken into account, the cold spiral of Fig. 4 unwinds and connects the hot spiral of Fig. 5 as sketched in Fig. 6.

As a result, a relativistic isothermal cluster in a box behaves very differently from a truncated relativistic isothermal cluster. This remark may question the physical relevance of the “hot spiral” at high energies for realistic isothermal star clusters except if one can justify a form of confinement playing the role of the box. This may be the case for relativistic stars. The box could be caused by a medium exerting a tremendously large pressure on an ultrarelativistic gas, like in the supernova phenomenon where the system achieves a “core–halo” structure. The core could correspond to the ultrarelativistic isothermal gas, and the halo could play the role of the box. In that case, the core would be sustained by the external pressure of the halo. However, this scenario remains to be put on a more rigorous basis (see [29, 30, 151, 153,154,155] and [156, 157] for different scenarios of supernova formation).

Appendix C: Entropy and free energy as functionals of the density for nonrelativistic self-gravitating classical particles

We consider a nonrelativistic system of self-gravitating classical particles. In Appendix C.2 of Paper I, we have introduced entropy and free energy functionals of the distribution function \(f({\mathbf{r}},\mathbf{v})\). In Sect. 2.2 of this paper and in Appendix C.1.a of Paper I, we have introduced entropy and free energy functionals of the local density n(r) and local kinetic energy \(\epsilon _{{\mathrm{kin}}}(r)\). In this Appendix, we introduce entropy and free energy functionals of the local density n(r).

1.1 Microcanonical ensemble

In the microcanonical ensemble, the statistical equilibrium state is obtained by maximizing the Boltzmann entropy S[f] at fixed energy E and particle number N. To solve this maximization problem, we proceed in two steps. We first maximize S[f] at fixed E, Nand particle density n(r). Since n(r) determines the particle number N[n] and the gravitational energy W[n], this is equivalent to maximizing S[f] at fixed kinetic energy \(E_{\mathrm{kin}}[f]\) and particle density n(r). The variational problem for the first variations (extremization) can be written as

$$\begin{aligned} \frac{\delta S}{k_\mathrm{B}}-\beta \delta E_{{\mathrm{kin}}}+\int \alpha (r)\delta n \, \hbox {d}{\mathbf{r}}=0, \end{aligned}$$
(C1)

where \(\beta \) is a global (uniform) Lagrange multiplier and \(\alpha (r)\) is a local (position dependent) Lagrange multiplier. This variational problem yields

$$\begin{aligned} f({\mathbf{r}},{\mathbf{p}})=\frac{g}{h^3}\hbox {e}^{-\beta p^2/2m+\alpha (r)}. \end{aligned}$$
(C2)

As in Appendix C.1 of Paper I, we can show that this distribution is the global maximum of S[f] at fixed \(E_{{\mathrm{kin}}}[f]\) and n(r). Substituting Eq. (C2) into Eqs. (I-18), (I-19) and (I-21), we get

$$\begin{aligned} n(r)= & {} \frac{g}{h^3}( 2\pi m k_\mathrm{B} T)^{3/2}\hbox {e}^{\alpha (r)}, \end{aligned}$$
(C3)
$$\begin{aligned} \epsilon _{{\mathrm{kin}}}(r)= & {} \frac{3}{2}n(r)k_\mathrm{B} T, \end{aligned}$$
(C4)
$$\begin{aligned} P(r)= & {} \frac{2}{3}\epsilon _{{\mathrm{kin}}}(r)=n(r)k_\mathrm{B} T. \end{aligned}$$
(C5)

The Lagrange multiplier \(\alpha (r)\) is determined by the density n(r) according to Eq. (C3). As a result, the distribution function (C2) can be written in terms of the density as

$$\begin{aligned} f({\mathbf{r}},{\mathbf{p}})=n({r}) \left( \frac{\beta }{2\pi m}\right) ^{3/2} \hbox {e}^{-\beta {p}^2/2m}. \end{aligned}$$
(C6)

On the other hand, since T is uniform, we see from Eq. (C5) that the equation of state is isothermal: \(P(r)=n(r) k_\mathrm{B} T\). The temperature T is determined by the kinetic energy \(E_{\mathrm{kin}}[n(r),T]=E-W[n(r)]\) using Eq. (C4) integrated over the volume giving \(E_{{\mathrm{kin}}}=(3/2)Nk_\mathrm{B} T\). In other words, the temperature is determined by the energy constraint

$$\begin{aligned} E=\frac{3}{2}Nk_\mathrm{B} T+W[n(r)]. \end{aligned}$$
(C7)

We note that T is a functional of the density n(r) but, for brevity, we shall not write this dependence explicitly. Substituting the Maxwell–Boltzmann distribution (C6) into the entropy density (1), we obtain

$$\begin{aligned} \frac{s(r)}{k_\mathrm{B}}=\frac{5}{2}n(r)-n(r)\ln n(r)-n(r)\ln \left( \frac{h^3}{g}\right) +\frac{3}{2}n(r)\ln (2\pi m k_\mathrm{B} T), \end{aligned}$$
(C8)

which is equivalent to the integrated Gibbs–Duhem relation (9). The entropy is given by

$$\begin{aligned} \frac{S}{k_\mathrm{B}}=\frac{5}{2}N-\int n(r)\ln n(r)\, 4\pi r^2\, \hbox {d}r-N\ln \left( \frac{h^3}{g}\right) +\frac{3}{2}N\ln (2\pi m k_\mathrm{B} T). \end{aligned}$$
(C9)

Finally, the statistical equilibrium state in the microcanonical ensemble is obtained by maximizing the entropy S[n(r), T] at fixed particle number N, the energy constraint being taken into account in the determination of the temperature T[n] through Eq. (C7). The variational problem for the first variations (extremization) can be written as

$$\begin{aligned} \frac{\delta S}{k_\mathrm{B}}+\alpha _0\delta N=0. \end{aligned}$$
(C10)

The conservation of energy implies [see Eq. (C7)]:

$$\begin{aligned} 0=\delta E_{{\mathrm{kin}}}+\int m\varPhi \delta n\, \hbox {d}{\mathbf{r}}. \end{aligned}$$
(C11)

Using Eqs. (C1) and (C11), and proceeding as in Appendix F1 of Paper I, or performing the variations over \(\delta n\) and \(\delta T\) directly on the explicit expressions (C7) and (C9), we obtain

$$\begin{aligned} \alpha (r)=\alpha _0-\beta m\varPhi (r)\qquad \mathrm{and}\qquad n(r)=\frac{g}{h^3}( 2\pi m k_\mathrm{B} T)^{3/2}\hbox {e}^{\alpha _0-\beta m\varPhi (r)}. \end{aligned}$$
(C12)

We then recover all the results of Sect. 2. The interest of this formulation is that it allows us to solve more easily the stability problem related to the sign of the second variations of entropy. This problem has been studied in detail in [112, 113, 122]. It has also been studied in [16] within the framework of special relativity.

1.2 Canonical ensemble

In the canonical ensemble, the statistical equilibrium state is obtained by minimizing the Boltzmann free energy \(F[f]=E[f]-TS[f]\) at fixed particle number N, or equivalently, by maximizing the Massieu function \(J[f]=S[f]/k_\mathrm{B}-\beta E[f]\) at fixed particle number N. To solve this maximization problem, we proceed in two steps. We first maximize \(J[f]=S[f]/k_\mathrm{B}-\beta E[f]\) at fixed Nand particle density n(r). Since n(r) determines the particle number N[n] and the gravitational energy W[n], this is equivalent to maximizing \(S[f]/k_\mathrm{B}-\beta E_{{\mathrm{kin}}}[f]\) at fixed particle density n(r). The variational problem for the first variations (extremization) can be written as

$$\begin{aligned} \delta \left( \frac{S}{k_\mathrm{B}}-\beta E_{{\mathrm{kin}}}\right) +\int \alpha (r)\delta n \, \hbox {d}{\mathbf{r}}=0, \end{aligned}$$
(C13)

where \(\alpha (r)\) is a local (position dependent) Lagrange multiplier. Since \(\beta \) is constant in the canonical ensemble, this is equivalent to the condition (C1) yielding the distribution function (C2). This distribution is the global maximum of \(S[f]/k_\mathrm{B}-\beta E_{{\mathrm{kin}}}[f]\) at fixed n(r). We then obtain the same results as in Appendix C 1, except that T is fixed while it was previously determined by the conservation of energy.

We can now simplify the expression of the free energy. The entropy is given by Eq. (C9) and the energy by Eq. (C7). Since \(F=E-TS\), we obtain

$$\begin{aligned} F= & {} W[n(r)]-Nk_\mathrm{B} T+k_\mathrm{B} T\int n(r)\ln n(r)\, 4\pi r^2\, \hbox {d}r+Nk_\mathrm{B}T\ln \left( \frac{h^3}{g}\right) \nonumber \\&-\frac{3}{2}Nk_\mathrm{B}T\ln (2\pi m k_\mathrm{B} T). \end{aligned}$$
(C14)

Therefore, the free energy can be written as a functional of the density as

$$\begin{aligned} F[n(r)]=U[n(r)]+W[n(r)], \end{aligned}$$
(C15)

where U[n(r)] is the internal energy given by

$$\begin{aligned} U[n(r)]= & {} -Nk_\mathrm{B} T+k_\mathrm{B} T\int n(r)\ln n(r)\, 4\pi r^2\, \hbox {d}r+Nk_\mathrm{B}T\ln \left( \frac{h^3}{g}\right) \nonumber \\&-\frac{3}{2}Nk_\mathrm{B}T\ln (2\pi m k_\mathrm{B} T). \end{aligned}$$
(C16)

Finally, the statistical equilibrium state in the canonical ensemble is obtained by minimizing the free energy F[n] at fixed particle number N. The variational problem for the first variations (extremization) can be written as

$$\begin{aligned} \delta J+\alpha _0\delta N=0. \end{aligned}$$
(C17)

Decomposing the Massieu function as \(J[f]=S[f]/k_\mathrm{B}-\beta E_{\mathrm{kin}}[f]-\beta W[n]\) and proceeding as in Appendix F2 of Paper I, or performing the variations over \(\delta n\) directly on the explicit expression (C14) of the free energy, we get

$$\begin{aligned} \alpha (r)=\alpha _0-\beta m\varPhi (r)\qquad \mathrm{and}\qquad n(r)=\frac{g}{h^3}( 2\pi m k_\mathrm{B} T)^{3/2}\hbox {e}^{\alpha _0-\beta m\varPhi (r)}. \end{aligned}$$
(C18)

We then recover all the results of Sect. 2. The interest of this formulation is that it allows us to solve more easily the stability problem related to the sign of the second variation of free energy. This problem has been studied in detail in [120, 122]. It has also been studied in [16] within the framework of special relativity.

Remark

We note that the free energy \(F[\rho ]\) coincides with the energy functional associated with the Euler–Poisson equations describing a gas with an isothermal equation of state \(P=\rho k_\mathrm{B} T/m\) (see Sec. V.A.1 of Paper I). As a result, the thermodynamical stability of a classical self-gravitating system in the canonical ensemble is equivalent to the dynamical stability of the corresponding isothermal gas described by the Euler–Poisson equations [120]. This is a particular case of the general result established in [139] which is valid for an arbitrary form of entropy. According to the Poincaré turning point criterion, the series of equilibria becomes both thermodynamically unstable (in the canonical ensemble) and dynamically unstable with respect to the Euler–Poisson equations at the first turning point of temperature (or, equivalently, at the first turning point of mass).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chavanis, PH. Statistical mechanics of self-gravitating systems in general relativity: II. The classical Boltzmann gas. Eur. Phys. J. Plus 135, 310 (2020). https://doi.org/10.1140/epjp/s13360-020-00291-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-020-00291-1

Navigation