This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Caloric curves of self-gravitating fermions in general relativity

Giuseppe Alberti Laboratoire de Physique Théorique, Université Paul Sabatier, 118 route de Narbonne 31062 Toulouse, France Living Systems Research, Roseggerstraße 27/2, A-9020 Klagenfurt am Wörthersee, Austria    Pierre-Henri Chavanis Laboratoire de Physique Théorique, Université Paul Sabatier, 118 route de Narbonne 31062 Toulouse, France
Abstract

We study the nature of phase transitions between gaseous and condensed states in the self-gravitating Fermi gas at nonzero temperature in general relativity. The condensed states can represent compact objects such as white dwarfs, neutron stars, or dark matter fermion balls. The caloric curves depend on two parameters: the system size RR and the particle number NN. When N<NOVN<N_{\rm OV}, where NOVN_{\rm OV} is the Oppenheimer-Volkoff limit, there exists an equilibrium state for any value of the temperature TT and of the energy EE as in the nonrelativistic case [P.H. Chavanis, Int. J. Mod. Phys. B 20, 3113 (2006)]. Gravitational collapse is prevented by quantum mechanics (Pauli’s exclusion principle). When N>NOVN>N_{\rm OV}, there is no equilibrium state below a critical energy and below a critical temperature. In that case, the system is expected to collapse towards a black hole. We plot the caloric curves of the general relativistic Fermi gas, study the different types of phase transitions that occur in the system, and determine the phase diagram in the (R,N)(R,N) plane. The nonrelativistic results are recovered for NNOVN\ll N_{\rm OV} and RROVR\gg R_{\rm OV} with NR3NR^{3} fixed. The classical results are recovered for NNOVN\gg N_{\rm OV} and RROVR\gg R_{\rm OV} with N/RN/R fixed. We discuss the commutation of the limits c+c\rightarrow+\infty and 0\hbar\rightarrow 0. We study the relativistic corrections to the nonrelativistic caloric curves and the quantum corrections to the classical caloric curves. We highlight a situation of physical interest where a gaseous Fermi gas, by cooling, first undergoes a phase transition towards a compact object (white dwarf, neutron star, dark matter fermion ball), then collapses into a black hole. This situation occurs in the microcanonical ensemble when NOV<N<3.73NOVN_{\rm OV}<N<3.73\,N_{\rm OV}. We also relate the phase transitions from a gaseous state to a core-halo state in the microcanonical ensemble to the onset of red-giant structure and to the supernova phenomenon.

pacs:
95.30.Sf, 95.35.+d, 04.40.Dg, 67.85.Lm, 05.70.-a, 05.70.Fh

I Introduction

The study of phase transitions is an important problem in physics. Some examples include solid-liquid-gas phase transitions, superconducting and superfluid transitions, Bose-Einstein condensation, liquid-glass phase transition in polymers, liquid crystal phases, Kosterlitz-Thouless transition etc. Self-gravitating systems also undergo phase transitions but they are special due to the unshielded long-range attractive nature of the interaction paddy . This leads to unusual phenomena such as negative specific heats, ensembles inequivalence, long-lived metastable states, and gravitational collapse. A strict equilibrium state can exist only if the system is confined within a box, otherwise it has the tendency to evaporate (this is already the case for an ordinary gas). On the other hand, in order to define a condensed phase we need to introduce a short-range repulsion between the particles that opposes itself to the gravitational attraction.111Without small-scale regularization, there is no equilibrium state (global entropy maximum) in a strict sense antonov . There can exist, however, metastable gaseous states (local entropy maxima) that are insensistive to the small-scale regularization antonov ; lbw . These metastable states have a very long lifetime, scaling as eNe^{N}, where NN is the number of particles in the system lifetime . In practice this lifetime is much larger than the age of the Universe, making the metastable states fully relevant in astrophysics ijmpb . In this paper, we consider the case of self-gravitating fermions where an effective short-range repulsion is due to quantum mechanics (Pauli’s exclusion principle). The object of this paper is to present a complete description of phase transitions in the self-gravitating Fermi gas in general relativity. This study can have applications in relation to the formation of compact objects such as white dwarfs, neutron stars, dark matter stars, black holes etc. On the other hand, the phase transition from a gaseous state to a condensed state may be related to the onset of red-giant structure and to the supernova phenomenon. We first start by reviewing the literature on the subject. We focus our review on papers that study phase transitions in the box-confined self-gravitating Fermi gas at nonzero temperature.222The case of completely degenerate self-gravitating fermions at T=0T=0 and the case of classical (nondegenerate) self-gravitating systems are considered in our companion papers paper1 ; paper2 where a detailed review of the literature is made. We do not review the immensely vast literature related to self-gravitating fermions as models of white dwarfs, neutron stars, and dark matter halos. For a connection to this literature, we refer to chandrabook ; shapiroteukolsky ; btv ; vss ; rar ; clm1 ; clm2 ; urbano ; rsu and references therein. For a connection to the general literature on the statistical mechanics of self-gravitating systems and systems with long-range interactions we refer to the introduction of sd and to the reviews paddy ; houches ; katzrevue ; ijmpb ; cdr ; campabook .

The statistical mechanics of nonrelativistic self-gravitating fermions at nonzero temperature enclosed within a box of radius RR was first studied by Hertel & Thirring (1971) htf . They worked in the canonical ensemble and rigorously proved that the mean field approximation (or effective field approximation) and the Thomas-Fermi (TF) approximation (which amounts to neglecting the quantum potential) become exact in a suitable thermodynamic limit N+N\rightarrow+\infty where RN1/3R\sim N^{-1/3}, TN4/3T\sim N^{4/3}, EN7/3E\sim N^{7/3}, SNS\sim N, and FN7/3F\sim N^{7/3} (the scaling FN7/3F\sim N^{7/3} was first obtained by Lévy-Leblond (1969) levyleblond for the ground state).333This is also equivalent to the usual thermodynamic limit N+N\rightarrow+\infty where RN1/3R\sim N^{1/3}, T1T\sim 1, ENE\sim N, SNS\sim N and FNF\sim N with GN2/3G\sim N^{-2/3} (see Appendix A). This leads to the temperature-dependent TF equation.444It can be obtained by combining the fundamental equation of hydrostatic equilibrium with the Fermi-Dirac equation of state or, equivalently, by substituting the Fermi-Dirac density into the Poisson equation; see, e.g., Hertel (1977) hseul . For that reason, the temperature-dependent TF equation is sometimes called the Fermi-Dirac-Poisson equation. The existence of the TF limit for the thermodynamic functions of self-gravitating fermions was proven by Hertel et al. (1972) hnt for the microcanonical and canonical ensembles and by Messer (1979) messer for the grand canonical ensemble. The convergence of the quantum-statistically defined particle density towards the TF density was proven by Baumgartner (1976) baumgartner . He also showed that there are no correlations in the thermodynamic limit. Narnhofer and Sewell (1980) ns1 showed that when N+N\rightarrow+\infty the equilibrium Gibbs distribution becomes a tensor product of density functions of an ideal Fermi gas which minimize the TF free energy functional. These density functions can be stable (global minima) or metastable (local minima). Finally, Narnhofer and Sewell (1982) ns2 showed that when N+N\rightarrow+\infty a quantum system of self-gravitating fermions is described by the classical Vlasov equation bh .

Hertel & Thirring (1971) ht studied numerically phase transitions in the nonrelativistic self-gravitating Fermi gas in relation with the structure of neutron stars.555The possibility of phase transitions in the self-gravitating Fermi gas at nonzero temperature was suggested in the Appendix IV of Lynden-Bell and Wood (1969) lbw . They assumed that the gas is enclosed within a box and worked in the canonical ensemble. For a given number of particles NN, they showed that a canonical first order phase transition arising from a multiplicity of solutions in the TF equation appears if the radius of the box is larger than a certain value RCCP(N)=12.82/(N1/3Gm3)R_{\rm CCP}(N)=12.8\,\hbar^{2}/(N^{1/3}Gm^{3}). This phase transition is characterized by a jump of energy (the energy E=(βF)/βE=\partial(\beta F)/\partial\beta, the first derivative of βF\beta F with respect to β\beta, becomes discontinuous) at a transition temperature TtT_{t} determined by a Maxwell construction like in the theory of the van der Waals gas.666The phase transition arises because the TF equation has two stable solutions at the same temperature that minimize the TF free energy. A rigorous analytical proof for the existence of this phase transition was given by Messer (1981a,1981b) messerpt1 ; messerpt2 following numerical calculations by Hertel (1977) hseul . When there are multiple solutions in the TF equation, they argue that one must choose the one with the smallest value of free energy. This corresponds to a transition between a nearly homogeneous phase of medium mass density (gaseous phase) and a phase with a high density core surrounded by an atmosphere of low density (condensed phase) when the system cools down below TtT_{t}. Hertel & Thirring (1971) ht explained that this phase transition replaces the region of negative specific heats in the microcanonical ensemble (or the piece of convex curvature in the entropy curve S(E)S(E)) which is associated with unstable equilibrium states in the canonical ensemble. Therefore, the microcanonical and canonical ensembles are not equivalent hnt . The region of negative specific heat in the microcanonical ensemble is bridged by a phase transition in the canonical ensemble.777Canonical phase transitions, associated with negative specific heats, have also been found by Thirring (1970) thirring in a toy model of self-gravitating systems, by Aronson and Hansen (1972) ah for a self-gravitating hard spheres gas, by Carlitz (1972) carlitz for hadronic matter, and by Hawking (1976) hawking for black holes. Hertel & Thirring (1971) ht applied their crude model of neutron stars to a system of N=1057N=10^{57} neutrons (the corresponding mass being of the order of the solar mass) initially contained in a sphere of radius R=100kmR=100\,{\rm km}. The critical radius is RCCP=43.1kmR_{\rm CCP}=43.1\,{\rm km}. For R=100km>RCCPR=100\,{\rm km}>R_{\rm CCP}, the system undergoes a first order phase transition below a critical temperature Tt=7.03×1010KT_{t}=7.03\times 10^{10}\,{\rm K}, collapses, and forms a compact object (neutron star) containing almost all the mass. This compact object has approximately the same size, RC=4.512/(N1/3Gm3)=15.1kmR_{C}=4.51\,\hbar^{2}/(N^{1/3}Gm^{3})=15.1\,{\rm km}, as a completely degenerate Fermi gas at T=0T=0 (equivalent, in their nonrelativistic model, to a polytrope of index n=3/2n=3/2) but it is surrounded by a small isothermal atmosphere. This gravitational phase transition could account for the implosion of the core in the supernova phenomenon where the energy is carried quickly by neutrinos.888Thirring (1970) thirring , Hertel and Thirring (1971) ht and Messer (1981) messerpt2 mention the analogy between this phase transition and the formation of red giants and supernovae. However, this analogy may not be fully correct because the phase transition that they obtain just corresponds to an implosion. This is because they work in the canonical ensemble and consider relatively small systems while the phase transition leading to an implosion-explosion phenomenon, associated with a core-halo structure, occurs in the microcanonical ensemble for larger systems (see Ref. supernova and Sec. XIII). Lynden-Bell and Wood (1968) lbw , considering a classical self-gravitating gas in the microcanonical ensemble, find the emergence of a core-halo structure and relate it to the onset of red giants.

Gravitational phase transitions of fermionic matter were also studied by Bilic & Viollier (1997) bvn in a cosmological setting. They considered weakly interacting massive fermions of mass 17.2keV/c217.2\,{\rm keV/c^{2}} in the presence of a large radiation-density background fixing the temperature. They studied a halo of mass M=109MM=10^{9}\,M_{\odot} and radius R=1.68×102pc>RCCP=6.00×103pcR=1.68\times 10^{-2}\,{\rm pc}>R_{\rm CCP}=6.00\times 10^{-3}\,{\rm pc}. When the system cools down below a transition temperature Tt=4.80×105KT_{t}=4.80\times 10^{5}\,{\rm K}, a condensed phase emerges consisting of quasidegenerate supermassive fermion stars of mass M109MM\sim 10^{9}\,M_{\odot} and radius RC=2.10×103pcR_{C}=2.10\times 10^{-3}\,{\rm pc}. They argued that these compact dark objects could play an important role in structure formation in the early Universe. In particular, these fermion stars could explain, without resorting to the black hole hypothesis, some of the features observed around supermassive compact dark objects which are reported to exist at the centers of a number of galaxies including our own and quasistellar objects (QSOs). On a technical point of view, their study is analogous to the one carried out by Hertel & Thirring (1971) ht for neutron stars, i.e., they described the canonical first order phase transtion between a “gaseous” phase and a “condensed” phase that appears below a transition temperature when the size of the object is sufficiently large.

A detailed theoretical description of phase transitions in the nonrelativistic self-gravitating Fermi gas at nonzero temperature was given by Chavanis (2002) ijmpb (see also Refs. csmnras ; pt ; dark ; ispolatov ; rieutord ; ptd ).999In these papers, the statistical equilibrium state is obtained by maximizing the Fermi-Dirac entropy SS at fixed mass MM and energy EE in the microcanonical ensemble and by minimizing the Fermi-Dirac free energy F=ETSF=E-TS at fixed mass MM in the canonical ensemble, where SS is obtained from a combinatorial analysis taking into account the Pauli exclusion principle. This leads to the TF (or Fermi-Dirac-Poisson) equation in a direct manner. The study of the self-gravitating Fermi gas has also applications in the statistical theory of violent relaxation developed by Lynden-Bell lb that also leads to a Fermi-Dirac-type distribution csmnras . He showed that the caloric curves T(E)T(E) depend on a single control parameter μ=η0512π4G3MR3\mu=\eta_{0}\sqrt{512\pi^{4}G^{3}MR^{3}} with η0=gm4/h3\eta_{0}=gm^{4}/h^{3} (gg is the spin multiplicity of the quantum states). For a fixed particle number NN, this paramerer can be seen as a measure of the size of the system since μR3/2\mu\propto R^{3/2}. Chavanis ijmpb studied in detail the nature of phase transitions in the nonrelativistic self-gravitating Fermi gas in both microcanonical and canonical ensembles. He showed that there exist two critical points (one in each ensemble) at which zeroth and first order phase transitions appear. The canonical critical point μCCP=83\mu_{\rm CCP}=83 at which canonical phase transitions appear is equivalent to the one previously found by Hertel and Thirring (1971) ht . The microcanonical critical point μMCP=2670\mu_{\rm MCP}=2670 at which microcanonical phase transitions appear was not found previously. For μ+\mu\rightarrow+\infty, one recovers the caloric curve of a nonrelativistic self-gravitating classical gas paddy . Chavanis ijmpb ; lifetime argued that first order phase transitions do not take place in practice, contrary to previous claims ht ; bvn , because of the very long lifetime of metastable states for systems with long-range interactions. Therefore, only zeroth order phase transitions take place at the spinodal points where the metastable branches disappear. Recently, this study of phase transitions was extended to the nonrelativistic fermionic King model clm1 ; clm2 . This model is more realistic as it avoids the need of an artificial box to confine the system.

Gravitational phase transitions of fermionic matter in general relativity were studied by Bilic and Viollier (1999) bvr .101010In that case, the suitable thermodynamic limit corresponds to N+N\rightarrow+\infty where RN2/3R\sim N^{2/3}, TN1/3T\sim N^{-1/3}, EN2/3E\sim N^{2/3}, SNS\sim N and FN2/3F\sim N^{2/3} with mN1/3m\sim N^{-1/3} bvr . This is also equivalent to the usual thermodynamic limit N+N\rightarrow+\infty where RN1/3R\sim N^{1/3}, T1T\sim 1, ENE\sim N, SNS\sim N and FNF\sim N with GN2/3G\sim N^{-2/3} (see Appendix A). They showed that, at some critical temperature TtT_{t}, weakly interacting massive fermionic matter with a total mass below the Oppenheimer-Volkoff (OV) limit ov undergoes a first order gravitational phase transition from a diffuse to a clustered state, i.e., a nearly completely degenerate fermion star. This is an extension of their previous paper bvr in the Newtonian approximation. This relativistic extension allowed them to consider situations where the mass of the system is close to the OV limit so that the fermion star is strongly relativistic. For fermions masses of 1010 to 25keV/c225\,{\rm keV/c^{2}} they argued that these fermions stars may well provide an alternative explanation for the supermassive compact dark objects that are observed at galactic centers. Indeed, a few Schwarzschild radii away from the object, there is little difference between a supermassive black hole and a fermion star of the same mass near the OV limit.111111Some difficulties with the “fermion ball” scenario to provide an alternative to supermassive black holes at the centers of the galaxies are pointed out in genzel . In their paper, they considered fermionic particles of mass m=17.2keV/c2m=17.2\,{\rm keV/c^{2}} for which NOV=1.4254×1071N_{\rm OV}=1.4254\times 10^{71}, NOVm=2.1973×109MN_{\rm OV}m=2.1973\times 10^{9}\,M_{\odot}, MOV=2.1186×109MM_{\rm OV}=2.1186\times 10^{9}\,M_{\odot} and ROV=8.88×104pcR_{\rm OV}=8.88\times 10^{-4}\,{\rm pc}. They studied a system of N=0.95350NOVN=0.95350\,N_{\rm OV} fermions, corresponding to a rest mass Nm=2.0951×109MNm=2.0951\times 10^{9}\,M_{\odot} which is slightly below the OV limit, in a sphere of size R=29.789ROV=2.6391×102pcR=29.789\,R_{\rm OV}=2.6391\times 10^{-2}\,{\rm pc}. The transition occurs at Tt=0.0043951mc2=8.7725×105KT_{t}=0.0043951\,mc^{2}=8.7725\times 10^{5}\,{\rm K}. This leads to a fermion star containing almost all the particles surrounded by a small atmosphere. If we approximate the fermion star by a Fermi gas at T=0T=0 containing all the rest mass 2.0951×109M\sim 2.0951\times 10^{9}\,M_{\odot}, we find a radius RC=1.220ROV=1.0809×103pcR_{C}=1.220\,R_{\rm OV}=1.0809\times 10^{-3}\,{\rm pc} and a mass MC=0.9577MOV=2.0290×109MM_{C}=0.9577\,M_{\rm OV}=2.0290\times 10^{9}\,M_{\odot}.

The study of Bilic and Viollier bvr is restricted to a unique value of RR and NN, with N<NOVN<N_{\rm OV}, leading to a canonical phase transition. The object of this paper is to perform a more general study of phase transitions in the self-gravitating Fermi gas in general relativity for arbitrary values of RR and NN. In particular, we would like to determine what happens when N>NOVN>N_{\rm OV}, or what happens for larger values of RR where a microcanonical phase transition is expected.

The paper is organized as follows. In Sec. II, we present the basic equations describing a general relativistic Fermi gas at statistical equilibrium in a box. In Sec. III, we expose general notions concerning the construction of the caloric curves and the description of phase transitions. In Sec. IV, we recall the results previously obtained in the nonrelativistic and classical limits. In Sec. V, we consider the case RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP} where the system undergoes a canonical phase transition from a gaseous phase to a condensed phase when NCCP<N<NeN_{\rm CCP}<N<N_{e}. In Sec. VI, we consider the case R>RMCPR>R_{\rm MCP} where the system undergoes a canonical phase transition when NCCP<N<NeN_{\rm CCP}<N<N_{e} and a microcanonical phase transition when NMCP<N<NfN_{\rm MCP}<N<N_{f} (we find that NeNOVN_{e}\sim N_{\rm OV} and Nf3.73NOVN_{f}\sim 3.73\,N_{\rm OV}). In Sec. VII, we consider the case of very large radii RRMCPR\gg R_{\rm MCP} where extreme core-halo configurations with a high central density appear. They correspond to the solutions computed in btv ; rar ; clm2 in connection to the “fermion ball” scenario. However, following clm2 , we point out that these solutions are thermodynamically unstable (hence very unlikely). In Secs. VIII and IX, we consider the cases ROV<R<RCCPR_{\rm OV}<R<R_{\rm CCP} and R<ROVR<R_{\rm OV} where there is no phase transition. In Sec. X, we present the complete phase diagram of the general relativistic Fermi gas in the (R,N)(R,N) plane. In Sec. XI, we recover the nonrelativistic and classical results as particular limits of our general study and we discuss the commutation of the limits 0\hbar\rightarrow 0 and c+c\rightarrow+\infty. In Sec. XII, we study the relativistic corrections to the nonrelativistic caloric curves and the quantum corrections to the classical caloric curves. In Sec. XIII, we consider astrophysical applications of our results in relation to the formation of white dwarfs, neutron stars, dark matter fermion stars, and black holes. We also connect the phase transitions found in our study with the onset of the red-giant structure and with the supernova phenomenon.

II Basic equations of a general relativistic Fermi gas

In this section, we give the basic equations describing the structure of a general relativistic Fermi gas at nonzero temperature (see bvcqg ; bvr ; papiertheorique for their derivation). Using the normalized variables introduced in Appendix B, the local number density n(r)n(r), the energy density ϵ(r)\epsilon(r), the pressure P(r)P(r) and the temperature T(r)T(r) are related to the gravitational potential Φ(r)\Phi(r) by

n(r)=1π20+y2dy1+eαe|α|(y2+1)/(Φ(r)+1),n(r)=\frac{1}{\pi^{2}}\int_{0}^{+\infty}\frac{y^{2}\,{\rm d}y}{1+e^{-\alpha}e^{|\alpha|\sqrt{(y^{2}+1)/(\Phi(r)+1)}}}, (1)
ϵ(r)=1π20+y21+y2dy1+eαe|α|(y2+1)/(Φ(r)+1),\epsilon(r)=\frac{1}{\pi^{2}}\int_{0}^{+\infty}\frac{y^{2}\sqrt{1+y^{2}}\,{\rm d}y}{1+e^{-\alpha}e^{|\alpha|\sqrt{(y^{2}+1)/(\Phi(r)+1)}}}, (2)
P(r)=13π20+y4dy1+y2[1+eαe|α|(y2+1)/(Φ(r)+1)],P(r)=\frac{1}{3\pi^{2}}\int_{0}^{+\infty}\frac{y^{4}\,{\rm d}y}{\sqrt{1+y^{2}}\left[1+e^{-\alpha}e^{|\alpha|\sqrt{(y^{2}+1)/(\Phi(r)+1)}}\right]}, (3)
T(r)=1|α|Φ(r)+1,T(r)=\frac{1}{|\alpha|}\sqrt{\Phi(r)+1}, (4)

where

α=μ(r)T(r)=μT\alpha=\frac{\mu(r)}{T(r)}=\frac{\mu_{\infty}}{T_{\infty}} (5)

is a quantity that is uniform throughout the system. These equations define the equation of state of a relativistic Fermi gas in parametric form.

The Tolman-Oppenheimer-Volkoff (TOV) equations, which correspond to the equations of hydrostatic equilibrium in general relativity, can be written as

dΦdr=2[Φ(r)+1]M(r)+4πP(r)r3r2[12M(r)r],\frac{{\rm d}\Phi}{{\rm d}r}=-2\left[\Phi(r)+1\right]\frac{M(r)+4\pi P(r)r^{3}}{r^{2}\left[1-\frac{2M(r)}{r}\right]}, (6)
dMdr=4πϵ(r)r2,\frac{{\rm d}M}{{\rm d}r}=4\pi\epsilon(r)r^{2}, (7)

where M(r)M(r) is the mass-energy within the sphere of radius rr. They have to be solved with the boundary conditions

M(0)=0,Φ(0)=Φ0>1.M(0)=0,\qquad\Phi(0)=\Phi_{0}>-1. (8)

We assume that the system is confined within a box of radius RR. The total mass of the gas and the total particle number are given by

M=M(R)=0Rϵ(r)4πr2dr,M=M(R)=\int_{0}^{R}\epsilon(r)4\pi r^{2}\,{\rm d}r, (9)
N=0Rn(r)[12M(r)r]1/24πr2dr.N=\int_{0}^{R}n(r)\left[1-\frac{2M(r)}{r}\right]^{-1/2}4\pi r^{2}\,{\rm d}r. (10)

The temperature at infinity is given by

T=T(R)(12MR)1/2,T_{\infty}=T(R)\left(1-\frac{2M}{R}\right)^{1/2}, (11)

where T(R)T(R) is the temperature of the system on the edge of the box. Using Eq. (4), we obtain

T=1|α|Φ(R)+1(12MR)1/2.T_{\infty}=\frac{1}{|\alpha|}\sqrt{\Phi(R)+1}\left(1-\frac{2M}{R}\right)^{1/2}. (12)

The entropy is given by

S=0RP+ϵT[12M(r)r]1/24πr2drαN.S=\int_{0}^{R}\frac{P+\epsilon}{T}\left[1-\frac{2M(r)}{r}\right]^{-1/2}4\pi r^{2}\,{\rm d}r-\alpha N. (13)

Finally, the free energy is given by

F=ETS,F=E-T_{\infty}S, (14)

where E=MNE=M-N is the binding energy.121212The binding energy is usually defined as Eb=Nmc2Mc2E_{\rm b}=Nmc^{2}-Mc^{2}. Here, for convenience, we define it with the opposite sign, i.e., E=Mc2Nmc2E=Mc^{2}-Nmc^{2}. In the Newtonian limit, MNmM\simeq Nm and EE reduces to the usual energy E=K+WE=K+W which is the sum of the kinetic and potential (gravitational) energies.

III Caloric curves and phase transitions

In order to study the phase transitions in the general relativistic Fermi gas we have to determine the caloric curves T(E)T_{\infty}(E) relating the temperature at infinity TT_{\infty} to the energy EE. These caloric curves depend on two parameters RR and NN. The manner to obtain these caloric curves is detailed in Appendix C. In order to make the connection with the nonrelativistic results ijmpb , we shall plot the caloric curves in terms of the dimensionless parameters η\eta (inverse temperature) and Λ\Lambda (minus energy) defined by

η=βGNm2RandΛ=ERGN2m2,\eta=\frac{\beta GNm^{2}}{R}\qquad{\rm and}\qquad\Lambda=\frac{-ER}{GN^{2}m^{2}}, (15)

where β=1/(kBT)\beta=1/(k_{\rm B}T_{\infty}) and E=Mc2Nmc2E=Mc^{2}-Nmc^{2}. In terms of our normalized variables, they reduce to

η=βNRandΛ=ERN2,\eta=\frac{\beta N}{R}\qquad{\rm and}\qquad\Lambda=\frac{-ER}{N^{2}}, (16)

where β=1/T\beta=1/T_{\infty} and E=MNE=M-N. We shall therefore plot the caloric curves η(Λ)\eta(\Lambda) as a function of RR and NN.

We recall that for systems with long-range interactions, such as self-gravitating systems, the statistical ensembles are not equivalent. In this paper, we shall consider the microcanonical and canonical ensembles separately.

In the microcanonical ensemble, the system is isolated so that its energy EE is conserved. It serves as a control parameter. A stable equilibrium state is a (local) maximum of entropy SS at fixed energy EE and particle number NN. A minimum, or a saddle point, of entropy is unstable. The global maximum of entropy corresponds to the most probable state (the one that is the most represented at the microscopic level). The microcanonical caloric curve gives the temperature at infinity 1/T=S/E1/T_{\infty}=\partial S/\partial E as a function of the energy EE.

In the canonical ensemble, the system is in contact with a heat bath so that its temperature at infinity TT_{\infty} is fixed. It serves as a control parameter. A stable equilibrium state is a (local) minimum of free energy FF at fixed temperature TT_{\infty} and particle number NN. A maximum, or a saddle point, of free energy is unstable. The global minimum of free energy corresponds to the most probable state. The caloric curve gives the average energy E=(βF)/βE=\partial(\beta F)/\partial\beta as a function of the temperature at infinity TT_{\infty}.

The equilibrium states are the same in the microcanonical and canonical ensembles. This is because an extremum (first variations) of entropy at fixed energy and particle number coincides with an extremum of free energy at fixed particle number. However, their stability (second variations) may differ in the microcanonical and canonical ensembles. A configuration that is stable in the canonical ensemble is necessarily stable in the microcanoniocal ensemble but the converse is wrong. As a corollary we recall that the specific heat C=dE/dT=NkBη2dΛ/dηC=dE/dT_{\infty}=Nk_{B}\eta^{2}d\Lambda/d\eta of stable equilibrium states is always positive in the canonical ensemble while it can be positive or negative in the microcanonical ensemble (for systems with long-range interactions).

The stability of the solutions can be determined by using the Poincaré turning point criterion poincare . We refer to the papers of Katz katzpoincare1 ; katzpoincare2 for a presentation and a generalization of this criterion, and for its application to the nonrelativistic classical self-gravitating gas. This method was applied to the nonrelativistic self-gravitating Fermi gas in ijmpb . We use the same method in the present paper.

In the discussion of the caloric curves, we shall only consider stable states. An equilibrium state that is a local, but not a global, extremum of the relevant thermodynamical potential (entropy in the microcanonial ensemble and free energy in the canonical ensemble) is said to be metastable. A global extremum of the thermodynamical potential is said to be fully stable. For systems with short-range interactions, metastable states have a short lifetime so that the caloric curve should contain only fully stable states. However, for systems with long-range interactions, the metastable states have a very long lifetime scaling as eNe^{N} which is usually much longer than the age of the Universe. As a result, metastable states can be as much, or even more, relevant than fully stable states lifetime . The selection between a fully stable state or a metastable state depends on the initial condition and on a notion of basin of attraction. In this paper, we shall not distinguish between metastable and fully stable states. The physical caloric curve should contain all types of stable equilibrium states.131313The existence, or nonexistence, of fully stable states for self-gravitating fermions in general relativity is an interesting problem by itself but it will not be considered in the present paper (see the Remark at the end of Sec. V.3 showing that this problem is not trivial).

For real systems, that are not in a box, the natural evolution proceeds along the series of equilibria towards larger and larger density contrasts.141414The reason is that, for real systems (globular clusters, dark matter halos…) such as those described by the King model, the Boltzmann or Fermi-Dirac entropy (resp. the Boltzmann or Fermi-Dirac free energy) increases (resp. decreases) with the concentration parameter; see Fig. 5 of cohn and Fig. 46 of clm2 . Note that, surprisingly, for box-confined systems this is the opposite; see Fig 3 of pt . In general, this corresponds to lower and lower temperatures and energies.151515This is explicitly shown in Figs. 12 and 15 below. Note that this result is valid only for mid and low energies and temperatures. At very high energies and temperatures, where the system behaves as a self-gravitating radiation, the density contrast increases with the energy and the temperature (see Figs. 2 and 3 of paper2 ) implying that the natural evolution of the system is towards higher and higher energies and temperatures. This situation has been discussed in paper2 and will not be considered here. Therefore, in the discussion of the caloric curves, we shall describe the evolution of the system starting from high energies and high temperatures, and reducing the temperature and the energy until an instability takes place.

IV Particular limits

In this section, we briefly recall well-known results that correspond to particular limits of the general relativistic Fermi gas.

IV.1 The nonrelativistic ++ classical limit

The thermodynamics of a nonrelativistic classical self-gravitating gas has been studied in detail in antonov ; lbw ; katzpoincare1 ; paddyapj ; aa . The caloric curve η(Λ)\eta(\Lambda) forms a spiral (see Fig. 1). In the microcanonical ensemble, there is no equilibrium state below a critical energy EcE_{c} corresponding to Λc=0.335\Lambda_{c}=0.335. In that case, the system undergoes a gravothermal catastrophe (core collapse) leading to a binary star surrounded by a hot halo lbe ; inagaki ; cohn . In the canonical ensemble, there is no equilibrium state below a critical temperature TcT_{c}, corresponding to ηc=2.52\eta_{c}=2.52. In that case, the system undergoes an isothermal collapse leading to a Dirac peak containing all the mass post .

Refer to caption
Figure 1: Caloric curve of the nonrelativistic classical self-gravitating gas.

IV.2 The nonrelativistic limit

The thermodynamics of the nonrelativistic self-gravitating Fermi gas has been studied in detail in ijmpb . It is shown that the caloric curves η(Λ)\eta(\Lambda) depend on a single control parameter (it should not be confused with the chemical potential):

μ=η0512π4G3NmR3,η0=gm4h3.\mu=\eta_{0}\sqrt{512\pi^{4}G^{3}NmR^{3}},\qquad\eta_{0}=\frac{gm^{4}}{h^{3}}. (17)

It can be written as ijmpb :

μ=17.3(RR0)3/2,R0=0.181h2Gm8/3g2/3M1/3,\mu=17.3\left(\frac{R}{R_{0}}\right)^{3/2},\quad R_{0}=0.181\frac{h^{2}}{Gm^{8/3}g^{2/3}M^{1/3}}, (18)

or as

μ=17.3(MM0)1/2,M0=5.97×103h6G3m8g2R3,\mu=17.3\left(\frac{M}{M_{0}}\right)^{1/2},\quad M_{0}=5.97\times 10^{-3}\frac{h^{6}}{G^{3}m^{8}g^{2}R^{3}}, (19)

where R0R_{0} (resp. M0M_{0}) is the radius (resp. mass) of a fermion star of mass MM (resp. radius RR) at T=0T=0 (see Appendix F). Introducing the normalized variables of Appendix B, this parameter becomes

μ=42π(NR3)1/2.\mu=\frac{4\sqrt{2}}{\pi}(NR^{3})^{1/2}. (20)

Some caloric curves are represented in Fig. 2. They display a canonical critical point at μCCP=83\mu_{\rm CCP}=83 and a microcanonical critical point at μMCP=2670\mu_{\rm MCP}=2670. When μ<μCCP=83\mu<\mu_{\rm CCP}=83 there is no phase transition. When μCCP=83<μ<μMCP=2670\mu_{\rm CCP}=83<\mu<\mu_{\rm MCP}=2670 the system displays zeroth and first order canonical phase transitions. When μ>μMCP=2670\mu>\mu_{\rm MCP}=2670 the system displays zeroth and first order canonical and microcanonical phase transitions. When μ+\mu\rightarrow+\infty we recover the caloric curve of the nonrelativistic classical self-gravitating gas (spiral) represented in Fig. 1. When μ<+\mu<+\infty there is a statistical equilibrium state for any accessible value of energy and temperature. The gravitational collapse of the nonrelativistic classical self-gravitating gas (gravothermal catastrophe in the microcanonical ensemble and isothermal collapse in the canonical ensemble) is prevented by quantum mechanics (Pauli’s exclusion principle).

Refer to caption
Figure 2: Caloric curves of the nonrelativistic self-gravitating Fermi gas for different values of μ=(42/π)(NR3)1/2\mu=(4\sqrt{2}/{\pi})(NR^{3})^{1/2}.

For a given box radius, the nonrelativistic canonical phase transition appears when

N>NCCP(R)(πμCCP42)21R3=2125R3.N>N_{\rm CCP}(R)\equiv\left(\frac{\pi\mu_{\rm CCP}}{4\sqrt{2}}\right)^{2}\frac{1}{R^{3}}=\frac{2125}{R^{3}}. (21)

If we consider the general relativistic problem, we must require NNOVN\ll N_{\rm OV}, where NOV=0.39853N_{\rm OV}=0.39853 is the OV limit, for the validity of the nonrelativistic treatment. Therefore, we will see the nonrelativistic canonical phase transition for NCCP(R)<NNOVN_{\rm CCP}(R)<N\ll N_{\rm OV} provided that

RRCCPapprox.(πμCCP42)2/31NOV1/3=17.5.R\gg R^{\rm approx.}_{\rm CCP}\equiv\left(\frac{\pi\mu_{\rm CCP}}{4\sqrt{2}}\right)^{2/3}\frac{1}{N_{\rm OV}^{1/3}}=17.5. (22)

In comparison ROV=3.3569R_{\rm OV}=3.3569. This argument just provides an order of magnitude of the radius RCCPR_{\rm CCP} above which a canonical phase transition appears for N>NCCP(R)N>N_{\rm CCP}(R). By solving the general relativistic equations, we find that the exact value is RCCPexact=12.0R_{\rm CCP}^{\rm exact}=12.0 (see Sec. X).

For a given box radius, the nonrelativistic microcanonical phase transition appears when

N>NMCP(R)(πμMCP42)21R3=2.20×106R3.N>N_{\rm MCP}(R)\equiv\left(\frac{\pi\mu_{\rm MCP}}{4\sqrt{2}}\right)^{2}\frac{1}{R^{3}}=\frac{2.20\times 10^{6}}{R^{3}}. (23)

If we consider the general relativistic problem, using the same argument as before, we will see the nonrelativistic microcanonical phase transition for NMCP(R)<NNOVN_{\rm MCP}(R)<N\ll N_{\rm OV} provided that

RRMCPapprox.(πμMCP42)2/31NOV1/3=177.R\gg R^{\rm approx.}_{\rm MCP}\equiv\left(\frac{\pi\mu_{\rm MCP}}{4\sqrt{2}}\right)^{2/3}\frac{1}{N_{\rm OV}^{1/3}}=177. (24)

This argument just provides an order of magnitude of the radius RMCPR_{\rm MCP} above which a microcanonical phase transition appear for N>NMCP(R)N>N_{\rm MCP}(R). By solving the general relativistic equations, we find that the exact value is RMCPexact=92.0R_{\rm MCP}^{\rm exact}=92.0 (see Sec. X).

IV.3 The classical limit

The thermodynamics of a classical self-gravitating gas in general relativity has been studied in detail in Refs. roupas and paper2 . This corresponds to the nondegenerate limit of the general relativistic Fermi gas. It is shown that the caloric curves η(Λ)\eta(\Lambda) depend on a single control parameter

ν=GNmRc2.\nu=\frac{GNm}{Rc^{2}}. (25)

It can be written as

ν=RS2R,RS=2GNmc2,\nu=\frac{R^{*}_{S}}{2R},\qquad R^{*}_{S}=\frac{2GNm}{c^{2}}, (26)

or as

ν=N2NS,NS=Rc22Gm,\nu=\frac{N}{2N^{*}_{S}},\qquad N^{*}_{S}=\frac{Rc^{2}}{2Gm}, (27)

where RSR^{*}_{S} can be interpreted as a sort of Schwarzschild radius defined with the rest mass NmNm instead of the mass MM (reciprocally, NSmN^{*}_{S}m is a sort of Schwarzschild rest mass of an object of radius RR). Introducing the normalized variables of Appendix B this parameter becomes

ν=NR.\nu=\frac{N}{R}. (28)
Refer to caption
Figure 3: Caloric curves of the classical self-gravitating gas in general relativity for different values of ν=N/R\nu=N/R.

Some caloric curves are represented in Fig. 3. When ν0\nu\rightarrow 0 (NNSN\ll N_{S}^{*} or RRSR\gg R_{S}^{*}), we recover the caloric curve of the nonrelativistic classical self-gravitating gas (spiral) represented in Fig. 1. When 0<ν<νS=0.1280<\nu<\nu^{\prime}_{S}=0.128 the caloric curve has the form of a double spiral exhibiting a collapse at low energies and low temperatures (cold spiral) and at high energies and high temperatures (hot spiral).161616The hot spiral corresponds to an ultrarelativistic classical gas roupas which is similar to a form of radiation described by an equation of state P=ϵ/3P=\epsilon/3 sorkin ; aarelat1 ; aarelat2 (see paper2 for a detailed discussion). When νS=0.128<ν<νS=0.1415\nu^{\prime}_{S}=0.128<\nu<\nu_{S}=0.1415 the two spirals are amputed (truncated) and touch each other. When νS=0.1415<ν<νmax=0.1764\nu_{S}=0.1415<\nu<\nu_{\rm max}=0.1764 the two spirals disappear and the caloric curve makes a loop resembling to the symbol “\infty”. As ν\nu increases, the loop shrinks more and more and, when ν=νmax=0.1764\nu=\nu_{\rm max}=0.1764, it reduces to a point located at (Λ,η)=(0.9829,1.2203)(\Lambda_{*},\eta_{*})=(-0.9829,1.2203). When ν>νmax=0.1764\nu>\nu_{\rm max}=0.1764, no equilibrium state is possible.

For a given box radius, the spirals touch each other when

N>NS(R)=νSR=0.128RN>N^{\prime}_{S}(R)=\nu^{\prime}_{S}R=0.128\,R (29)

and they form a loop when

N>NS(R)=νSR=0.1415R.N>N_{S}(R)=\nu_{S}R=0.1415\,R. (30)

The caloric curve reduces to a point when

N=Nmax(R)=νmaxR=0.1764R.N=N_{\rm max}(R)=\nu_{\rm max}R=0.1764\,R. (31)

If we consider the truly quantum problem, we must require NNOV=0.39853N\gg N_{\rm OV}=0.39853 for the validity of the classical (nondegenerate) treatment. Therefore, we will see the double spiral and its evolution described previously for NOVN<NmaxN_{\rm OV}\ll N<N_{\rm max} provided that

RNOVνmax=2.259.R\gg\frac{N_{\rm OV}}{\nu_{\rm max}}=2.259. (32)

We note that NOV/νmax=2.259N_{\rm OV}/\nu_{\rm max}=2.259 is of the order of ROV=3.3569R_{\rm OV}=3.3569.

Remark: For a given box radius RR, coming back to dimensional variables, equilibrium states exist only when NNmax(R)=0.1764Rc2/Gm=0.3528NSN\leq N_{\rm max}(R)=0.1764\,Rc^{2}/Gm=0.3528\,N_{S}^{*}. Inversely, for a given number of particles NN, equilibrium states exist only when RRmin(N)=5.669GNm/c2=2.834RSR\geq R_{\rm min}(N)=5.669\,GNm/c^{2}=2.834\,R^{*}_{S}. The nonrelativistic limit corresponds to NNmax(R)Rc2/2GmNSN\ll N_{\rm max}(R)\sim Rc^{2}/2Gm\sim N_{S}^{*} or RRmin(N)2GNm/c2RSR\gg R_{\rm min}(N)\sim 2GNm/c^{2}\sim R^{*}_{S}. These results are valid in the classical limit. For small systems, quantum effects will come into play. If we argue that Nmax=νmaxRc2/GmNOVN_{\rm max}=\nu_{\rm max}Rc^{2}/Gm\sim N_{\rm OV} when RROVR\sim R_{\rm OV}, or equivalently Rmin=GNm/νmaxc2ROVR_{\rm min}=GNm/\nu_{\rm max}c^{2}\sim R_{\rm OV} when NNOVN\sim N_{\rm OV}, we find that νmaxGNOVm/ROVc20.1187\nu_{\rm max}\sim GN_{\rm OV}m/R_{\rm OV}c^{2}\sim 0.1187. This may justify the order of magnitude of this constant. Alternatively, we may just remark that νmax=GNmaxm/Rc2=0.1764\nu_{\rm max}=GN_{\rm max}m/Rc^{2}=0.1764 is of the same order as GNOVm/ROVc2=0.1187GN_{\rm OV}m/R_{\rm OV}c^{2}=0.1187.

IV.4 Summary

Before treating the general case, let us summarize the previous results.

Nonrelativistic ++ classical limit. For a given box radius RR and particle number NN the system undergoes a catastrophic collapse towards a singularity at low temperatures in the canonical ensemble and at low energies in the microcanonical ensemble.

Nonrelativistic limit. For a given box radius RR there is no phase transition when N<NCCP(R)N<N_{\rm CCP}(R), the system can undergo a canonical phase transition when NCCP(R)<N<NMCP(R)N_{\rm CCP}(R)<N<N_{\rm MCP}(R), and the system can undergo a canonical and a microcanonical phase transition when N>NMCP(R)N>N_{\rm MCP}(R). For a given particle number NN, there is no phase transition when R<RCCP(N)R<R_{\rm CCP}(N), the system can undergo a canonical phase transition when RCCP(N)<R<RMCP(N)R_{\rm CCP}(N)<R<R_{\rm MCP}(N), and the system can undergo a canonical and a microcanonical phase transition when R>RMCP(N)R>R_{\rm MCP}(N). Here, RCCP(N)=12.9/N1/3R_{\rm CCP}(N)=12.9/N^{1/3} and RMCP(N)=130/N1/3R_{\rm MCP}(N)=130/N^{1/3} are the reciprocal of NCCP(R)=2125/R3N_{\rm CCP}(R)=2125/R^{3} and NMCP(R)=2.20×106/R3N_{\rm MCP}(R)=2.20\times 10^{6}/R^{3}. There is an equilibrium state at all temperatures T0T\geq 0 in the canonical ensemble and at all accessible energies EEminE\geq E_{\rm min} (where EminE_{\rm min} is the energy of the ground state) in the microcanonical ensemble.

Classical limit. For a given box radius RR, the caloric curve has the form of a double spiral when N<NS(R)N<N_{\rm S}^{\prime}(R), the spirals touch each other when NS(R)<N<NS(R)N_{\rm S}^{\prime}(R)<N<N_{\rm S}(R), the caloric curve makes a loop when NS(R)<N<Nmax(R)N_{\rm S}(R)<N<N_{\rm max}(R), and there is no equilibrium state when N>Nmax(R)N>N_{\rm max}(R). For a given particle number NN, the caloric curve has the form of a double spiral when R>RS(N)R>R_{\rm S}^{\prime}(N), the spirals touch each other when RS(N)<R<RS(N)R_{\rm S}(N)<R<R^{\prime}_{\rm S}(N), the caloric curve makes a loop when Rmin(N)<R<RS(N)R_{\rm min}(N)<R<R_{\rm S}(N), and there is no equilibrium state when R<Rmin(N)R<R_{\rm min}(N). Here, RS(N)=7.81NR_{\rm S}^{\prime}(N)=7.81\,N, RS(N)=7.07NR_{\rm S}(N)=7.07\,N and Rmin(N)=5.67NR_{\rm min}(N)=5.67\,N are the reciprocal of NS(R)=0.128RN_{\rm S}^{\prime}(R)=0.128\,R, NS(R)=0.1415RN_{\rm S}(R)=0.1415\,R and Nmax(R)=0.1764RN_{\rm max}(R)=0.1764\,R. The system undergoes a catastrophic collapse towards a singularity at both low and high temperatures in the canonical ensemble and at both low and high energies in the microcanonical ensemble.

V The case RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP}

In this section, we study the general relativistic Fermi gas in the case RCCP=12.0<R<RMCP=92.0R_{\rm CCP}=12.0<R<R_{\rm MCP}=92.0 where only a canonical phase transition may occur (see Fig. 47 below). For illustration, we select R=50R=50. For this value of RR, the canonical phase transition occurs above NCCP=0.0170N_{\rm CCP}=0.0170.

V.1 The case N<NCCPN<N_{\rm CCP}

In Fig. 4 we have plotted the caloric curve for N<NCCP=0.0170N<N_{\rm CCP}=0.0170. Since NNOV=0.39853N\ll N_{\rm OV}=0.39853, this caloric curve coincides with the one obtained in the nonrelativistic limit ijmpb except at very high energies and very high temperatures (see the Remark at the end of this section).171717As discussed in Sec. XI the nonrelativistic limit corresponds to N0N\rightarrow 0 and R+R\rightarrow+\infty in such a way that NR3NR^{3} is fixed (in more physical terms NNOVN\ll N_{\rm OV} and RROVR\gg R_{\rm OV} with NR3NR^{3} fixed).

Refer to caption
Figure 4: Caloric curve for N<NCCP=0.0170N<N_{\rm CCP}=0.0170 (specifically R=50R=50 and N=0.012N=0.012).

The series of equilibria η(Λ)\eta(\Lambda) is monotonic. According to the Poincaré theory of linear series of equilibria, all the equilibrium states are stable. The statistical ensembles (microcanonical and canonical) are equivalent. The caloric curve presents the following features:

(i) There is no phase transition and no gravitational collapse.

(ii) The specific heat is always positive. The entropy versus energy curve (not represented) is concave.

The evolution of the system is the following. At high energies and high temperatures, the system is nondegenerate (Boltzmannian). As the energy and the temperature are reduced, the system becomes more and more centrally condensed. At intermediate energies and intermediate temperatures, the Fermi gas is partially degenerate (see Appendix D). At T=0T=0, the Fermi gas is completely degenerate. This cold nonrelativistic fermion ball, equivalent to a polytrope of index n=3/2n=3/2, is similar to a nonrelativistic white dwarf. This is the state of minimum energy EminE_{\rm min} (ground state). Since there is a stable equilibrium state at T=0T=0 (i.e. η+\eta\rightarrow+\infty) with a finite energy EminE_{\rm min}, the caloric curve η(Λ)\eta(\Lambda) presents a vertical asymptote at Λ=Λmax\Lambda=\Lambda_{\rm max}.181818In the nonrelativistic limit Λmax=0.0950N1/3R\Lambda_{\rm max}=0.0950N^{1/3}R (see Appendix F). More generally, a complete characterization of the ground state of the self-gravitating Fermi gas, in the nonrelativistic and relativistic regimes, taking into account the presence of the box is given in paper1 .

Remark: At very high energies and very high temperatures, the system is relativistic even though NNOVN\ll N_{\rm OV}. In that case, we recover the hot spiral studied in roupas ; paper2 . As a result, the complete caloric curve of the general relativistic Fermi gas presents a region of negative specific heat and a region of ensemble inequivalence at very high energies and very high temperatures. The system undergoes a gravitational collapse above EmaxE_{\rm max} in the microcanonocal ensemble and above TmaxT_{\rm max} in the canonical ensemble. We note that quantum mechanics cannot prevent such a gravitational collapse since it takes place at very high energies and very high temperatures where the system is nondegenerate. As a result, the system is expected to collapse towards a black hole. For N0N\rightarrow 0, it is shown in paper2 that Λmin0.246/N2\Lambda_{\rm min}\sim-0.246/N^{2}\rightarrow-\infty and ηmin18.3N20\eta_{\min}\sim 18.3N^{2}\rightarrow 0 so that the hot spiral is rejected at infinity.191919In terms of dimensional variables this correponds to Emax0.24631Rc4/GE_{\rm max}\rightarrow 0.24631Rc^{4}/G and kBTmax0.0547Rc4/NGk_{B}T_{\rm max}\sim 0.0547Rc^{4}/NG. For small values of NN (NNmaxN\ll N_{\rm max}) the hot spiral occurs at very negative values of Λ\Lambda and at very small values of η\eta. This is why we do not see it in Fig. 4 (it is outside of the frame since Λmin1708\Lambda_{\rm min}\simeq-1708 and ηmin2.63×103\eta_{\min}\simeq 2.63\times 10^{-3}). The hot spiral becomes visible only for larger values of NN (NNmaxN\lesssim N_{\rm max}) as in Fig. 22 below. In this paper, we shall not discuss the hot spiral specifically since it has been described in detail in roupas ; paper2 .

V.2 The case NCCP<N<N1N_{\rm CCP}<N<N_{1}

In Fig. 5 we have plotted the caloric curve for NCCP=0.0170<N<N1=0.18131N_{\rm CCP}=0.0170<N<N_{1}=0.18131. Since NNOV=0.39853N\ll N_{\rm OV}=0.39853, the caloric curve coincides with the one obtained in the nonrelativistic limit ijmpb . The novely with respect to the previous case is that the caloric curve has a NN-shape structure leading to canonical phase transitions and ensembles inequivalence. This NN-shape structure appears at N=NCCP=0.0170N=N_{\rm CCP}=0.0170 where the caloric curve presents a horizontal inflexion point. Let us consider the microcanonical and canonical ensembles successively (see ijmpb for a more detailed discussion).

Refer to caption
Figure 5: Caloric curve for NCCP=0.0170<N<N1=0.18131N_{\rm CCP}=0.0170<N<N_{1}=0.18131 (specifically R=50R=50 and N=0.15N=0.15).

V.2.1 Microcanonical ensemble

The curve η(Λ)\eta(\Lambda) is univalued. According to the Poincaré theory, the whole series of equilibria is stable. The caloric curve presents the following features:

(i) There is no phase transition and no gravitational collapse.

(ii) There is a region of negative specific heats between Λgas\Lambda_{\rm gas} and Λcond\Lambda_{\rm cond}. In this range of intermediate energies the system is purely self-gravitating, i.e., it almost does not feel the quantum pressure (Pauli exclusion principle) nor the pressure of the box. The negative specific heat leads to a convex intruder (dip) in the entropy versus energy curve (see Fig. 25 of ijmpb ).

The evolution of the system in the microcanonical ensemble is the following. Let us start from high energy states and decrease the energy. At high energies, the system is almost homogeneous. As energy decreases, and especially when we enter in the region of negative specific heats, the system becomes more and more concentrated and partially degenerate. At the minimum energy EminE_{\rm min} (ground state) the system is completely degenerate. There is no phase transition, just a progressive clustering of the system until the ground state is reached.

V.2.2 Canonical ensemble

The curve Λ(η)\Lambda(\eta) is multivalued leading to the possibility of phase transitions in the canonical ensemble. The left branch up to ηc\eta_{c} corresponds to the gaseous phase and the right branch after η\eta_{*} corresponds to the condensed phase. According to the Poincaré turning point criterion, these equilibrium states are stable while the equilibrium states on the intermediate branch between ηc\eta_{c} and η\eta_{*} are unstable. These equilibrium states have a core-halo structure (see below) and a negative specific heat. This is a sufficient (but not necessary) condition of instability in the canonical ensemble. The caloric curve presents the following features:

(i) When η<η\eta<\eta_{*} there are only gaseous states. When η>ηc\eta>\eta_{c} there are only condensed states. When η<η<ηc\eta_{*}<\eta<\eta_{c} there exist gaseous and condensed states at the same temperature. A first order phase transition is expected at a transition temperature ηt\eta_{t} determined by the Maxwell construction (see Fig. 5) or by the equality of the free energy of the gaseous and condensed phases (see Fig. 28 of ijmpb ). When η<η<ηt\eta_{*}<\eta<\eta_{t} the gaseous states have a lower free energy than the condensed states. When ηt<η<ηc\eta_{t}<\eta<\eta_{c} the condensed states have a lower free energy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of the metastable states.

(ii) There is a zeroth order phase transition at ηc\eta_{c} from the gaseous phase to the condensed phase. It corresponds to a gravitational collapse (isothermal collapse) ultimately halted by quantum degeneracy.

(iii) There is a zeroth order phase transition at η\eta_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

The evolution of the system in the canonical ensemble is the following. Let us start from high temperature states and decrease the temperature. At high temperatures the system is in the gaseous phase. At η=ηt\eta=\eta_{t}, the system is expected to undergo a first order phase transition from the gaseous phase to the condensed phase. However, in practice, this phase transition does not take place because the metastable gaseous states have a very long lifetime. At η=ηc\eta=\eta_{c} the system collapses towards the condensed phase. Complete gravitational collapse is prevented by quantum mechanics. The system reaches an equilibrium state similar to a nonrelativistic white dwarf (fermion ball). If we now increase the temperature the system remains in the condensed phase until the point η\eta_{*} (again, the first order phase transition expected at ηt\eta_{t} does not take place because the metastable condensed states have a very long lifetime) at which it explodes and returns to the gaseous phase. We have thus decribed an hysteretic cycle in the canonical ensemble ijmpb .

V.2.3 Density profiles

In Fig. 6 we have plotted the density profiles of the gaseous (G), core-halo (CH) and condensed (C) states at the transition point ηt\eta_{t}. We note that the energy density is very low confirming that we are in the nonrelativistic regime.

Refer to caption
Figure 6: Energy density profiles of gaseous, core-halo and condensed states at η=ηt\eta=\eta_{t} (specifically ηt=1.9285\eta_{t}=1.9285). The dashed line represents the density profile of a nonrelativistic fermion ball at T=0T=0 (similar to a nonrelativistic white dwarf).

(i) In the gaseous phase (high energies and high temperatures), quantum mechanics is negligible and the density profile is dilute. The equilibrium state results from the competition between the gravitational attraction and the thermal pressure. The gaseous equilibrium state (G) is almost uniform because the temperature is high so that the thermal pressure overcomes the gravitational attraction. In that case, the gas is held by the walls of the box.

(ii) In the condensed phase (low energies and low temperatures), thermal effects are negligible and the density profile is very compact. The equilibrium state results from the competition between the gravitational attraction and the quantum pressure arising from the Pauli exclusion principle. The condensed equilibrium state (C) almost coincides with a nonrelativistic fermion ball at T=0T=0 containing all the mass (see ijmpb and Appendix E.2.1). It is similar to a nonrelativistic white dwarf corresponding to a polytrope n=3/2n=3/2. In that case, gravitational collapse is prevented by quantum mechanics and the confining box is not necessary. At small but finite temperatures, we see in Fig. 6 that the dashed line corresponding to a polytrope n=3/2n=3/2 provides a good fit to the core of the distribution. There is a small isothermal atmosphere that becomes thiner and thiner as the temperature is reduced.

(iii) The intermediate state (CH) has a sort of core-halo structure with a degenerate core and an isothermal atmosphere. The equilibrium state results from the competition between the gravitational attraction, the thermal pressure, and the quantum pressure. The pressure of the box and the quantum pressure have a weak effect on the equilibrium of the system so it essentially behaves as a self-gravitating isothermal gas. This is why it presents a negative specific heat.

Let us recall that the these three equilibrium states have the same temperature but different energies. The core-halo state (CH) is unstable in the canonical ensemble while it is stable in the microcanonical ensemble. It lies in a region of negative specific heats. The gaseous and condensed states (G) and (C) are stable in both ensembles.

V.3 The case N1<N<NOVN_{1}<N<N_{\rm OV}

In Fig. 7 we have plotted the caloric curve for N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853. The novelty with respect to the previous case is the existence of a secondary branch presenting an asymptote at Λmax\Lambda^{\prime}_{\rm max}. This secondary branch appears suddently at N=N1=0.18131N=N_{1}=0.18131 (at that point Λmax=0.536R\Lambda^{\prime}_{\rm max}=-0.536\,R and Λmax=0.0570R\Lambda_{\rm max}=0.0570\,R). As detailed in paper1 , for N1<N<NOVN_{1}<N<N_{\rm OV}, there exists another equilibrium state at T=0T=0 (i.e. η+\eta\rightarrow+\infty) corresponding to a completely degenerate fermion ball distinct from the ground state. This secondary equilibrium state is unstable.202020Actually, for N>N1N>N_{1}, there can exist several unstable equilibrium states at T=0T=0 (up to an infinity) that have more and more modes of instability. They are related to the spiral structure of the mass-radius relation of the general relativistic Fermi gas at T=0T=0 shapiroteukolsky ; paper1 . They give rise to additional branches (with vertical asymptotes) in the caloric curve. We shall not consider these unstable solutions here, except for the less unstable one already mentioned. Its mass is larger than the mass of the stable ground state so that ΛmaxΛmax\Lambda^{\prime}_{\rm max}\leq\Lambda_{\rm max}. According to the Poincaré theory, all the configurations of the secondary branch are unstable.212121The spiral present on the left of this secondary branch will ultimately become the cold spiral of Refs. roupas ; paper2 when NN will be sufficiently large (see below). Therefore, the presence of this secondary branch does not qualitatively change the description of the caloric curve made in Sec. V.2.

Refer to caption
Figure 7: Caloric curve for N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853 (specifically R=50R=50 and N=0.29N=0.29).

However, for N>N1N>N_{1}, relativistic effects start to become important. This has some consequences on the interpretation of the density profiles. In Fig. 8 we have plotted the different density profiles at ηt\eta_{t}. We see that the energy density is low for the gaseous state (G) and for the core-halo state (CH) indicating that we are in the nonrelativistic regime. By contrast, the energy density is relatively high for the stable condensed state (C) and for the unstable condensed state (U) indicating that we are in the relativistic regime. The condensed states almost coincide with a general relativistic fermion ball at T=0T=0 containing all the mass (see Appendix E.2.1). They are similar to stable and unstable neutron stars ov . At small but finite temperatures, we see in Fig. 8 that the dashed line obtained from the OV theory provides a good fit to the core of the distribution. There is a small atmosphere (containing a little mass) that becomes thinner and thinner as the temperature is reduced.

Refer to caption
Figure 8: Density profiles of gaseous, core-halo and (stable and unstable) condensed states at η=ηt\eta=\eta_{t} (specifically ηt=1.5722\eta_{t}=1.5722). The dashed lines represent the density profiles of stable and unstable general relativistic fermion balls at T=0T=0 (similar to neutron stars).

Remark: In Fig. 7, when the temperature is low enough, we find four solutions. The solutions (G) and (C) are stable (local minima of free energy) while the solutions (CH) and (U) are unstable (saddle points of free energy). Since we have an even number of extrema, this suggests that there is no global minimum of free energy (naively, this results from simple topological arguments if we plot a curve f(x)f(x) with two minima and two maxima). The stable equilibrium state with the lowest value of free energy may be only metastable, not fully stable. This is consistent with the result of Zel’dovich zel446 who showed that, at T=0T=0, the OV equilibrium states are only metastable. In Fig. 5, when η<η<ηc\eta_{*}<\eta<\eta_{c}, we find three solutions. The solutions (G) and (C) are stable (local minima of free energy) while the solution (CH) is unstable (saddle point of free energy). Since we have an odd number of extrema, this suggests that the solution with the lowest value of free energy is a global minimum. This is the case in Newtonian gravity ijmpb . However, this is not quite clear in general relativity since the result of Zel’dovich zel446 still applies for N<N1N<N_{1}. Therefore, the existence of a global minimum of free energy (fully stable state) in general relativity is not trivial and would require a more careful study. Anyway, for practical purposes, metastable states are very relevant (possibly more relevant than fully stable states) so we shall determine all types of stable equilibrium states, disregarding whether they are fully stable or just metastable.

V.4 The case NOV<N<NeN_{\rm OV}<N<N_{e}

In Fig. 9 we have plotted the caloric curve for NOV=0.39853<N<Ne=0.40002N_{\rm OV}=0.39853<N<N_{e}=0.40002. The novelty with respect to the previous case is that the two branches have merged. The merging occurs at N=NOVN=N_{\rm OV} at which the two asymptotes Λmax\Lambda^{\prime}_{\rm max} and Λmax\Lambda_{\rm max} coincide (at that point Λmax=Λmax=0.08985R\Lambda_{\rm max}=\Lambda^{\prime}_{\rm max}=0.08985\,R). This is the highest value of NN at which there exist an equilibrium state at T=0T=0 (ground state). When N>NOVN>N_{\rm OV} there is no equilibrium state at T=0T=0 (no ground state) anymore ov . In that case, the caloric curve presents a turning point of temperature at ηc\eta^{\prime}_{c} and a turning point of energy at Λc\Lambda^{\prime}_{c}. As a result, there is no equilibrium state at η>ηc\eta>\eta^{\prime}_{c} in the canonical ensemble, i.e., below a critical temperature. Similarly, there is no equilibrium state at Λ>Λc\Lambda>\Lambda^{\prime}_{c} in the microcanonical ensemble, i.e., below a critical energy. This means that when the system becomes strongly relativistic (i.e. when N>NOVN>N_{\rm OV}) quantum mechanics is not able to prevent gravitational collapse at low temperatures and low energies. This is a generalization of the result first obtained at T=0T=0 by Oppenheimer and Volkoff ov in the context of neutron stars.

Refer to caption
Figure 9: Caloric curve for NOV=0.39853<N<Ne=0.40002N_{\rm OV}=0.39853<N<N_{e}=0.40002 (specifically R=50R=50 and N=0.399N=0.399).

V.4.1 Microcanonical ensemble

Let us first consider the microcanonical ensemble. The curve η(Λ)\eta(\Lambda) is multivalued. According to the Poincaré turning point criterion, the series of equilibria is stable up to Λc\Lambda_{c}^{\prime} and then becomes unstable. The caloric curve presents the following features:

(i) There is no phase transition (there is only one stable equilibrium state for each Λ<Λc\Lambda<\Lambda^{\prime}_{c}).

(ii) There are two regions of negative specific heats, one between Λgas\Lambda_{\rm gas} and Λcond\Lambda_{\rm cond} (as before) and another one between Λgas\Lambda^{\prime}_{\rm gas} (the energy corresponding to ηc\eta^{\prime}_{c}) and Λc\Lambda^{\prime}_{c}. We note that this second region of negative specific heats is extremely tiny. In Fig. 10 we clearly see the convex intruder (dip) associated with the first region of specific heat. The convex intruder associated with the second region of specific heat is imperceptible.

(iii) There is a catastrophic collapse at Λc\Lambda_{c}^{\prime} towards a black hole.222222For simplicity, when there is no equilibrium state, we shall say that the system forms a black hole. Actually, as discussed in Paper II, it is not completely clear that the system will always form a black hole in that case. We leave this interesting problem open to future works.

Refer to caption
Figure 10: Entropy per fermion as a function of the normalized energy for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399). We can check that the unstable equilibrium states (saddle points of entropy) have an entropy lower than the stable equilibrium states (entropy maxima).

In Fig. 11 we have plotted the relation Λ(Φ0)\Lambda(\Phi_{0}) between the normalized energy and the central potential. We can see that Φ0\Phi_{0} increases monotonically along the series of equilibria. The curve Λ(Φ0)\Lambda(\Phi_{0}) presents a peak at Λc\Lambda^{\prime}_{c} then displays damped oscillations. These oscillations correspond to the unstable equilibrium states forming the spiral of the caloric curve.

Refer to caption
Figure 11: Normalized energy as a function of the central potential Φ0\Phi_{0} for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399).

In Fig. 12 we have plotted the relation Λ()\Lambda({\cal R}) between the normalized energy and the energy density contrast =ϵ0/ϵR{\cal R}=\epsilon_{0}/\epsilon_{R}. We can see that {\cal R} increases monotonically along the series of equilibria up to Λc\Lambda^{\prime}_{c}. Then, on the unstable branch, it displays a more complicated behavior.

Refer to caption
Figure 12: Normalized energy as a function of the energy density contrast =ϵ0/ϵR\mathcal{R}=\epsilon_{0}/\epsilon_{R} for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399).

The evolution of the system in the microcanonical ensemble is the following. Let us start from high energy states and decrease the energy. As energy decreases, the system becomes more and more concentrated. The central potential and the density contrast increase. If we keep decreasing the energy there comes a point EcE^{\prime}_{c} at which the system undergoes a gravitational collapse towards a black hole. This is an instability of general relativistic origin which has no counterpart in the Newtonian theory.

V.4.2 Canonical ensemble

We now consider the canonical ensemble. The function Λ(η)\Lambda(\eta) is multivalued. According to the Poincaré turning point criterion, the series of equilibria is stable up to ηc\eta_{c}, becomes unstable between ηc\eta_{c} and η\eta_{*}, is stable again between η\eta_{*} and ηc\eta^{\prime}_{c} and becomes unstable again after ηc\eta^{\prime}_{c}. The caloric curve presents the following features:

(i) When η<η\eta<\eta_{*} there are only gaseous states. When ηc<η<ηc\eta_{c}<\eta<\eta^{\prime}_{c} there are only condensed states. When η<η<ηc\eta_{*}<\eta<\eta_{c} there exist gaseous and condensed states at the same temperature. A first order phase transition is expected at a transition temperature ηt\eta_{t} determined by the Maxwell construction (see Fig. 9) or by the equality of the free energy of the gaseous and condensed phases (see Fig. 13). When η<η<ηt\eta_{*}<\eta<\eta_{t} the gaseous states have a lower free energy than the condensed states. When ηt<η<ηc\eta_{t}<\eta<\eta_{c} the condensed states have a lower free energy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of the metastable states.

(ii) There is a zeroth order phase transition at ηc\eta_{c} from the gaseous phase to the condensed phase. It corresponds to a gravitational collapse (isothermal collapse) ultimately halted by quantum degeneracy.

(iii) There is a zeroth order phase transition at η\eta_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

(iv) There is a catastrophic collapse at ηc\eta^{\prime}_{c} from the condensed phase to a black hole.

Refer to caption
Figure 13: Normalized free energy as a function of the normalized inverse temperature for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399). The first derivative of βF\beta F with respect to β\beta is discontinuous at the transition temperature βt\beta_{t}. This corresponds to a first order phase transition, connecting the gaseous phase to the condensed phase, which is associated with a jump of energy E=(βF)/βE=\partial(\beta F)/\partial\beta in the caloric curve. On the other hand, βF\beta F is discontinuous at the spinodal points ηc\eta_{c} and η\eta_{*}. This corresponds to zeroth order phase transitions which are associated with a jump of free energy. We can check that the unstable equilibrium states (saddle points of free energy) between ηc\eta_{c} and η\eta_{*} have a free energy higher than the stable equilibrium states (minima of free energy). However, the unstable equilibrium states after ηc\eta^{\prime}_{c} can have a free energy lower than the stable equilibrium states before ηc\eta_{c}.

In Fig. 14 we have plotted the relation η(Φ0)\eta(\Phi_{0}) between the inverse temperature and the central potential. We see that Φ0\Phi_{0} increases monotonically along the series of equilibria. The curve η(Φ0)\eta(\Phi_{0}) presents a first peak at ηc\eta_{c} and a second peak at ηc\eta^{\prime}_{c}. Then, it displays damped oscillations. They correspond to unstable equilibrium states associated with the spiral of the caloric curve.

Refer to caption
Figure 14: Normalized inverse temperature as a function of the central potential Φ0\Phi_{0} for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399).

In Fig. 15 we have plotted the relation η()\eta({\cal R}) between the normalized inverse temperature and the energy density contrast =ϵ0/ϵR{\cal R}=\epsilon_{0}/\epsilon_{R}. We can see that {\cal R} increases monotonically along the series of equilibria up to ηc\eta^{\prime}_{c}. Then, on the second unstable branch, it displays a more complicated behavior.

Refer to caption
Figure 15: Normalized inverse temperature as a function of the energy density contrast =ϵ0/ϵR\mathcal{R}=\epsilon_{0}/\epsilon_{R} for NOV<N<NeN_{\rm OV}<N<N_{e} (specifically R=50R=50 and N=0.399N=0.399).

The evolution of the system in the canonical ensemble in the following. Let us start from high temperature states and decrease the temperature. At high temperatures, the system is in the gaseous phase. At η=ηt\eta=\eta_{t}, we expect the system to undergo a first order phase transition from the gaseous phase to the condensed phase. However, in practice, this phase transition does not take place because the metastable gaseous states have a very long lifetime. The physical transition occurs at the critical temperature ηc\eta_{c} (spinodal point) at which the gaseous phase disappears. At that point the system undergoes a zeroth order phase transition (collapse) from the gaseous phase to the condensed phase. If we keep decreasing the temperature there comes another critical point ηc\eta^{\prime}_{c} at which the system undergoes a catastrophic collapse from the condensed phase to a black hole. This is an instability of general relativistic origin which has no counterpart in the Newtonian theory. Inversely, if we increase the temperature, the system displays a zeroth order phase transition (explosion) at η\eta_{*} from the condensed phase to the gaseous phase.

V.5 The case Ne<N<NeN_{e}<N<N^{\prime}_{e}

In Fig. 16 we have plotted the caloric curve for Ne=0.40002<N<Ne=0.40469N_{e}=0.40002<N<N^{\prime}_{e}=0.40469. The novelty with respect to the previous case is that now ηc\eta^{\prime}_{c} is smaller than ηc\eta_{c} (they become equal when N=Ne=0.40002N=N_{e}=0.40002).

Refer to caption
Figure 16: Caloric curve for Ne=0.40002<N<Ne=0.40469N_{e}=0.40002<N<N^{\prime}_{e}=0.40469 (specifically R=50R=50 and N=0.401N=0.401).
Refer to caption
Figure 17: Normalized free energy as a function of the inverse temperature for Ne<N<NeN_{e}<N<N^{\prime}_{e} (specifically R=50R=50 and N=0.401N=0.401).

The description in the microcanonical ensemble is the same as before.

In the canonical ensemble, the caloric curve presents the following features:

(i) When η<η\eta<\eta_{*} and when ηc<η<ηc\eta^{\prime}_{c}<\eta<\eta_{c} there are only gaseous states. When η<η<ηc\eta_{*}<\eta<\eta^{\prime}_{c} there exist gaseous and condensed states at the same temperature. A first order phase transition is expected at a transition temperature ηt\eta_{t} determined by the Maxwell construction (see Fig. 16) or by the equality of the free energy of the two phases (see Fig. 17). When η<η<ηt\eta_{*}<\eta<\eta_{t} the gaseous states have a lower free energy than the condensed states. When ηt<η<ηc\eta_{t}<\eta<\eta^{\prime}_{c} the condensed states have a lower free energy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of the metastable states.

(ii) There is a catastrophic collapse at ηc\eta_{c} from the gaseous phase to a black hole.

(iii) There is a catastrophic collapse at ηc\eta^{\prime}_{c} from the condensed phase to a black hole.

(iv) There is a zeroth order phase transition at η\eta_{*} from the condensed phase to the gaseous phase. It correspond to an explosion ultimately halted by the boundary of the box.

The evolution of the system in the canonical ensemble is the following. Let us start from high temperature states and decrease the temperature. At high temperatures, the system is in the gaseous phase. At η=ηt\eta=\eta_{t} the system is expected to undergo a first order phase transition from the gaseous phase to the condensed phase. However, this phase transition does not take place in practice. At η=ηc\eta=\eta_{c} the system undergoes a catastrophic collapse towards a black hole. A condensed phase exists for η<η<ηc\eta_{*}<\eta<\eta^{\prime}_{c} but it is not clear how it can be reached in practice.

V.6 The case Ne<N<NN^{\prime}_{e}<N<N_{*}

In Fig. 18 we have plotted the caloric curve for Ne=0.40469<N<N=0.41637N^{\prime}_{e}=0.40469<N<N_{*}=0.41637, where NeN^{\prime}_{e} is defined such that ηc=ηt\eta^{\prime}_{c}=\eta_{t}.

Refer to caption
Figure 18: Caloric curve for Ne=0.40469<N<N=0.41637N^{\prime}_{e}=0.40469<N<N_{*}=0.41637 (specifically R=50R=50 and N=0.41N=0.41).
Refer to caption
Figure 19: Normalized free energy as a function of the inverse temperature for Ne<N<NN^{\prime}_{e}<N<N_{*} (specifically R=50R=50 and N=0.41N=0.41).

The description in the microcanonical ensemble is the same as before.

In the canonical ensemble, the caloric curve presents the following features:

(i) When η<η\eta<\eta_{*} and when ηc<η<ηc\eta^{\prime}_{c}<\eta<\eta_{c} there are only gaseous states. When η<η<ηc\eta_{*}<\eta<\eta^{\prime}_{c} there exist gaseous and condensed states at the same temperature. However, there is no first order phase transition, even in theory, because we cannot satisfy the Maxwell construction (see Fig. 18) or the equality of the free energy of the gaseous and condensed phases (see Fig. 19). When η<η<ηc\eta_{*}<\eta<\eta^{\prime}_{c} the gaseous states always have a lower free energy than the condensed states (see Fig. 19). Therefore, although there are several stable equilibrium states when η<η<ηc\eta_{*}<\eta<\eta^{\prime}_{c} there is no phase transition from one phase to the other. This is a particularity of the relativistic situation.

(ii) There is a catastrophic collapse at ηc\eta_{c} from the gaseous phase to a black hole.

(iii) There is a catastrophic collapse at ηc\eta^{\prime}_{c} from the condensed phase to a black hole.

(iv) There is a zeroth order phase transition at η\eta_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

The evolution of the system is the same as described previously.

V.7 The case N>NN>N_{*}

In Fig. 20 we have plotted the caloric curve for N>N=0.41637N>N_{*}=0.41637, where NN_{*} is defined such that ηc=η\eta^{\prime}_{c}=\eta_{*}. From that moment, we denote the minimum energy by Λc\Lambda_{c} instead of Λc\Lambda^{\prime}_{c}.

Refer to caption
Figure 20: Caloric curve for N=0.41637<N<NS=6.40N_{*}=0.41637<N<N^{\prime}_{S}=6.40 (specifically R=50R=50 and N=0.45N=0.45).

V.7.1 Microcanonical ensemble

Let us first consider the microcanonical ensemble. The curve η(Λ)\eta(\Lambda) is multivalued. According to the Poincaré turning point criterion, the series of equilibria is stable up to Λc\Lambda_{c} and then becomes unstable. The caloric curve presents the following features:

(i) There is no phase transition (there is only one stable equilibrium state for each Λ<Λc\Lambda<\Lambda_{c}).

(ii) There is a region of negative specific heats between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c}.

(iii) There is a catastrophic collapse at Λc\Lambda_{c} towards a black hole.

The evolution of the system is the same as described previously.

V.7.2 Canonical ensemble

We now consider the canonical ensemble. The function Λ(η)\Lambda(\eta) is multivalued. According to the Poincaré turning point criterion, the series of equilibria is stable up to ηc\eta_{c} and then becomes unstable. The caloric curve presents the following features:

(i) There is no phase transition (there is only one stable equilibrium state for each η<ηc\eta<\eta_{c}).

(ii) There is a catastrophic collapse at ηc\eta_{c} towards a black hole.

The evolution of the system is the same as described previously. The only difference is that the condensed phase has disappeared.

V.8 Larger values of NN

In Figs. 21 and 22 we have plotted the caloric curves for larger values of NN. When NNOV=0.39853N\gg N_{\rm OV}=0.39853, the system is nondegenerate and we recover the results of roupas ; paper2 for a classical general relativistic gas described by the Boltzmann distribution.232323As discussed in Sec. XI the classical limit corresponds to N+N\rightarrow+\infty and R+R\rightarrow+\infty in such a way that N/RN/R is fixed (in more physical terms NNOVN\gg N_{\rm OV} and RROVR\gg R_{\rm OV} with N/RN/R fixed). The caloric curve exhibits a double spiral. When N<NS=6.40N<N^{\prime}_{S}=6.40 (see Fig. 7 of paper2 ) the two spirals are separated. When NS=6.40<N<NS=7.08N^{\prime}_{S}=6.40<N<N_{S}=7.08 (see Fig. 8 of paper2 ) the two spirals are amputed (truncated) and touch each other. When NS=7.08<N<Nmax=8.821N_{S}=7.08<N<N_{\rm max}=8.821 (see Fig. 9 of paper2 ) the spirals disappear and the caloric curve makes a “loop”. When NNmaxN\rightarrow N_{\rm max}, the caloric curve reduces to a “point” located at (Λ,η)=(0.9829,1.2203)(\Lambda_{*},\eta_{*})=(-0.9829,1.2203).

Refer to caption
Figure 21: Caloric curve for N<N<NSN_{*}<N<N^{\prime}_{S} (specifically R=50R=50 and N=1.5N=1.5).
Refer to caption
Figure 22: Caloric curve for N<N<NSN_{*}<N<N^{\prime}_{S} (specifically R=50R=50 and N=4N=4).

V.9 The canonical phase diagram

In Figs. 23 and 24 we have represented the canonical phase diagram corresponding to RCCP=12.0<R<RMCP=92.0R_{\rm CCP}=12.0<R<R_{\rm MCP}=92.0 (specifically R=50R=50). It shows the evolution of the critical temperatures ηmin\eta_{\rm min}, ηc\eta_{c}, η\eta_{*}, ηt\eta_{t}, ηc\eta^{\prime}_{c} with NN. We can clearly see the canonical critical point at NCCP=0.0170N_{\rm CCP}=0.0170 at which the canonical phase transition appears. We also see the point NOV=0.39853N_{\rm OV}=0.39853 above which quantum mechanics is not able to prevent gravitational collapse above ηc(N)\eta^{\prime}_{c}(N). Finally, we see the point Nmax=8.821N_{\rm max}=8.821 above which there is no equilibrium state anymore.

The nonrelativistic limit ijmpb corresponds to the dashed lines. It provides a very good approximation of ηc\eta_{c}, η\eta_{*} and ηt\eta_{t} for NNOVN\ll N_{\rm OV}. As we approach NOVN_{\rm OV} general relativity must be taken into account.

The classical limit roupas ; paper2 corresponds to the dotted lines. It provides a very good approximation of ηmin\eta_{\rm min} (hot spiral) for any NN. It also provides a very good approximation of ηc\eta_{c} (cold spiral) for NNOVN\gg N_{\rm OV}. As we approach NOVN_{\rm OV} quantum mechanics must be taken into account.

Refer to caption
Figure 23: Canonical phase diagram for RCCP=12.0<R<RMCP=92.0R_{\rm CCP}=12.0<R<R_{\rm MCP}=92.0 (specifically R=50R=50).
Refer to caption
Figure 24: Zoom of the canonical phase diagram for RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP} (specifically R=50R=50) in the region near NOVN_{\rm OV}. For NNOV+N\rightarrow N_{\rm OV}^{+}, we find that ηc0.104(NNOV)1/2\eta^{\prime}_{c}\sim 0.104\,(N-N_{\rm OV})^{-1/2}.

V.10 The microcanonical phase diagram

In Figs. 25 and 26 we have represented the microcanonical phase diagram corresponding to RCCP=12.0<R<RMCP=92.0R_{\rm CCP}=12.0<R<R_{\rm MCP}=92.0 (specifically R=50R=50). It shows the evolution of the critical energies Λmin\Lambda_{\rm min}, Λmax\Lambda_{\rm max}, Λmax\Lambda^{\prime}_{\rm max}, Λgas\Lambda_{\rm gas}, Λcond\Lambda_{\rm cond}, Λc\Lambda^{\prime}_{c}, Λc\Lambda_{c} with NN. We can clearly see the canonical critical point at NCCP=0.0170N_{\rm CCP}=0.0170 at which the region of negative specific heat (associated with the canonical phase transition) appears. We also see the point NOV=0.39853N_{\rm OV}=0.39853 above which quantum mechanics is not able to prevent gravitational collapse above Λc(N)\Lambda^{\prime}_{c}(N), and the point Nmax=8.821N_{\rm max}=8.821 above which there is no equilibrium state anymore.

The nonrelativistic limit ijmpb corresponds to the dashed lines. It provides a very good approximation of Λmax\Lambda_{\rm max}, Λgas\Lambda_{\rm gas} and Λcond\Lambda_{\rm cond} for NNOVN\ll N_{\rm OV}. As we approach NOVN_{\rm OV} general relativity must be taken into account.

The classical limit roupas ; paper2 corresponds to the dotted lines. It provides a very good approximation of Λmin\Lambda_{\rm min} (hot spiral) for any NN. It also provides a very good approximation of Λc\Lambda_{c} (cold spiral) for NNOVN\gg N_{\rm OV}. As we approach NOVN_{\rm OV} quantum mechanics must be taken into account.

Refer to caption
Figure 25: Microcanonical phase diagram for RCCP=12.0<R<RMCP=92.0R_{\rm CCP}=12.0<R<R_{\rm MCP}=92.0 (specifically R=50R=50).
Refer to caption
Figure 26: Zoom of the microcanonical phase diagram for RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP} (specifically R=50R=50) in the region near NOVN_{\rm OV}.

Remark: we recall that the minimum energy above which equilibrium states exist is Λmax\Lambda_{\rm max} (ground state) when N<NOVN<N_{\rm OV} and Λc\Lambda^{\prime}_{c} or Λc\Lambda_{c} when N>NOVN>N_{\rm OV}. From Fig. 25 we note that Λmax(N)\Lambda_{\rm max}(N) increases with NN while Λc(N)\Lambda^{\prime}_{c}(N) and Λc(N)\Lambda_{c}(N) decrease with NN. We also note that the system would be a black hole if M>Rc2/2GM>Rc^{2}/2G, i.e., M>R/2M>R/2 in terms of dimensionless variables. Using Eq. (15), this leads to the condition

Λ<(R2N)RN2ΛBH(N,R).\Lambda<-\frac{\left(\frac{R}{2}-N\right)R}{N^{2}}\equiv\Lambda_{\rm BH}(N,R). (33)

One can locate the black hole energy curve ΛBH(N,R)\Lambda_{\rm BH}(N,R) in Fig. 25. It behaves as ΛBHR2/2N2\Lambda_{\rm BH}\sim-R^{2}/2N^{2}\rightarrow-\infty when N0N\rightarrow 0 and as ΛBHR/N0+\Lambda_{\rm BH}\sim R/N\rightarrow 0^{+} when N+N\rightarrow+\infty. It vanishes at N=R/2N=R/2 and has a maximum (ΛBH)max=1/2(\Lambda_{\rm BH})_{\rm max}=1/2 at N=RN=R. One can show that the black hole energy curve never intersects the curves of Fig. 25 so that the system is never a black hole (see paper2 for a detailed discussion).

VI The case R>RMCPR>R_{\rm MCP}

We now study the case R>RMCP=92.0R>R_{\rm MCP}=92.0 where the system can display a canonical phase transition (as before) and a microcanonical phase transition (see Fig. 47 below). For illustration we take R=600R=600. In that case, the canonical phase transition appears above NCCP=9.84×106N_{\rm CCP}=9.84\times 10^{-6} and the microcanonical phase transition appears above NMCP=1.02×102N_{\rm MCP}=1.02\times 10^{-2}.

The description of the caloric curves in the canonical ensemble is the same as before. Therefore, in the following, we only consider the microcanonical ensemble. In addition, we focus on what is new and do not treat in detail the situations that are similar to those described previously.

VI.1 The case N<NMCPN<N_{\rm MCP}

When N<NMCP=1.02×102N<N_{\rm MCP}=1.02\times 10^{-2}, the discussion is the same as in Sec. V (R=50R=50). The canonical phase transition appears at NCCP=9.84×106N_{\rm CCP}=9.84\times 10^{-6}. Since NMCPNOVN_{\rm MCP}\ll N_{\rm OV}, we are in the nonrelativistic regime ijmpb .

VI.2 The case NMCP<N<N1N_{\rm MCP}<N<N_{\rm 1}

In Fig. 27 we have plotted the caloric curve for NMCP=1.02×102<N<NOV=0.39853N_{\rm MCP}=1.02\times 10^{-2}<N<N_{\rm OV}=0.39853. Since NNOVN\ll N_{\rm OV}, the caloric curve coincides with the one obtained in the nonrelativistic limit ijmpb . It has a ZZ-shape structure leading to a microcanonical phase transition.242424The caloric curve resembles a dinosaur’s neck ijmpb . However, in Fig. 27 the dinosaur has no “chin”. The “chin” appears at Nchin=0.5062N_{\rm chin}=0.5062 as explained in Appendix C.2. The presence, or not, of the “chin” has no physical consequence since it concerns a region of the caloric curve where the equilibrium states are unstable. This ZZ-shape structure appears at N=NMCP=1.02×102N=N_{\rm MCP}=1.02\times 10^{-2} at which the caloric curve presents a vertical inflexion point. The caloric curve continues up to Λmax\Lambda_{\rm max} (outside the frame of the figure) at which it presents an asymptote.

Refer to caption
Figure 27: Caloric curve for NMCP=1.02×102<N<NOV=0.39853N_{\rm MCP}=1.02\times 10^{-2}<N<N_{\rm OV}=0.39853 (specifically R=600R=600 and N=0.29N=0.29).

The curve η(Λ)\eta(\Lambda) is multivalued leading to the possibility of phase transitions in the microcanonical ensemble. The upper branch up to Λc\Lambda_{c} corresponds to the gaseous phase and the lower branch after Λ\Lambda_{*} corresponds to the condensed phase. According to the Poincaré turning point criterion, these equilibrium states are stable while the equilibrium states on the intermediate branch between Λc\Lambda_{c} and Λ\Lambda_{*} are unstable. The caloric curve presents the following features:

(i) When Λ<Λ\Lambda<\Lambda_{*} there are only gaseous states. When Λ>Λc\Lambda>\Lambda_{c} there are only condensed states. When Λ<Λ<Λc\Lambda_{*}<\Lambda<\Lambda_{c} there exist gaseous and condensed states with the same energy. A first order microcanonical phase transition is expected at a transition energy Λt\Lambda_{t} determined by the Maxwell construction (see Fig. 27) or by the equality of the entropy of the gaseous and condensed phases (see Fig. 18 of ijmpb ). When Λ<Λ<Λt\Lambda_{*}<\Lambda<\Lambda_{t} the gaseous states have a higher entropy than the condensed states. When Λt<Λ<Λc\Lambda_{t}<\Lambda<\Lambda_{c} the condensed states have a higher entropy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of the metastable states.

(ii) There is a zeroth order phase transition at Λc\Lambda_{c} from the gaseous phase to the condensed phase. It corresponds to a gravitational collapse (gravothermal catastrophe) ultimately halted by quantum degeneracy.

(iii) There is a zeroth order phase transition at Λ\Lambda_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

(iv) There are two regions of negative specific heats, one between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c} and another one between Λ\Lambda_{*} and Λcond\Lambda_{\rm cond}.

The evolution of the system in the microcanonical ensemble is the following. Let us start from high energies and decrease the energy. At high energies, the system is in the gaseous phase. At Λ=Λt\Lambda=\Lambda_{t} we expect the system to undergo a first order phase transition from the gaseous phase to the condensed phase. However, in practice, this phase transition does not take place because the metastable gaseous states have a very long lifetime. At Λ=Λc\Lambda=\Lambda_{c} the system collapses towards the condensed phase. Complete gravitational collapse is prevented by quantum mechanics. The system reaches an equilibrium state similar to a nonrelativistic white dwarf (fermion ball) surrounded by an isothermal atmosphere. If we now increase the energy the system remains in the condensed phase (again, the first order phase transition expected at Λt\Lambda_{t} does not take place because the metastable condensed states have a very long lifetime) until the point Λ\Lambda_{*} at which it explodes and returns to the gaseous phase. We have thus described an hysteretic cycle in the microcanonical ensemble ijmpb .

In Fig. 28 we have plotted the density profiles of the gaseous (G), core-halo (CH) and condensed (C) states at the transition point Λt\Lambda_{t}. We note that the energy density is very low confirming that we are in the nonrelativistic regime. The discussion is essentially the same as in Sec. V.2.3 with the difference that the fermion ball (similar to a nonrelativistic cold white dwarf) that forms in the condensed phase contains only a fraction (1/4\sim 1/4) of the mass (see ijmpb , Sec. XIII and Appendix E.2.2). The rest of the mass is diluted in a hot halo. This core-halo structure is reminiscent of a red-giant (see Sec. XIII).

Refer to caption
Figure 28: Energy density profiles of gaseous, core-halo and condensed states at Λ=Λt\Lambda=\Lambda_{t} (specifically Λt=0.151\Lambda_{t}=0.151).

VI.3 The case N1<N<NOVN_{1}<N<N_{\rm OV}

The second branch with an asymptote at Λmax\Lambda^{\prime}_{\rm max} appears at N1=0.18131N_{1}=0.18131 but this does not change the discussion since this branch is made of unstable equilibrium states. From that moment, the system starts to be strongly relativistic.

VI.4 The case NOV<N<NfN_{\rm OV}<N<N_{f}

In Fig. 29 we have plotted the caloric curve for NOV=0.39853<N<Nf=1.4854N_{\rm OV}=0.39853<N<N_{f}=1.4854.252525We note that the “chin” of the dinosaur has appeared since N=1.3>Nchin=0.5062N=1.3>N_{\rm chin}=0.5062. The novelty with respect to the previous case is that the two branches have merged. As a result there is no ground state anymore (see Sec. V.4). The caloric curve presents a turning point of energy which corresponds to the minimum energy. When N<N=0.405N<N_{*}=0.405 we call it Λc\Lambda^{\prime}_{c} and when N>NN>N_{*} we call it Λc′′\Lambda^{\prime\prime}_{c} (see Sec. V.7 for the definition of NN_{*}). In the following, to be specific, we assume that N>NN>N_{*} but the discussion is essentially the same for N<NN<N_{*}.

Refer to caption
Figure 29: Caloric curve for NOV=0.39853<N<Nf=1.4854N_{\rm OV}=0.39853<N<N_{f}=1.4854 (specifically R=600R=600 and N=1.3N=1.3).

According to the Poincaré turning point criterion, the series of equilibria is stable up to Λc\Lambda_{c}, becomes unstable between Λc\Lambda_{c} and Λ\Lambda_{*}, becomes stable again between Λ\Lambda_{*} and Λc′′\Lambda^{\prime\prime}_{c} and becomes unstable again after Λc′′\Lambda^{\prime\prime}_{c}. The caloric curve presents the following features:

Refer to caption
Figure 30: Entropy per fermion as a function of the normalized energy for NOV<N<NfN_{\rm OV}<N<N_{f} (specifically R=600R=600 and N=1.3N=1.3). The first derivative of SS with respect to EE is discontinuous at the transition energy EtE_{t}. This corresponds to a first order phase transition, connecting the gaseous phase to the condensed phase, which is associated with a jump of temperature β=S/E\beta=\partial S/\partial E in the caloric curve. On the other hand, SS is discontinuous at the spinodal points EcE_{c} and EE_{*}. This corresponds to zeroth order phase transitions which are associated with a jump of entropy. We can check that the unstable equilibrium states (saddle points of entropy) between Λc\Lambda_{c} and Λ\Lambda_{*} have an entropy lower than the stable equilibrium states (maxima of entropy). However, the unstable equilibrium states after Λc′′\Lambda^{\prime\prime}_{c} can have an entropy higher than the stable equilibrium states before Λc\Lambda_{c}.

(i) When Λ<Λ\Lambda<\Lambda_{*} there are only gaseous states. When Λc<Λ<Λc′′\Lambda_{c}<\Lambda<\Lambda^{\prime\prime}_{c} there are only condensed states. When Λ<Λ<Λc\Lambda_{*}<\Lambda<\Lambda_{c} there exist gaseous and condensed states with the same energy. A first order phase transition is expected at a transition energy Λt\Lambda_{t} determined by the Maxwell construction (see Fig. 29) of by the equality of the entropy of the gaseous and condensed phases (see Fig. 30). When Λ<Λ<Λt\Lambda_{*}<\Lambda<\Lambda_{t} the gaseous states have a higher entropy than the condensed states. When Λt<Λ<Λc\Lambda_{t}<\Lambda<\Lambda_{c} the condensed states have a higher entropy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of metastable states.

(ii) There is a zeroth order phase transition at Λc\Lambda_{c} from the gaseous phase to the condensed phase. It corresponds to a gravitational collapse (gravothermal catastrophe) ultimately halted by quantum degeneracy.

(iii) There is a zeroth order phase transition at Λ\Lambda_{*} from the condensed phase to the gaseous phase. It corresponds to a explosion ultimately halted by the boundary of the box.

(iv) There is a catastrophic collapse at Λc′′\Lambda^{\prime\prime}_{c} from the condensed phase to a black hole.

(v) There are two regions of negative specific heats, one between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c} and another one between Λ\Lambda_{*} and Λc′′\Lambda^{\prime\prime}_{c}.

The evolution of the system in the microcanonical ensemble in the following. Let us start from high energies and decrease the energy. At high energies, the system is in the gaseous phase. At Λ=Λt\Lambda=\Lambda_{t}, we expect the system to undergo a first order phase transition from the gaseous phase to the condensed phase. However, in practice, this phase transition does not take place because the metastable gaseous states have a very long lifetime. The physical transition occurs at the critical energy Λc\Lambda_{c} (spinodal point) at which the gaseous phase disappears. At that point the system undergoes a zeroth order phase transition (collapse) from the gaseous phase to the condensed phase. If we keep decreasing the energy there comes another critical point Λc′′\Lambda^{\prime\prime}_{c} at which the system undergoes a catastrophic collapse from the condensed phase to a black hole. This is an instability of general relativistic origin which has no counterpart in the Newtonian theory. Inversely, if we increase the energy, the system displays a zeroth order phase transition (explosion) at Λ\Lambda_{*} from the condensed phase to the gaseous phase.

Refer to caption
Figure 31: Density profiles of gaseous, core-halo and (stable and unstable) condensed states at Λ=Λt\Lambda=\Lambda_{t} (specifically Λt=0.0510\Lambda_{t}=-0.0510).

In Fig. 31 we have plotted the different density profiles at Λt\Lambda_{t}. We see that the energy density is low for the gaseous state (G) and for the core-halo state (CH) indicating that we are in the nonrelativistic regime. By contrast, the energy density is relatively high for the stable condensed state (C) and for the unstable condensed state (U) indicating that we are in the relativistic regime. The discussion is essentially the same as in Sec. V.3 with the difference that the fermion ball (similar to a general relativistic cold neutron star) that forms in the condensed phase contains only a fraction (1/4\sim 1/4) of the mass (see Sec. XIII and Appendix E.2.2). The rest of the mass is diluted in a hot halo. This core-halo structure is reminiscent of a supernova (see Sec. XIII).

VI.5 The case Nf<N<NfN_{f}<N<N^{\prime}_{f}

In Fig. 32 we have plotted the caloric curve for Nf=1.4854<N<Nf=1.619N_{f}=1.4854<N<N^{\prime}_{f}=1.619. The novelty with respect to the previous case is that now Λc′′\Lambda^{\prime\prime}_{c} is smaller than Λc\Lambda_{c} (they become equal when N=NfN=N_{f}).

Refer to caption
Figure 32: Caloric curve for Nf=1.4854<N<Nf=1.619N_{f}=1.4854<N<N^{\prime}_{f}=1.619 (specifically R=600R=600 and N=1.5N=1.5).
Refer to caption
Figure 33: Entropy per fermion as a function of the normalized energy for Nf=1.4854<N<Nf=1.619N_{f}=1.4854<N<N^{\prime}_{f}=1.619 (specifically R=600R=600 and N=1.5N=1.5).

The caloric curve presents the following features:

(i) When Λ<Λ\Lambda<\Lambda_{*} and Λc′′<Λ<Λc\Lambda^{\prime\prime}_{c}<\Lambda<\Lambda_{c} there are only gaseous states. When Λ<Λ<Λc′′\Lambda_{*}<\Lambda<\Lambda^{\prime\prime}_{c} there exist gaseous and condensed states with the same energy. A first order phase transition is expected at a transition energy Λt\Lambda_{t} determined by the Maxwell construction (see Fig. 32) or by the equality of the entropy of the two phases (see Fig. 33). When Λ<Λ<Λt\Lambda_{*}<\Lambda<\Lambda_{t} the gaseous states have a higher entropy than the condensed states. When Λt<Λ<Λc′′\Lambda_{t}<\Lambda<\Lambda^{\prime\prime}_{c} the condensed states have a higher entropy than the gaseous states. However, the first order phase transition does not take place in practice because of the very long lifetime of the metastable states.

(ii) There is a catastrophic collapse at Λc\Lambda_{c} from the gaseous phase to a black hole.

(iii) There is a catastrophic collapse at Λc′′\Lambda^{\prime\prime}_{c} from the condensed phase to a black hole.

(iv) There is a zeroth order phase transition at Λ\Lambda_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

(v) There are two regions of negative specific heats, one between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c} and another one between Λ\Lambda_{*} and Λc′′\Lambda^{\prime\prime}_{c}.

The evolution of the system in the microcanonical ensemble is the following. Let us start from high energies and decrease the energy. At high energies, the system is in the gaseous phase. At Λ=Λt\Lambda=\Lambda_{t} the system is expected to undergo a first order phase transition from the gaseous phase to the condensed phase. However, this phase transition does not take place in practice. At Λ=Λc\Lambda=\Lambda_{c} the system undergoes a catastrophic collapse towards a black hole. A condensed phase exists for Λ<Λ<Λc′′\Lambda_{*}<\Lambda<\Lambda^{\prime\prime}_{c} but it is not clear how it can be reached in practice.

VI.6 The case Nf<N<NN^{\prime}_{f}<N<N^{\prime}_{*}

In Fig. 34 we have plotted the caloric curve for Nf=1.619<N<N=1.9000N^{\prime}_{f}=1.619<N<N^{\prime}_{*}=1.9000, where NfN^{\prime}_{f} is defined such that Λc′′=Λt\Lambda^{\prime\prime}_{c}=\Lambda_{t}.

Refer to caption
Figure 34: Caloric curve for Nf=1.619<N<N=1.9000N^{\prime}_{f}=1.619<N<N^{\prime}_{*}=1.9000 (specifically R=600R=600 and N=1.65N=1.65).
Refer to caption
Figure 35: Entropy per fermion as a function of the normalized energy for Nf=1.619<N<N=1.9000N^{\prime}_{f}=1.619<N<N^{\prime}_{*}=1.9000 (specifically R=600R=600 and N=1.65N=1.65).

The caloric curve presents the following features:

(i) When Λ<Λ\Lambda<\Lambda_{*} and when Λc′′<Λ<Λc\Lambda^{\prime\prime}_{c}<\Lambda<\Lambda_{c} there are only gaseous states. When Λ<Λ<Λc′′\Lambda_{*}<\Lambda<\Lambda^{\prime\prime}_{c} there exist gaseous and condensed states with the same energy. However, there is no first order phase transition, even in theory, because we cannot satisfy the Maxwell construction (see Fig. 34) or the equality of the entropy of the gaseous and condensed phases (see Fig. 35). When Λ<Λ<Λc′′\Lambda_{*}<\Lambda<\Lambda^{\prime\prime}_{c} the gaseous states always have an entropy higher than the condensed states. Therefore, although there are several stable equilibrium states when Λ<Λ<Λc′′\Lambda_{*}<\Lambda<\Lambda^{\prime\prime}_{c} there is no phase transition from one phase to the other. This is a particularity of the relativistic situation.

(ii) There is a catastrophic collapse at Λc\Lambda_{c} from the gaseous phase to a black hole.

(iii) There is a catastrophic collapse at Λc′′\Lambda^{\prime\prime}_{c} from the condensed phase to a black hole.

(iv) There is a zeroth order phase transition at Λ\Lambda_{*} from the condensed phase to the gaseous phase. It corresponds to an explosion ultimately halted by the boundary of the box.

(v) There are two regions of negative specific heats, one between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c} and another one between Λ\Lambda_{*} and Λc′′\Lambda^{\prime\prime}_{c}.

The evolution of the system is the same as described previously.

VI.7 The case N<N<NmaxN^{\prime}_{*}<N<N_{\rm max}

In Fig. 36 we have plotted the caloric curve for N=1.9000<N<Nmax=105.9N^{\prime}_{*}=1.9000<N<N_{\rm max}=105.9, where NN^{\prime}_{*} is defined such that Λc′′=Λ\Lambda^{\prime\prime}_{c}=\Lambda_{*}. From that moment, we denote the minimum energy by Λc\Lambda_{c} instead of Λc′′\Lambda^{\prime\prime}_{c}. The novelty with respect to the previous case is that there is no condensed phase anymore. The discussion is the same as in Secs. V.7 and V.8.

Refer to caption
Figure 36: Caloric curve for N=1.9000<N<Nmax=105.9N^{\prime}_{*}=1.9000<N<N_{\rm max}=105.9 (specifically R=600R=600 and N=5N=5).

VI.8 The microcanonical phase diagram

In Figs. 37 and 38 we have represented the microcanonical phase diagram corresponding to R>RMCPR>R_{\rm MCP}. It shows the evolution of the critical energies Λmin\Lambda_{\rm min}, Λmax\Lambda_{\rm max}, Λgas\Lambda_{\rm gas}, Λcond\Lambda_{\rm cond}, Λc\Lambda_{c}, Λ\Lambda_{*}, Λt\Lambda_{t}, Λmax\Lambda^{\prime}_{\rm max}, Λc\Lambda^{\prime}_{c} and Λc′′\Lambda^{\prime\prime}_{c} with NN. We can clearly see the canonical critical point at NCCP=9.84×106N_{\rm CCP}=9.84\times 10^{-6} at which the region of negative specific heat (associated with the canonical phase transition) appears and the microcanonical critical point at NMCP=1.02×102N_{\rm MCP}=1.02\times 10^{-2} at which the microcanonical phase transition appears. We also see the point NOV=0.39853N_{\rm OV}=0.39853 above which quantum mechanics is not able to prevent gravitational collapse above Λc(N)\Lambda^{\prime}_{c}(N) or Λc′′(N)\Lambda^{\prime\prime}_{c}(N). Finally, we see the point Nmax=105.9N_{\rm max}=105.9 above which there is no equilibrium state anymore.

The nonrelativistic limit ijmpb corresponds to the dashed lines. It provides a very good approximation of Λmax\Lambda_{\rm max}, Λgas\Lambda_{\rm gas}, Λcond\Lambda_{\rm cond}, Λc\Lambda_{c}, Λ\Lambda_{*}, and Λt\Lambda_{t} for NNOVN\ll N_{\rm OV}. As we approach NOVN_{\rm OV} general relativity must be taken into account.

The classical limit paper2 corresponds to the dotted lines. It provides a very good approximation of Λmin\Lambda_{\rm min} (hot spiral) for any NN. It also provides a very good approximation of Λc\Lambda_{c} (cold spiral) for NNOVN\gg N_{\rm OV}. As we approach NOVN_{\rm OV} quantum mechanics must be taken into account.

Refer to caption
Figure 37: Microcanonical phase diagram for R>RMCP=92.0R>R_{\rm MCP}=92.0 (specifically R=600R=600).
Refer to caption
Figure 38: Zoom of the microcanonical phase diagram for R>RMCPR>R_{\rm MCP} (specifically R=600R=600) in the region near NOVN_{\rm OV}.

Remark: From Fig. 37 we note that Λmax(N)\Lambda_{\rm max}(N) increases with NN while Λc(N)\Lambda_{c}(N), Λc(N)\Lambda^{\prime}_{c}(N) and Λc′′(N)\Lambda^{\prime\prime}_{c}(N) decrease with NN.

VII The case RRMCPR\gg R_{\rm MCP}

For very large radii (RRMCPR\gg R_{\rm MCP}), a spiral, winding then unwinding, appears in the caloric curve at the location of the “head” of the dinosaur (this is similar to Fig. 22 of ijmpb and Fig. 44 of clm2 ). However, this spiral is made of unstable states. Therefore, if we restrict ourselves to stable equilibrium states, this mathematical complication (the proliferation of unstable states associated with the spiral) does not change the previous discussion.

Remark: The equilibrium states that are deep in the spiral have a pronounced “core-halo” structure with a large central density (see Fig. 45 of clm2 ). These core-halo states correspond to the configurations computed by Bilic et al. btv and, more recently, by Ruffini et al. rar and Chavanis et al. clm2 . They consist in a large nondegenerate isothermal atmosphere harboring a small “fermion ball”. These solutions look very attractive at first sight because they could provide a self-consistent model of DM halos in which the fermion ball would mimic the presence of a supermassive black hole at the centers of the galaxies (an idea originally proposed in btv ). However, as argued in clm2 , these extreme core-halo structures are thermodynamically unstable (see Secs. VI-VIII of clm2 for a detailed discussion).262626By contrast, less extreme core-halo configurations, such as the solution (CH) computed in Fig. 6, can be stable in the microcanonical ensemble. They have a negative specific heat. These core-halo states are dynamically (Vlasov) stable meaning that if we artificially prepare the system in such a state, it will remain in this state for a long time. However, since these extreme core-halo states are thermodynamically unstable, they are very unlikely (from a thermodynamical point of view) to appear spontaneously. The fermion ball is like a “critical droplet” in nucleation processes. This may be a problem for the fermion ball scenario to mimic the effect of a black hole, as mentioned in clm2 . Other problems with the fermion ball scenario are pointed out in genzel .

VIII The case ROV<R<RCCPR_{\rm OV}<R<R_{\rm CCP}

We now study the case R<RCCP=12.0R<R_{\rm CCP}=12.0 where there is no phase transition (see Fig. 47 below). In this section, we assume R>ROV=3.3569R>R_{\rm OV}=3.3569 so that NOVN_{\rm OV} and NmaxN_{\rm max} are relatively well separated. For illustration, we take R=10R=10.

VIII.1 The case N<N1N<N_{1}

When N<N1=0.18131N<N_{1}=0.18131 the caloric curve is similar to that shown in Fig. 4. It is monotonic and presents an asymptote at Λmax\Lambda_{\rm max}. The discussion is similar to that given in Sec. V.1.

VIII.2 The case N1<N<NOVN_{1}<N<N_{\rm OV}

In Fig. 39 we have plotted the caloric curve for N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853. The difference with the case treated in Sec. V.3 is that there is no canonical phase transition. The caloric curve is monotonic272727We see a sort of inflexion of the curve which signals the canonical first order phase transition that appears at larger radii R>RCCPR>R_{\rm CCP}. and presents an asymptote at Λmax\Lambda_{\rm max}. There is another branch presenting an asymptote at Λmax\Lambda^{\prime}_{\rm max} but it is made of unstable states. The series of equilibria of the main branch is stable in both ensembles. The specific heat is always positive. There is no phase transition and no gravitational collapse. The ensembles are equivalent.

Refer to caption
Figure 39: Caloric curve for N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853 (specifically R=10R=10 and N=0.36N=0.36).

VIII.3 The case NOV<N<NmaxN_{\rm OV}<N<N_{\rm max}

In Fig. 40 we have plotted the caloric curve for NOV=0.39853<N<Nmax=1.764N_{\rm OV}=0.39853<N<N_{\rm max}=1.764. The difference with the cases treated in Secs. V.4-V.8 is that there is no phase transition. When N>NOVN>N_{\rm OV} the two asymptotes have merged leading to a turning point of temperature at ηc\eta_{c} and a turning point of energy at Λc\Lambda_{c}. According to the Poincaré turning point criterion, the series of equilibria is stable up to ηc\eta_{c} in the canonical ensemble and up to Λc\Lambda_{c} in the microcanonical ensemble.

Refer to caption
Figure 40: Caloric curve for NOV=0.39853<N<Nmax=1.764N_{\rm OV}=0.39853<N<N_{\rm max}=1.764 (specifically R=10R=10 and N=0.4N=0.4).

In the microcanonical ensemble, the caloric curve presents the following features:

(i) There is no phase transition.

(ii) There is a region of negative specific heats between Λgas\Lambda_{\rm gas} and Λc\Lambda_{c}.

(iii) There is a catastrophic collapse at Λc\Lambda_{c} towards a black hole.

In the canonical ensemble, the caloric curve presents the following features:

(i) There is no phase transition.

(ii) There is a catastrophic collapse at ηc\eta_{c} towards a black hole.

VIII.4 The phase diagrams

In Fig. 41 we have represented the canonical phase diagram corresponding to R<RCCPR<R_{\rm CCP}. It shows the evolution of the critical temperatures ηmax\eta_{\rm max} and ηc\eta_{c} with NN. We see the point NOVN_{\rm OV} above which quantum mechanics is not able to prevent gravitational collapse above ηc\eta_{c}. We also see the point NmaxN_{\rm max} above which there is no equilibrium state anymore.

The classical limit paper2 corresponds to the dotted lines. It provides a very good approximation of ηmax\eta_{\rm max} (hot spiral) for any NN. It also provides a very good approximation of ηc\eta_{c} (cold spiral) for NNOVN\gg N_{\rm OV}. As we approach NOVN_{\rm OV} quantum mechanics must be taken into account.

Refer to caption
Figure 41: Canonical phase diagram for ROV<R<RCCPR_{\rm OV}<R<R_{\rm CCP} (specifically R=10R=10). For NNOV+N\rightarrow N_{\rm OV}^{+}, we find that ηc0.516(NNOV)1/2\eta_{c}\sim 0.516\,(N-N_{\rm OV})^{-1/2}.

In Fig. 42 we have represented the microcanonical phase diagram corresponding to R<RCCPR<R_{\rm CCP}. It shows the evolution of the critical energies Λmin\Lambda_{\rm min}, Λmax\Lambda_{\rm max}, Λc\Lambda_{c} and Λmax\Lambda^{\prime}_{\rm max} with NN. We see the point NOVN_{\rm OV} above which quantum mechanics is not able to prevent gravitational collapse above Λc\Lambda_{c}. We also see the point NmaxN_{\rm max} above which there is no equilibrium state anymore.

The nonrelativistic limit ijmpb corresponds to the dashed lines. It provides a very good approximation of Λmax\Lambda_{\rm max} for NNOVN\ll N_{\rm OV} (this is not apparent in the figure but the curves coincide for smaller values of NN). As we approach NOVN_{\rm OV} general relativity must be taken into account.

The classical limit paper2 corresponds to the dotted lines. It provides a very good approximation of Λmin\Lambda_{\rm min} (hot spiral) for any NN. It also provides a very good approximation of Λc\Lambda_{c} (cold spiral) for NNOVN\gg N_{\rm OV}. As we approach NOVN_{\rm OV} quantum mechanics must be taken into account.

Refer to caption
Figure 42: Microcanonical phase diagram for ROV<R<RCCPR_{\rm OV}<R<R_{\rm CCP} (specifically R=10R=10).

IX The case R<ROVR<R_{\rm OV}

We now study the case R<ROV=3.3569R<R_{\rm OV}=3.3569. In that case, NOVbNmaxN_{\rm OV}^{b}\simeq N_{\rm max} and N1bNSN_{1}^{b}\simeq N^{\prime}_{S} (see Fig. 47 below).282828We note that for small values of RR, the values of N1bN_{1}^{b} and NOVbN_{\rm OV}^{b} (as well as Λmaxb\Lambda_{\rm max}^{b}, (Λmax)b(\Lambda^{\prime}_{\rm max})^{b}…) are affected by the presence of the box. This is because the fermion ball at T=0T=0 and E=EminE=E_{\rm min} (ground state) is confined by the walls of the box instead of being self-confined (see paper1 for a detailed study). This is why we have added the letter bb on these quantities. For illustration, we take R=1R=1.

IX.1 The case N<N1bNSN<N_{1}^{b}\simeq N^{\prime}_{S}

In Fig. 43 we have plotted the caloric curve for N<N1b=0.13627NSN<N_{1}^{b}=0.13627\simeq N^{\prime}_{S}. It is similar to that shown in Fig. 4. Since NN is close to NmaxN_{\rm max} (see below), we clearly see the hot spiral that was outside the frame of Fig. 4. According to the Poincaré turning point criterion, the series of equilibria is stable up to the maximum temperature TmaxT_{\rm max} (corresponding to ηmin\eta_{\rm min}) in the the canonical ensemble and up to the maximum energy EmaxE_{\rm max} (corresponding to Λmin\Lambda_{\rm min}) in the microcanonical ensemble. Above TmaxT_{\rm max} and EmaxE_{\rm max} the system collapses into a black hole as discussed in roupas ; paper2 . If we restrict ourselves to small and mid temperatures and energies (as in the preceding sections), there is no phase transition and no gravitational collapse. The specific heat is always positive and the ensembles are equivalent.

Refer to caption
Figure 43: Caloric curve for N<N1b=0.13627NSN<N_{1}^{b}=0.13627\simeq N^{\prime}_{S} (specifically R=1R=1 and N=0.12N=0.12).

IX.2 The case N1bNS<N<NOVbNmaxN_{1}^{\rm b}\simeq N^{\prime}_{S}<N<N_{\rm OV}^{\rm b}\simeq N_{\rm max}

In Fig. 44 we have plotted the caloric curve for N1b=0.13627NS<N<NOVb=0.2015NmaxN_{1}^{b}=0.13627\simeq N^{\prime}_{S}<N<N_{\rm OV}^{b}=0.2015\simeq N_{\rm max}. In that case, we have two asymptotes at Λmaxb\Lambda_{\rm max}^{\rm b} and (Λmaxb)(\Lambda_{\rm max}^{\rm b})^{\prime} and a turning point of energy and temperature at Λmin\Lambda_{\rm min} and ηmin\eta_{\rm min}. According to the Poincaré turning point criterion, the series of equilibria is stable along the main branch up to the maximum temperature TmaxT_{\rm max} (corresponding to ηmin\eta_{\rm min}) in the the canonical ensemble and up to the maximum energy EmaxE_{\rm max} (corresponding to Λmin\Lambda_{\rm min}) in the microcanonical ensemble. As before, if we restrict ourselves to small and mid temperatures and energies, we conclude that there is no phase transition and no gravitational collapse. The specific heat is always positive and the ensembles are equivalent.

Refer to caption
Figure 44: Caloric curve for N1b=0.13627NS<N<NOVb=0.2015NmaxN_{1}^{b}=0.13627\simeq N^{\prime}_{S}<N<N_{\rm OV}^{b}=0.2015\simeq N_{\rm max} (specifically R=1R=1 and N=0.14N=0.14).

IX.3 The phase diagrams

Refer to caption
Figure 45: Canonical phase diagram for R<ROVR<R_{\rm OV} (specifically R=1R=1). For NNmaxN\rightarrow N_{\rm max}^{-}, we find that ηmin0.01(NmaxN)1\eta_{\rm min}\sim 0.01\,(N_{\rm max}-N)^{-1}.
Refer to caption
Figure 46: Microcanonical phase diagram for R<ROVR<R_{\rm OV} (specifically R=1R=1).

In Fig. 45 we have represented the canonical phase diagram corresponding to R<ROVR<R_{\rm OV}. It shows the evolution of the critical temperature ηmin\eta_{\rm min} with NN. We note that ηmin\eta_{\rm min} diverges to ++\infty when NNOVb=0.2015NmaxN\rightarrow N_{\rm OV}^{b}=0.2015\simeq N_{\rm max} meaning that the caloric curve goes up to infinity and disappears. The classical limit paper2 corresponds to the dotted line. It provides a good approximation for NNmaxN\ll N_{\rm max}. It is not valid for NNmaxN\sim N_{\rm max}. This shows that when R<ROVR<R_{\rm OV} quantum effects are still important close to NmaxN_{\rm max}. This is because, in the present case, NmaxNOVbN_{\rm max}\simeq N_{\rm OV}^{b} while in the previous examples NmaxNOVN_{\rm max}\gg N_{\rm OV}.

In Fig. 46 we have represented the microcanonical phase diagram corresponding to R<ROVR<R_{\rm OV}. It shows the evolution of the critical energies Λmin\Lambda_{\rm min}, Λmaxb\Lambda^{b}_{\rm max} and (Λmax)b(\Lambda^{{}^{\prime}}_{\rm max})^{b} with NN. The classical limit paper2 corresponds to the dotted line. It provides a good approximation for NNmaxN\ll N_{\rm max}. It is not valid for NNmaxN\sim N_{\rm max}. When R<ROVR<R_{\rm OV} quantum effects are still important close to NmaxN_{\rm max}.

X The (R,N)(R,N) phase diagram

We can summarize our results by plotting the characteristic particle numbers NX(R)N_{\rm X}(R) encountered in our study as a function of the box radius RR. This leads to the (R,N)(R,N) phase diagram represented in Fig. 47.

X.1 Characteristic particle numbers and characteristic radii

Let us briefly recall the meaning of the characteristic particle numbers appearing on this diagram.292929We present the characteristic particle numbers by order of appearance in the caloric curves as we increase NN for a given value of RR. To fix the ideas, we take a large radius so that all kinds of phase transitions are present. We start from a small value of NN. In that case, the caloric curve is monotonic with an asymptote at Λmax(N,R)\Lambda_{\rm max}(N,R) corresponding to the stable ground state of the self-gravitating Fermi gas at T=0T=0 (η=\eta=\infty). We then increase NN until we meet the different characteristic particle numbers signaling a topological change of the caloric curve.

(i) When R>RCCP=12.0R>R_{\rm CCP}=12.0, NCCP(R)N_{\rm CCP}(R) is the particle number at which the canonical phase transition appears, i.e., the particle number at which ηc\eta_{c} and η\eta_{*} appear in the caloric curve. When RRCCPR\gg R_{\rm CCP}, the function NCCP(R)N_{\rm CCP}(R) is given by the relation NCCP(R)2.12×103R3N_{\rm CCP}(R)\sim 2.12\times 10^{3}R^{-3} obtained in the nonrelativistic study of ijmpb .

(ii) When R>RMCP=92.0R>R_{\rm MCP}=92.0, NMCP(R)N_{\rm MCP}(R) is the particle number at which the microcanonical phase transition appears, i.e., the particle number at which Λc\Lambda_{c} and Λ\Lambda_{*} appear in the caloric curve. When RRMCPR\gg R_{\rm MCP}, the function NMCP(R)N_{\rm MCP}(R) is given by the relation NMCP(R)2.20×106R3N_{\rm MCP}(R)\sim 2.20\times 10^{6}R^{-3} obtained in the nonrelativistic study of ijmpb .

(iii) N1(R)N_{1}(R) is the particle number at which the unstable equilibrium states at T=0T=0 appear, i.e., the particle number at which the second branch with an asymptote at Λmax(N,R)\Lambda^{\prime}_{\rm max}(N,R) appears in the caloric curve. The function N1(R)N_{1}(R) is studied in paper1 . When R>R1=2.0556R>R_{1}=2.0556, the fermion star is self-confined and we have the standard value N1=0.18131N_{1}=0.18131 of the OV theory. When R<R1R<R_{1}, the fermion star is box-confined and we find that N1b(R)N_{1}^{\rm b}(R) decreases as RR decreases. When R0R\rightarrow 0, we find that N1b(R)0.2492R3/2N_{1}^{\rm b}(R)\sim 0.2492\,R^{3/2} paper1 .

(iv) NOV(R)N_{\rm OV}(R) is the particle number above which there is no equilibrium state at T=0T=0 (no ground state) anymore. At N=NOV(R)N=N_{\rm OV}(R) the asymptotes Λmax(N,R)\Lambda_{\rm max}(N,R) and Λmax(N,R)\Lambda^{\prime}_{\rm max}(N,R) merge. When N>NOV(R)N>N_{\rm OV}(R) they are replaced by a turning point ηc\eta^{\prime}_{c} in temperature and by a turning point Λc\Lambda^{\prime}_{c} in energy. The function NOV(R)N_{\rm OV}(R) is studied in paper1 . When R>ROV=3.3569R>R_{\rm OV}=3.3569, the fermion star is self-confined and we have the standard value NOV=0.39853N_{\rm OV}=0.39853 of the OV theory. When R<ROVR<R_{\rm OV}, the fermion star is box-confined, and we find that NOVb(R)N_{\rm OV}^{\rm b}(R) decreases as RR decreases. When R0R\rightarrow 0, we find that NOVb(R)0.3104R3/2N_{\rm OV}^{\rm b}(R)\sim 0.3104\,R^{3/2} paper1 .

(v) When R>RCCP=12.0R>R_{\rm CCP}=12.0, Ne(R)N_{e}(R) is the particle number at which the zeroth order phase transition in the canonical ensemble disappears, i.e., the particle number at which ηc=ηc\eta^{\prime}_{c}=\eta_{c}.

(vi) When R>RCCP=12.0R>R_{\rm CCP}=12.0, Ne(R)N^{\prime}_{e}(R) is the particle number at which the first order phase transition in the canonical ensemble disappears, i.e., the particle number at which ηc=ηt\eta^{\prime}_{c}=\eta_{t}.

(vii) When R>RCCP=12.0R>R_{\rm CCP}=12.0, N(R)N_{*}(R) is the particle number at which the condensed phase disappears in the canonical ensemble, i.e., the particle number at which ηc=η\eta^{\prime}_{c}=\eta_{*}.

Refer to caption
Figure 47: The (R,N)(R,N) phase diagram of the general relativistic Fermi gas. The characteristic particle numbers and radii are defined in the text. The dashed lines correspond to the nonrelativistic self-gravitating Fermi gas ijmpb . The dotted lines correspond to the radii R=1R=1, R=10R=10, R=50R=50 and R=600R=600 considered in the paper. Note that the quantities Ne(R)N_{e}(R) and Ne(R)N^{\prime}_{e}(R) have not been represented because they are extremely close to N(R)N_{*}(R). Similarly, the quantities Nf(R)N_{f}(R) and Nf(R)N^{\prime}_{f}(R) have not been represented because they are extremely close to N(R)N^{\prime}_{*}(R).

(viii) When R>RMCP=92.0R>R_{\rm MCP}=92.0, Nf(R)N_{f}(R) is the particle number at which the zeroth order phase transition in the microcanonical ensemble disappears, i.e., the particle number at which Λc′′=Λc\Lambda^{\prime\prime}_{c}=\Lambda_{c}.

(ix) When R>RMCP=92.0R>R_{\rm MCP}=92.0, Nf(R)N^{\prime}_{f}(R) is the particle number at which the first order microcanonical phase transition disappears, i.e., the particle number at which Λc′′=Λt\Lambda^{\prime\prime}_{c}=\Lambda_{t}.

(x) When R>RMCP=92.0R>R_{\rm MCP}=92.0, N(R)N_{*}^{\prime}(R) is the particle number at which the condensed phase disappears in the microcanonical ensemble, i.e., the particle number at which Λc′′=Λ\Lambda^{\prime\prime}_{c}=\Lambda_{*}.

(xi) When R>R1=2.0556R>R_{1}=2.0556, NS(R)N_{\rm S}^{\prime}(R) is the value of the particle number above which the two spirals of the caloric curve are amputed (truncated) and touch each other. When RR1R\gg R_{1}, the function NS(R)N_{\rm S}^{\prime}(R) is given by the relation NS(R)0.128RN_{\rm S}^{\prime}(R)\sim 0.128\,R obtained in the classical study of paper2 . When R<R1=2.0556R<R_{1}=2.0556, we find that NS(R)N1b(R)N_{\rm S}^{\prime}(R)\simeq N_{1}^{b}(R) paper1 . Above that value, the caloric curve looks like Fig. 44 instead of looking like a double spiral.

(xii) When R>R1=2.0556R>R_{1}=2.0556, NS(R)N_{\rm S}(R) is the value of the particle number above which there is no spiral anymore and the caloric curve makes a “loop”. When RROVR\gg R_{\rm OV}, the function NS(R)N_{\rm S}(R) is given by the relation NS(R)0.1415RN_{\rm S}(R)\sim 0.1415\,R obtained in the classical study of paper2 . When R<R1=2.0556R<R_{1}=2.0556, we find that NS(R)NS(R)N1b(R)N_{\rm S}(R)\simeq N_{\rm S}^{\prime}(R)\simeq N_{1}^{b}(R) paper1 .

(xiii) Nmax(R)N_{\rm max}(R) is the maximum particle number below which an equilibrium state may exist for certain values of energy and temperature. For N>Nmax(R)N>N_{\rm max}(R) there is no equilibrium state, whatever the energy and the temperature. When RROVR\gg R_{\rm OV}, the function Nmax(R)N_{\rm max}(R) is given by the relation Nmax(R)0.1764RN_{\rm max}(R)\sim 0.1764\,R obtained in the classical study of roupas ; paper2 . When R<ROV=3.3569R<R_{\rm OV}=3.3569, we find that Nmax(R)NOVb(R)N_{\rm max}(R)\simeq N_{\rm OV}^{b}(R) paper1 . We note that NmaxNOV=0.39853N_{\rm max}\gg N_{\rm OV}=0.39853 when RROV=3.3569R\gg R_{\rm OV}=3.3569 while NmaxNOVN_{\rm max}\ll N_{\rm OV} when RROVR\ll R_{\rm OV}. The change of regime takes place at RROVR\sim R_{\rm OV} where NmaxNOVN_{\rm max}\sim N_{\rm OV} (see the Remark at the end of Sec. IV.3).

X.2 Summary of the main results when RR is fixed and NN is varied

The (R,N)(R,N) phase diagram exhibits two critical points at (RCCP,NCCP)=(12.0,0.424)(R_{\rm CCP},N_{\rm CCP})=(12.0,0.424) and (RMCP,NMCP)=(92.0,1.25)(R_{\rm MCP},N_{\rm MCP})=(92.0,1.25). RCCP=12.0R_{\rm CCP}=12.0 is the radius above which the system experiences a canonical phase transition when NCCP(R)<N<Ne(R)N_{\rm CCP}(R)<N<N_{e}(R). RMCP=92.0R_{\rm MCP}=92.0 is the radius above which the system experiences a microcanonical phase transition when NMCP(R)<N<Nf(R)N_{\rm MCP}(R)<N<N_{f}(R). Below, we summarize the essential features of the microcanonical and canonical phase transitions found for the self-gravitating Fermi gas in general relativity. In this section, we consider the situation where RR is fixed and NN is varied. We recall that there is a possible equilibrium state only for N<Nmax(R)N<N_{\rm max}(R). We have

Nmax(R)0.3104R3/2(RROV),N_{\rm max}(R)\sim 0.3104\,R^{3/2}\qquad(R\ll R_{\rm OV}), (34)
Nmax(R)0.1764R(RROV).N_{\rm max}(R)\sim 0.1764\,R\qquad(R\gg R_{\rm OV}). (35)

As in the previous sections, we do not consider the case of very high energies and very high temperatures which has been treated in roupas ; paper2 .

X.2.1 R<RCCPR<R_{\rm CCP}

When N<NOV(R)N<N_{\rm OV}(R), there is no phase transition and no catastrophic collapse (see Fig. 39). When NOV(R)<N<Nmax(R)N_{\rm OV}(R)<N<N_{\rm max}(R), there is no phase transition but there is a catastrophic collapse towards a black hole at ηc(N,R)\eta_{c}(N,R) in the canonical ensemble and at Λc(N,R)\Lambda_{c}(N,R) in the microcanonical ensemble (see Fig. 40).

X.2.2 RCCP<R<RCCPR_{\rm CCP}<R<R^{\prime}_{\rm CCP}

This case, in which NCCP>NOVN_{\rm CCP}>N_{\rm OV} (see Fig. 48), was not treated explicitly in Secs. V-IX.

In the canonical ensemble when N<NOVN<N_{\rm OV}, there is no phase transition and no catastrophic collapse. When NOV<N<NCCP(R)N_{\rm OV}<N<N_{\rm CCP}(R), there is no phase transition but there is a catastrophic collapse toward a black hole at ηc(N,R)\eta_{c}(N,R). When NCCP(R)<N<NeN_{\rm CCP}(R)<N<N_{e}, there is a zeroth order phase transition from the gaseous phase to the condensed phase at ηc(N,R)\eta_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at ηc(N,R)\eta^{\prime}_{c}(N,R). When N>Ne(R)N>N_{e}(R), there is no phase transition but there is a catastrophic collapse from the gaseous phase to a black hole at ηc(N,R)\eta_{c}(N,R).

In the microcanonical ensemble, the situation is the same as before.

Refer to caption
Figure 48: Zoom of the (R,N)(R,N) phase diagram of the general relativistic Fermi gas. When RCCP=12.0<R<RCCP=12.6R_{\rm CCP}=12.0<R<R^{\prime}_{\rm CCP}=12.6 (resp. R>RCCP=12.6R>R^{\prime}_{\rm CCP}=12.6) the canonical phase transition appears after (resp. before) NOVN_{\rm OV}. When RMCP=92.0<R<RMCP=173R_{\rm MCP}=92.0<R<R^{\prime}_{\rm MCP}=173 (resp. R>RMCP=173R>R^{\prime}_{\rm MCP}=173) the microcanonical phase transition appears after (resp. before) NOVN_{\rm OV}.

X.2.3 RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP}

In the canonical ensemble when N<NCCP(R)N<N_{\rm CCP}(R) there is no phase transition and no catastrophic collapse (see Fig. 4). When NCCP(R)<N<NOVN_{\rm CCP}(R)<N<N_{\rm OV} there is a zeroth order phase transition from the gaseous phase to the condensed phase at ηc(N,R)\eta_{c}(N,R) and no catastrophic collapse (see Figs. 5 and 7). When NOV<N<Ne(R)N_{\rm OV}<N<N_{e}(R) there is a zeroth order phase transition from the gaseous phase to the condensed phase at ηc(N,R)\eta_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at ηc(N,R)\eta^{\prime}_{c}(N,R) (see Fig. 9). When N>Ne(R)N>N_{e}(R) there is no phase transition but there is a catastrophic collapse from the gaseous phase to a black hole at ηc(N,R)\eta_{c}(N,R) (see Figs. 16, 18, 20, 21 and 22).

In the microcanonical ensemble, the situation is the same as before.

X.2.4 RMCP<R<RMCPR_{\rm MCP}<R<R_{\rm MCP}^{\prime}

This case, in which NMCP>NOVN_{\rm MCP}>N_{\rm OV} (see Fig. 48), was not treated explicitly in Secs. V-IX.

In the canonical ensemble, the situation is the same as before so we focus on the microcanonical ensemble. When N<NOVN<N_{\rm OV} there is no phase transition and no catastrophic collapse. When NOV<N<NMCP(R)N_{\rm OV}<N<N_{\rm MCP}(R) there is no phase transition but there is a catastrophic collapse towards a black hole at Λc(N,R)\Lambda_{c}(N,R). When NMCP(R)<N<Nf(R)N_{\rm MCP}(R)<N<N_{f}(R) there is a zeroth order phase transition from the gaseous phase to the condensed phase at Λc(N,R)\Lambda_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at Λc(N,R)\Lambda^{\prime}_{c}(N,R) or Λc′′(N,R)\Lambda^{\prime\prime}_{c}(N,R). When N>Nf(R)N>N_{f}(R) there is no phase transition but there is a catastrophic collapse from the gaseous phase to a black hole at Λc(N,R)\Lambda_{c}(N,R).

X.2.5 R>RMCPR>R_{\rm MCP}^{\prime}

In the canonical ensemble, the situation is the same as before so we focus on the microcanonical ensemble. When N<NMCP(R)N<N_{\rm MCP}(R) there is no phase transition and no catastrophic collapse. When NMCP(R)<N<NOVN_{\rm MCP}(R)<N<N_{\rm OV} there is a zeroth order phase transition from the gaseous phase to the condensed phase at Λc(N,R)\Lambda_{c}(N,R) and no catastrophic collapse (see Fig. 27). When NOV<N<Nf(R)N_{\rm OV}<N<N_{f}(R) there is a zeroth order phase transition from the gaseous phase to the condensed phase at Λc(N,R)\Lambda_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at Λc′′(N,R)\Lambda^{\prime\prime}_{c}(N,R) (see Fig. 29). When N>Nf(R)N>N_{f}(R) there is no phase transition but there is a catastrophic collapse from the gaseous phase to a black hole at Λc(N,R)\Lambda_{c}(N,R) (see Figs. 32, 34 and 36).

X.3 Summary of the main results when NN is fixed and RR is varied

We now consider the situation where NN is fixed and RR is varied.303030This is, for example, the viewpoint adopted by Hertel and Thirring ht as recalled in the introduction (see also Sec. XIII). These results can be deduced from the (R,N)(R,N) phase diagram of Fig. 47 by taking the inverse of the functions NX(R)N_{\rm X}(R). This leads to the (N,R)(N,R) phase diagram. We note that there is a possible equilibrium state only for R>Rmin(N)R>R_{\rm min}(N) where Rmin(N)R_{\rm min}(N) is the inverse function of Nmax(R)N_{\rm max}(R). We have

Rmin(N)2.181N2/3(NNOV),R_{\rm min}(N)\sim 2.181\,N^{2/3}\qquad(N\ll N_{\rm OV}), (36)
Rmin(N)5.669N(NNOV).R_{\rm min}(N)\sim 5.669\,N\qquad(N\gg N_{\rm OV}). (37)

X.3.1 N<NOVN<N_{\rm OV}

In the canonical ensemble when Rmin(N)<R<RCCP(N)R_{\rm min}(N)<R<R_{\rm CCP}(N) there is no phase transition and no catastrophic collapse. When R>RCCP(N)R>R_{\rm CCP}(N) (with RCCP(N)12.8N1/3R_{\rm CCP}(N)\sim 12.8\,N^{-1/3} in the nonrelativistic limit NNOVN\ll N_{\rm OV}) there is a zeroth order phase transition from the gaseous phase to the condensed phase at ηc(N,R)\eta_{c}(N,R) and no catastrophic collapse.

In the microcanonical ensemble when Rmin(N)<R<RMCP(N)R_{\rm min}(N)<R<R_{\rm MCP}(N) (with RMCP(N)130N1/3R_{\rm MCP}(N)\sim 130\,N^{-1/3} in the nonrelativistic limit NNOVN\ll N_{\rm OV}) there is no phase transition and no catastrophic collapse. When R>RMCP(N)R>R_{\rm MCP}(N) there is a zeroth order phase transition from the gaseous phase to the condensed phase at Λc(N,R)\Lambda_{c}(N,R) and no catastrophic collapse.

This is essentially like in the nonrelativistic limit ijmpb . Relativistic corrections occur only close to NOVN_{\rm OV} and/or close to Rmin(N)R_{\rm min}(N).

X.3.2 NOV<N<NCCPNeN_{\rm OV}<N<N_{\rm CCP}\simeq N_{e}

In the canonical ensemble when Rmin(N)<R<RCCP(N)R_{\rm min}(N)<R<R_{\rm CCP}(N) there is no phase transition but there is a catastrophic collapse towards a black hole at ηc(N,R)\eta_{c}(N,R). When R>RCCP(N)R>R_{\rm CCP}(N) there is a zeroth order phase transition from the gaseous phase to the condensed phase at ηc(N,R)\eta_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at ηc(N,R)\eta^{\prime}_{c}(N,R).

In the microcanonical ensemble when Rmin(N)<R<RMCP(N)R_{\rm min}(N)<R<R_{\rm MCP}(N) there is no phase transition but there is a catastrophic collapse towards a black hole at Λc(N,R)\Lambda_{c}(N,R). When R>RMCP(N)R>R_{\rm MCP}(N) there is a zeroth order phase transition from the gaseous phase to the condensed phase at Λc(N,R)\Lambda_{c}(N,R) and a catastrophic collapse from the condensed phase to a black hole at Λc′′(N,R)\Lambda^{\prime\prime}_{c}(N,R).

X.3.3 NCCPNe<N<NMCPNfN_{\rm CCP}\simeq N_{e}<N<N_{\rm MCP}\simeq N_{f}

In the canonical ensemble, when R>Rmin(N)R>R_{\rm min}(N) there is no phase transition but there is a catastrophic collapse to a black hole at ηc(N,R)\eta_{c}(N,R).

In the microcanonical ensemble, the situation is the same as before.

X.3.4 N>NMCPN>N_{\rm MCP}

In the canonical ensemble, the situation is the same as before. In the microcanonical ensemble when R>Rmin(N)R>R_{\rm min}(N) there is no phase transition but there is a catastrophic collapse towards a black hole at Λc(N,R)\Lambda_{c}(N,R).

This is essentially like in the classical limit roupas ; paper2 .

XI The nonrelativistic and classical limits

In this section, we consider the nonrelativistic (c+c\rightarrow+\infty) and classical (0\hbar\rightarrow 0) limits and study the commutation of these limits.

XI.1 An apparent paradox related to the commutation of the limits 0\hbar\rightarrow 0 and c+c\rightarrow+\infty

The commutation of the limits 0\hbar\rightarrow 0 and c+c\rightarrow+\infty leads to an apparent paradox. This can be seen from the expression of the maximum OV particle number given by

NOV=0.398532g(cG)3/21m3.N_{\rm OV}=0.39853\,\sqrt{\frac{2}{g}}\left(\frac{\hbar c}{G}\right)^{3/2}\frac{1}{m^{3}}. (38)

(i) If we take the nonrelativistic limit c+c\rightarrow+\infty first ijmpb , we find that NOV+N_{\rm OV}\rightarrow+\infty. Therefore, we always have N<NOVN<N_{\rm OV}. As a result, there is always an equilibrium state at low temperatures and low energies, whatever the value of \hbar, i.e., even if we consider the classical limit 0\hbar\rightarrow 0.313131In the canonical ensemble, when 1\hbar\ll 1 and T<TcT<T_{c}, the equilibrium state corresponds to a fermion ball containing most of the mass. When 0\hbar\rightarrow 0 the fermion ball contains all the mass and its radius goes to zero. In that case, we get a Dirac peak of mass MM (see Appendix E.2.1). In the microcanonical ensemble, when 1\hbar\ll 1 and E<EcE<E_{c}, the equilibrium state corresponds to a fermion ball containing a fraction of the total mass surrounded by a hot halo. When 0\hbar\rightarrow 0 the mass of the fermion ball and its radius go to zero while its potential energy goes to -\infty. As a result, the temperature of the halo goes to ++\infty. In that case, we get a Dirac peak of zero mass and infinite potential energy (binary) surrounded by an infinitely hot halo (see Appendix E.2.2). Note that for >0\hbar>0, as small as one pleases, there is always a regular equilibrium state.

(ii) If we take the classical limit 0\hbar\rightarrow 0 first roupas ; paper2 , we find that NOV0N_{\rm OV}\rightarrow 0. Therefore, we always have N>NOVN>N_{\rm OV}. As a result, the system undergoes a catastrophic collapse at low temperatures and low energies, whatever the value of cc, i.e., even if we consider the nonrelativistic limit c+c\rightarrow+\infty.

Therefore, if we consider a nonrelativistic classical gas (c+c\rightarrow+\infty and 0\hbar\rightarrow 0), the first argument tells us that there is an equilibrium state at low temperatures and low energies while the second argument tells us that there is no equilibrium state at low temperatures and low energies. How can we reconcile these two apparent contradictory situations? In the next two subsections, we re-express these results in terms of dimensionless variables, and in the third subsection we provide a solution of this apparent paradox.

XI.2 When the nonrelativistic limit c+c\rightarrow+\infty is taken before the classical limit 0\hbar\rightarrow 0

In this subsection, we consider the situation where the nonrelativistic limit (c+c\rightarrow+\infty) is taken before the classical limit (0\hbar\rightarrow 0). Using the dimensionless variables of Appendix B, the nonrelativistic limit corresponds to N0N\rightarrow 0 and R+R\rightarrow+\infty in such a way that NR3NR^{3} is fixed.323232This scaling is obtained in order to keep the parameter μ\mu defined in ijmpb fixed (see also Sec. IV.2). Coming back to dimensional variables, the nonrelativistic limit corresponds to NNOV(c/G)3/2/m3N\ll N_{\rm OV}\sim(\hbar c/G)^{3/2}/m^{3} and RROV(3/Gc)1/2/m2R\gg R_{\rm OV}\sim(\hbar^{3}/Gc)^{1/2}/m^{2} with μ2NR3m9G3/6\mu^{2}\sim NR^{3}m^{9}G^{3}/\hbar^{6} fixed. This is consistent with the fact that NOV+N_{\rm OV}\rightarrow+\infty and ROV0R_{\rm OV}\rightarrow 0 when c+c\rightarrow+\infty. This scaling defines an ensemble of parallel lines of constant μ=(42/π)(NR3)1/2\mu=(4\sqrt{2}/\pi)(NR^{3})^{1/2} in the bottom right panel of Fig. 49. In the nonrelativistic limit, the caloric curves are the same for any couple of points (R,N)(R,N) belonging to a given μ\mu-line. As μ\mu increases, the μ\mu-lines move to the right and the system becomes more and more classical ijmpb .

The phase transitions occuring in a nonrelativistic self-gravitating Fermi gas have been studied in ijmpb . When μ<μCCP=83\mu<\mu_{\rm CCP}=83 there is no phase transition (see Fig. 14 of ijmpb ). When μCCP=83<μ<μMCP=2670\mu_{\rm CCP}=83<\mu<\mu_{\rm MCP}=2670 there is a canonical phase transition (see Fig. 31 of ijmpb ). When μ>μMCP=2670\mu>\mu_{\rm MCP}=2670 there are canonical and microcanonical phase transitions (see Fig. 21 of ijmpb ). The classical limit corresponds to μ+\mu\rightarrow+\infty (see Fig. 22 of ijmpb )

Refer to caption
Figure 49: Lines of constant μ\mu in the (R,N)(R,N) phase diagram of Fig. 47 characterizing the nonrelativistic limit N0N\rightarrow 0 and R+R\rightarrow+\infty with fixed μ=(42/π)(NR3)1/2\mu=(4\sqrt{2}/\pi)(NR^{3})^{1/2}. We have chosen μ=10\mu=10, μ=103\mu=10^{3} and μ=104\mu=10^{4} (brown lines) corresponding to the values appearing in Fig. 14 of ijmpb .
Refer to caption
Figure 50: Caloric curve of nonrelativistic self-gravitating fermions in the classical limit μ+\mu\rightarrow+\infty (here μ=107\mu=10^{7}).

For large but finite values of μ\mu (see Fig. 50), the series of equilibria forms a spiral which finally unwinds, progresses backward along an inverse spiral until Λ(μ)\Lambda_{*}(\mu), turns right, forms a lower branch, and finally tends towards an asymptote at Λmax(μ)=0.0642μ2/3\Lambda_{\rm max}(\mu)=0.0642\,\mu^{2/3} where η+\eta\rightarrow+\infty (ground state). When μ+\mu\rightarrow+\infty, the direct spiral tends to a limit curve (ηc(μ)2.52\eta_{c}(\mu)\rightarrow 2.52, Λc(μ)0.335\Lambda_{c}(\mu)\rightarrow 0.335), the inverse spiral coincides with the direct spiral, the turning point Λ(μ)\Lambda_{*}(\mu) is pushed towards -\infty, the turning point η(μ)\eta_{*}(\mu) is pushed towards 0, the lower branch coincides with the xx-axis (η=0\eta=0) and the asymptote at Λmax(μ)\Lambda_{\rm max}(\mu) is pushed towards ++\infty. In this limit, we recover the standard nonrelativistic classical caloric curve of Fig. 1 plus a singular branch at η=0\eta=0 corresponding to a Dirac peak of zero mass but infinite potential energy surrounded by a halo of infinite temperature, and a singular branch at Λ=Λmax+\Lambda=\Lambda_{\rm max}\rightarrow+\infty corresponding to a Dirac peak containing all the mass (ground state).

XI.3 When the classical limit 0\hbar\rightarrow 0 is taken before the nonrelativistic limit c+c\rightarrow+\infty

In this section, we consider the situation where the classical limit (0\hbar\rightarrow 0) is taken before the nonrelativistic limit (c+c\rightarrow+\infty). Using the dimensionless variables of Appendix B, the classical limit corresponds to N+N\rightarrow+\infty and R+R\rightarrow+\infty in such a way that N/RN/R is fixed.333333This scaling is obtained in order to keep the parameter ν\nu defined in roupas ; paper2 fixed (see also Sec. IV.3). Coming back to dimensional variables, the classical limit corresponds to NNOV(c/G)3/2/m3N\gg N_{\rm OV}\sim(\hbar c/G)^{3/2}/m^{3} and RROV(3/Gc)1/2/m2R\gg R_{\rm OV}\sim(\hbar^{3}/Gc)^{1/2}/m^{2} with νGNm/Rc2\nu\sim GNm/Rc^{2} fixed. This is consistent with the fact that NOV0N_{\rm OV}\rightarrow 0 and ROV0R_{\rm OV}\rightarrow 0 when 0\hbar\rightarrow 0. This scaling defines an ensemble of parallel lines of constant ν=N/R\nu=N/R in the upper right panel of Fig. 51. In the classical limit, the caloric curves are the same for any couple of points (R,N)(R,N) belonging to a given ν\nu-line. As ν\nu decreases, the ν\nu-lines move to the right and the system becomes less and less relativistic roupas ; paper2 .

Refer to caption
Figure 51: Lines of constant ν\nu in the (R,N)(R,N) phase diagram of Fig. 47 characterizing the classical limit N+N\rightarrow+\infty and R+R\rightarrow+\infty with fixed ν=N/R\nu=N/R. We have chosen ν=0.005\nu=0.005, ν=0.05\nu=0.05 and ν=0.1\nu=0.1 (brown lines) corresponding to the values appearing in Fig. 12 of paper2 .

The classical general relativistic self-gravitating gas has been studied in roupas ; paper2 . When ν<νS=0.128\nu<\nu^{\prime}_{S}=0.128 the caloric curve displays a double spiral (see Fig. 7 of paper2 ). When νS=0.128<ν<νS=0.1415\nu^{\prime}_{S}=0.128<\nu<\nu_{S}=0.1415 the two spirals are truncated (see Fig. 8 of paper2 ). When νS=0.1415<ν<νmax=0.1764\nu_{S}=0.1415<\nu<\nu_{\rm max}=0.1764 the caloric curve makes a loop (see Fig. 9 of paper2 ). When ν=νmax=0.1764\nu=\nu_{\rm max}=0.1764 the caloric curve reduces to a point and disappears (see Fig. 15 of paper2 ). There is a gravitational collapse at low energies and low temperatures (cold spiral) and at high energies and high temperatures (hot spiral). The nonrelativistic limit corresponds to ν0\nu\rightarrow 0 (see Fig. 12 of paper2 ).

Refer to caption
Figure 52: Caloric curve of classical self-gravitating systems in the nonrelativistic limit ν0\nu\rightarrow 0 (here ν=0.01\nu=0.01). This figures zooms on the cold spiral.
Refer to caption
Figure 53: Caloric curve of classical self-gravitating systems in the nonrelativistic limit ν0\nu\rightarrow 0 (here ν=0.01\nu=0.01). This figures zooms on the hot spiral.

For small but nonzero values of ν\nu (see Figs. 52 and 53), the series of equilibria forms two spirals very distant to each other. When ν0\nu\rightarrow 0, the cold spiral tends to a limit curve (ηc(ν)2.52\eta_{c}(\nu)\rightarrow 2.52, Λc(ν)0.335\Lambda_{c}(\nu)\rightarrow 0.335) while the hot spiral is rejected to the left at infinity (Λmin(ν)0.24631/ν2\Lambda_{\rm min}(\nu)\sim-0.24631/\nu^{2}\rightarrow-\infty and ηmin(ν)18.27ν20\eta_{\rm min}(\nu)\sim 18.27\,\nu^{2}\rightarrow 0) paper2 . In this limit, we recover the standard nonrelativistic classical caloric curve of Fig. 1 plus a spiral at very high energies and very high temperatures (Λmin\Lambda_{\rm min}\rightarrow-\infty and ηmin0\eta_{\rm min}\rightarrow 0).

XI.4 The solution of the apparent paradox

In the two processes described previously, in which c+c\rightarrow+\infty and 0\hbar\rightarrow 0, we recover the standard classical nonrelativistic spiral of Fig. 1. However, the manner to obtain it is different depending on the order in which the limits are taken.

When the nonrelativistic limit c+c\rightarrow+\infty is taken first, there is no “hot” spiral at Λmin\Lambda_{\rm min} since the hot spiral is a general relativity result associated with a form of self-gravitating radiation. By contrast, there is always an asymptote at Λmax\Lambda_{\rm max} corresponding to an equilibrium state at T=0T=0 (the ground state of the Fermi gas) because N<NOV=+N<N_{\rm OV}=+\infty. In the classical limit 0\hbar\rightarrow 0, the asymptote at Λmax\Lambda_{\rm max} is rejected at ++\infty (while Λ\Lambda_{*} is pushed towards -\infty and η\eta_{*} towards zero) so the caloric curve has the form of a single spiral. For \hbar infinitely small but finite, we get the classical caloric curve plus singular branches η0\eta\simeq 0 (horizontal) and Λmax+\Lambda_{\rm max}\simeq+\infty (vertical) as described previously. According to the results of Appendix E, we have the scalings Emin2E_{\min}\propto-\hbar^{-2}, E2(ln)7/3E_{*}\propto\hbar^{-2}(-\ln\hbar)^{-7/3} and T2(ln)1T_{*}\propto\hbar^{-2}(-\ln\hbar)^{-1} for 0\hbar\rightarrow 0.

When the classical limit 0\hbar\rightarrow 0 is taken first, there is always a “hot” spiral at Λmin\Lambda_{\rm min} since the system is relativistic. By contrast, there is no asymptote at Λmax\Lambda_{\rm max}, i.e., there is no equilibrium state at T=0T=0 (ground state) because N>NOV=0N>N_{\rm OV}=0. In the nonrelativistic limit c+c\rightarrow+\infty, the “hot” spiral is rejected at infinity so the caloric curve has the form of a single spiral. For cc infinitely large but finite, we get the nonrelativistic caloric curve plus a spiral at very high energies Λmin\Lambda_{\rm min} and temperatures ηmin\eta_{\rm min} as described previously. According to the results of paper2 (see also footnote 19), we have the scalings Emaxc4E_{\max}\sim c^{4} and Tmaxc4T_{\rm max}\sim c^{4} for c+c\rightarrow+\infty.

The previous considerations lead to the following conclusion.

The nonrelativistic limit ijmpb corresponds to RROVR\gg R_{\rm OV} and NNOVN\ll N_{\rm OV}. This corresponds to the lower panel QNR of Fig. 54 below NOVN_{\rm OV}. In that case, for a fixed radius RR, the caloric curve shows no phase transition below NCCPN_{\rm CCP}, a canonical phase transition above NCCPN_{\rm CCP} and a microcanonical phase transition (in addition to the canonical phase transition) above NMCPN_{\rm MCP}. The classical limit corresponds to NMCPNNOVN_{\rm MCP}\ll N\ll N_{\rm OV}. This corresponds to the lower right panel CNR1 of Fig. 54 (far on the right) below NOVN_{\rm OV}.

The classical limit roupas ; paper2 corresponds to RROVR\gg R_{\rm OV} and NNOVN\gg N_{\rm OV}. This corresponds to the upper panel CR of Fig. 54 above NOVN_{\rm OV}. In that case, for a fixed radius RR, the caloric curve shows a double spiral below NSN^{\prime}_{S}, a truncated double spiral above NSN^{\prime}_{S}, a loop above NSN_{S} and no equilibrium states above NmaxN_{\rm max}. The nonrelativistic limit corresponds to NOVNNSN_{\rm OV}\ll N\ll N^{\prime}_{S}. This corresponds to the upper right panel CNR2 of Fig. 54 (far on the right) above NOVN_{\rm OV}.

Therefore, the nonrelativistic ++ classical limit corresponds to two distinct regions in the right panel of Fig. 54, below or above NOVN_{\rm OV}, depending on the order in which the limits are taken. Note also that quantum and relativistic effects are both important only close to ROVR_{\rm OV} or only close to NOVN_{\rm OV}. This corresponds to the region denoted QR in Fig. 54.

Refer to caption
Figure 54: Schematic representation of the different regimes of the self-gravitating Fermi gas. QR: quantum relativistic; QNR: quantum nonrelativistic; CR: classical relativistic; CNR: classical nonrelativistic (CNR1 is when the nonrelativistic limit cc\rightarrow\infty limit is taken before the classical limit 0\hbar\rightarrow 0; CNR2 is when the classical limit 0\hbar\rightarrow 0 is taken before the nonrelativistic limit cc\rightarrow\infty limit).

XII Relativistic and quantum corrections

XII.1 Relativistic corrections to the nonrelativistic caloric curves

We have seen in Sec. XI.2 that the nonrelativistic caloric curves of the self-gravitating Fermi gas correspond to parallel lines of constant μ=(42/π)(NR3)1/2\mu=(4\sqrt{2}/\pi)(NR^{3})^{1/2} in the bottom right panel of Fig. 49. On a line of constant μ\mu, the nonrelativistic limit is valid when R+R\rightarrow+\infty and N0N\rightarrow 0 (physically RROVR\gg R_{\rm OV} and NNOVN\ll N_{\rm OV}). For small values of RR and large values of NN, i.e., at the top of a μ\mu-line, relativistic corrections come into play.

XII.1.1 μ=103\mu=10^{3}

Let us first consider the case μ=103\mu=10^{3} corresponding to μCCP=83<μ<μMCP=2670\mu_{\rm CCP}=83<\mu<\mu_{\rm MCP}=2670 (see Figs. 55 and 56). When N0N\rightarrow 0, we recover the nonrelativistic caloric curve plotted in Fig. 31 of ijmpb . It has a NN-shape structure leading to canonical phase transitions. The hot spiral is rejected at infinity. Let us increase the number of particles NN at fixed μ\mu, hence decreasing the box radius RR accordingly, in order to see the relativisitic corrections. The description in the change of the caloric curves as relativistic effects become more and more important is qualitatively similar to that given in Sec. V for R=50R=50 when N>NCCPN>N_{\rm CCP}. The only difference is that we work at fixed μ\mu (with μCCP<μ<μMCP\mu_{\rm CCP}<\mu<\mu_{\rm MCP}) instead of fixed RR (with RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP}). Therefore, in the (R,N)(R,N) diagram, we follow the μ=103\mu=10^{3} oblique line (see Fig. 49) instead of the R=50R=50 vertical line (see Fig. 47). As a result, when N0N\rightarrow 0, we tend towards a limit curve (the nonrelativistic caloric curve with μ=103\mu=10^{3} of ijmpb ) which presents a canonical phase transition while in the case R=50R=50 studied in Sec. V the canonical phase transition disappears when N<NCCPN<N_{\rm CCP}.

Remark: The characteristic particle numbers NXN_{\rm X} described in Sec. X.1 now depend on μ\mu instead of RR. They can be obtained by considering the intersection between the curves NX(R)N_{\rm X}(R) and the curve N=π2μ2/(32R3)N=\pi^{2}\mu^{2}/(32R^{3}) with fixed μ\mu. In this manner, we obtain NS(μ)=0.159μN_{\rm S}^{\prime}(\mu)=0.159\sqrt{\mu}, NS(μ)=0.172μN_{\rm S}(\mu)=0.172\sqrt{\mu}, and Nmax(μ)=0.203μN_{\rm max}(\mu)=0.203\sqrt{\mu}.

Refer to caption
Figure 55: Caloric curves for different values of NN at fixed μ=(42/π)(NR3)1/2=103\mu=(4\sqrt{2}/\pi)(NR^{3})^{1/2}=10^{3}. When N0N\rightarrow 0 (black curve), we recover the nonrelativistic caloric curve with an NN-shape structure obtained in Fig. 31 of ijmpb . The present figure illustrates the effect of general relativity on that caloric curve as NN increases.
Refer to caption
Figure 56: Same as Fig. 55 for larger values of NN.

XII.1.2 μ=105\mu=10^{5}

Let us now consider the case μ=105\mu=10^{5} corresponding to μ>μMCP=2670\mu>\mu_{\rm MCP}=2670 (see Figs. 57-59). When N0N\rightarrow 0, we recover the nonrelativistic caloric curve plotted in Fig. 21 of ijmpb . It has a ZZ-shape structure leading to microcanonical phase transitions.343434The resemblance with a dinosaur’s neck in clear on this figure ijmpb . The vertical asymptote at Λmax\Lambda_{\rm max} is outside of the frame. The hot spiral is rejected at infinity. Let us increase the number of particles NN at fixed μ\mu, hence decreasing the box radius RR accordingly, in order to see the relativisitic corrections. The description in the change of the caloric curves as relativistic effects become more and more important is qualitatively similar to that given in Sec. VI for R=600R=600 when N>NMCPN>N_{\rm MCP}. The only difference is that we work at fixed μ\mu (with μ>μMCP\mu>\mu_{\rm MCP}) instead of fixed RR (with R>RMCPR>R_{\rm MCP}). Therefore, in the (R,N)(R,N) diagram, we follow the μ=105\mu=10^{5} oblique line (see Fig. 49) instead of the R=600R=600 vertical line (see Fig. 47). As a result, when N0N\rightarrow 0, we tend towards a limit curve (the nonrelativistic caloric curve with μ=105\mu=10^{5} of ijmpb ) which presents a microcanonical phase transition while in the case R=600R=600 studied in Sec. VI the microcanonical phase transition disappears when N<NMCPN<N_{\rm MCP}.

Refer to caption
Figure 57: Caloric curves for different values of NN at fixed μ=(42/π)(NR3)1/2=105\mu=(4\sqrt{2}/\pi)(NR^{3})^{1/2}=10^{5}. When N0N\rightarrow 0 (black curve), we recover the nonrelativistic caloric curve with a ZZ-shape structure obtained in Fig. 21 of ijmpb . The present figure illustrates the effect of general relativity on that caloric curve as NN increases.
Refer to caption
Figure 58: Same as Fig. 57 for larger values of NN.
Refer to caption
Figure 59: Same as Figs. 57 and 58 for larger values of NN.

XII.2 Quantum corrections to the classical caloric curve

We have seen in Sec. XI.3 that the classical caloric curves of the general relativistic self-gravitating gas correspond to parallel lines of constant ν=N/R\nu=N/R in the upper right panel of Fig. 51. On a line of constant ν\nu, the classical limit is valid when R+R\rightarrow+\infty and N+N\rightarrow+\infty (physically RROVR\gg R_{\rm OV} and NNOVN\gg N_{\rm OV}). For small values of RR and small values of NN, i.e., on the left of a ν\nu-line, quantum corrections come into play.

Let us consider the case ν=0.1\nu=0.1 (see Figs. 60 and 61). When N+N\rightarrow+\infty (black curve) we recover the classical general relativistic caloric curve plotted in Fig. 1 of paper2 . It has the form of a double spiral leading to a gravitational collapse for both cold and hot systems.353535When N+N\rightarrow+\infty the caloric curves of Fig. 60 and 61 tend towards a limit curve (the classical general relativistic caloric curve with ν=0.1\nu=0.1 of paper2 ) which presents a double spiral while in the cases studied in Secs. V-IX the two spirals merge and disappear when NNmaxN\rightarrow N_{\rm max}. This is because in Secs. V-IX we work at fixed radius RR while in the present case we work at fixed ν\nu so that the radius RR increases as NN increases. Let us decrease the number of particles NN at fixed ν\nu, hence decreasing the box radius RR accordingly, in order to see the quantum corrections. When N>NOV=0.39853N>N_{\rm OV}=0.39853 (blue and green curves), the caloric curve keeps a similar structure. When N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853 (red curve) the caloric curve is made of two branches, each presenting an asymptote (right) and a spiral (left). When N<N1=0.18131N<N_{1}=0.18131 (purple curve) the caloric curve has just one branch presenting an asymptote (right) and a spiral (left).

Remark: For smaller values of ν\nu,363636This case is specifically investigated in rc . we have a richer variety of caloric curves as NN decreases with the appearance of canonical and microcanonical phase transitions. This can be seen on the phase diagram of Fig. 51. The characteristic particle numbers NXN_{\rm X} described in Sec. X.1 now depend on ν\nu instead of RR. They can be obtained by considering the intersection between the curves NX(R)N_{\rm X}(R) and the line N=νRN=\nu R. In this manner, we obtain NCCP(ν)=6.79ν3/4N_{\rm CCP}(\nu)=6.79\,\nu^{3/4} and NMCP(ν)=38.5ν3/4N_{\rm MCP}(\nu)=38.5\,\nu^{3/4}. There is no microcanonical phase transition for ν>νMCP=0.0136\nu>\nu_{\rm MCP}=0.0136 and there is no canonical phase transition for ν>νCCP=0.0353\nu>\nu_{\rm CCP}=0.0353.

Refer to caption
Figure 60: Caloric curves for different values of NN at fixed ν=N/R=0.1\nu=N/R=0.1. When N+N\rightarrow+\infty (black curve), we recover the classical caloric curve with a double spiral obtained in Fig. 1 of paper2 . The present figure illustrates the effect of quantum mechanics on that caloric curve as NN decreases. This figure is focused on the evolution of the cold spiral.
Refer to caption
Figure 61: Same as Fig. 60. This figure is focused on the evolution of the hot spiral.

XIII Astrophysical applications

XIII.1 Nonrelativistic model: collapse towards a fermion star

In this section, we discuss astrophysical applications of the nonrelativistic self-gravitating Fermi gas model ijmpb . This model exhibits a phase transition from a gaseous star to a compact fermion star. The fermion star can be a white dwarf, a neutron star or a DM fermion ball.373737In this section we take g=2g=2 in the numerical applications. We relate this phase transition to the onset of red-giant structure and to the supernova phenomenon.

XIII.1.1 Canonical ensemble

Let us consider a system of nonrelativistic self-gravitating fermions in the canonical ensemble. The canonical phase transition appears for μμCCP=83\mu\geq\mu_{\rm CCP}=83 (canonical critical point) ijmpb hence for

RRCCPNR(N)=0.517h2Gm8/3M1/3g2/3.R\geq R^{\rm NR}_{\rm CCP}(N)=0.517\frac{h^{2}}{Gm^{8/3}M^{1/3}g^{2/3}}. (39)

We assume that we are in this situation. In that case, the caloric curve has the form of Fig. 5. The natural evolution proceeds along the series of equilibria towards lower and lower temperatures (see Sec. III). We assume that the system is initially in the metastable gaseous phase. As its temperature decreases it remains in this phase up to the critical point TcT_{c} at which the metastable gaseous branch disappears.383838We recall that the collapse takes place at the critical (spinodal) point TcT_{c}, not at the transition point TtT_{t}, because of the tremendously long lifetime of metastable gaseous states. This critical temperature is not very sensitive on quantum effects (when μ1\mu\gg 1) so it can be approximated by its classical value ηc=βcGMm/R=2.52\eta_{c}=\beta_{c}GMm/R=2.52 yielding

kBTc=0.397GMmR.k_{B}T_{c}=0.397\,\frac{GMm}{R}. (40)

At that point, the system collapses and forms a compact fermion star. As explained in Appendix E.2.1, in the canonical ensemble, the fermion star contains almost all the mass (MCMM_{C}\simeq M) and is surrounded by a tenuous atmosphere. If we approximate the fermion star by a Fermi gas at T=0T=0 (polytrope of index n=3/2n=3/2) containing all the mass, its radius is given by (see Appendix E.2.1):

RC=0.181h2Gm8/3M1/3g2/3.R_{C}=0.181\frac{h^{2}}{Gm^{8/3}M^{1/3}g^{2/3}}. (41)

The energy of the gaseous phase at the point of isothermal collapse is not very sensitive on quantum effects (when μ1\mu\gg 1) so it can be approximated by its classical value Λgas=EgasR/GM2=0.199\Lambda_{\rm gas}=-E_{\rm gas}R/GM^{2}=0.199 yielding

Egas=0.199GM2R.E_{\rm gas}=-0.199\frac{GM^{2}}{R}. (42)

The energy of the condensed object can be approximated by its value at T=0T=0 yielding (see Appendix E.2.1):

Emin=2.36G2m8/3M7/3g2/3h2.E_{\rm min}=-2.36\frac{G^{2}m^{8/3}M^{7/3}g^{2/3}}{h^{2}}. (43)

Finally, the collapse time in the canonical ensemble is of the order of the dynamical time

tDR3/2(GM)1/2.t_{D}\sim\frac{R^{3/2}}{(GM)^{1/2}}. (44)

We note that the collapse in the canonical ensemble corresponds to a pure implosion since almost all the mass is in the condensate (fermion star), not in the halo. The thermodynamical reason for this implosion phenomenon is explained in Appendix E.1. It is also corroborated by the following arguments. If we calculate the density perturbation δρ\delta\rho that triggers the instability at TcT_{c}, we find that it has only one node (see Fig. 10 of aa ). Therefore, the instability develops itself in such a way that the density in the core increases while the density in the halo decreases. We also find that the velocity profile δv\delta v has no node (see Fig. 12 of aa ) so that it is purely inward. These two results confirm the implosion phenomenon.

We make below numerical applications to illustrate the preceding results. They correspond to the nonrelativistic models of neutron stars (made of neutrons of mass mn=0.940GeV/c2m_{n}=0.940\,{\rm GeV/c^{2}}) and dark matter halos (made of fermions of mass m=17.2keV/c2m=17.2\,{\rm keV/c^{2}}) respectively studied by Hertel and Thirring ht and Bilic and Viollier bvn as recalled in the Introduction.

Neutron stars (crude model) ht : We consider a gas of neutrons of total mass M=1MM=1\,M_{\odot}.393939The maximum mass and minimum radius of an ideal neutron star set by general relativity are MOV=0.710MM_{\rm OV}=0.710\,M_{\odot} and ROV=9.16kmR_{\rm OV}=9.16\,{\rm km}. Therefore, the value of the mass chosen by ht is larger than the maximum mass. If general relativity were taken into account (see below) the system would not form a neutron star but would collapse towards a black hole (assuming that all the initial mass goes in the compact object). It contains about N1057N\sim 10^{57} neutrons. The canonical phase transition appears for RRCCPNR(N)=43.1kmR\geq R^{\rm NR}_{\rm CCP}(N)=43.1\,{\rm km}. For a gaseous star of size R=100kmR=100\,{\rm km}, corresponding to μ=294\mu=294, the collapse temperature is Tc=6.39×1010KT_{c}=6.39\times 10^{10}\,{\rm K} (the transition temperature is Tt=7.03×1010KT_{t}=7.03\times 10^{10}\,{\rm K} ht ). The radius of the neutron star of mass MC1MM_{C}\simeq 1\,M_{\odot} resulting from the collapse of the gaseous star is RC=15.1kmR_{C}=15.1\,{\rm km}. The energy of the gaseous star is Egas=5.25×1051ergE_{\rm gas}=-5.25\times 10^{51}\,{\rm erg} and the energy of the neutron star is Emin=7.49×1052ergE_{\rm min}=-7.49\times 10^{52}\,{\rm erg}. The energy released during the collapse is ΔE=EgasEmin=6.96×1052erg\Delta E=E_{\rm gas}-E_{\rm min}=6.96\times 10^{52}\,{\rm erg}. The collapse time is a multiple of tD2.74×103st_{D}\sim 2.74\times 10^{-3}\,{\rm s} which is very short on an astrophysical timescale.404040Note that the prefactor of the collapse time is uncertain and could be of order 10310^{3} or larger. As a result, our estimate of the collapse time is not inconsistent with the duration of the supernova phenomenon which can be as short as a few seconds. For supernovae, the energy ΔE\Delta E may be carried quickly by neutrinos. The release of gravitational energy in a supernova (WGM2/RNmc2W\sim GM^{2}/R\sim Nmc^{2}) is comparable with the energy of fusion processses which kept the star shining during the first 101010^{10} years of its life. However, this takes place in a few seconds (or days) instead of 101010^{10} years leading to a huge luminosity. This explains why a star can become as bright as the whole galaxy.

Fermion ball bvn : We consider a gas of DM fermions of total mass M=10MM=10\,M_{\odot}.414141This is the typical mass of a DM halo surrounding a baryonic star of mass 1M\sim 1\,M_{\odot} since the present fraction of baryons and dark matter are Ωb,0=0.0487273\Omega_{\rm b,0}=0.0487273 and Ωdm,0=0.2645\Omega_{\rm dm,0}=0.2645 respectively. Here and in the following we consider DM fermions of mass m=17.2keV/c2m=17.2\,{\rm keV/c^{2}}. The maximum mass and minimum radius set by general relativity are MOV=2.12×109MM_{\rm OV}=2.12\times 10^{9}\,M_{\odot} and ROV=8.86×104pcR_{\rm OV}=8.86\times 10^{-4}\,{\rm pc}. Since MMOVM\ll M_{\rm OV}, the nonrelativistic model is justified in that case. The canonical phase transition appears for RRCCPNR(N)=2.78pcR\geq R^{\rm NR}_{\rm CCP}(N)=2.78\,{\rm pc}. For a gaseous halo of size R=41.3pcR=41.3\,{\rm pc}, corresponding to μ=4747\mu=4747, the collapse temperature is Tc=9.18×107KT_{c}=9.18\times 10^{-7}\,{\rm K} (the transition temperature is Tt=2.83×106KT_{t}=2.83\times 10^{-6}\,{\rm K} bvn ). The radius of the fermion ball of mass MC10MM_{C}\simeq 10\,M_{\odot} resulting from the collapse of the gaseous halo is RC=0.974pcR_{C}=0.974\,{\rm pc}. The energy of the gaseous halo is Egas=4.12×1040ergE_{\rm gas}=-4.12\times 10^{40}\,{\rm erg} and the energy of the fermion ball is Emin=3.76×1042ergE_{\rm min}=-3.76\times 10^{42}\,{\rm erg}. The energy released during the collapse is ΔE=EgasEmin=3.72×1042erg\Delta E=E_{\rm gas}-E_{\rm min}=3.72\times 10^{42}\,{\rm erg}. The collapse time is a multiple of tD1.25Gyrst_{D}\sim 1.25\,{\rm Gyrs} which is quite long (possibly irrelevant).

Supermassive fermion ball (crude model) bvn : We consider a gas of DM fermions of total mass M=109MM=10^{9}\,M_{\odot}.424242In that case, general relativity should be taken into account (see below) since MMmaxM\sim M_{\rm max}. The canonical phase transition appears for RRCCPNR(N)=6.00×103pcR\geq R^{\rm NR}_{\rm CCP}(N)=6.00\times 10^{-3}\,{\rm pc}. For a gaseous halo of size R=1.68×102pcR=1.68\times 10^{-2}\,{\rm pc}, corresponding to μ=389\mu=389, the collapse temperature is Tc=2.26×105KT_{c}=2.26\times 10^{5}\,{\rm K} (the transition temperature is Tt=3.02×105KT_{t}=3.02\times 10^{5}\,{\rm K} bvn ). The radius of the fermion ball of mass MC109MM_{C}\simeq 10^{9}\,M_{\odot} resulting from the collapse of the gaseous halo is RC=2.10×103pcR_{C}=2.10\times 10^{-3}\,{\rm pc}. The energy of the gaseous halo is Egas=1.01×1060ergE_{\rm gas}=-1.01\times 10^{60}\,{\rm erg} and the energy of the fermion ball is Emin=1.74×1061ergE_{\rm min}=-1.74\times 10^{61}\,{\rm erg}. The energy released during the collapse is ΔE=EgasEmin=1.64×1061erg\Delta E=E_{\rm gas}-E_{\rm min}=1.64\times 10^{61}\,{\rm erg}. The collapse time is a multiple of tD1.03yrst_{D}\sim 1.03\,{\rm yrs}, which is very short on a cosmological timescale.

XIII.1.2 Microcanonical ensemble

We now consider a system of nonrelativistic self-gravitating fermions in the microcanonical ensemble. The microcanonical phase transition appears for μμMCP=2670\mu\geq\mu_{\rm MCP}=2670 (microcanonical critical point) ijmpb hence for

RRMCPNR(N)=5.22h2Gm8/3M1/3g2/3.R\geq R^{\rm NR}_{\rm MCP}(N)=5.22\frac{h^{2}}{Gm^{8/3}M^{1/3}g^{2/3}}. (45)

We assume that we are in this situation.434343When RRMCPNR(N)R\leq R^{\rm NR}_{\rm MCP}(N) the whole series of equilibria is stable. When RCCPNR(N)RRMCPNR(N)R^{\rm NR}_{\rm CCP}(N)\leq R\leq R^{\rm NR}_{\rm MCP}(N) the system evolves, as energy decreases, from the gaseous states to the core-halo states (with a negative specific heat) without instability or phase transition. In that case, the caloric curve has the form of Fig. 27. The natural evolution proceeds along the series of equilibria towards lower and lower energies (see Sec. III). We assume that the system is initially in the metastable gaseous phase. As its energy decreases it remains in this phase up to the critical point EcE_{c} at which the metastable gaseous branch disappears.444444We recall that the collapse takes place at the critical (spinodal) point EcE_{c}, not at the transition point EtE_{t}, because of the tremendously long lifetime of metastable gaseous states. This critical energy is not very sensitive on quantum effects (when μ1\mu\gg 1) so it can be approximated by its classical value Λc=EcR/GM2=0.335\Lambda_{c}=-E_{c}R/GM^{2}=0.335 yielding

Ec=0.335GM2R.E_{c}=-0.335\frac{GM^{2}}{R}. (46)

At that point, the system collapses and forms a compact fermion star. As explained in Appendix E.2.2, in the microcanonical ensemble, the fermion star contains only a fraction of the total mass (MC<MM_{C}<M) and is surrounded by a massive and very hot atmosphere. If we approximate the fermion star by a Fermi gas at T=0T=0 containing a mass MC=αCMM_{C}=\alpha_{C}M, its radius is given by (see Appendix E.2.2):

RC=0.181h2Gm8/3αC1/3M1/3g2/3.R_{C}=0.181\frac{h^{2}}{Gm^{8/3}\alpha_{C}^{1/3}M^{1/3}g^{2/3}}. (47)

On the other hand, the temperature of the halo in the condensed phase is given by (see Appendix E.2.2):

kBTcond=1.57αC7/31αCG2M4/3m11/3g2/3h2.k_{B}T_{\rm cond}=1.57\frac{\alpha_{C}^{7/3}}{1-\alpha_{C}}\frac{G^{2}M^{4/3}m^{11/3}g^{2/3}}{h^{2}}. (48)

From the analytical model developed in pt one finds that the fraction of mass in the core is approximately given by (see Appendix E.2.2):

αC74lnμ.\alpha_{C}\simeq\frac{7}{4\ln\mu}. (49)

In many applications, it is sufficient to consider that αC1/4\alpha_{C}\simeq 1/4 (see footnote 45). The temperature of the gaseous phase at the point of gravothermal catastrophe is not very sensitive on quantum effects (when μ1\mu\gg 1) so it can be approximated by its classical value ηgas=βgasGMm/R=2.03\eta_{\rm gas}=\beta_{\rm gas}GMm/R=2.03 yielding

kBTgas=0.493GMmR.k_{B}T_{\rm gas}=0.493\frac{GMm}{R}. (50)

The relaxation time depends on the physical process governing the dynamical evolution of the system so it will not be discussed here.

We note that the collapse in the microcanonical ensemble corresponds to an implosion of the core and a heating of the halo. In the box model, the atmosphere is held by the walls of the box. Without the box, it would be expelled at large distances (see Figs. 38 and 41 of clm2 for an illustration in the context of the fermionic King model). Therefore, in the microcanonical ensemble, the collapse leads to the formation of a fermion star and to the expulsion of a hot atmosphere. This core-halo structure is reminiscent of the onset of a red-giant before the white dwarf stage. The implosion of the core and the explosion of the halo is also similar to the supernova phenomenon leading to the formation of a neutron star. These ideas are further developed and illustrated in supernova . The thermodynamical reason for this implosion-explosion phenomenon is explained in Appendix E.1.2. It is also corroborated by the following arguments. If we calculate the density perturbation δρ\delta\rho that triggers the instability at EcE_{c}, we find that it has two nodes, corresponding to a core-halo structure (see Fig. 6b of paddyapj ). Therefore, the instability develops itself in such a way that the density increases in the core and in the halo while it decreases between them (the intermediate shell is depopulated). We also find that the velocity profile δv\delta v has one node (see Fig. 4.b of supernova ) so the velocity is directed inward in the core and outward in the halo. These two results confirm the implosion-explosion phenomenon similar to the red-giant structure and to the supernova phenomenon previously described.

We make below numerical applications to illustrate the preceding results.454545In the numerical applications, we have chosen the values of MM and RR in order to have μ=105\mu=10^{5}. From Fig. 21 of ijmpb we find that ηcond=0.290\eta_{\rm cond}=0.290. From the relation ηcond7(1αC)/(2λμ2/3αC7/3)\eta_{\rm cond}\sim 7(1-\alpha_{C})/(2\lambda\mu^{2/3}\alpha_{C}^{7/3}) with λ=0.149736\lambda=0.149736..., equivalent to Eq. (48), we find that αC=0.220\alpha_{C}=0.220. This can be compared to the approximate value αC0.125\alpha_{C}\simeq 0.125 obtained from Eq. (49). The agreement is reasonable in view of the numerous approximations and the logarithmic corrections.

Neutron stars (crude model) ht : We consider a gas of neutrons of total mass M=1MM=1\,M_{\odot}. The microcanonical phase transition appears for RRMCPNR(N)=435kmR\geq R^{\rm NR}_{\rm MCP}(N)=435\,{\rm km}. For a gaseous star of size R=4875kmR=4875\,{\rm km} the collapse energy is Ec=1.81×1050ergE_{c}=-1.81\times 10^{50}\,{\rm erg} (the transition energy is Et=1.51×1050ergE_{t}=1.51\times 10^{50}\,{\rm erg}). The mass of the neutron star resulting from the collapse of the gaseous star is MC=0.220MM_{C}=0.220M_{\odot} and its radius is RC=25.0kmR_{C}=25.0\,{\rm km}. The temperature of the gaseous star is Tgas=1.63×109KT_{\rm gas}=1.63\times 10^{9}\,{\rm K} and the temperature of the halo surrounding the neutron star is Tcond=1.14×1010KT_{\rm cond}=1.14\times 10^{10}\,{\rm K}.

Fermion ball bvn : We consider a gas of DM fermions of total mass M=10MM=10\,M_{\odot}. The microcanonical phase transition appears for RRMCPNR(N)=28.1pcR\geq R^{\rm NR}_{\rm MCP}(N)=28.1\,{\rm pc}. For a gaseous halo of size R=315pcR=315\,{\rm pc} the collapse energy is Ec=9.10×1039ergE_{c}=-9.10\times 10^{39}\,{\rm erg} (the transition energy is Et=7.61×1039ergE_{t}=7.61\times 10^{39}\,{\rm erg}). The mass of the fermion ball resulting from the collapse of the gaseous halo is MC=2.20MM_{C}=2.20M_{\odot} and its radius is RC=1.61pcR_{C}=1.61\,{\rm pc}. The temperature of the gaseous halo is Tgas=1.50×107KT_{\rm gas}=1.50\times 10^{-7}\,{\rm K} and the temperature of the halo surrounding the fermion ball is Tcond=1.05×106KT_{\rm cond}=1.05\times 10^{-6}\,{\rm K}.

Supermassive fermion ball (crude model) bvn : We consider a gas of DM fermions of total mass M=109MM=10^{9}\,M_{\odot}. The microcanonical phase transition appears for RRMCPNR(N)=6.06×102pcR\geq R^{\rm NR}_{\rm MCP}(N)=6.06\times 10^{-2}\,{\rm pc}. For a gaseous halo of size R=0.679pcR=0.679\,{\rm pc} the collapse energy is Ec=4.22×1058ergE_{c}=-4.22\times 10^{58}\,{\rm erg} (the transition energy is Et=3.53×1058ergE_{t}=3.53\times 10^{58}\,{\rm erg}). The mass of the fermion ball resulting from the collapse of the gaseous halo is MC=2.20×108MM_{C}=2.20\times 10^{8}\,M_{\odot} and its radius is RC=3.48×103pcR_{C}=3.48\times 10^{-3}\,{\rm pc}. The temperature of the gaseous halo is Tgas=6.94×103KT_{\rm gas}=6.94\times 10^{3}\,{\rm K} and the temperature of the halo surrounding the fermion ball is Tcond=4.86×104KT_{\rm cond}=4.86\times 10^{4}\,{\rm K}.

XIII.2 Relativistic model with N<NOVN<N_{\rm OV}: Collapse towards a fermion star

We now consider the truly general relativistic Fermi gas model. We first assume that N<NOVN<N_{\rm OV} so that the collapse always leads to a fermion star (not a black hole). The discussion is essentially the same as before. However, we make below new numerical applications to see the effect of relativistic corrections when NNOVN\lesssim N_{\rm OV}. These numerical applications are based on the general relativistic models of dark matter halos studied by Bilic and Viollier bvr as recalled in the Introduction.464646In terms of the dimensionless variables defined in Appendix B they take N=0.38N=0.38, which is slightly below the OV limit NOV=0.39853N_{\rm OV}=0.39853, and R=100R=100. The radius of the completely degenerate fermion ball corresponding to N=0.38N=0.38 is RC=4.095=1.22ROVR_{C}=4.095=1.22R_{\rm OV} paper1 . We consider their adaptation to the case of neutron stars. We restrict ourselves to the canonical ensemble.

Supermassive fermion ball bvr : For a fermionic particle of mass m=17.2keV/c2m=17.2\,{\rm keV/c^{2}} the OV limits are NOV=1.4254×1071N_{\rm OV}=1.4254\times 10^{71}, NOVm=2.1973×109MN_{\rm OV}m=2.1973\times 10^{9}\,M_{\odot}, MOV=2.1186×109MM_{\rm OV}=2.1186\times 10^{9}\,M_{\odot} and ROV=8.8859×104pcR_{\rm OV}=8.8859\times 10^{-4}\,{\rm pc}. We consider a gas of N=0.95350NOVN=0.95350\,N_{\rm OV} fermions, corresponding to a rest mass Nm=2.0951×109MNm=2.0951\times 10^{9}\,M_{\odot}. According to Fig. 47, the canonical phase transition appears for RRCCP(N)=3.93ROV=3.4818×103pcR\geq R_{\rm CCP}(N)=3.93R_{\rm OV}=3.4818\times 10^{-3}\,{\rm pc} (the nonrelativistic value is RCCPNR(N)=4.6894×103pcR_{\rm CCP}^{\rm NR}(N)=4.6894\times 10^{-3}\,{\rm pc}). For a system of initial size R=29.789ROV=2.6391×102pcR=29.789\,R_{\rm OV}=2.6391\times 10^{-2}\,{\rm pc}, the collapse temperature is Tc=3.0112×105KT_{c}=3.0112\times 10^{5}\,{\rm K} (the transition temperature is Tt=0.0043951mc2=8.7725×105KT_{t}=0.0043951\,mc^{2}=8.7725\times 10^{5}\,{\rm K} bvr ). The collapse of the gaseous halo leads to a supermassive fermion ball containing almost all the particles surrounded by a tenuous atmosphere. Since the particle number is slightly below the OV limit, the fermion ball is strongly relativistic. If we approximate the fermion ball by a Fermi gas at T=0T=0 containing all the rest mass Nm2.0951×109MNm\sim 2.0951\times 10^{9}\,M_{\odot}, we find a radius RC=1.220ROV=1.0809×103pcR_{C}=1.220\,R_{\rm OV}=1.0809\times 10^{-3}\,{\rm pc} and a mass MC=0.9577MOV=2.0290×109MM_{C}=0.9577\,M_{\rm OV}=2.0290\times 10^{9}\,M_{\odot} paper1 (the nonrelativistic values are RCNR=1.6399×103pcR_{C}^{\rm NR}=1.6399\times 10^{-3}\,{\rm pc} and MCNR=2.0951×109MM_{C}^{\rm NR}=2.0951\times 10^{9}\,M_{\odot}). The energy of the gaseous halo is Egas=2.8324×1060ergE_{\rm gas}=-2.8324\times 10^{60}\,{\rm erg} and the energy of the fermion ball is Emin=(MCNm)c2=1.1822×1062ergE_{\rm min}=(M_{C}-Nm)c^{2}=-1.1822\times 10^{62}\,{\rm erg} (the nonrelativistic value is EminNR=9.7925×1061ergE_{\rm min}^{\rm NR}=-9.7925\times 10^{61}\,{\rm erg}). The energy released during the collapse is ΔE=EgasEmin=1.1539×1062erg\Delta E=E_{\rm gas}-E_{\rm min}=1.1539\times 10^{62}\,{\rm erg} (the nonrelativistic value is ΔENR=9.5092×1061erg\Delta E^{\rm NR}=9.5092\times 10^{61}\,{\rm erg}). The collapse time is a multiple of tD1.3973yrst_{D}\sim 1.3973\,{\rm yrs}.

Neutron stars: For the neutrons of mass mn=0.940GeV/c2m_{n}=0.940\,{\rm GeV/c^{2}}, the OV limits are NOV=8.7448×1056N_{\rm OV}=8.7448\times 10^{56}, NOVm=0.73636MN_{\rm OV}m=0.73636\,M_{\odot}, MOV=0.71000MM_{\rm OV}=0.71000\,M_{\odot} and ROV=9.1614kmR_{\rm OV}=9.1614\,{\rm km}. We consider a gas of N=0.95350NOVN=0.95350\,N_{\rm OV} neutrons, corresponding to a rest mass Nm=0.70212MNm=0.70212\,M_{\odot}. The canonical phase transition appears for RRCCP(N)=3.93ROV=36.0kmR\geq R_{\rm CCP}(N)=3.93R_{\rm OV}=36.0\,{\rm km} (the nonrelativistic value is RCCPNR(N)=48.5kmR_{\rm CCP}^{\rm NR}(N)=48.5\,{\rm km}). For a system of initial size R=29.789ROV=272.91kmR=29.789\,R_{\rm OV}=272.91\,{\rm km}, the collapse temperature is Tc=1.6449×1010KT_{c}=1.6449\times 10^{10}\,{\rm K} (the transition temperature is Tt=0.0043951mc2=4.7921×1010KT_{t}=0.0043951\,mc^{2}=4.7921\times 10^{10}\,{\rm K}). The collapse of the gaseous star leads to a neutron star containing almost all the particles surrounded by a tenuous atmosphere. Since the particle number is slightly below the OV limit, the system is strongly relativistic. If we approximate the neutron star by a Fermi gas at T=0T=0 containing all the rest mass Nm0.70212MNm\sim 0.70212\,M_{\odot}, we find a radius RC=1.220ROV=11.177kmR_{C}=1.220R_{\rm OV}=11.177\,{\rm km} and a mass MC=0.9577MOV=0.67996MM_{C}=0.9577\,M_{\rm OV}=0.67996\,M_{\odot} (the nonrelativistic values are RCNR=16.957kmR_{C}^{\rm NR}=16.957\,{\rm km} and MCNR=0.70212MM_{C}^{\rm NR}=0.70212\,M_{\odot}). The energy of the gaseous star is Egas=9.4919×1050ergE_{\rm gas}=-9.4919\times 10^{50}\,{\rm erg} and the energy of the neutron star is Emin=(MCNm)c2=3.9618×1052ergE_{\rm min}=(M_{C}-Nm)c^{2}=-3.9618\times 10^{52}\,{\rm erg} (the nonrelativistic value is EminNR=3.28167×1052ergE_{\rm min}^{\rm NR}=-3.28167\times 10^{52}\,{\rm erg}). The energy released during the collapse is ΔE=EgasEmin=3.8669×1052erg\Delta E=E_{\rm gas}-E_{\rm min}=3.8669\times 10^{52}\,{\rm erg} (the nonrelativistic value is ΔENR=3.1867×1052erg\Delta E^{\rm NR}=3.1867\times 10^{52}\,{\rm erg}). The collapse time is a multiple of tD1.4767×102st_{D}\sim 1.4767\times 10^{-2}\,{\rm s}.

Remark: For the value of NN considered in the previous examples, we find from Fig. 47 that the microcanonical phase transition appears for RRMCP(N)=52.4ROVR\geq R_{\rm MCP}(N)=52.4\,R_{\rm OV} (the nonrelativistic value is RMCPNR(N)=53.4ROVR_{\rm MCP}^{\rm NR}(N)=53.4\,R_{\rm OV}). Since RR is below this critical value, the system does not display any phase transition in the microcanonical ensemble.

XIII.3 Relativistic model with NOV<N<NcN_{\rm OV}<N<N_{c}: Collapse towards a fermion star followed by a collapse towards a black hole

We now assume that NOV<N<NcN_{\rm OV}<N<N_{c} (where NcCE=NeN_{c}^{\rm CE}=N_{e} and NcMCE=NfN_{c}^{\rm MCE}=N_{f}) so that, by cooling, the system undergoes two successive collapses: a collapse towards a fermion star followed by a collapse towards a black hole.

We consider a system of relativistic self-gravitating fermions in the canonical ensemble. We assume that NOV<N<Ne(R)N_{\rm OV}<N<N_{e}(R). The caloric curve has the form of Fig. 9. We assume that the system is initially in the gaseous phase. At its temperature decreases, the system collapses from the gaseous phase to the condensed phase at TcT_{c} then undergoes a catastrophic collapse from the condensed phase to a black hole at TcT^{\prime}_{c}. We note that the interval (ΔN)CE=Ne(R)NOV(\Delta N)_{\rm CE}=N_{e}(R)-N_{\rm OV} is extremely narrow since NOV=0.39853N_{\rm OV}=0.39853 and Ne=0.40002N_{e}=0.40002 for R=50R=50 (we see in Fig. 47 that Ne(R)N_{e}(R) does not change much with RR). We have (ΔN)CE/NOV=3.81×1031(\Delta N)_{\rm CE}/N_{\rm OV}=3.81\times 10^{-3}\ll 1 so that

NcCENOV.N_{c}^{\rm CE}\simeq N_{\rm OV}. (51)

The reason why (ΔN)CE/NOV1(\Delta N)_{\rm CE}/N_{\rm OV}\ll 1 is easy to understand. We have previously explained that the fermion star contains almost all the particles (NCNN_{C}\sim N). Therefore, as soon as NN is larger than NOVN_{\rm OV} the fermion star becomes unstable (NC>NOVN_{C}>N_{\rm OV}) and collapses towards a black hole (see Appendix G).

We consider a system of relativistic self-gravitating fermions in the microcanonical ensemble. We assume that NOV<N<Nf(R)N_{\rm OV}<N<N_{f}(R). The caloric curve has the form of Fig. 29. We assume that the system is initially in the gaseous phase. At its energy decreases, the system collapses from the gaseous phase to the condensed phase at EcE_{c} then undergoes a catastrophic collapse from the condensed phase to a black hole at EcE^{\prime}_{c}. The interval (ΔN)MCE=Nf(R)NOV(\Delta N)_{\rm MCE}=N_{f}(R)-N_{\rm OV} is much larger than in the canonical ensemble since NOV=0.39853N_{\rm OV}=0.39853 and Nf1.4854N_{f}\simeq 1.4854 for R=600R=600 (we see in Fig. 47 that Nf(R)N_{f}(R) remains in the range 121-2). We have (ΔN)MCE/NOV=2.73(\Delta N)_{\rm MCE}/N_{\rm OV}=2.73 so that

NcMCE3.73NOV.N^{\rm MCE}_{c}\simeq 3.73\,N_{\rm OV}. (52)

Again, the reason why (ΔN)MCE/NOV1(\Delta N)_{\rm MCE}/N_{\rm OV}\sim 1 is easy to understand. We have previously explained that the fermion star contains only a fraction of the particles (NCN/4N_{C}\sim N/4). Therefore, if NN is only slightly larger than NOVN_{\rm OV}, the fermion ball is stable (NC<NOVN_{C}<N_{\rm OV}). It is only when NN is substantially larger than NOVN_{\rm OV} (by a factor of 4\sim 4) that the fermion ball becomes unstable (NC>NOVN_{C}>N_{\rm OV}) and collapses towards a black hole.

XIII.4 Relativistic model with N>NcN>N_{c}: Direct collapse towards a black hole

We finally assume that N>NcN>N_{c} so that, by cooling, the system directly collapses towards a black hole, without forming a fermion star.

We consider a system of relativistic self-gravitating fermions in the canonical ensemble. We assume that N>Ne(R)N>N_{e}(R). The caloric curve has the form of Figs. 16, 18, 20, 21, and 22. We assume that the system is initially in the gaseous phase. As its temperature decreases, the system undergoes a catastrophic collapse from the gaseous phase to a black hole at TcT_{c}. This situation corresponds to N>NcCEN>N_{c}^{\rm CE}, where NcCENOVN_{c}^{\rm CE}\simeq N_{\rm OV} [see Eq. (51)]. This result shows that, in the canonical ensemble, there is no condensed configurations in the general relativisitic Fermi gas at nonzero temperature as soon as NN is slightly larger than NOVN_{\rm OV}. The reason is that almost of all the particles are in the degenerate core, the rest forming a tenuous isothermal atmosphere.

We consider a system of relativistic self-gravitating fermions in the microcanonical ensemble. We assume that N>Nf(R)N>N_{f}(R). The caloric curve has the form of Figs. 32, 34 and 36. We assume that the system is initially in the gaseous phase. At its energy decreases, the system undergoes a catastrophic collapse from the gaseous phase to a black hole at EcE_{c}. This situation corresponds to N>NcMCEN>N_{c}^{\rm MCE}, where NcMCE3.73NOVN_{c}^{\rm MCE}\simeq 3.73\,N_{\rm OV} [see Eq. (52)]. This result shows that, in the microcanonical ensemble, there exist condensed configurations in the general relativisitic Fermi gas at nonzero temperature with NOV<N<3.73NOVN_{\rm OV}<N<3.73\,N_{\rm OV}. The reason is that only about 1/41/4 of the particles are in the degenerate core (so that NC<NOVN_{C}<N_{\rm OV}), the rest forming an isothermal atmosphere.

XIII.5 Summary

In this section, we summarize the possible evolution of a gaseous star when its temperature (canonical ensemble) or its energy (microcanonical ensemble) is reduced below a critical value.

XIII.5.1 Canonical ensemble

When N<NOVN<N_{\rm OV} and Rmin(N)<R<RCCP(N)R_{\rm min}(N)<R<R_{\rm CCP}(N), there is no collapse.

When N<NOVN<N_{\rm OV} and R>RCCP(N)R>R_{\rm CCP}(N), the gaseous star collapses below TcT_{c} towards a fermion star (white dwarf, neutron star, DM fermion ball). The fermion star contains almost all the mass and is surrounded by a very tenuous atmosphere. Therefore, the collapse corresponds just to an implosion.

When NOV<N<NcCENOVN_{\rm OV}<N<N^{\rm CE}_{c}\simeq N_{\rm OV} and R>RCCP(N)R>R_{\rm CCP}(N) the gaseous star first collapses below TcT_{c} toward a fermion star, then the fermion star collapses below TcT^{\prime}_{c} towards a black hole.

When NOV<N<NcCENOVN_{\rm OV}<N<N^{\rm CE}_{c}\simeq N_{\rm OV} and Rmin(N)<R<RCCP(N)R_{\rm min}(N)<R<R_{\rm CCP}(N) or when N>NcCENOVN>N^{\rm CE}_{c}\simeq N_{\rm OV} the gaseous star directly collapses towards a black hole.

XIII.5.2 Microcanonical ensemble

When N<NOVN<N_{\rm OV} and Rmin(N)<R<RMCP(N)R_{\rm min}(N)<R<R_{\rm MCP}(N), there is no collapse.

When N<NOVN<N_{\rm OV} and R>RMCP(N)R>R_{\rm MCP}(N), the gaseous star collapses below EcE_{c} towards a fermion star (white dwarf, neutron star, DM fermion ball). The fermion star contains only a fraction of the total mass and is surrounded by a very hot atmosphere. Therefore, the system has a core-halo structure which is similar to the onset of red-giants or to the supernova phenomenon. This core-halo structure results from an implosion of the core and an explosion of the halo.

When NOV<N<NcMCE3.73NOVN_{\rm OV}<N<N^{\rm MCE}_{c}\simeq 3.73\,N_{\rm OV} and R>RMCP(N)R>R_{\rm MCP}(N) the gaseous star first collapses below EcE_{c} toward a fermion star ++ halo, then the fermion star collapses below EcE^{\prime}_{c} towards a black hole.

When NOV<N<NcMCE3.73NOVN_{\rm OV}<N<N^{\rm MCE}_{c}\simeq 3.73\,N_{\rm OV} and Rmin(N)<R<RMCP(N)R_{\rm min}(N)<R<R_{\rm MCP}(N) or when N>NcMCE3.73NOVN>N^{\rm MCE}_{c}\simeq 3.73\,N_{\rm OV} the gaseous star directly collapses towards a black hole.

XIV Conclusion

In this paper we have studied the nature of phase transitions in the general relativistic Fermi gas. This is the most general situation that we can imagine since both quantum and relativistic effects are taken into account in a rigorous manner. We have obtained the complete phase diagram of the system in the (R,N)(R,N) plane (see Fig. 47).

When N<NOVN<N_{\rm OV}, the results are similar to those obtained in the nonrelativistic limit (recovered for NNOVN\ll N_{\rm OV} and RROVR\gg R_{\rm OV} with NR3NR^{3} fixed) ijmpb . In that case, there exists an equilibrium state for any value of the temperature T0T\geq 0 and any value of the energy EEminE\geq E_{\rm min}. Catastrophic collapse towards a black hole is prevented by quantum mechanics (Pauli’s exclusion principle). Small systems (Rmin<R<RCCPR_{\rm min}<R<R_{\rm CCP}) do not experience phase transition. Intermediate size systems (RCCP<R<RMCPR_{\rm CCP}<R<R_{\rm MCP}) experience a canonical phase transition. Large systems (R>RMCPR>R_{\rm MCP}) experience canonical and microcanonical phase transitions. A zeroth order phase transition takes place, below a critical temperature TcT_{c} or below a critical energy EcE_{c}, from a gaseous phase to a condensed phase. The gaseous phase corresponds to a radiative star, a molecular cloud or a primordial DM nebula. The condensed phase corresponds to a compact object (fermion star) such as a white dwarf, a neutron star or a DM fermion ball. In the canonical ensemble the fermion star contains almost all the mass (or is surrounded by a tenuous isothermal atmosphere). The phase transition corresponds to an implosion. In the microcanonial ensemble the compact object contains only a fraction of the total mass (1/4\sim 1/4 to fix the ideas) and is surrounded by a hot isothermal atmosphere that contains the remaining mass. Therefore, the condensed phase has a core-halo structure. In the box model the atmosphere is held by the walls of the box. Without the box it would be expelled at large distances. The phase transition corresponds to an an implosion of the core and an explosion of the halo.

When Nc<N<NmaxN_{c}<N<N_{\rm max} (where NcNOVN_{c}\simeq N_{\rm OV} in the canonical ensemble and Nc3.73NOVN_{c}\simeq 3.73\,N_{\rm OV} in the microcanonical ensemble), the results are similar to those obtained in the classical limit (recovered for NNOVN\gg N_{\rm OV} and RROVR\gg R_{\rm OV} with N/RN/R fixed) roupas ; paper2 . In that case, there is no condensed phase. Below a critical temperature TcT_{c} or below a critical energy EcE_{c}, the system undergoes a catastrophic collapse from the gaseous phase to a black hole (presumably). Indeed, quantum mechanics cannot prevent this singularity.

When NOV<N<NcN_{\rm OV}<N<N_{c}, the results are new and more complex because the system is both relativistic and quantum. In that case, the system can experience two successive collapses: a zeroth order phase transition at TcT_{c} (when R>RCCPR>R_{\rm CCP}) or EcE_{c} (when R>RMCPR>R_{\rm MCP}) from the gaseous phase to the condensed phase (representing a white dwarf, a neutron star, or a DM fermion ball) followed by a catastrophic collapse at TcT^{\prime}_{c} or EcE^{\prime}_{c} from the condensed phase to a black hole (presumably). This behavior occurs in a very narrow range of parameters in the canonical ensemble (NcNOVN_{c}\simeq N_{\rm OV}) and in a wider range of parameters (Nc3.73NOVN_{c}\simeq 3.73\,N_{\rm OV}) in the microcanonical ensemble.

The previous results apply to mid and low values of energy and temperature. At very high energies and very high temperatures, the system is ultrarelativistic and behaves like a form of self-gravitating radiation. Above a maximum energy EmaxE_{\rm max} or above a maximum temperature TmaxT_{\rm max}, there is no equilibrium state and the system is expected to collapse towards a black hole (presumably) roupas ; paper2 .

The astrophysical applications of our results remain limited by the introduction of an artificial confining box. This is necessary in order to have isothermal equilibrum states with a finite mass and thus investigate phase transitions rigorously. A more astrophysically relevant model with a finite mass is provided by the general relativistic fermionic King model. Phase transitions in the nonrelativistic fermionic King model have been studied in detail in clm1 ; clm2 and give results that are qualitatively similar to those obtained with the box model ijmpb . We believe that similar results would be obtained with the general relativistic fermionic King model.

We have suggested that the microcanonial phase transitions occurring in the self-gravitating Fermi gas may be related to the onset of red-giant structure or to the supernova phenomenon. In these spectacular events, the collapse of the core of the system (resulting ultimately in the formation of a white dwarf or a neutron star) is accompanied by the explosion and the expulsion of a hot envelope.474747Newtonian gravity is sufficient to describe white dwarfs (and the planetary nebula) that follow the red-giant stage while general relativity is necessary to describe neutron stars or black holes formed from supernova explosion. Similarly, in the self-gravitating Fermi gas, the microcanonical phase transition from a gaseous state to a condensed state corresponds to the implosion of the core (leading to a fermion star) and the explosion of the halo. These analogies are further developed in supernova . We may wonder if similar phenomena can occur in the context of DM as suggested in clm2 . It would also be interesting to develop dynamical models of gravitational collapse towards a black hole when no equilibrium state is possible to ensure that the system really forms a black hole. These topics will be considered in future works.

Acknowledgement: One of us (PHC) would like to dedicate this paper to the memory of Donald Lynden-Bell (1935-2018) who was a pioneer in the statistical mechanics of self-gravitating systems.

Appendix A Thermodynamic limit

The thermodynamic limit of nonrelativistic self-gravitating fermions corresponds to N+N\rightarrow+\infty in such a way that η=βGMm/R\eta=\beta GMm/R, Λ=ER/GM2\Lambda=-ER/GM^{2} and μ=(gm4/h3)512π4G3NmR3\mu=(gm^{4}/h^{3})\sqrt{512\pi^{4}G^{3}NmR^{3}} are O(1)O(1). Taking mhG1m\sim h\sim G\sim 1, this corresponds to N+N\rightarrow+\infty with RN1/3R\sim N^{-1/3}, TN4/3T\sim N^{4/3}, EN7/3E\sim N^{7/3}, SNS\sim N, and FN7/3F\sim N^{7/3}. These scalings are given in htf ; ijmpb . Taking mhN/V1m\sim h\sim N/V\sim 1 (with VR3V\sim R^{3}), this corresponds to N+N\rightarrow+\infty with RN1/3R\sim N^{1/3}, GN2/3G\sim N^{-2/3}, T1T\sim 1, ENE\sim N, SNS\sim N, and FNF\sim N.

The thermodynamic limit of relativistic classical self-gravitating systems corresponds to N+N\rightarrow+\infty in such a way that η=βGNm2/R\eta=\beta GNm^{2}/R, Λ=ER/GN2m2\Lambda=-ER/GN^{2}m^{2} (with E=(MNm)c2E=(M-Nm)c^{2}) and ν=GNm/Rc2\nu=GNm/Rc^{2} are O(1)O(1). Taking mcG1m\sim c\sim G\sim 1 this corresponds to N+N\rightarrow+\infty with RNR\sim N, T1T\sim 1, ENE\sim N, SNS\sim N, and FNF\sim N. To the best of our knowledge, these scalings have not been given previously. Taking mcN/V1m\sim c\sim N/V\sim 1 (with VR3V\sim R^{3}) this corresponds to N+N\rightarrow+\infty with RN1/3R\sim N^{1/3}, GN2/3G\sim N^{-2/3}, T1T\sim 1, ENE\sim N, SNS\sim N, and FNF\sim N.

The thermodynamic limit of relativistic self-gravitating fermions corresponds to N+N\rightarrow+\infty in such a way that η=βGNm2/R\eta=\beta GNm^{2}/R, Λ=ER/GN2m2\Lambda=-ER/GN^{2}m^{2} (with E=(MNm)c2E=(M-Nm)c^{2}), μ=(gm4/h3)512π4G3NmR3\mu=(gm^{4}/h^{3})\sqrt{512\pi^{4}G^{3}NmR^{3}} and ν=GNm/Rc2\nu=GNm/Rc^{2} are O(1)O(1). Taking hcG1h\sim c\sim G\sim 1 this corresponds to N+N\rightarrow+\infty with RN2/3R\sim N^{2/3}, mN1/3m\sim N^{-1/3}, TN1/3T\sim N^{-1/3}, EN2/3E\sim N^{2/3}, SNS\sim N, and FN2/3F\sim N^{2/3}. These scalings were given in bvr . Taking mchN/V1m\sim c\sim h\sim N/V\sim 1 (with VR3V\sim R^{3}) this corresponds to N+N\rightarrow+\infty with RN1/3R\sim N^{1/3}, GN2/3G\sim N^{-2/3}, T1T\sim 1, ENE\sim N, SNS\sim N, and FNF\sim N.

The thermodynamic limit of nonrelativistic classical self-gravitating systems is discussed in Appendix A of aakin where several possible scalings are given.

We note that the different situations considered above can be unified by considering a thermodynamic limit of the form N+N\rightarrow+\infty with RN1/3R\sim N^{1/3}, GN2/3G\sim N^{-2/3}, T1T\sim 1, ENE\sim N, SNS\sim N, and FNF\sim N, corresponding to mchN/V1m\sim c\sim h\sim N/V\sim 1 (with VR3V\sim R^{3}). This is the standard thermodynamic limit with a renormalized gravitational constant. To the best of our knowledge this result has not been highlighted previously.

Appendix B Dimensionless quantities

According to the OV theory ov , the maximum mass, the maximum particle number and the minimum radius of a general relativistic fermion star at T=0T=0 are484848Qualitatively, the scaling of the maximum mass MOV(c/G)3/2/m2M_{\rm OV}\sim({\hbar c}/{G})^{3/2}/{m^{2}} can be obtained from the mass-radius relation MR36/(m8G3)MR^{3}\sim\hbar^{6}/(m^{8}G^{3}) of nonrelativistic fermion stars (see Appendix F) by determining when the radius RR of the configuration becomes comparable to the Schwarzschild radius RS=2GM/c2R_{S}=2GM/c^{2}.

MOV=0.384262g(cG)3/21m2,M_{\rm OV}=0.38426\,\sqrt{\frac{2}{g}}\left(\frac{\hbar c}{G}\right)^{3/2}\frac{1}{m^{2}}, (53)
NOV=0.398532g(cG)3/21m3,N_{\rm OV}=0.39853\,\sqrt{\frac{2}{g}}\left(\frac{\hbar c}{G}\right)^{3/2}\frac{1}{m^{3}}, (54)
ROV=8.7360GMOVc2=3.35692g(3Gc)1/21m2.R_{\rm OV}=8.7360\,\frac{GM_{\rm OV}}{c^{2}}=3.3569\,\sqrt{\frac{2}{g}}\left(\frac{\hbar^{3}}{Gc}\right)^{1/2}\frac{1}{m^{2}}. (55)

We note that WOVGMOV2/ROVMOVc2W_{\rm OV}\sim GM_{\rm OV}^{2}/R_{\rm OV}\sim M_{\rm OV}c^{2}. We introduce the mass, particle number and length scales

M=2gMP3m2=(23c3gm4G3)1/2,M_{*}=\sqrt{\frac{2}{g}}\frac{M_{P}^{3}}{m^{2}}=\left(\frac{2\hbar^{3}c^{3}}{gm^{4}G^{3}}\right)^{1/2}, (56)
N=Mm=2gMP3m3=(23c3gm6G3)1/2,N_{*}=\frac{M_{*}}{m}=\sqrt{\frac{2}{g}}\frac{M_{P}^{3}}{m^{3}}=\left(\frac{2\hbar^{3}c^{3}}{gm^{6}G^{3}}\right)^{1/2}, (57)
R=2gMPm2c=2gMP2m2lP=(23gm4cG)1/2,R_{*}=\sqrt{\frac{2}{g}}\frac{\hbar M_{P}}{m^{2}c}=\sqrt{\frac{2}{g}}\frac{M_{P}^{2}}{m^{2}}l_{P}=\left(\frac{2\hbar^{3}}{gm^{4}cG}\right)^{1/2}, (58)

where MP=(c/G)1/2M_{P}=(\hbar c/G)^{1/2} is the Planck mass and lP=(G/c3)1/2l_{P}=(\hbar G/c^{3})^{1/2} is the Planck length. We then define

r=Rr~,M=MM~,ϵ=Mc2R3ϵ~,r=R_{*}{\tilde{r}},\qquad M=M_{*}{\tilde{M}},\qquad\epsilon=\frac{M_{*}c^{2}}{R_{*}^{3}}{\tilde{\epsilon}}, (59)
N=NN~,n=NR3n~,Φ=c2Φ~,N=N_{*}{\tilde{N}},\qquad n=\frac{N_{*}}{R_{*}^{3}}{\tilde{n}},\qquad\Phi=c^{2}{\tilde{\Phi}}, (60)
P=Mc2R3P~,T=mc2kBT~,μ=mc2μ~,P=\frac{M_{*}c^{2}}{R_{*}^{3}}{\tilde{P}},\qquad T=\frac{mc^{2}}{k_{B}}{\tilde{T}},\qquad\mu=mc^{2}{\tilde{\mu}}, (61)
S=NkBS~,F=Mc2F~,S=N_{*}k_{B}{\tilde{S}},\qquad F=M_{*}c^{2}{\tilde{F}}, (62)

where the tilde variables are dimensionless. We note that

Mc2R3=gm4c523,NR3=gm3c323,\frac{M_{*}c^{2}}{R_{*}^{3}}=\frac{gm^{4}c^{5}}{2\hbar^{3}},\qquad\frac{N_{*}}{R_{*}^{3}}=\frac{gm^{3}c^{3}}{2\hbar^{3}}, (63)
NR3=46g2m9G3,GMRc2=1.N_{*}R_{*}^{3}=\frac{4\hbar^{6}}{g^{2}m^{9}G^{3}},\qquad\frac{GM_{*}}{R_{*}c^{2}}=1. (64)

In the main text, in order to simplify the notations, we do not write the tildes anymore. This amounts to taking =c=G=m=g/2=1\hbar=c=G=m=g/2=1 in the dimensional expressions. In particular, we have

MOV=0.38426,M_{\rm OV}=0.38426, (65)
NOV=0.39853,N_{\rm OV}=0.39853, (66)
ROV=3.3569.R_{\rm OV}=3.3569. (67)

Appendix C Connexion between the Nα(Φ0)N_{\alpha}(\Phi_{0}) curves and the caloric curves η(Λ)\eta(\Lambda)

In this Appendix, we explain how we obtained the caloric curves of self-gravitating fermions in general relativity following the method given by Bilic and Viollier bvr .494949This Appendix is technical and can be skipped for a first reading. It is nevertheless important to understand where the critical values of NN (such as N1N_{1}, NOVN_{\rm OV}, NCCPN_{\rm CCP}, NN_{*} etc) obtained in the main text come from.

Let us illustrate this procedure with a simple example. In order to construct the caloric curve η(Λ)\eta(\Lambda) corresponding to R=50R=50 and N=0.15N=0.15 (see Fig. 5), we proceed as follows. For a given value of α\alpha and Φ0\Phi_{0}, we can solve the differential equations (6) and (7) up to r=Rr=R and determine NN from Eq. (10) [we can also determine η\eta and Λ\Lambda with the aid of Eqs. (9), (10), (12) and (16)]. By varying the central potential Φ0\Phi_{0} from 1-1 to ++\infty, we can obtain the curve Nα(Φ0)N_{\alpha}(\Phi_{0}). It is usually nonmonotonic and displays damped oscillations for large values of Φ0\Phi_{0}. As example is represented in Fig. 62 for R=50R=50 and α=750.24\alpha=750.24 (see also Figs. 65 and 66 below).

Refer to caption
Figure 62: Particle number NN as a function of the central potential Φ0\Phi_{0} for R=50R=50 and α=750.24\alpha=750.24. The intersections between the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) and the line level N=0.15N=0.15 determine three equilibrium states with central potentials (Φ0)1=0.9954(\Phi_{0})_{1}=0.9954, (Φ0)2=1.0069(\Phi_{0})_{2}=1.0069 and (Φ0)3=1.0495(\Phi_{0})_{3}=1.0495. Their corresponding energy and temperature are (Λ1,η1)=(0.0242,2.2725)(\Lambda_{1},\eta_{1})=(-0.0242,2.2725), (Λ2,η2)=(0.5067,2.2447)(\Lambda_{2},\eta_{2})=(0.5067,2.2447) and (Λ3,η3)=(2.4468,2.2794)(\Lambda_{3},\eta_{3})=(2.4468,2.2794). Each solution is represented by a bullet in the caloric curve of Fig. 64. The first solution (black) belongs to the gaseous phase, the second solution (red) belongs to the core-halo phase and the third solution (green) belongs to the condensed phase.
Refer to caption
Figure 63: Evolution of α(R)\alpha_{*}(R) as a function of RR. The dashed line represents the scaling law α(R)=5.01192lnR\alpha_{*}(R)=5.0119-2\ln R obtained in paper2 for classical systems. In the case of fermions, this law is asymptotically valid for R1R\gg 1. We have also indicated the radius Rt=12.255R_{t}=12.255 at which α(R)\alpha_{*}(R) passes from negative values to positive values as the box radius decreases.
Refer to caption
Figure 64: Caloric curve for R=50R=50 and N=0.15N=0.15. By varying α\alpha in Fig. 62, the first intersections (Λ1(α),η1(α))(\Lambda_{1}(\alpha),\eta_{1}(\alpha)) form the gaseous branch (black), the second intersections (Λ2(α),η2(α))(\Lambda_{2}(\alpha),\eta_{2}(\alpha)) form the core-halo branch (red) and the third intersections (Λ3(α),η3(α))(\Lambda_{3}(\alpha),\eta_{3}(\alpha)) form the condensed branch (green). It turns out that two branches merge at an extremum of temperature: ηc\eta_{c} corresponds to the merging of the first and second intersections (for α=αM\alpha=\alpha_{M}) and η\eta_{*} corresponds to the merging of the second and third intersections (for α=αm\alpha=\alpha_{m}).

Let us introduce some notations that will be useful in the following. We call 𝒩(α){\cal N}(\alpha) the maximum value of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) and we denote by Ψ(α)\Psi(\alpha) the value of the central potential Φ0\Phi_{0} corresponding to this maximum. By varying α\alpha from -\infty to ++\infty, we find that the peaks 𝒩(α){\cal N}(\alpha) of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} reach a maximum Nmax(R)=𝒩(α(R))N_{\rm max}(R)={\cal N}(\alpha_{*}(R)) at α=α(R)\alpha=\alpha_{*}(R) (see Figs. 65 and 66 below). The evolution of α\alpha_{*} as a function of RR is plotted in Fig. 63.

Let us come back to the curve of Fig. 62. The intersections (Φ0)i{1,,n}(\Phi_{0})_{i\in\{1,...,n\}} between the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) and the line level N=0.15N=0.15, and the corresponding values of Λ\Lambda and η\eta at these intersections, determine nn points in the caloric curve η(Λ)\eta(\Lambda) of Fig. 64. In the present exemple, n=3n=3. By varying α\alpha these points form nn branches in the caloric curve η(Λ)\eta(\Lambda). These branches have been represented in color in Fig. 64. In the present exemple, they correspond to the gaseous, core-halo and condensed phases respectively. We have observed (but not proven mathematically) that when α>α\alpha>\alpha_{*}, the merging of two intersections between the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} and the line level NN, occuring at some αe(N,R)\alpha_{e}(N,R), corresponds to the merging of two branches in the caloric curve η(Λ)\eta(\Lambda) occurring at an extremum of temperature ηe\eta_{e}.505050By using this result, it is easy to obtain the curves ηc(N)\eta_{c}(N), η(N)\eta_{*}(N) and ηc(N)\eta^{\prime}_{c}(N) in the canonical phase diagram of Fig. 23, since they correspond to α>α\alpha>\alpha_{*} (see below). Indeed, each extremum of Nα(Φ)N_{\alpha}(\Phi) determines an extremum inverse temperature ηe(N)\eta_{e}(N) for the corresponding value of N=N(α)N=N(\alpha). Thus, by considering the first three extrema (when they exist) and running α\alpha from α\alpha_{*} to ++\infty, we get the full curves ηc(N)\eta_{c}(N), η(N)\eta_{*}(N) and ηc(N)\eta^{\prime}_{c}(N). Unfortunately, it is not possible to use a similar method to obtain the curves Λc(N)\Lambda_{c}(N), Λ(N)\Lambda_{*}(N), Λc(N)\Lambda^{\prime}_{c}(N) and Λc′′(N)\Lambda^{\prime\prime}_{c}(N) in the microcanonical phase diagram of Fig. 37. They have to be obtained “by hand” directly from the caloric curves. We have also observed (but not proven mathematically) that when α<α\alpha<\alpha_{*}, the merging of two intersections between the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} and the line level NN, occuring at some αe(N,R)\alpha^{\prime}_{e}(N,R), corresponds to the merging of two branches in the caloric curve η(Λ)\eta(\Lambda) occurring at an extremum of energy Λe\Lambda_{e}.515151By using this result, it is easy to obtain the curve Λmin(N)\Lambda_{\rm min}(N) in the microcanonical phase diagram of Fig. 25, since it corresponds to α<α\alpha<\alpha_{*} (see below). Indeed, each extremum of Nα(Φ)N_{\alpha}(\Phi) determines an extremum energy Λe(N)\Lambda_{e}(N) for the corresponding value of N=N(α)N=N(\alpha). Thus, by considering the first extremum and running α\alpha from -\infty to α\alpha_{*}, we get the full curve Λmin(N)\Lambda_{\rm min}(N). Unfortunately, it is not possible to use a similar method to obtain the curve ηmin(N)\eta_{\rm min}(N) in the canonical phase diagram of Fig. 23. It has to be obtained “by hand” directly from the caloric curves. In the present example, we have α>α\alpha>\alpha_{*}. As a result the merging of the first and second intersections in Fig. 62, occurring at some αM\alpha_{M}, corresponds to the temperature minimum ηc\eta_{c} in the caloric curve of Fig. 64. Similarly, the merging of the second and third intersections in Fig. 62, occurring at some αm\alpha_{m}, corresponds to the temperature maximum η\eta_{*} in the caloric curve of Fig. 64 .

We now generalize this procedure to different values of NN and RR. The main difference between the case of classical particles studied in paper2 and the case of fermions studied in the present paper is the following. When quantum mechanics is taken into account, we find that for α+\alpha\rightarrow+\infty the {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} curves tend towards an invariant curve correponding to the OV-curve NOV(Φ0)N_{\rm OV}(\Phi_{0}) paper1 . Close to this curve, i.e. for α\alpha large and NN small, the {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} curves can have a complex behavior with several oscillations responsible for the phase transitions studied in this paper. The nature of these oscillations, and consequently the nature of phase transitions, depends on the particle number NN and on the radius of the system RR. We consider below the cases treated in the main text.

C.1 R=50R=50

In this subsection, we consider a system of size R=50R=50 and make the link between the topological properties of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} and the caloric curves analysed in Sec. V.

We first consider the case α<0\alpha<0 (see Fig. 65). For α\alpha\rightarrow-\infty, we find that 𝒩(α)0{\cal N}(\alpha)\rightarrow 0 and Ψ(α)+\Psi(\alpha)\rightarrow+\infty. This corresponds to the ultrarelativistic regime. As α\alpha increases, 𝒩(α){\cal N}(\alpha) increases and Ψ(α)\Psi(\alpha) decreases: the peak of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) grows and moves towards the left. At α=α=2.75\alpha=\alpha_{*}=-2.75, 𝒩(α){\cal N}(\alpha) reaches its maximum value Nmax=8.821N_{\rm max}=8.821. For larger values of α\alpha, 𝒩(α){\cal N}(\alpha) and Ψ(α)\Psi(\alpha) both decrease: the peak of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) decays and moves towards the left. For α0\alpha\rightarrow 0^{-}, 𝒩(α)N0=8.408{\cal N}(\alpha)\rightarrow N_{0}=8.408 and Ψ(α)1\Psi(\alpha)\rightarrow-1: the peak of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) is squeezed near Φ0=1\Phi_{0}=-1.525252Apart from this mathematical property, α=0\alpha=0 does not play a particular role in the problem. If we had plotted NN as a function of b0=1/T0b_{0}=1/T_{0} instead of Φ0\Phi_{0} [see Eq. (4)] the specificity of the value α=0\alpha=0 would not have arisen.

Refer to caption
Figure 65: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α<0\alpha<0 for R=50R=50 (for illustration the curves go from α=18.5\alpha=-18.5 to α=103\alpha=-10^{-3}). We have indicated different characteristics values of NN: N0=8.408N_{0}=8.408 and Nmax=8.821N_{\rm max}=8.821.

We now consider the case α>0\alpha>0 (see Fig. 66). For α>0\alpha>0, we find that 𝒩(α){\cal N}(\alpha) decreases and Ψ(α)\Psi(\alpha) increases: the peak of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) decays and moves towards the right. For α+\alpha\rightarrow+\infty, the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) tends towards the OV-curve NOV(Φ0)N_{\rm OV}(\Phi_{0}): 𝒩(α){\cal N}(\alpha) tends towards NOV=0.39853N_{\rm OV}=0.39853 and Ψ(α)\Psi(\alpha) tends towards 0.6950.695.

Refer to caption
Figure 66: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α>0\alpha>0 for R=50R=50 (for illustration the curves go from α=103\alpha=10^{-3} to α=901\alpha=901). The OV curve has been plotted in green. We have indicated different characteristics values of NN: NOV=0.39853N_{\rm OV}=0.39853, N0=8.408N_{0}=8.408 and Nmax=8.821N_{\rm max}=8.821.

The curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} with α<α\alpha<\alpha_{*} (i.e. the ones that go up as α\alpha increases) are associated with the hot spiral (radiation) studied in paper2 . The hot spiral corresponds to the ultrarelativistic limit valid for high values of the energy and of the temperature. This spiral is present for any value of NN, except for NN close to NmaxN_{\rm max} where the caloric curve presents a different behavior described in Sec. VI of paper2 and in Sec. IX.2 of the present paper.

The curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} with α>α\alpha>\alpha_{*} (i.e. the ones that go down as α\alpha increases) are associated with the part of the caloric curve corresponding to mid and low values of the energy and of the temperature. Figure 67 is a zoom of Fig. 66 close to the OV curve, i.e. for small values of NN. This is the region of interest where canonical phase transitions appear (they are related to the appearance of turning points of temperature). We have indicated on this figure the different characteristic values of NN that have been identified in Sec. V. They can be related to the topological properties of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} as follows:

(i) NOV=0.39853N_{\rm OV}=0.39853 and N1=0.18131N_{1}=0.18131 are intrinsic properties of the OV curve NOV(Φ0)N_{\rm OV}(\Phi_{0}). They correspond to its first maximum and to its first minimum.

(ii) NCCP=0.01697N_{\rm CCP}=0.01697 and N=0.41637N_{*}=0.41637 can be related to the first and second inflexion points of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} (see Figs. 68 and 71). Indeed, we have seen that an extremum of temperature in the caloric curve corresponds to a merging of two intersections in the N={Nα(Φ0)}N=\{N_{\alpha}(\Phi_{0})\} plots. As a result, the canonical phase transitions appear (at N=NCCPN=N_{\rm CCP}) and disappear (at N=NN=N_{*}) when the curve {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} presents an inflexion point. This is how we can precisely determine NCCPN_{\rm CCP} and NN_{*}.

Let us now describe in more detail the different intersections as a function of NN.

Refer to caption
Figure 67: Some curves Nα(Φ0)N_{\alpha}(\Phi_{0}) for R=50R=50 together with different characteristics values of NN: NCCP=0.01697N_{\rm CCP}=0.01697, N1=0.18131N_{1}=0.18131, NOV=0.39853N_{\rm OV}=0.39853, and N=0.41637N_{*}=0.41637.
Refer to caption
Figure 68: Illustration of the intersections in the case N<NCCP=0.01697N<N_{\rm CCP}=0.01697 (specifically R=50R=50 and N=0.012N=0.012).

For N<NCCPN<N_{\rm CCP} (see Fig. 68), we have just one intersection between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. This explains why the caloric curve of Fig. 4 is monotonic. We note that the intersection between the line level NN and the OV-curve (α+\alpha\rightarrow+\infty) corresponds to the ground state T=0T=0 (i.e. η+\eta\rightarrow+\infty). This leads to the vertical asymptote at Λ=Λmax\Lambda=\Lambda_{\rm max} in the caloric curve of Fig. 4. We note that, for N<NCCPN<N_{\rm CCP}, Φ01\Phi_{0}\ll 1 showing that we are in the Newtonian limit.

Refer to caption
Figure 69: Illustration of the intersections in the case NCCP=0.01697<N<N1=0.18131N_{\rm CCP}=0.01697<N<N_{1}=0.18131 (specifically R=50R=50 and N=0.15N=0.15).

For NCCP<N<N1N_{\rm CCP}<N<N_{1} (see Fig. 69), we are above the first inflexion point so we can have up to three intersections between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. This determines three branches in the caloric curve of Fig. 5. This is why it has an NN-shape. The first and second intersections merge at α=αM\alpha=\alpha_{M}. Correspondingly, the first and second branches in the caloric curve merge at ηc\eta_{c}, the first turning point of temperature. The second and third intersections merge at α=αm\alpha=\alpha_{m}. Correspondingly, the second and third branches in the caloric curve merge at η\eta_{*}, the second turing point of temperature.

Refer to caption
Figure 70: Illustration of the intersections in the case N1=0.18131<N<NOV=0.39853N_{1}=0.18131<N<N_{\rm OV}=0.39853 (specifically R=50R=50 and N=0.29N=0.29).

For N1<N<NOVN_{1}<N<N_{\rm OV} (see Fig. 70), the novelty is that there is a second intersection between the line level NN and the OV-curve NOV(Φ0)N_{\rm OV}(\Phi_{0}). This corresponds to an unstable equilibrium state at T=0T=0 (i.e. η+\eta\rightarrow+\infty). This gives rise to the second vertical asymptote at Λ=Λmax\Lambda=\Lambda^{\prime}_{\rm max} in the caloric curve of Fig. 7. There are also secondary intersections leading to the unstable spiral of Fig. 7 as discussed in the next paragraph.

Refer to caption
Figure 71: Illustration of the intersections in the case NOV=0.39853<N<N=0.41637N_{\rm OV}=0.39853<N<N_{*}=0.41637 (specifically R=50R=50 and N=0.408N=0.408).
Refer to caption
Figure 72: Illustration of the intersections in the case NOV=0.39853<N<N=0.41637N_{\rm OV}=0.39853<N<N_{*}=0.41637 (specifically R=50R=50 and N=0.408N=0.408).

For NOV<N<NN_{\rm OV}<N<N_{*} (see Figs. 71 and 72) we can have up to four fundamental intersections between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. This gives rise to four fundamental branches in the caloric curve of Fig. 16. The first and second intersections merge at α=αM\alpha=\alpha_{M}. Correspondingly, the first and second branches in the caloric curve merge at ηc\eta_{c}, the first turning point of temperature. The second and third intersections merge at α=αm\alpha=\alpha_{m}. Correspondingly, the second and third branches in the caloric curve merge at η\eta_{*}, the second turing point of temperature. The third and fourth intersections merge at α=αM\alpha=\alpha^{\prime}_{M}. Correspondingly, the third and fourth branches in the caloric curve merge at ηc\eta^{\prime}_{c}, the third turing point of temperature. Furthermore, there are additional intersections giving rise to the spiral (that will become the cold spiral for larger values of NN) in the caloric curve of Fig. 16. These intersections are less relevant since they correspond to unstable states. Note that there is no intersection with the OV curve so there is no vertical asymptote corresponding to η+\eta\rightarrow+\infty. Finally, we note that Φ01\Phi_{0}\sim 1 showing that we are in the relativistic regime.

Refer to caption
Figure 73: Illustration of the intersections in the case N>N=0.41637N>N_{*}=0.41637 (specifically R=50R=50 and N=0.45N=0.45).

For N>NN>N_{*} (see Fig. 73), we are above the second inflexion point so we can have at most two fundamental intersections between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. This determines two fundamental branches in the caloric curve of Fig. 20. The first and second intersections merge at α=αM\alpha=\alpha_{M}. Correspondingly, the first and second branches in the caloric curve merge ηc\eta_{c}, the first turning point of temperature. There are also secondary intersections giving rise to the spiral (made of unstable equilibrium states) as described in the previous paragraph.

For larger values of NN the discussion is similar to the one given in paper2 .

C.2 R=600R=600

We now consider a system of size R=600R=600 corresponding to the case analysed in Sec. VI. The novelty with respect to the previous situation is that microcanonical phase transitions can appear (they are related to the appearance of turning points of energy).

Some curves Nα(Φ0)N_{\alpha}(\Phi_{0}) are plotted in Fig. 74 for different values of α\alpha. We have indicated on this figure the different characteristic values of NN that have been identified in Sec. VI. As we have seen in the previous section Nmax=106.057N_{\rm max}=106.057, NOV=0.39853N_{\rm OV}=0.39853, N1=0.18131N_{1}=0.18131, NCCP=9.719×106N_{\rm CCP}=9.719\times 10^{-6} and N=0.418N_{*}=0.418 can be related to the topological properties of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. Unfortunately, NMCP=0.00965N_{\rm MCP}=0.00965 and N=1.5N^{\prime}_{*}=1.5 cannot be determined from a simple graphical construction because there does not seem to be a simple manner to relate a turning point of energy in the caloric curve η(Λ)\eta(\Lambda) to the topological properties of the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. Therefore, in Sec. VI, we had to determine NMCPN_{\rm MCP} and NN^{\prime}_{*} “by hand” directly from the study of the caloric curves as a function of NN.

Refer to caption
Figure 74: Some curves Nα(Φ0)N_{\alpha}(\Phi_{0}) for R=600R=600 together with different characteristics values of NN: NCCP=9.719×106N_{\rm CCP}=9.719\times 10^{-6}, NMCP=0.00965N_{\rm MCP}=0.00965, N1=0.18131N_{1}=0.18131, NOV=0.39853N_{\rm OV}=0.39853, N=0.418N_{*}=0.418, N=1.5N^{\prime}_{*}=1.5 and Nmax=106.057N_{\rm max}=106.057.

Apart from the occurence of microcanonical phase transitions, there is another novelty with respect to the previous situation (R=50R=50). A new inflexion point appears at Nchin=0.5062N_{\rm chin}=0.5062 (see Fig. 75). For N<NchinN<N_{\rm chin} we have two fundamental intersections in the N={Nα(Φ0)}N=\{N_{\alpha}(\Phi_{0})\} plot leading to two branches of solutions in the caloric curve η(Λ)\eta(\Lambda) that merge at the temperature minimum ηc\eta_{c} (see Fig. 27).535353In this paragraph, we do not consider the temperature maximum η\eta_{*} that is far away from the dinosaur’s head. For N>NchinN>N_{\rm chin} we have four fundamental intersections in the N={Nα(Φ0)}N=\{N_{\alpha}(\Phi_{0})\} plot leading to four branches of solutions that merge at the temperature minimum ηc\eta_{c}, at the temperature maximum η2\eta_{2} and at the temperature minimum η3\eta_{3} repectively (see Fig. 76). In that case, the dinosaur has a “chin”. This is essentially a curiousity since the solutions in this part of the caloric curve are unstable.

Refer to caption
Figure 75: Illustration of the intersections in the case N>Nchin=0.5062N>N_{\rm chin}=0.5062 (specifically R=600R=600 and N=1.6N=1.6).
Refer to caption
Figure 76: Caloric curve for N>Nchin=0.5062N>N_{\rm chin}=0.5062 (specifically R=600R=600 and N=1.6N=1.6).

C.3 R=10R=10

We consider a system of size R=10R=10 corresponding to the case analyzed in Sec. VIII. Some curves Nα(Φ0)N_{\alpha}(\Phi_{0}) are plotted in Figs. 77 and 78 for α<0\alpha<0 and α>0\alpha>0 respectively. The difference with the case R=50R=50 studied in Appendix C.1 is that there is no canonical phase transition. This is manifested by the absence of inflexion points in the family of curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. We also note that, for R=10R=10, NmaxN_{\rm max} is very close to N0N_{0}.545454They become equal when α=0\alpha_{*}=0 corresponding to R=Rt=12.255R=R_{t}=12.255 (see Fig. 63). However, this is essentially a mathematical curiosity without physical consequences.

Refer to caption
Figure 77: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α<0\alpha<0 for R=10R=10 (for illustration the curves go from α=10\alpha=-10 to α=103\alpha=-10^{-3}).
Refer to caption
Figure 78: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α>0\alpha>0 for R=10R=10 (for illustration the curves go from α=102\alpha=10^{-2} to α=100\alpha=100).

Let us describe the different intersections as a function of NN. As already mentioned, the intersections with the curves α<α0\alpha<\alpha_{*}\simeq 0 (see Fig. 77) give rise to the hot spiral. This case has already been studied in paper2 . Therefore, we consider below the intersections with the curves α>α0\alpha>\alpha_{*}\simeq 0 (see Fig. 78).

For N<N1N<N_{1}, we have just one intersection between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. Therefore, the caloric curve is monotonic and is similar to Fig. 4. The intersection between the line level NN and the OV-curve corresponds to the ground state T=0T=0 (i.e. η+\eta\rightarrow+\infty). This leads to the vertical asymptote at Λ=Λmax\Lambda=\Lambda_{\rm max} in the caloric curve.

For N1<N<NOVN_{1}<N<N_{\rm OV}, there is a second intersection between the line level NN and the OV-curve NOV(Φ0)N_{\rm OV}(\Phi_{0}). This corresponds to an unstable equilibrium state at T=0T=0 (i.e. η+\eta\rightarrow+\infty). This gives rise to the second vertical asymptote at Λ=Λmax\Lambda=\Lambda^{\prime}_{\rm max} in the caloric curve of Fig. 39. There are also secondary intersections leading to the cold spiral that is apparent on Fig. 39.

For N>NOVN>N_{\rm OV}, there are two fundamental intersections between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\}. This gives rise to two fundamental branches in the caloric curve of Fig. 40. The first and second intersections merge at α=αM\alpha=\alpha_{M}. Correspondingly, the first and second branches in the caloric curve merge at ηc\eta_{c}, the first turning point of temperature. Furthermore, there are additional intersections giving rise to the cold spiral in the caloric curve of Fig. 40. These intersections are less relevant since they correspond to unstable states. Note that there is no intersection with the OV curve so there is no vertical asymptote corresponding to η+\eta\rightarrow+\infty.

For larger values of NN the discussion is similar to the one given in paper2 .

C.4 R=1R=1

We consider a system of size R=1R=1 corresponding to the case analyzed in Sec. IX. Some curves Nα(Φ0)N_{\alpha}(\Phi_{0}) are plotted in Figs. 79 and 80 for α<0\alpha<0 and α>0\alpha>0 respectively.555555Since R=1<Rt=12.255R=1<R_{t}=12.255, implying α>0\alpha_{*}>0, we note that Nmax(Φ0)N_{\rm max}(\Phi_{0}) is reached after N0(Φ0)N_{0}(\Phi_{0}) (see Figs. 79 and 80) while it was reached before N0(Φ0)N_{0}(\Phi_{0}) in the case R=50R=50 (see Figs. 65 and 66). The difference with the previous case is that now NOV(Φ0)N_{\rm OV}(\Phi_{0}) is very close to Nmax(Φ0)N_{\rm max}(\Phi_{0}). This is because α1\alpha_{*}\gg 1 (see Fig. 63). Therefore NOVbN_{\rm OV}^{b} is very close to NmaxN_{\rm max} and N1bN_{1}^{b} is very close to NSN^{\prime}_{\rm S}.

Refer to caption
Figure 79: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α<0\alpha<0 for R=1R=1 (for illustration the curves go from α=10\alpha=-10 to α=103\alpha=-10^{-3}).
Refer to caption
Figure 80: Evolution of the curve Nα(Φ0)N_{\alpha}(\Phi_{0}) for different values of α>0\alpha>0 for R=1R=1 (for illustration the curves go from α=104\alpha=10^{-4} to α=103\alpha=10^{3}).

Let us describe the different intersections as a function of NN (note that most of the curves correspond to α<α\alpha<\alpha_{*}).

For N<NSN1bN<N^{\prime}_{S}\simeq N_{1}^{b} (see Fig. 81), we have one intersection between the line level NN and the OV-curve which corresponds to the ground state T=0T=0 (i.e. η+\eta\rightarrow+\infty). This leads to the vertical asymptote at Λ=Λmax\Lambda=\Lambda_{\rm max} in the caloric curve of Fig. 43. In addition, we can have up to an infinity of intersections with the curves Nα(Φ0)N_{\alpha}(\Phi_{0}) leading to the hot spiral displayed in the caloric curve of Fig. 43.

Refer to caption
Figure 81: Illustration of the intersections in the case N<NSN1bN<N^{\prime}_{S}\simeq N_{1}^{b} (specifically R=1R=1 and N=0.12N=0.12).
Refer to caption
Figure 82: Illustration of the intersections in the case NSN1b<N<NmaxNOVbN^{\prime}_{S}\simeq N_{1}^{b}<N<N_{\rm max}\simeq N_{\rm OV}^{b} (specifically R=1R=1 and N=0.14N=0.14).

For NSN1b<N<NmaxNOVbN^{\prime}_{S}\simeq N_{1}^{b}<N<N_{\rm max}\simeq N_{\rm OV}^{b} (see Fig. 82), we have two fundamental intersections between the line level NN and the OV-curve (corresponding to α+\alpha\rightarrow+\infty) leading to the asymptotes η=+\eta=+\infty at Λmax\Lambda_{\rm max} (stable ground state) and Λmax\Lambda^{\prime}_{\rm max} (unstable ground state) in the caloric curve of Fig. 44. For smaller values of α\alpha, the two fundamental intersections between the line level NN and the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} lead to two fundamental branches in the caloric curve of Fig. 44. These intersections merge at α=αn\alpha=\alpha_{n}. Since αn<α\alpha_{n}<\alpha_{*}, this is associated with a turning point of energy at Λmin\Lambda_{\rm min} in the caloric curve of Fig. 44. There may also be a third intersection with the OV-curve and secondary intersections with the curves {Nα(Φ0)}\{N_{\alpha}(\Phi_{0})\} forming a third branch exhibiting an asymptote at Λmax′′\Lambda^{\prime\prime}_{\rm max} and a spiral. However, these solutions are not represented on the caloric curve because they are associated with unstable states of high order.

Appendix D Condition to be degenerate

In the nonrelativistic limit, the system is degenerate when the thermal pressure P=ρkBT/mP=\rho k_{B}T/m is small as compared to the quantum pressure P=(1/5)(3/4πg)2/3h2ρ5/3/m8/3P=(1/5)(3/4\pi g)^{2/3}h^{2}\rho^{5/3}/m^{8/3} arising from the Pauli exclusion principle. This condition can be written as

kBT15(34πg)2/3h2m5/3ρ2/3k_{B}T\ll\frac{1}{5}\left(\frac{3}{4\pi g}\right)^{2/3}\frac{h^{2}}{m^{5/3}}\rho^{2/3} (68)

which is the nonrelativistic Sommerfeld criterion (the right hand side of Eq. (68) is of the order of the Fermi temperature TFT_{F}). To get an estimate of the importance of degeneracy, we replace the density ρ\rho by the average density ρ¯=3M/4πR3\overline{\rho}=3M/4\pi R^{3}. In this manner, we obtain the condition

η0.917μ2/3.\eta\gg 0.917\,\mu^{2/3}. (69)

As stated above this condition is only valid in an average sense. A system which does not satisfy this condition in average may still have a degenerate core and a nondegenerate halo as in Appendix E.2.

In the ultrarelativistic limit, the system is degenerate when the thermal pressure P=ρkBT/mP=\rho k_{B}T/m is small as compared to the quantum pressure P=(1/4)(3/4πg)1/3hcρ4/3/m4/3P=(1/4)(3/4\pi g)^{1/3}hc\rho^{4/3}/m^{4/3} arising from the Pauli exclusion principle. This can be written as

kBT14(34πg)1/3hcm1/3ρ1/3k_{B}T\ll\frac{1}{4}\left(\frac{3}{4\pi g}\right)^{1/3}\frac{hc}{m^{1/3}}\rho^{1/3} (70)

which is the ultrarelativistic Sommerfeld criterion. Proceeding as above, we obtain the condition

η1.10(MMOV)2/3,\eta\gg 1.10\,\left(\frac{M}{M_{\rm OV}}\right)^{2/3}, (71)

where MOVM_{\rm OV} is the OV critical mass given by Eq. (53).

Appendix E Thermodynamics of nonrelativistic self-gravitating systems

In this Appendix, we recall and complete important results concerning the thermodynamics of nonrelativistic self-gravitating systems (classical particles and fermions) that are needed in our analysis. We refer to Appendix F for useful formulae that are used throughout this Appendix.

E.1 Classical particles

We first consider a self-gravitating system of nonrelativistic classical particles confined within a spherical box of radius RR. We show below that there is no statistical equilibrium state in a strict sense but that long-lived metastable states can exist under certain conditions.

E.1.1 Canonical ensemble

In the canonical ensemble, one can make the free energy F=ETSF=E-TS diverge towards -\infty at fixed mass MM in the most efficient manner by approaching all the particles at the same point (see Appendix B of sc ). Indeed, let us consider a homogeneous sphere of radius aa containing all the particles. When a0a\rightarrow 0 the potential energy W=3GM2/(5a)W=-3GM^{2}/(5a) diverges to -\infty. The entropy S3NkBlnaS\sim 3Nk_{B}\ln a diverges to -\infty but it is subdominant. As a result, the free energy F=ETSW3GM2/(5a)F=E-TS\sim W\sim-3GM^{2}/(5a) diverges to -\infty. Therefore, there is no global minimum of free energy. In a sense, the most probable structure in the canonical ensemble is a Dirac peak containing all the mass.

On the other hand, there exist metastable gaseous states with a temperature T>Tc=0.397GMm/(kBR)T>T_{c}=0.397GMm/(k_{B}R) lbw ; aa . They have a density contrast <32.1{\cal R}<32.1 lbw ; aa and they are very long-lived lifetime . When T<TcT<T_{c}, or when T>TcT>T_{c} and >32.1{\cal R}>32.1, there are no metastable states anymore and the system collapses (isothermal collapse) aa . According to the previous thermodynamical argument, it is expected to form a Dirac peak containing all the mass. By solving the Smoluchowski-Poisson equations describing the canonical evolution of self-gravitating Brownian particles sc ; post it is found that the Dirac peak is formed in the postcollapse regime of the dynamics.

E.1.2 Microcanonical ensemble

In the microcanonical ensemble, one can make the entropy SS diverge towards ++\infty at fixed mass MM and energy EE by forming a core-halo structure and letting the size of the core go to zero (see Appendix A of sc ). Indeed, let us consider a homogeneous core made of NCN_{C} particles in a sphere of radius RCR_{C}. Its potential energy WC=3GMC2/(5RC)W_{C}=-3GM_{C}^{2}/(5R_{C}) tends to -\infty when RC0R_{C}\rightarrow 0. In order to conserve the total energy, the kinetic energy of the halo Kh=(3/2)NhkBTK_{h}=(3/2)N_{h}k_{B}T must tend to ++\infty like KhWCK_{h}\sim-W_{C} meaning that its temperature T2GMC2/(5RCNhkB)T\sim 2GM_{C}^{2}/(5R_{C}N_{h}k_{B}) tends to ++\infty. As a result, the entropy of the system behaves as S(3/2)kB(NhNC)lnRCS\sim-(3/2)k_{B}(N_{h}-N_{C})\ln R_{C} and tends to ++\infty when RC0R_{C}\rightarrow 0. Therefore, there is no global maximum of entropy at fixed energy. We note that the divergence of the entropy is the most efficient when the core contains a few particles (NCNhN_{C}\ll N_{h}). Actually, we only need to approach 22 particles to each other and make a tight binary. Its potential energy Gm2/a-Gm^{2}/a diverges towards -\infty. The released energy serves to heat the halo made of the N2N-2 other particles. This produces the most efficient divergence of entropy (note that the divergence of entropy is weak — logarithmic). In a sense, the most probable structure in the microcanonical ensemble is a tight binary surrounded by a hot halo. This can be seen as a Dirac peak of zero mass but infinite potential energy ++ a hot halo.

On the other hand, there exist metastable gaseous states with an energy E>Ec=0.335GM2/RE>E_{c}=-0.335GM^{2}/R lbw . They have a density contrast <709{\cal R}<709 antonov ; lbw and they are very long-lived lifetime . When E<EcE<E_{c}, or when E>EcE>E_{c} and >709{\cal R}>709, there are no metastable states anymore and the system collapses (gravothermal catastrophe) lbw . It is expected to form a binary surrounded by a hot halo. Dynamical models describing the collisional evolution of globular clusters (fluid equations, orbit-averaged Fokker-Planck equation…) show that the binary is formed in the postcollapse regime of the dynamics lbe ; inagaki ; cohn . The energy released by the binary can to stop the collapse and induce a re-expansion of the halo. Then, in principle, a series of gravothermal oscillations should follow oscillations ; hr .

E.2 Fermions

We now consider a self-gravitating system of nonrelativistic fermions confined within a spherical box of radius RR. In that case, there exists a statistical equilibrium state at any energy EE and temperature TT ijmpb . We consider a situation where we are close to the classical limit (μ+\mu\rightarrow+\infty or 0\hbar\rightarrow 0). When T>TcT>T_{c} and E>EcE>E_{c} there exist metastable gaseous states that are not affected by quantum mechanics (see Appendix E.1). When T<TcT<T_{c} and E<EcE<E_{c} there are no metastable states anymore. The system collapses and becomes very dense until quantum mechanics (Pauli’s exclusion principle) comes into play. Generically, the system forms a core-halo structure with a completely degenerate fermionic core of mass MCM_{C} and radius RCR_{C} surrounded by an (almost classical) isothermal halo of mass Mh=MMCM_{h}=M-M_{C} and radius RR. We can obtain the value of MCM_{C} from a simple analytical model developed in pt . We summarize this model below by using dimensional variables in order to understand more easily the physical mechanisms at play.

We model the core as a completely degenerate fermion ball. Its mass-radius relation is

MCRC3=χh6g2m8G3.\displaystyle M_{C}R_{C}^{3}=\chi\frac{h^{6}}{g^{2}m^{8}G^{3}}. (72)

Its energy (kinetic ++ potential) is

EC=37χ1/3G2MC7/3g2/3m8/3h2.E_{C}=-\frac{3}{7\chi^{1/3}}\frac{G^{2}M_{C}^{7/3}g^{2/3}m^{8/3}}{h^{2}}. (73)

Its entropy is zero: SC=0S_{C}=0.

We model the halo by a classical gas at temperature TT with a uniform density.565656We will see that its temperature is very large in the situations considered. This justifies the uniform density approximation. Its kinetic energy is

Kh=32NhkBT.K_{h}=\frac{3}{2}N_{h}k_{B}T. (74)

Its potential (gravitational) energy, taking into account the presence of the core, is

Wh=3GMCMh2R3GMh25R.W_{h}=-\frac{3GM_{C}M_{h}}{2R}-\frac{3GM_{h}^{2}}{5R}. (75)

Its entropy is

Sh=NhkBln(MhV)+32NhkBln(2πkBTm)\displaystyle S_{h}=-N_{h}k_{B}\ln\left(\frac{M_{h}}{V}\right)+\frac{3}{2}N_{h}k_{B}\ln\left(\frac{2\pi k_{B}T}{m}\right)
+52NhkB+NhkBlnη0.\displaystyle+\frac{5}{2}N_{h}k_{B}+N_{h}k_{B}\ln\eta_{0}. (76)

In the foregoing expressions, we have assumed RCRR_{C}\ll R which can be checked a posteriori.

The total mass of the system is M=MC+MhM=M_{C}+M_{h}, its total energy is E=EC+Kh+WhE=E_{C}+K_{h}+W_{h}, its total entropy is S=ShS=S_{h} and its total free energy is F=ETSF=E-TS. The mass MCM_{C} of the core is obtained by maximizing the entropy at fixed mass and energy in the microcanonical ensemble or by minimizing the free energy at fixed mass in the canonical ensemble. The extremization problem gives in both ensembles pt :

G2MC4/3g2/3m8/3χ1/3h23G(M2MC)2R\displaystyle-\frac{G^{2}M_{C}^{4/3}g^{2/3}m^{8/3}}{\chi^{1/3}h^{2}}-\frac{3G(M-2M_{C})}{2R}
+6G(MMC)5RkBTmln(MMCV)\displaystyle+\frac{6G(M-M_{C})}{5R}-\frac{k_{B}T}{m}\ln\left(\frac{M-M_{C}}{V}\right)
+32kBTmln(2πkBTm)+kBTmlnη0=0.\displaystyle+\frac{3}{2}\frac{k_{B}T}{m}\ln\left(\frac{2\pi k_{B}T}{m}\right)+\frac{k_{B}T}{m}\ln\eta_{0}=0. (77)

This equation may have several solutions that have been analyzed in detail in pt . Below, we restrict ourselves to the stable condensed state.

E.2.1 Canonical ensemble

In the canonical ensemble, when h0h\rightarrow 0, we expect that the core contains a large mass MC/M1M_{C}/M\sim 1 (see Appendix E.1.1). Guided by this ansatz, which can be checked a posteriori, one can see from Eq. (E.2) that the core mass is given by

1MCMη0VM(2πkBTm)3/2e3GMm2RkBTeG2M4/3g2/3m11/3χ1/3h2kBT.1-\frac{M_{C}}{M}\sim\frac{\eta_{0}V}{M}\left(\frac{2\pi k_{B}T}{m}\right)^{3/2}e^{\frac{3GMm}{2Rk_{B}T}}e^{-\frac{G^{2}M^{4/3}g^{2/3}m^{11/3}}{\chi^{1/3}h^{2}k_{B}T}}. (78)

When h0h\rightarrow 0, the core mass tends towards MM exponentially rapidly. Therefore the core contains almost all the mass: MCMM_{C}\sim M. Its radius is given by

RC=χ1/3h2g2/3m8/3GM1/3.\displaystyle R_{C}=\chi^{1/3}\frac{h^{2}}{g^{2/3}m^{8/3}GM^{1/3}}. (79)

When h0h\rightarrow 0, it tends to zero as h2h^{2}. The energy of the core is

EC=37χ1/3G2M7/3g2/3m8/3h2.E_{C}=-\frac{3}{7\chi^{1/3}}\frac{G^{2}M^{7/3}g^{2/3}m^{8/3}}{h^{2}}. (80)

When h0h\rightarrow 0, it tends to -\infty as h2-h^{-2}. The free energy, which is dominated by the energy of the core, FECF\sim E_{C}, behaves in a similar manner. In the classical limit h0h\rightarrow 0, we recover the Dirac peak containing all the mass. This structure leads to the divergence of the free energy in agreement with the arguments of Appendix E.1.1. We note that these results are independent of the presence, or not, of the box.

In terms of dimensionless variables pt , the preceding results can be written as

1αCπ6μe3η/2η3/2eλημ2/3,1-\alpha_{C}\sim\frac{\sqrt{\pi}}{6}\mu\frac{e^{3\eta/2}}{\eta^{3/2}}e^{-\lambda\eta\mu^{2/3}}, (81)
RCR1λμ2/3,\frac{R_{C}}{R}\sim\frac{1}{\lambda\mu^{2/3}}, (82)
ΛCΛmax=37λμ2/3,\Lambda_{C}\sim\Lambda_{\rm max}=\frac{3}{7}\lambda\mu^{2/3}, (83)

where αC=MC/M\alpha_{C}=M_{C}/M. We note that letting h0h\rightarrow 0 (classical limit) in the dimensional equations is equivalent to letting μ+\mu\rightarrow+\infty in the dimensionless equations.

Remark: Using the results of Ref. pt , the maximum temperature of the condensed phase, and the corresponding core mass, are given by

η2lnμλμ2/3,1α38lnμ.\eta_{*}\sim\frac{2\ln\mu}{\lambda\mu^{2/3}},\qquad 1-\alpha_{*}\sim\frac{3}{8\ln\mu}. (84)

Coming back to dimensional variables, we get575757Here and in the following, we give the logarithmic correction in the dominant approximation. Furthermore, lnh-\ln h should be understood as (1/3)lnμ(1/3)\ln\mu where μ=(gm4/h3)512π4G3MR3\mu=(gm^{4}/h^{3})\sqrt{512\pi^{4}G^{3}MR^{3}} is dimensionless.

kBT16χ1/3g2/3G2M4/3m11/3h2(lnh).k_{B}T_{*}\sim\frac{1}{6\chi^{1/3}}\frac{g^{2/3}G^{2}M^{4/3}m^{11/3}}{h^{2}(-\ln h)}. (85)

On the other hand, the temperature of transition, and the corresponding core mass, are given by

ηt14lnμ3λμ2/3,1αtπ6(3λ14)3/21μ8/3(lnμ)3/2.\eta_{t}\sim\frac{14\ln\mu}{3\lambda\mu^{2/3}},\qquad 1-\alpha_{t}\sim\frac{\sqrt{\pi}}{6}\left(\frac{3\lambda}{14}\right)^{3/2}\frac{1}{\mu^{8/3}(\ln\mu)^{3/2}}. (86)

Coming back to dimensional variables, we get

kBTt114χ1/3g2/3G2M4/3m11/3h2(lnh).k_{B}T_{t}\sim\frac{1}{14\chi^{1/3}}\frac{g^{2/3}G^{2}M^{4/3}m^{11/3}}{h^{2}(-\ln h)}. (87)

E.2.2 Microcanonical ensemble

In the microcanonical ensemble, we expect that the core contains a small mass MC/M1M_{C}/M\ll 1 (see Appendix E.1.2). Guided by this ansatz, which can be checked a posteriori, one can see from Eq. (E.2) that the core mass is given by

MCM712(lnh).\displaystyle\frac{M_{C}}{M}\sim\frac{7}{12(-\ln h)}. (88)

When h0h\rightarrow 0, the core mass tends towards 0 extremely slowly (logarithmically). Therefore, the core contains a fraction of the total mass and this fraction goes to zero as (lnh)1(-\ln h)^{-1} when h0h\rightarrow 0. The radius of the core is given by

RC(127)1/3χ1/3h2(lnh)1/3g2/3m8/3GM1/3.\displaystyle R_{C}\sim\left(\frac{12}{7}\right)^{1/3}\chi^{1/3}\frac{h^{2}(-\ln h)^{1/3}}{g^{2/3}m^{8/3}GM^{1/3}}. (89)

When h0h\rightarrow 0, it tends to zero as h2(lnh)1/3h^{2}(-\ln h)^{1/3}. The energy of the core is

EC37χ1/3(712)7/3G2M7/3g2/3m8/3h2(lnh)2.E_{C}\sim-\frac{3}{7\chi^{1/3}}\left(\frac{7}{12}\right)^{7/3}\frac{G^{2}M^{7/3}g^{2/3}m^{8/3}}{h^{2}(-\ln h)^{2}}. (90)

When h0h\rightarrow 0, it tends to -\infty as h2(lnh)2-h^{-2}(-\ln h)^{-2}. Since the energy of the core is very negative the kinetic energy of the halo must be very positive in order to conserve the total energy. It must behave as KhECK_{h}\sim-E_{C}. Therefore, the temperature of the halo must be very large:

kBTcond27χ1/3(712)7/3G2M4/3g2/3m11/3h2(lnh)2.k_{B}T_{\rm cond}\sim\frac{2}{7\chi^{1/3}}\left(\frac{7}{12}\right)^{7/3}\frac{G^{2}M^{4/3}g^{2/3}m^{11/3}}{h^{2}(-\ln h)^{2}}. (91)

When h0h\rightarrow 0, it diverges to ++\infty as h2(lnh)2h^{-2}(-\ln h)^{-2}. The entropy behaves as S6NkBlnhS\sim-6Nk_{B}\ln h. Subtracting the term 3NkBlnh-3Nk_{B}\ln h that we get even in the absence of gravity (see Appendix F), we obtain

ΔS3NkBlnh.\Delta S\sim-3Nk_{B}\ln h. (92)

When h0h\rightarrow 0, the entropy diverges to ++\infty as lnh-\ln h. In the classical limit h0h\rightarrow 0, we recover the core-halo structure made of a core having a small mass, a small radius and a huge potential energy (Dirac peak of zero mass) surrounded by a very hot halo.585858We note that the collapse at low energies in the microcanonical ensemble (gravothermal catastrophe) produces hot systems (T+T\rightarrow+\infty) with a core-halo structure. Actually, although the temperature is uniform throughout the system, the halo is hot while the core is cold. Indeed, the halo is nondegenerate (Boltzmannian) because TTFT\gg T_{F}, where TFT_{F} is the Fermi temperature (see Appendix D), while the core is completely degenerate because TTFT\ll T_{F}. Fundamentally, this core-halo structure is the consequence of the negative specific heat of self-gravitating systems as explained in Ref. lbw . This core-halo structure leads to the (logarithmic) divergence of the entropy in agreement with the results of Appendix E.1.2. We note that these results are independent of the presence, or not, of the box. They are also independent of the value of the energy EE provided that it is not too extreme.

In terms of dimensionless variables pt , the preceding results can be rewritten as

αC74lnμ,\alpha_{C}\sim\frac{7}{4\ln\mu}, (93)
RCR1λ(47)1/3(lnμ)1/3μ2/3,\frac{R_{C}}{R}\sim\frac{1}{\lambda}\left(\frac{4}{7}\right)^{1/3}\frac{(\ln\mu)^{1/3}}{\mu^{2/3}}, (94)
ΛC37λ(74)7/3μ2/3(lnμ)7/3,\Lambda_{C}\sim\frac{3}{7}\lambda\left(\frac{7}{4}\right)^{7/3}\frac{\mu^{2/3}}{(\ln\mu)^{7/3}}, (95)
ηcond72λ(47)7/3(lnμ)7/3μ2/3,\eta_{\rm cond}\sim\frac{7}{2\lambda}\left(\frac{4}{7}\right)^{7/3}\frac{(\ln\mu)^{7/3}}{\mu^{2/3}}, (96)

where αC=MC/M\alpha_{C}=M_{C}/M. More generally (without specifying the value of the core mass αC\alpha_{C}), in the case where the energy of the core tends to -\infty and the energy of the halo tends to ++\infty we have the relation

ηcond1αCαC7/372λμ2/3.\eta_{\rm cond}\sim\frac{1-\alpha_{C}}{\alpha_{C}^{7/3}}\frac{7}{2\lambda\mu^{2/3}}. (97)

When αC\alpha_{C} is given by Eq. (93) obtained from Eq. (E.2), we recover Eq. (94). Finally, we note that letting h0h\rightarrow 0 (classical limit) in the dimensional equations is equivalent to letting μ+\mu\rightarrow+\infty in the dimensionless equations.

Remark: Using the results of Ref. pt , the maximum energy of the condensed phase, and the corresponding core mass, are given by

Λ9λ28μ2/3(lnμ)7/3,α1lnμ.\Lambda_{*}\sim-\frac{9\lambda}{28}\frac{\mu^{2/3}}{(\ln\mu)^{7/3}},\qquad\alpha_{*}\sim\frac{1}{\ln\mu}. (98)

Coming back to dimensional variables, we get

E128(3χ)1/3g2/3G2M7/3m8/3h2(lnh)7/3.E_{*}\sim\frac{1}{28(3\chi)^{1/3}}\frac{g^{2/3}G^{2}M^{7/3}m^{8/3}}{h^{2}(-\ln h)^{7/3}}. (99)

The energy of transition Λt\Lambda_{t}, and the corresponding core mass αt\alpha_{t}, have the same scalings.

Appendix F Useful formulae

In this Appendix, we regroup basic formulae that are useful in our study.

F.1 Energy and entropy

The energy of a nonrelativistic self-gravitating system can be written as

E=fv22𝑑𝐫𝑑𝐯+12ρΦ𝑑𝐫=K+W,E=\int f\frac{v^{2}}{2}\,d{\bf r}d{\bf v}+\frac{1}{2}\int\rho\Phi\,d{\bf r}=K+W, (100)

where KK is the kinetic energy and WW the potential (gravitational) energy bt .

The Fermi-Dirac entropy functional is given by

S=kBη0m{fη0lnfη0+(1fη0)ln(1fη0)}𝑑𝐫𝑑𝐯,S=-k_{B}\frac{\eta_{0}}{m}\int\Biggl{\{}\frac{f}{\eta_{0}}\ln\frac{f}{\eta_{0}}+\left(1-\frac{f}{\eta_{0}}\right)\ln\left(1-\frac{f}{\eta_{0}}\right)\Biggr{\}}\,d{\bf r}d{\bf v}, (101)

where η0=gm4/h3\eta_{0}=g{m^{4}}/{h^{3}} is the maximum allowed value of the distribution function f(𝐫,𝐯)f({\bf r},{\bf v}) fixed by Pauli’s exclusion principle ptd . In the classical (nondegenerate) limit fη0f\ll\eta_{0}, it reduces to the Boltzmann entropy functional

S=kB(fmlnfη0fm)𝑑𝐫𝑑𝐯.S=-k_{B}\int\left(\frac{f}{m}\ln\frac{f}{\eta_{0}}-\frac{f}{m}\right)\,d{\bf r}d{\bf v}. (102)

The distribution function that maximizes the Boltzmann entropy at fixed density ρ\rho and energy EE is the Boltzmann distribution

f(𝐫,𝐯)=(m2πkBT)3/2ρ(𝐫)emv22kBT,f({\bf r},{\bf v})=\left(\frac{m}{2\pi k_{B}T}\right)^{3/2}\rho({\bf r})\,e^{-\frac{mv^{2}}{2k_{B}T}}, (103)

where TT is the temperature paddy ; aa . Using Eq. (103), the kinetic energy and the entropy of a nonrelativistic classical isothermal self-gravitating system are

K=32NkBTK=\frac{3}{2}Nk_{B}T (104)

and

S=kBρmlnρd𝐫+32NkBln(2πkBTm)\displaystyle S=-k_{B}\int\frac{\rho}{m}\ln\rho\,d{\bf r}+\frac{3}{2}Nk_{B}\ln\left(\frac{2\pi k_{B}T}{m}\right)
+52NkB+NkBlnη0.\displaystyle+\frac{5}{2}Nk_{B}+Nk_{B}\ln\eta_{0}. (105)

Remark: We note that the Boltzmann entropy defined by Eq. (102) diverges like SNkBlnη03NkBln+S\sim Nk_{B}\ln\eta_{0}\sim-3Nk_{B}\ln\hbar\rightarrow+\infty when 0\hbar\rightarrow 0. This divergence is present in the famous Sackur-Tetrode formula for the entropy of a perfect gas (without self-gravity). In order to see the absence of statistical equilibrium states for classical self-gravitating systems, marked by the divergence of the entropy when 0\hbar\rightarrow 0, we first have to subtract the term 3NkBlnh-3Nk_{B}\ln h from the total entropy (see Appendix E.2.2).

F.2 Homogeneous sphere

The potential (gravitational) energy of a spatially homogeneous sphere of mass MM and radius RR is bt :

W=3GM25R.W=-\frac{3GM^{2}}{5R}. (106)

Using Eqs. (100), (104) and (106), the total energy of a nonrelativistic classical isothermal self-gravitating system with a unifom density is

E=32NkBT3GM25R.E=\frac{3}{2}Nk_{B}T-\frac{3GM^{2}}{5R}. (107)

Using Eq. (F.1) its entropy is

SB=NkBln(MV)+32NkBln(2πkBTm)\displaystyle S_{B}=-Nk_{B}\ln\left(\frac{M}{V}\right)+\frac{3}{2}Nk_{B}\ln\left(\frac{2\pi k_{B}T}{m}\right)
+52NkB+NkBlnη0,\displaystyle+\frac{5}{2}Nk_{B}+Nk_{B}\ln\eta_{0}, (108)

where V=(4/3)πR3V=(4/3)\pi R^{3} is the volume of the system.

F.3 Completely degenerate nonrelativistic self-gravitating Fermi gas

The mass-radius relation of a completely degenerate fermion star (T=0T=0) in the nonrelativistic limit is

MCRC3=χh6g2m8G3,RC=χ1/3h2g2/3m8/3GMC1/3M_{C}R_{C}^{3}=\chi\frac{h^{6}}{g^{2}m^{8}G^{3}},\quad R_{C}=\chi^{1/3}\frac{h^{2}}{g^{2/3}m^{8/3}GM_{C}^{1/3}} (109)

with

χ=18(34π)2ω3/216π2=5.97241×103,\displaystyle\chi=\frac{1}{8}\left(\frac{3}{4\pi}\right)^{2}\frac{\omega_{3/2}}{16\pi^{2}}=5.97241\times 10^{-3}, (110)

where ω3/2=132.3843\omega_{3/2}=132.3843 chandrabook .

Its energy (kinetic ++ potential) is chandrabook :

EC=3GMC27RC.E_{C}=-\frac{3GM_{C}^{2}}{7R_{C}}. (111)

Combined with the mass-radius relation (109), we get

EC=37χ1/3G2MC7/3g2/3m8/3h2.E_{C}=-\frac{3}{7\chi^{1/3}}\frac{G^{2}M_{C}^{7/3}g^{2/3}m^{8/3}}{h^{2}}. (112)

This is the energy of the ground state.

Remark: In terms of dimensionless variables pt , the mass-radius relation can be written as

RCR=1λαC1/3μ2/3,\displaystyle\frac{R_{C}}{R}=\frac{1}{\lambda\alpha_{C}^{1/3}\mu^{2/3}}, (113)

where αC=MC/M\alpha_{C}=M_{C}/M and

λ=1(512π4χ)1/3=0.149736\lambda=\frac{1}{(512\pi^{4}\chi)^{1/3}}=0.149736... (114)

Similarly, the energy-mass relation can be written as

ΛC=37λαC7/3μ2/3.\displaystyle\Lambda_{C}=\frac{3}{7}\lambda\alpha_{C}^{7/3}\mu^{2/3}. (115)

In writing these expressions, we have implicitly assumed that the fermion star of mass MCM_{C} and radius RCR_{C} constitutes the core of a larger system of mass MM and radius RR as in Appendix E.

F.4 Ground state of a self-gravitating Fermi gas in a box

In terms of dimensionless variables pt , the minimum energy (ground state) of a nonrelativistic self-gravitating Fermi gas enclosed within a box is given by (see Eq. (115) with αC=1\alpha_{C}=1):

Λmax=37λμ2/3.\Lambda_{\rm max}=\frac{3}{7}\lambda\mu^{2/3}. (116)

This expression is valid for a self-confined fermion star such that RC<RR_{C}<R (i.e., the density of the fermion star vanishes before reaching the box). Using Eq. (113) with αC=1\alpha_{C}=1, we find that Eq. (116) is valid for μ>λ3/2=17.26\mu>\lambda^{-3/2}=17.26. When μ<17.26\mu<17.26, the fermion star at T=0T=0 (ground state) is box-confined (RC>RR_{C}>R) and its energy Λmaxb(μ)\Lambda_{\rm max}^{b}(\mu) is represented in Fig. 2 of ptd .

Introducing the normalized variables of Appendix B and using Eq. (20), we find from Eq. (116) that the minimum energy of a nonrelativistic self-gravitating Fermi gas is given by

ΛmaxR=37λ(42π)2/3N1/3=0.0950N1/3.\frac{\Lambda_{\rm max}}{R}=\frac{3}{7}\lambda\left(\frac{4\sqrt{2}}{\pi}\right)^{2/3}N^{1/3}=0.0950\,N^{1/3}. (117)

This expression is valid for N>91.9/R3N>91.9/R^{3} so that the fermion star is self-confined (RC<RR_{C}<R). This equation can be used to locate the vertical asymptote Λmax\Lambda_{\rm max} in the caloric curves of this paper. However, it is only valid in the nonrelativistic regime NNOVN\ll N_{\rm OV}. In the relativistic regime, the minimum energy Λmax\Lambda_{\rm max} of the self-gravitating Fermi gas, as well as the energy Λmax\Lambda^{\prime}_{\rm max} of the unstable fermion star at T=0T=0, are represented in Fig. 14 of paper1 . At the point N=N1N=N_{1} where the second asymptote (corresponding to the unstable fermion star at T=0T=0) appears, we find that

ΛmaxR=0.53617,ΛmaxR=0.0570.\frac{\Lambda^{\prime}_{\rm max}}{R}=-0.53617,\qquad\frac{\Lambda_{\rm max}}{R}=0.0570. (118)

At the point N=NOVN=N_{\rm OV} where the two asymptotes meet each other, we find that

ΛmaxR=ΛmaxR=0.08985.\frac{\Lambda^{\prime}_{\rm max}}{R}=\frac{\Lambda_{\rm max}}{R}=0.08985. (119)

Appendix G Temperature-dependent OV maximum particle number

For R=50R=50 and NNOV+N\rightarrow N_{\rm OV}^{+}, we find from Fig. 24 that

ηc(N)0.104(NNOV)1/2.\displaystyle\eta^{\prime}_{c}(N)\sim 0.104\,(N-N_{\rm OV})^{-1/2}. (120)

For a given normalized temperature η>ηc2.52\eta>\eta_{c}\simeq 2.52, the system collapses towards a black hole when ηc(N)<η\eta^{\prime}_{c}(N)<\eta, i.e., when NNOV(η)N\geq N_{\rm OV}(\eta) with

NOV(η)=NOV+0.0108/η2.\displaystyle N_{\rm OV}(\eta)=N_{\rm OV}+0.0108/\eta^{2}. (121)

This can be seen as a temperature-dependent OV maximum particle number. We note that NOV(η)N_{\rm OV}(\eta) is very close to NOVN_{\rm OV} since NOV(ηc)=NcCE=1.00427NOVN_{\rm OV}(\eta_{c})=N_{c}^{\rm CE}=1.00427\,N_{\rm OV} (see Sec. XIII.3). The relation (120) remains valid, with a different prefactor, for other values of R>RCCPR>R_{\rm CCP}. On the other hand, for ROV<R<RCCPR_{\rm OV}<R<R_{\rm CCP}, we have a similar relation for ηc(N)\eta_{c}(N) close to NOVN_{\rm OV} (see Fig. 41 for R=10R=10):

ηc(N)0.516(NNOV)1/2.\displaystyle\eta_{c}(N)\sim 0.516\,(N-N_{\rm OV})^{-1/2}. (122)

Using the same argument as before, this yields

NOV(η)=NOV+0.266/η2.\displaystyle N_{\rm OV}(\eta)=N_{\rm OV}+0.266/\eta^{2}. (123)

More generally, writing Eqs. (120) and (122) under the form

ηc()(N)a(R)(NNOV)1/2,\displaystyle\eta_{c}^{(^{\prime})}(N)\sim a(R)\,(N-N_{\rm OV})^{-1/2}, (124)

we get

NOV(η)=NOV+a(R)2/η2.\displaystyle N_{\rm OV}(\eta)=N_{\rm OV}+a(R)^{2}/\eta^{2}. (125)

If we substitute η=βGNm2/R\eta=\beta GNm^{2}/R into Eq. (125) and replace NN by NOVN_{\rm OV} at leading order, we obtain

NOV(T)=NOV+a(R)2R2(kBT)2G2NOV2m4.\displaystyle N_{\rm OV}(T_{\infty})=N_{\rm OV}+a(R)^{2}\frac{R^{2}(k_{B}T_{\infty})^{2}}{G^{2}N_{\rm OV}^{2}m^{4}}. (126)

We make the guess that the product a(R)Ra(R)R in Eq. (126) is independent of RR. This can be checked on the two values that we have computed since the products 0.104×50=5.20.104\times 50=5.2 and 0.516×10=5.160.516\times 10=5.16 are almost the same. As a result, we guess that a(R)(R/ROV)5.2/3.35691.5a(R)(R/R_{\rm OV})\simeq 5.2/3.3569\simeq 1.5, i.e., a(R)1.5(ROV/R)a(R)\simeq 1.5(R_{\rm OV}/R). Substituting this relation into Eq. (126), we finally obtain (using the results of Appendix B):

NOV(T)NOV+160(kBTmc2)2.\displaystyle N_{\rm OV}(T_{\infty})\simeq N_{\rm OV}+160\left(\frac{k_{B}T_{\infty}}{mc^{2}}\right)^{2}. (127)

This relation is expected to be valid for kBTmc2k_{B}T_{\infty}\ll mc^{2}. It gives the first order correction to the OV maximum number due to thermal effects.

References

  • (1) T. Padmanabhan, Phys. Rep. 188, 285 (1990)
  • (2) V.A. Antonov, Vest. Leningr. Gos. Univ. 7, 135 (1962)
  • (3) D. Lynden-Bell, R. Wood, Mon. Not. R. Astron. Soc. 138, 495 (1968)
  • (4) P.H. Chavanis, Astron. Astrophys. 432, 117 (2005)
  • (5) P.H. Chavanis, Int. J. Mod. Phys. B 20, 3113 (2006)
  • (6) G. Alberti, P.H. Chavanis, in preparation
  • (7) G. Alberti, P.H. Chavanis, in preparation
  • (8) S. Chandrasekhar, An Introduction to the Theory of Stellar Structure (Dover, 1942)
  • (9) S.L. Shapiro, S.A. Teukolsky Black Holes, White Dwarfs, and Neutron Stars (Wiley Interscience, 1983)
  • (10) N. Bilić, G.B. Tupper, R.D. Viollier, Lect. Notes Phys. 616, 24 (2003)
  • (11) H.J. de Vega, P. Salucci, N.G. Sanchez, Mon. Not. R. Astron. Soc. 442, 2717 (2014)
  • (12) R. Ruffini, C.R. Argüelles, J.A. Rueda, Mon. Not. R. Astron. Soc. 451, 622 (2015)
  • (13) P.H. Chavanis, M. Lemou, F. Méhats, Phys. Rev. D 91, 063531 (2015)
  • (14) P.H. Chavanis, M. Lemou, F. Méhats, Phys. Rev. D 92, 123527 (2015)
  • (15) V. Domcke, A. Urbano, JCAP 01, 002 (2015)
  • (16) L. Randall, J. Scholtz, J. Unwin, Mon. Not. R. Astron. Soc. 467, 1515 (2017)
  • (17) P.H. Chavanis, Eur. Phys. J. B 87, 9 (2014)
  • (18) Dynamics and thermodynamics of systems with long range interactions, edited by T. Dauxois, S. Ruffo, E. Arimondo, M. Wilkens, Lecture Notes in Physics 602, (Springer, 2002)
  • (19) J. Katz, Found. Phys. 33, 223 (2003)
  • (20) A. Campa, T. Dauxois, S. Ruffo, Physics Reports 480, 57 (2009)
  • (21) A. Campa, T. Dauxois, D. Fanelli, S. Ruffo, Physics of long-range interacting systems (Oxford University Press, 2014)
  • (22) P. Hertel and W. Thirring, Commun. Math. Phys. 24, 22 (1971)
  • (23) J.M. Lévy-Leblond, J. Math. Phys. 10, 806 (1969)
  • (24) P. Hertel, H. Narnhofer and W. Thirring, Commun. Math. Phys. 28, 159 (1972)
  • (25) J. Messer, Z. Physik 33, 313 (1979)
  • (26) B. Baumgartner, Commun. Math. Phys. 48, 207 (1976)
  • (27) P. Hertel, Acta Phys. Austr. Suppl. 17, 209 (1977)
  • (28) H. Narnhofer and G.L. Sewell, Commun. Math. Phys. 71, 1 (1980)
  • (29) H. Narnhofer and G.L. Sewell, Commun. Math. Phys. 79, 9 (1981)
  • (30) W. Braun and K. Hepp, Commun. Math. Phys. 56, 101 (1977)
  • (31) P. Hertel and W. Thirring, Thermodynamic Instability of a System of Gravitating Fermions. In: H.P. Dürr (Ed.): Quanten und Felder (Brauschweig: Vieweg 1971)
  • (32) J. Messer, J. Math. Phys. 22, 2910 (1981)
  • (33) J. Messer, Phys. Lett. 83A, 304 (1981)
  • (34) W. Thirring, Z. Physik 235, 339 (1970)
  • (35) E.B. Aronson, C.J. Hansen, Astrophys. J. 177, 145 (1972)
  • (36) R.D. Carlitz, Phys. Rev. D 5, 3231 (1972)
  • (37) S.W. Hawking, Phys. Rev. D 13, 191 (1976)
  • (38) P.H. Chavanis, B. Denet, M. Le Berre and Y. Pomeau, preprint
  • (39) N. Bilić and R.D. Viollier, Phys. Lett. B 408, 75 (1997)
  • (40) P.H. Chavanis, J. Sommeria, Mon. Not. R. Astron. Soc. 296, 569 (1998)
  • (41) P.H. Chavanis, Phys. Rev. E 65, 056123 (2002)
  • (42) P.H. Chavanis, The self-gravitating Fermi gas, in Dark Matter in Astro- and Particle Physics, edited by H.V. Klapdor-Kleingrothaus and R.D. Viollier (Springer, 2002)
  • (43) P.H. Chavanis, I. Ispolatov, Phys. Rev. E 66, 036109 (2002)
  • (44) P.H. Chavanis, M. Rieutord, Astron. Astrophys. 412, 1 (2003)
  • (45) P.H. Chavanis, Phys. Rev. E 69, 066126 (2004)
  • (46) D. Lynden-Bell, Mon. Not. R. Astron. Soc. 136, 101 (1967)
  • (47) N. Bilić and R.D. Viollier, Eur. Phys. J. C 11, 173 (1999)
  • (48) P.H. Chavanis, G. Alberti, in preparation
  • (49) J.R. Oppenheimer, G.M. Volkoff, Phys. Rev. 55, 374 (1939)
  • (50) R. Genzel, F. Eisenhauer, S. Gillessen, Rev. Mod. Phys. 82, 3121 (2010)
  • (51) N. Bilić and R.D. Viollier, Gen. Rel. Grav. 31, 1105 (1999)
  • (52) H. Poincaré, Acta Math. 7, 259 (1885)
  • (53) J. Katz, Mon. Not. R. Astron. Soc. 183, 765 (1978)
  • (54) J. Katz, Mon. Not. R. Astron. Soc. 189, 817 (1979)
  • (55) H. Cohn, Astrophys. J. 242, 765 (1980)
  • (56) T. Padmanabhan, Astrophys. J. Supp. 71, 651 (1989)
  • (57) P.H. Chavanis, Astron. Astrophys. 381, 340 (2002)
  • (58) D. Lynden-Bell, P.P. Eggleton, Mon. Not. R. Astron. Soc. 191, 483 (1980)
  • (59) S. Inagaki, D. Lynden-Bell, Mon. Not. R. Astron. Soc. 205, 913 (1983)
  • (60) C. Sire, P.H. Chavanis, Phys. Rev. E 69, 066109 (2004)
  • (61) Z. Roupas, Class. Quantum Grav. 32, 135023 (2015)
  • (62) R.D. Sorkin, R.M. Wald, Z.Z. Jiu, Gen. Relativ. Gravit. 13, 1127 (1981)
  • (63) P.H. Chavanis, Astron. Astrophys. 381, 709 (2002)
  • (64) P.H. Chavanis, Astron. Astrophys. 483, 673 (2008)
  • (65) Ya. B. Zel’dovich, Soviet Physics JETP 15, 446 (1962)
  • (66) Z. Roupas, P.H. Chavanis, in preparation
  • (67) P.H. Chavanis, Astron. Astrophys. 556, A93 (2013)
  • (68) C. Sire, P.H. Chavanis, Phys. Rev. E 66, 046133 (2002)
  • (69) D. Sugimoto, E. Bettwieser, Mon. Not. R. Astron. Soc. 204, 19 (1983)
  • (70) D. Heggie, N. Ramamani, Mon. Not. R. Astron. Soc. 237, 757 (1989)
  • (71) J. Binney, S. Tremaine, Galactic Dynamics (Princeton Series in Astrophysics, 1987)