A&A 396, 315-329 (2002)
DOI: 10.1051/0004-6361:20021374
P. H. Chavanis
Laboratoire de Physique Quantique, Université Paul Sabatier, 118 route de Narbonne, 31062 Toulouse, France
Received 15 April 2002 / Accepted 12 September 2002
Abstract
We discuss the statistical mechanics of rotating
self-gravitating systems by allowing properly for the conservation
of angular momentum. We study analytically the case of slowly
rotating isothermal spheres by expanding the solutions of the
Boltzmann-Poisson equation in a series of Legendre polynomials,
adapting the procedure introduced by Chandrasekhar (1933) for
distorted polytropes. We show how the classical spiral of
Lynden-Bell & Wood (1967) in the temperature-energy plane is
deformed by rotation. We find that gravitational instability occurs
sooner in the microcanonical ensemble and later in the canonical
ensemble. According to standard turning point arguments, the onset
of the collapse coincides with the minimum energy or minimum
temperature state in the series of equilibria. Interestingly, it
happens to be close to the point of maximum flattening. We
generalize the singular isothermal solution to the case of a slowly rotating
configuration. We also consider slowly rotating configurations of
the self-gravitating Fermi gas at non-zero temperature.
Key words: hydrodynamics - instabilities - stars: formation
Recently, the statistical mechanics of self-gravitating systems has attracted considerable attention (Chavanis et al. 1996; Chavanis & Sommeria 1998; de Vega et al. 1998; Youngkins & Miller 2000; Follana & Laliena 2000; Semelin et al. 2001; Cerruti-Sola et al. 2001; Ispolatov & Cohen 2001; Chavanis 2002a; Taruya & Sakagami 2002; de Vega & Sanchez 2002; Huber & Pfenniger 2002...). This topic was introduced in the 1960s by Antonov (1962) and Lynden-Bell & Wood (1968) and further developed by Hertel & Thirring (1971), Horwitz & Katz (1978), Katz (1978) and Padmanabhan (1989) among others (see a complete list of references in the review of Padmanabhan 1990). These authors pointed out the particularity of self-gravitating systems to possess negative specific heats. They showed that this strange property is responsible for the inequivalence of statistical ensembles (microcanonical/canonical) and the occurence of giant phase transitions associated with gravitational collapse. For a long time, these topics were only discussed in the astrophysical literature and were considered as a curiosity (not to say a fallacy) by statistical mechanicians. The situation is changing lately as these properties are re-discovered for other physical systems with long-range interactions which can be studied in the laboratory (see, e.g., Gross 2001). For that reason, the statistical mechanics of self-gravitating systems comes back to fashion with new perspectives.
The statistical mechanics of self-gravitating systems is far from being completely understood and rests on simplifying idealizations. The first idealization is to enclose the system within a box so as to prevent evaporation. It is only under this condition (or by introducing more realistic truncated models) that a rigorous statistical mechanics of self-gravitating systems can be carried out. Second, for most astrophysical systems, the relaxation time by two-body encounters is much larger than the age of the universe so that a more subtle, collisionless, relaxation must be advocated to explain the structure of galaxies. This is the concept of violent relaxation formalized by Lynden-Bell in 1967. Then, it is implicitly assumed that the relaxation towards statistical equilibrium proceeds to completion, which is not necessarily the case in reality. Indeed, it is possible that the relaxation stops before the maximum entropy state is attained (see Lynden-Bell 1967; Tremaine et al. 1987; Chavanis et al. 1996). This problem of incomplete relaxation must be approached with extensive numerical simulations. Of course, this program was started long ago (e.g., van Albada 1982) but only recently are N-body simulations carefully compared with the predictions of the statistical mechanics approach (Cerruti-Sola et al. 2001; Huber & Pfenniger 2002) with variable success.
The statistical equilibrium of a non-rotating classical gas enclosed
within a box was first investigated by Antonov (1962). He worked in
the microcanonical ensemble and found that thermodynamical equilibrium
exists only above a critical energy
.
Below
that energy, the system is expected to collapse and overheat; this is
the so-called "gravothermal catastrophe'' (Lynden-Bell & Wood
1968). An isothermal collapse also occurs below a critical temperature
in the canonical ensemble. This
thermodynamical instability is closely related to the dynamical Jeans
instability (Semelin et al. 2001; Chavanis 2002a). In a recent series
of papers, we considered some extensions of the Antonov problem to the
case of self-gravitating fermions (Chavanis & Sommeria 1998; Chavanis
2002c), special and general relativity (Chavanis 2002b) and confined
polytropes (Chavanis 2002d). We also introduced a simple dynamical
model of Brownian particles in gravitational interaction (Chavanis et al. 2002;
Sire & Chavanis 2002). By adding by hands a
friction and a noise, we force the system to increase entropy
continuously, thereby avoiding the problem of incomplete
relaxation. This model can be used to test precisely some ideas of
statistical mechanics (inequivalence of ensembles, phase transitions,
gravitational instabilities, basin of attraction,...) and is
sufficiently simple to allow for a thorough analytical investigation
of the collapse regime when an equilibrium state does not exist.
In all these studies, the system is assumed to be non-rotating so that the conservation of angular momentum is trivially satisfied. The object of the present paper is to extend the statistical mechanics approach to the case of rotating self-gravitating systems. This problem has been considered previously by Lagoute & Longaretti (1996), Laliena (1999), Lynden-Bell (2000) and Fliegans & Gross (2002) with different types of models. Clearly, the most interesting situation is the case of rapidly rotating systems since a wide variety of structures can emerge as maximum entropy states (Votyakov et al. 2002). However, we shall restrict ourselves in the present paper to the case of slowly rotating systems for which the problem can be tackled analytically. We shall adapt to the case of isothermal spheres the classical procedure developed by Chandrasekhar (1933) for distorted polytropes, i.e. we shall expand the solutions of the Boltzmann-Poisson equation in terms of Legendre polynomials. A similar procedure was performed by Lagoute & Longaretti (1996) for rotating globular clusters subject to tidal forces and described by an extended Michie-King model. We believe that it is useful to consider the case of a gas enclosed within a spherical box so as to make a clear connexion with the Antonov model when the rotation is set to zero. In particular, we shall derive the expression of the thermodynamical parameters for slowly rotating isothermal spheres and show how the classical spiral of Lynden-Bell & Wood (1968) in the E-T plane is modified by rotation. We shall show that rotation advances the onset of gravothermal catastrophe in the microcanonical ensemble and delays the isothermal collapse in the canonical ensemble. Using the turning point criterion of Katz (1978), we argue that the series of equilibria becomes unstable at the point of minimum energy (in the microcanonical ensemble) or at the point of minimum temperature (in the canonical ensemble). Interestingly, these instabilities happen to be close, in each ensemble, to the point of maximum flattening. We generalize the singular isothermal solution to the case of a slowly rotating configuration. We also consider the case of slowly rotating self-gravitating fermions. This system exhibits phase transitions between "gaseous'' states with an almost uniform distribution of matter and "condensed'' states with a core-halo structure. By cooling below a critical temperature, an almost nonrotating gaseous sphere can collapse into a rotating "fermion ball'' containing a large fraction of mass and angular momentum.
Consider a system of N particles, each of mass m, interacting via
Newtonian gravity. The system is enclosed within a spherical box of
radius R which preserves the rotational symmetry of the original
Hamiltonian. We allow the system to have a non vanishing angular
momentum. Let
denote the distribution function
of the system, i.e.
gives the mass of particles whose position and velocity are in the
cell
at
time t. The integral of f over the velocity determines the spatial
density
Introducing Lagrange multipliers ,
and
for each constraint, we find that the critical points of entropy are given by
To determine the structure of rotating isothermal spheres, we first introduce the function
,
where
is the gravitational potential at r=0. Then, the density field can be written
So far, we have made no approximation regarding the value of the angular velocity. We shall now consider the case of slowly rotating structures and let
.
Assuming the following form for our solution
So far, the Aj are arbitrary. They will be determined by requiring that the gravitational potential and its radial derivative are continuous across the sphere at r=R. Now, outside the sphere the potential is given by the Laplace equation
For
,
the solution of the Emden Eq. (16) behaves like (Chandrasekhar 1942)
In spherical coordinates, the mass is given by
Using the Maxwell-Boltzmann distribution (9), we can rewrite the angular momentum (3) in the form
Quite generally, the potential energy of a self-gravitating system can be written in the form (see, e.g., Binney & Tremaine 1987)
Using Eqs. (9), (10) and (13), the distribution function of a rotating isothermal gas can be written
The value of the potential at
for a non-rotating configuration is
.
For a rotating configuration, the equation of the surface with the same value of the potential is given, to first order in v, by
Coming back to Eq. (74), we find that the flattening behaves with the distance (for a given value of )
as
The microcanonical ensemble corresponds to isolated systems
characterized by their energy
and their angular momentum
.
In order to determine the
caloric curve
for different values of
,
we need to solve Eqs. (16), (31) and (32)
numerically. Expanding the functions
,
and
in Taylor series for
,
we get
In Fig. 5, we have represented the caloric curve
.
It has a classical spiral behaviour as noted by a
number of authors in the non-rotating case. There is no equilibrium
state (i.e., no critical point of entropy) above the value
.
In that case, the system will collapse and
overheat (gravothermal catastrophe). It is also at this point that
the series of equilibria becomes unstable (saddle points of entropy) in
the microcanonical ensemble (Katz 1978). We see that rotation tends to
favour the instability, i.e., the gravothermal catastrophe occurs
sooner than in the non-rotating case (see also
Fig. 2).
![]() |
Figure 1:
Moment of inertia for a non-rotating isothermal sphere along the series of equilibria (parametrized by ![]() |
Open with DEXTER |
![]() |
Figure 2:
Normalized energy of an isothermal sphere along the series of equilibria (parametrized by ![]() ![]() |
Open with DEXTER |
![]() |
Figure 3:
Normalized inverse temperature of an isothermal sphere along the series of equilibria (parametrized by ![]() ![]() |
Open with DEXTER |
![]() |
Figure 4:
Iso-density contours of the rotating singular isothermal sphere with an angular momentum
![]() |
Open with DEXTER |
![]() |
Figure 5:
Caloric curve of rotating spheres giving the inverse temperature ![]() ![]() ![]() ![]() |
Open with DEXTER |
In Fig. 6, we have represented the angular velocity of the
system as a function of energy for different values of the angular
momentum. We observe that the curve has a spiral behaviour similar to
the
diagram, but reversed. We might expect that the
moment of inertia decreases as the system becomes more and more
concentrated, resulting in an increase of angular velocity along
the series of equilibria. This is true for moderate density contrasts
(up to
709), coinciding with the region of stability, but not
for larger density contrasts. Indeed, although the central density
tends to diverge, the mass contained in the core is low and does not
dominate the moment of inertia. Therefore, the moment of inertia and the
angular velocity have a non monotonic (in fact oscillatory) behaviour
with
and saturate to finite values
and
as
.
In Fig. 7, we plot the flattening function
defined by Eq. (76). As expected, the flattening is a monotonic function of the distance. For
,
and for
,
.
In Fig. 8, we plot the flattening at the edge of the configuration
as a function of
in the microcanonical ensemble (see Eq. (79)). We observe that the curve displays damped oscillations towards the asymptotic value
.
In particular, the flattening (by unit of
)
is maximum for
Interestingly, this value lies precisely in the range of values at which the gravothermal catastrophe sets in (compare Fig. 8 with Fig. 2 giving the energy along the series of equilibria).
![]() |
Figure 6:
Angular velocity ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 7:
Spatial dependance of the flattening function ![]() |
Open with DEXTER |
![]() |
Figure 8:
Flattening at the edge of the configuration (per unit angular momentum squared) along the series of equilibria in the microcanonical ensemble. The flattening is maximum for
![]() |
Open with DEXTER |
The canonical ensemble is characterized by the specification of the
inverse temperature
and the angular velocity
.
In
Fig. 9, we have represented the curve
for different values of
.
There is no
equilibrium state (i.e., no critical point of free energy) above the
value
.
In that case, the system will undergo an
isothermal collapse. It is also at this point that the series of
equilibria becomes unstable (saddle points of free energy) in the
canonical ensemble (Katz 1978). We see that rotation tends to delay
the instability, i.e., the isothermal collapse occurs later than in
the non-rotating case.
![]() |
Figure 9:
Caloric curve of rotating spheres giving minus the energy ![]() ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 10:
Flattening at the edge of the configuration (per unit of angular velocity squared) along the series of equilibria in the canonical ensemble. The flattening is maximum for
![]() |
Open with DEXTER |
![]() |
Figure 11:
Normalized inverse temperature of an isothermal sphere along the series of equilibria (parametrized by ![]() ![]() |
Open with DEXTER |
In Fig. 10, we plot the flattening at the edge of
the configuration
as a function of
in the
canonical ensemble (see Eq. (78)). The curve displays damped
oscillations towards the value
.
The flattening (by
unit of
)
is maximum for
Interestingly,
this value is close to the typical values at which the isothermal
collapse sets in (compare Fig. 10 with
Fig. 11 giving the inverse temperature along the
series of equilibria).
The previous results can be easily generalized to the case of self-gravitating fermions. This extension is relatively straightforward and we shall just give the main steps of the calculations. The thermodynamical parameters of a non-rotating Fermi gas at finite temperature have been calculated by Chavanis & Sommeria (1998) and we shall adopt a similar presentation. For a rotating configuration, the Fermi-Dirac distribution can be written
We shall now consider the case of slowly rotating structures and let
.
Repeating the steps of Sect. 2.2, we
find that
is given by Eq. (30) with the new functions
and
defined by
We now determine the thermodynamical parameters of a slowly rotating
self-gravitating Fermi gas. We can check that the inverse temperature
is still given by Eq. (47) and that the relation (46) remains valid. On the other hand, eliminating the
central density in Eq. (26), using Eq. (86),
we find that
The equilibrium phase diagram can be obtained in the following manner. For given k,
and
,
we can solve Eqs. (89),
(92) and (93) until the value
at which the relation (97) is satisfied. Then,
Eqs. (47) and (99) determine the temperature and the energy of the configuration. If the angular momentum is fixed instead of the angular velocity, we must use Eq. (53) with Eq. (98) to express
in terms of
.
By varying the parameter k (for a fixed value of the degeneracy parameter
), we can cover the whole diagram in parameter space. A complete description of this diagram has been given by Chavanis (2002c) in the non-rotating case.
In Fig. 12, we represent the caloric curve of self-gravitating fermions for a degeneracy parameter
and for different values of angular momentum. We observe that degeneracy has the effect of unwinding the spiral of Fig. 5. For sufficiently large values of the degeneracy parameter, there is still gravitational collapse at
accompanied by a rise of temperature, but this "gravothermal catastrophe'' stops when the core of the system becomes degenerate. This leads to the formation of a "fermion ball'' which contains a moderately large fraction of mass q M (at point D, we typically have
). In the microcanonical ensemble, the decrease of potential energy in the core is compensated by an increase of temperature. Therefore, the mass (1-q)M contained in the halo undergoes an expansion which, in our model, is arrested by the walls of the box. As a result, the density of the halo is almost uniform. Typical density profiles are given by Chavanis & Sommeria (1998) in the non rotating case. Because of the expansion of the halo, the moment of inertia of the system increases during the collapse despite the formation of a massive nucleus. Therefore, the angular velocity decreases contrary to what might have been expected. The angular velocity is represented as a function of energy in Fig. 13. It has a complicated behaviour which corresponds to the unwindement of the spiral of Fig. 6.
![]() |
Figure 12:
Equilibrium phase diagram of self-gravitating fermions giving the inverse temperature ![]() ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 13:
Angular velocity vs energy for self-gravitating fermions with a degeneracy parameter
![]() ![]() |
Open with DEXTER |
For smaller values of the degeneracy parameter, the gravitational
phase transition is suppressed (Chavanis & Sommeria 1998; Chavanis
2002c) and the caloric curve has the structure of
Fig. 14. This diagram is similar to the one found by
Fliegans & Gross (2002) in their two-dimensional model of rotating
self-gravitating systems with a relatively large cut-off radius (which
plays the role of the inverse of our degeneracy parameter). Indeed,
when the cut-off radius (or degeneracy) is sufficiently large, the
spiral of Fig. 12 unwinds and the
curve is univalued like in Fig. 14. For high energies,
the density is homogeneous. For low energies, the equilibrium states
have a core-halo structure with a partially degenerate nucleus and a
dilute envelope. As energy decreases any further, the nucleus contains
more and more mass and becomes smaller and smaller. This is a property
of the
law of degenerate configurations (like white
dwarfs). For rotating systems described in the microcanonical ensemble
(fixed E and L), the resulting decrease of moment of inertia is
accompanied by an increase of angular velocity as shown in
Fig. 15. In this diagram, the spiral of
Fig. 6 is completely unwound and the angular velocity
increases monotonically with
.
![]() |
Figure 14:
Same as Fig. 12 for
![]() |
Open with DEXTER |
![]() |
Figure 15:
Same as Fig. 13 for
![]() |
Open with DEXTER |
In Fig. 16, we have focused our attention to what
happens close to the minimum energy
corresponding to
T=0. At that point, the system has the same structure as a cold
white dwarf star (see Appendix
B). For our perturbative analysis to be valid even after
collapse, we have taken a very small angular momentum
.
We shall describe the diagram of Fig. 16 in the
(L,T) ensemble in which the angular momentum and the temperature are
fixed (this is a situation intermediate between microcanonical and
canonical ensembles). For
,
the system is in a gaseous phase
with an almost homogeneous density profile. For
,
the system
undergoes an isothermal collapse that only stops when gravity is
balanced by the degeneracy pressure. The result of this phase
transition is a "fermion ball'' which contains almost all the mass
(
)
unlike in the microcanonical ensemble at the point of
gravothermal catastrophe. These fermion balls can be of astrophysical
interest in Dark Matter models consisting of massive neutrinos (Bilic
& Viollier 1997; Chavanis 2002e).
For rotating systems, the isothermal collapse is
accompanied by a discontinuous rise of angular velocity at the
critical temperature
(see
Fig. 17). Therefore, even if the initial
rotation of the system is negligible in the gaseous phase, after
collapse the "fermion ball'' can have appreciable rotation as
suggested in Fig. 17. Its structure is then
similar to a distorted polytrope of index n=3/2 as computed by
Chandrasekhar (1933) in the limit of slow rotation.
![]() |
Figure 16:
Equilibrium phase diagram for self-gravitating fermions with a degeneracy parameter
![]() ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 17:
Angular velocity vs. temperature plot for
![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
In this paper, we have considered the effect of a small rotation on the thermodynamic stability of self-gravitating systems. We have worked in a finite spherical box in order to make a clear connexion with the Antonov problem for non-rotating systems and render the statistical mechanics of these objects rigorous. Physically, this idealization means that our isothermal system is surrounded by a medium which fixes its size. We have found that for rotating systems, the well-known inequivalence of statistical ensembles for self-gravitating systems manifests itself in a striking manner: the instability is advanced in the microcanonical ensemble and delayed in the canonical one. In addition, we have found a connexion between the onset of instability and the configuration of maximum flattening in the series of equilibria. These results have been generalized to the case of self-gravitating fermions.
On the other hand, the case of rapidly rotating isothermal configurations is interesting because new, non trivial, structures can emerge as maximum entropy states (i.e., most probable states). The classification of such structures is complicated because many bifurcations can occur depending on the values of the control parameters (M,E,L). In particular, we must be careful to select only entropy maxima, discarding the critical points of entropy which are only saddle points. This can be obtained either by solving the Boltzmann-Poisson equation (Votyakov et al. 2002) and checking the stability of the solutions or by using relaxation equations towards the maximum entropy state (Chavanis et al. 1996). These studies are important in order to obtain a classification of the most probable configurations of self-gravitating systems. Such a classification scheme is also relevant in the very similar context of two-dimensional vortices described as maximum entropy structures (e.g., Chavanis & Sommeria 1996, 1998). The idea of this classification was proposed in an earlier paper (Chavanis et al. 1996) and will be given further attention elsewhere (Chavanis & Rieutord 2002). We must be aware, however, that the statistical approach is conditioned by an ergodicity hypothesis which may not be completely fulfilled in practice. This complicated problem of incomplete relaxation demands further discussion.
Acknowledgements
I acknowledge interesting discussions with J. Katz. I have also benefited from stimulating interactions with the participants of the Les Houches School of Physics on "Dynamics and thermodynamics of systems with long range interactions'' (February 2002).
The Fermi-Dirac entropy (85) can be written in the equivalent form
In this Appendix, we derive a simple analytic formula for the potential energy of a slowly rotating polytrope of index n. The index n=3/2 describes a completely degenerate Fermi gas at zero temperature, which is a particular limit of the model studied in Sect. 5.
For a self-gravitating system rotating with constant angular velocity ,
the condition of hydrostatic equilibrium in the rotating frame can be written
From the above results, we can obtain an explicit expression for the potential energy of a slowly rotating polytrope. For an axisymmetric system, we have the relation
where I is the axial moment of inertia (51). To our order of approximation, we just need to determine the value of I for a non-rotating polytrope. Using the relations
and
,
expressing the central density as a function of R0 by the relation (B.14) and using the mass-radius relation (B.19), we obtain
We now consider the case of a non relativistic degenerate Fermi gas at
zero temperature. As is well-known, this system is equivalent to a
polytrope of index n=3/2. In addition, the constant K is
explicitly given by