Brought to you by:
Paper

Truncated γ-exponential models for tidal stellar systems

and

Published 8 April 2014 © 2014 IOP Publishing Ltd and SISSA Medialab srl
, , Citation Y J Gomez-Leyton and L Velazquez J. Stat. Mech. (2014) P04006 DOI 10.1088/1742-5468/2014/04/P04006

1742-5468/2014/4/P04006

Abstract

We introduce a parametric family of models in order to characterize the properties of astrophysical systems in quasi-stationary evolution under the influence of evaporation. We start from a one-particle distribution fγ q, pβ, εe that is based on an appropriate deformation of Maxwell–Boltzmann form with inverse temperature β and, in particular, a power-law truncation at the escape energy εe with exponent γ > 0. This deformation is implemented using a generalized γ-exponential function obtained from the fractional integration of an ordinary exponential. As shown in this work, this proposal generalizes models of tidal stellar systems that predict particle distributions with isothermal cores and polytropic haloes, e.g. Michie–King models. We perform an analysis of the thermodynamic features of these models and their associated distribution profiles. A nontrivial consequence of this study is that profiles with isothermal cores and polytropic haloes are only obtained for low energies when the deformation parameter γ < γc ≃ 2.13.

Export citation and abstract BibTeX RIS

1. Introduction

Gravitation is an example of an attractive interaction that is unable to confine particles. It is always possible that some particles acquire sufficient energy (via encounters or any other mechanism) and escape from the gravitational influence of the system. Evaporation is an unavoidable process that drives the dynamical evolution of astrophysical systems1. As discussed elsewhere [1]–[8], the presence of evaporation is a very important ingredient in explaining the behavior of stellar structures observed in Nature. The effects of evaporation crucially depend on the relaxation mechanisms that are present in the microscopic dynamics, which depend on the concrete conditions of a given particular system. Among such mechanisms, one could mention (i) particle collisions [3]–[5], which are present in stellar systems that are sufficiently dense, such as globular clusters; and (ii) general mechanisms like parametric resonance, which is the origin of the chaotic behavior of microscopic dynamics in nonlinear many-body Hamiltonian systems with bound motions in configuration space [9].

We propose in this work a parametric family of astrophysical models that account for the influence of evaporation under different relaxation regimes: the truncated γ-exponential models. These phenomenological models constitute a suitable generalization of some models of tidal stellar systems available in the literature, such as the Wooley and Dickens truncated isothermal model [8], as well as King's models of globular clusters [4]. This proposal enables us to analyze the influence of the evaporation on the thermodynamic behavior and distribution profiles. Moreover, it provides a unification framework for the known isothermal and polytropic profiles considered in hydrodynamic models of stellar systems. Remarkably, these models are sufficiently tractableto allow an analytical derivation of most of their associated hydrodynamic quantities.

The paper is organized into sections as follows. Section 2 is devoted to introducing the parametric family of quasi-stationary distributions, and the derivation of their associated hydrodynamic quantities, as well as some details about the integration of the associated nonlinear Poisson equation. Afterwards, a third section is devoted to discussing numerical results obtained from this model: information concerning thermodynamical descriptions as well as distribution profiles. Some conclusions are drawn in the fourth section. Finally, mathematical properties concerning the so-called γ-exponential function are discussed in the appendix.

2. A parametric family of models

Both observational evidence and theoretical analysis suggest that the distribution profiles of many stellar systems can be explained by considering a quasi-stationary one-particle distribution with a mathematical form close to a Maxwell–Boltzmann distribution [10]:

Equation (1)

where ε(q, p) = p2/2m + q is the mechanical energy of a given particle with momentum p located at the position q with a mean field gravitational potential ϕ q. The presence of evaporation introduces a deformation of the above distribution—specifically, a truncation for energies above a certain escape energy εe. Two examples of distributions that introduce this kind of deformation are the truncated isothermal model proposed by Wooley and Dickens [8]:

Equation (2)

and the Michie–King models [3, 4]:

Equation (3)

which vanish for energies ε(q, p) > εe, where A represents a normalization constant. The first of these corresponds to a Maxwell–Boltzmann distribution (1) that is discontinuously truncated at the escape energy εe, while the second one considers a progressive vanishing at this point.

Our interest is in proposing a generalization of the evaporation truncation scheme considered in the above astrophysical models. Firstly, let us express them in a unifying form as follows:

Equation (4)

where x = β εeε(q, p), while Eγ(x) is a certain function obtained from the truncation of the exponential function exp(x), with γ being a certain parameter. In particular, distributions (2) and (3) are described by the functions

Equation (5)

These two truncation schemes can be generalized to any nonnegative integer γ = n by subtracting the first n terms of the power expansion of the exponential function:

Equation (6)

It is worthy of note that this ansatz contains the functions (5) as particular cases with γ = 0 and γ = 1, which now guarantees the vanishing of its n − 1 first derivatives at x = 0:

Equation (7)

For positive values of the argument x, the function (6) can be rewritten into a more convenient integral representation as follows:

Equation (8)

This integral representation enables a direct extension of the above truncation scheme to any nonnegative real number γ, on replacing the factorial (n − 1)! by the known Gamma function Γ(γ):

Equation (9)

The resulting function

Equation (10)

modifies the usual exponential function using γ as a deformation parameter. Hereinafter, this function will be referred to as a γ-exponential function, and it will be equivalently denoted as Eγ(x) or E x; γ. Definition (10) is simply the Riemann–Liouville fractional integral of the ordinary exponential function [20]. Considering the fractional differentiation rule for the power-function series:

Equation (11)

conditions (7) are now generalized as follows:

Equation (12)

where the order of the fractional differentiation α belongs to the interval 0 ≤ α < γ. Readers can find some useful properties of the γ-exponential function E x; γ in the appendix.

The proposed family of quasi-stationary distributions (4) describe the influence of evaporation with two independent parameters: the escape energy εe and the deformation parameter γ. This family of distributions is shown schematically in figure 1. They also exhibit a Maxwell–Boltzmann profile (1) for energies ε(q, p) that are sufficiently below the escape energy εe. Besides this, they show a power-law truncation at εe with exponent γ. The increasing of the deformation parameter γ characterizes a larger deviation from the Maxwell–Boltzmann profile (1). From a phenomenological viewpoint, a larger value of the deformation parameter γ describes a stronger influence of the evaporation, or equivalently, a weaker influence of the relaxation mechanism. As shown below, the ansatz (4) can describe distribution profiles with isothermal cores and polytropic haloes, where the deformation parameter γ determines the polytropic structure of haloes. The escape energy εe = e can be derived from the so-called tidal potential ϕe, which accounts for the gravitational influence of neighboring systems [1]–[3], [5]. According to this energetic truncation, a particle that is trapped under the system gravitational influence will be confined inside a finite region of the space, where the gravitational potential ϕ(q) fulfills the inequality ϕ(q) ≤ ϕe. For a system with spherical symmetry, the boundary of this region is the sphere of radius Rt:

Equation (13)

which is referred to as the tidal radius [1], with M = Nm being the total mass of the astrophysical system.

Figure 1.

Figure 1. One-particle distributions versus mechanical energy ε(q, p) corresponding to a truncated γ-exponential model (4) for some values of the deformation parameter γ. We have employed pre-fixed values εe = −1 and β = 1 for the escape energy and the inverse temperature, respectively.

Standard image High-resolution image

Let us now obtain some hydrodynamic observables associated with the family of distributions (4). Introducing the dimensionless variable τ = βp2/2m as well as the dimensionless potential Φ = Φ q:

Equation (14)

the particle density n(q) can be calculated as follows:

Equation (15)

where A' = 2πA 2m/β3/2. The integration can be performed by using the convolution formula (A.11) with parameter ν = 3/2, which leads to an expression for the particle density in terms of the γ-exponential function as follows:

Equation (16)

Here, C = A 2/β3/2 and $\Gamma (3/2)=\sqrt {\pi }/2$, in accordance with the gamma function for half-integer values:

Equation (17)

The dependence of particle density on the dimensionless potential, equation (17), is shown in figure 2. Another important observable is the kinetic energy density υ(q):

Equation (18)

which is also expressed in terms of the γ-exponential function using the convolution formula (A.11) with parameter ν = 5/2:

Equation (19)

Finally, we can introduce the magnitude epsilon(q) = υ(q)/n(q):

Equation (20)

which represents the kinetic energy per particle.

Figure 2.

Figure 2. Dependence of the particle density n(Φ, γ) on the dimensionless potential Φ for some values of the deformation parameter γ. One can notice the asymptotic convergence of all these dependences towards an exponential growth for large values of Φ, n(Φ, γ) ∝exp(Φ), which is related to a dominance of isothermal conditions. We show these same dependences in the inset panel, in such a way as to show the power-law behavior for small values of Φ, n(Φ) ∝ Φγ, which is related to a dominance of polytropic conditions as a consequence of evaporation. This type of dependence can describe distributions with isothermal cores and polytropic haloes.

Standard image High-resolution image

According to the kinetic theory for ideal gases, this magnitude epsilon(q) can be related to the fluid pressure p(q) as follows:

Equation (21)

The hydrostatic equilibrium arises in a fluid when the gravitational force and pressure gradient are balanced as follows:

Equation (22)

with p and ρ being its pressure and mass density, respectively. The gravitation intensity vector g(q) can be expressed in terms of the gravitational potential ϕ(q):

Equation (23)

Considering the total differentiations for Φ(q) and p(q):

Equation (24)

the hydrodynamic equilibrium condition (22) can also be rewritten as

Equation (25)

where n Φ; γ and p Φ; , γ are the particle density and the pressure expressed in terms of the dimensionless potential Φ. Considering the differentiation identity (A.9), it is straightforwardly verified that the hydrodynamic quantities (16) and (21) derived from the truncated γ-exponential model satisfy the condition for hydrostatic equilibrium (22).

The mass density ρ(q) = mn(q) and gravitational potential ϕ(q) are related via the Poisson equation, Δϕ(q) = 4πGρ(q). Restricting to situations with spherical symmetry, one obtains the following nonlinear Poisson problem in terms of the dimensionless potential (14):

Equation (26)

We have considered here the dimensionless radius

Equation (27)

where r is the physical radius, κ = Gm2βCR2 is an auxiliary constant, and Rt is the tidal radius. Problem (26) can be solved numerically by demanding the following conditions at the origin:

Equation (28)

and integratinguntil the vanishing of a dimensionless potential at the surface of the system with dimensionless radius $\xi _{\rm c}=\sqrt {\kappa }$ is achieved:

Equation (29)

This constant allows us to express the mass density ρ in units of the characteristic density $\rho _{\rm c} = M/R^{3}_{t}$ as follows:

Equation (30)

where η is the dimensionless inverse temperature:

Equation (31)

The total energy U = K + V can be obtained from the total kinetic K and the total potential energy V :

Equation (32)

which are expressed in units of the characteristic energy Uc = GM2/Rt.

3. Results and discussion

3.1. Generalities

Many studies concerning astrophysical systems and cosmological problems [1, 5] start from assuming a polytropic dependence between the pressure p and the particle density ρ:

Equation (33)

where C is a certain constant and γ is the so-called polytropic index. The phenomenological state equation (33) can be combined with the condition for hydrostatic equilibrium (22) to derive a power-law relation between the density and the dimensionless potential Φ:

Equation (34)

where K = (γ− 2)/γmβCn and the exponent n ≡ 1/(γ− 1) ≥ 0. The marginal case γ = 1 is fully permitted. The condition for hydrostatic equilibrium now predicts an exponential dependence between ρ and Φ:

Equation (35)

where K is a certain integration constant. This behavior can be associated with a system under isothermal conditions, on imposing the constraint C ≡ 1/, where m is the mass of the constituent particles and β is the inverse temperature. The ansatz (4) proposed in this study allows us to consider the above dependences in a unifying fashion. According to the asymptotic behavior of the truncated γ-exponential function (A.7), the particle density (30) describes a power-law dependence for Φ small with exponent n = γ + 3/2, while it describes an exponential dependence for Φ large enough:

Equation (36)

Accordingly, the deformation parameter γ can be related to the polytropic index γ of the polytropic state equation (33) as γ = (2γ + 5)/(2γ + 3). According to the nonlinear Poisson problem (26), the dimensionless potential Φ decreases from the inner region towards the outer region of the system. This means that the truncated γ-exponential models could describe distribution profiles with isothermal cores and polytropic haloes. Since the particle density (30) should vanish at the system surface, where the dimensionless potential Φ(ξc) = 0, one should expect the external dimensionless radius ξc to be finite: 0 < ξc < +. However, polytropic profiles with exponent n > 5 are infinitely extended in space and exhibit an infinite mass: ξc→ + and M→ + . Such profiles cannot describe any realistic situation [5]. On the other hand, a finite character of the quasi-stationary distribution function fqe(q, p) is only possible if the deformation parameter γ ≥ 0. Consequently, the admissible values of the deformation parameter γ must be restricted to the following interval:

Equation (37)

Accordingly, these models can only describe polytropic dependences with exponent n restricted to the interval 3/2 ≤ n < 5. As shown below, this restriction will be manifested in both thermodynamic quantities and distribution profiles.

The nonlinear Poisson problem (26) is integrated using γ and Φ0 as independent integration parameters. This task was accomplished using the Runge–Kutta fourth-order method, which was implemented using FORTRAN 90 programming. Results from this integration are shown in figure 3—in particular, the dependences of ξc and η versus the central value of the dimensionless potential Φ0 for some values of the deformation parameter γ. According to these results, all dependences of the dimensionless radius ξc diverge when the dimensionless potential Φ0 approaches zero, which means that the admissible values of the parameter Φ0 are nonnegative. Curiously, the dependences of the dimensionless radius ξc corresponding to the deformation parameter γ = 2.5 and γ = 3.0 diverge at a certain value $\Phi ^{\infty }_{0}$ of the dimensionless potential Φ0 that depends on the deformation parameter γ, while the corresponding dependences of the dimensionless inverse temperature η simultaneously vanish. This means that the values of the dimensionless potential Φ0 above the point $\Phi ^{\infty }_{0}$ are also nonphysical. A better understanding of the physical meaning of the behaviors observed in the dependences of figure 3 is achieved by analyzing the dependences of the thermodynamic quantities.

Figure 3.

Figure 3. Behavior of some numerical parameters ξc and η versus the central value of the dimensionless potential Φ0 derived from the integration of the nonlinear Poisson problem (26).

Standard image High-resolution image

3.2. The thermodynamical behavior

Dependences of the inverse temperature η and central particle densities ρ(0) versus the dimensionless energy U are shown in figure 4 for different values of the deformation parameter γ. All quasi-stationary configurations obtained from these models always have negative values for the energy U. Moreover, one can recognize the existence of three notablepoints:

  • The critical point of gravothermal collapse UA: the quasi-stationary configuration with minimum energy UA. There are no quasi-stationary configurations for energies below this point. If a system is initially prepared with an energy below this point, it will experience an instability process that leads to a sudden contraction of the system under its own gravitational field, a phenomenon commonly referred to in the literature as gravothermal collapse [21].
  • The critical point of isothermal collapse UB: the quasi-stationary configuration with minimum temperature TB. There are no stable configurations for temperatures T < TB. If a system under isothermal conditions (in the presence of a thermostat at constant temperature) is initially put in thermal contact with a heat reservoir with T < TB, this system will experience an instability process fully analogous to gravitational collapse, which is referred to as isothermal collapse [21]. This type of thermodynamical instability is less relevant than the gravothermal collapse because considering the presence of a thermostat is actually unrealistic in most of astrophysical situations.
  • The critical point of evaporation disruption UC: the quasi-stationary configuration with maximal energy, that is, there are no quasi-stationary configurations for energies U > UC. The existence of this superior bound is a direct consequence of the presence of evaporation, which imposes a maximum value for the individual mechanical energies of the system constituents, ε < εe. If the system is initially prepared with an energy above this, it will experience a sudden evaporation in order to release its excess of energy [7]. Note that the inverse temperature ηC always vanishes at this point regardless of the value of the deformation parameter γ.
Figure 4.

Figure 4. Behavior of some thermodynamic quantities for some values of the deformation parameter γ. Panel (a): caloric curves η versus U; panel (b): central particle density ρ0 versus U. Additionally, we have included the following notable points: (squares) the critical point of gravothermal collapse UA; (circles) the critical point of isothermal collapse UB; (triangles) the critical point of evaporation disruption UC.

Standard image High-resolution image

All quasi-stationary configurations are located inside the energy region UAUUC. Moreover, there is more than oneadmissible value for the dimensionless inverse temperature η for a given energy U near the point of gravothermal collapse UA. According to results shown in figure 5, stable quasi-stationary configurations belong to the superior branch A–B–C, since these configurations exhibit a higher value of the entropy S for a given total energy. The energy dependence of this thermodynamic potential was evaluated from numerical integration of the expression

Equation (38)

which employs as a reference the value SC corresponding to the critical point of evaporation instability UC. Stable quasi-stationary configurations inside the energy range UAUUB exhibit negative heat capacities C < 0. The existence of this thermodynamic anomaly is a remarkable consequence of the long-range character of gravitation—in particular, because of the short-range divergence of its interaction potential energy:

Equation (39)

when the particle separation distance rjri drops to zero. While such configurations are unstable if the system is put into thermal contact with an environment at constant temperature (canonical ensemble), they are stable if the system is put into energetic isolation (microcanonical ensemble).

Figure 5.

Figure 5. Energy dependence of the entropy difference ΔS = SSC obtained from numerical integration using definition (38). Only the upper branches of the microcanonical caloric curves shown in figure 4 correspond to stable quasi-stationary configurations, because they exhibit higher entropy.

Standard image High-resolution image

To provide a better understanding of the influence of the deformation parameter γ, we have calculated the dependences of some thermodynamic quantities at notable points. Results are shown in figure 6. It is clearly evident that increase of the deformation parameter γ provokes a systematic decrease of the inverse temperatures (ηA, ηB), and an increase of the absolute values of the energies UA, UB, UC and their associated central particle densities ρ0A, ρ0B, ρ0C. Interestingly, the inverse temperature ηA at the notable point of gravothermal collapse vanishes when γγc ≃ 2.1. This means that both the total energy UA and the temperature TA diverge at this point as a consequence of the divergence of the central density ρA. To be precise, the existence of this divergence was also manifested as a divergence of the dimensionless radius ξc, which was shown in figure 3 for the particular cases with deformation parameters γ = 2.5 and γ = 3.0.

Figure 6.

Figure 6. Thermodynamic quantities associated with the notable points versus the deformation parameter γ: panel (a), the inverse temperatures ηA, ηB and the modulus of energies UA, UB, UC; panel (b), the central particle densities ρ0A, ρ0B, ρ0C. Inset panel: fit of the dependence ηA(γ) near the critical value γc using the power-law form (40).

Standard image High-resolution image

The fact that the energy of gravothermal collapse diverges, UA→−, when γγc significantly reduces the dramatic character of this phenomenon. For admissible values of the deformation parameter γ below the point γc, the system develops a gravothermal collapse at the finite energy UA, which should evolve in a discontinuous way towards a certain collapsed structure that is not describable with the present model. For admissible values of the deformation parameter γ above the point γc, the system should release an infinite amount of energy to reach a collapsed structure with a divergent central density associated with the point of gravothermal collapse. However, this collapsed structure is now described within the present models and the transition is developed in a continuous way with the decreasing of the internal energy.

It is noteworthy that an analogous divergence is observed in thermodynamic parameters of other notable points when the deformation parameter γ approaches its maximum admissible value γm = 3.5. As expected, this second divergence point is related to the nonphysical character of polytropic dependences when n > 5. A more precise estimation for the critical value γc can be obtained from considering an adjustment of the dependence ηA(γ)with a power-law form:

Equation (40)

As shown in the inset panel of figure 6, the proposed form (40) exhibits a great agreement with numerical results for the following set of parameters: A = 0.292 ± 0.004, p = 1.24 ± 0.01 and γc = 2.1307 ± 0.003.

3.3. Distribution profiles

We show in figure 7 a distribution profile with deformation parameter γ = 1 at the point of gravothermal collapse, which corresponds to the Michie–King profile with lowest energy. As a consequence of gravitation, the highest concentration of the particles is always located in the inner region of the system, while the particle density gradually decays with increase of the radius r until it vanishes at the tidal radius Rt. As clearly shown in this figure, the inner region—the core—can be well-fitted with an isothermal profile [19], while the outer one—the halo—is well-fitted with a polytropic profile [5].

Figure 7.

Figure 7. Panel (a): truncated γ-exponential profile with γ = 1 (Michie–King profile) corresponding to the point of gravothermal collapse and its comparison with isothermal and polytropic profiles using log–log scales. Panel (b): the same dependences but now using linear–log scales, to allow appreciation of the polytropic fit of the halo. Accordingly, the proposed family of models can describe distribution profiles that exhibit isothermal cores and polytropic haloes.

Standard image High-resolution image

Dependences of the distribution profiles on the deformation parameter γ and the internal energy U are illustrated in figure 8—specifically, twelve profiles corresponding to three notable points and four different values of the deformation parameter γ. The particle concentration in the inner regions decreases with increase of the internal energy U. Increase of the deformation parameter γ produces distribution profiles with more dilute haloes, and consequently, denser cores. Not all admissible profiles derived from the present family of models can exhibit isothermal cores. However, all these profiles exhibit polytropic haloes. In fact, distribution profiles near the point of evaporation disruption can be regarded as everywhere polytropic with high accuracy. Finally, distribution profiles corresponding to the point of gravothermal collapse with γ > γc are divergent at the origin.

Figure 8.

Figure 8. Dependence of the distribution profiles on the deformation parameter γ for three notable values of the internal energy U. While increase of the deformation parameter γ produces profiles with denser cores and more dilute haloes, increase of the energy produces the opposite effect. Interestingly, the qualitative forms of the haloes are the same for a given value of the deformation parameter γ. Panels (a)–(c): distribution profiles at the energy of gravothermal collapse UA for three values of the deformation parameter with γ < γc. All of them exhibit isothermal cores and polytropic haloes. Panel (d): a distribution profile at gravothermal collapse with deformation parameter γ > γc. Note that this profile exhibits not an isothermal core, but a divergence in the central density. Panels (i)–(l): distribution profiles very near the energy of evaporation disruption UC, which are everywhere polytropic. Panels (e)–(h): transitional profiles at the energy of isothermal collapse UB. These profiles hardly differ from the polytropic profiles of (i)–(l) because they exhibit denser cores.

Standard image High-resolution image

We have shown in figure 9 the phase diagram of truncated γ-exponential models in the plane of integration parameters γΦ0 of the nonlinear Poisson problem (26). For each value of the deformation parameter γ, the admissible values of the central dimensionless potential Φ0 are located inside the interval 0 ≤ Φ0ΦA(γ), where ΦA(γ) corresponds to the critical point of gravothermal collapse UA. Central values of the dimensionless potential Φ0 above the dependence ΦA(γ) correspond to nonphysical or unstable configurations (white region). Additionally, we have included the dependence ΦB(γ) associated with the point of isothermal collapse UB. Configurations between the dependences, ΦB(γ) ≤ Φ0ΦA(γ), exhibit negative heat capacities (dark gray region), while those with the central dimensionless potential Φ0 belonging to the interval 0 < Φ0 < ΦB(γ) exhibit positive heat capacities (light gray region). It is remarkable that the dependence ΦA(γ) is weakly modified by a change in the deformation parameter γ for values below the critical point γc. However, this function experiences a sudden change above this critical value. As expected, this behavior accounts for the sudden change in behavior of the distribution profiles: the proposed models can exhibit profiles with isothermal cores for γ < γc, while they only exhibit profiles without isothermal cores for γγc.

Figure 9.

Figure 9. Phase diagram of truncated γ-exponential models in the plane of integration parameters γΦ0. The dark gray region corresponds to quasi-stationary configurations with negative heat capacities, while the light gray region corresponds to positive heat capacities. White regions are nonphysical or unstable configurations. We have emphasized inside the dark gray region those profiles that exhibit isothermal cores and polytropic haloes. Configurations corresponding to isothermal collapse are always outside this region. Quasi-stationary configurations on the line of gravothermal collapse (thick solid line) exhibit a weak dependence on the deformation parameter γ if γ < γc. This dependence exhibits a significant change for γcγγm, which accounts for the divergence in the central density.

Standard image High-resolution image

According to equation (36), profiles with isothermal cores are directly related to asymptotic dependence of the particle distribution following an exponential law with regard to the local value of the dimensionless potential Φ(ξ). Such an asymptotic behavior is better described in terms of the deviation function δ(x, γ) with respect to the exponential function, which is introduced in the appendix. For the sake of convenience, we have denoted the dependence Φic(γ) as follows:

Equation (41)

where the convergence error epsilon was fixed at the value epsilon = 1.6 × 10−4. This small value guarantees the matching of this dependence at the critical value of the deformation parameter γc with the central dimensionless potential Φ0 associated with the point of gravothermal collapse, Φic(γc) = ΦA(γc). Quasi-stationary configurations located inside the region with Φic(γ) ≤ Φ0ΦA(γ) and 0 < γ < γc exhibit isothermal cores, that is, the inner regions of the particle distribution can be fitted with an isothermal profile.

4. Summary and open questions

We have introduced truncated γ-exponential models to characterize the properties of astrophysical systems in quasi-stationary evolution under the influence of evaporation. Our proposal generalizes models of tidal stellar systems available in the literature, such as Michie–King models. These models exhibit many features observed in other astrophysical models. Due to the truncation of the particle distribution in the configuration space fqe(q, p), the distribution of particles in the physical space is located inside a finite region limited by the tidal radio Rt. Moreover, the total energy U is also restricted to a finite region UAUUC. The lower bound UA represents the energy of gravitational collapse, while the upper one UC is the energy of evaporation disruption. Caloric curves for these models exhibit an anomalous branch with negative heat capacities for energies UAUUB, where UB is the value of the energy corresponding to the point of isothermal collapse, where the system reaches a minimum temperature TB. Distribution profiles with lower energy could exhibit isothermal cores and polytropic haloes with exponent n = γ + 3/2. Moreover, polytropic profiles are always obtained for energies near the point of evaporation disruption UC. The admissible values of the deformation parameter γ are restricted to the interval 0 ≤ γ < γm = 7/2. This means that this family of models describes polytropic profiles with exponent n inside the interval 3/2 ≤ n < 5. A nontrivial result obtained from these models is that the existence of distribution profiles with isothermal cores is no longer possible if the deformation parameter γγc ≃ 2.13. The existence of a notable value γc for deformation parameter indicates a drastic change in the behavior of the thermodynamic quantities and distribution profiles. Specifically, these cases exhibit a simultaneous divergence of the energy of gravitational collapse UA, its corresponding temperature TA and the central density ρ0A. For larger energies, these values of the deformation parameter γ describe distribution profiles with an almost polytropic form.

Before ending this section, let us briefly refer to some open questions. A realistic improvement of the present models would be the consideration of the mass spectrum for the constituent particles, which would enable us to study the influence of mass segregation on the thermodynamic behavior and distribution profiles [7]. A second improvement would be the consideration of factors that lead to a breakdown of spherical symmetry, such as system rotation or the tidal field of a neighboring astrophysical system [11]–[15]. Finally, we shall perform a comparison of results obtained from these models with experimental data—in particular, particle distributions of tidal stellar systems such as globular clusters and elliptical galaxies. These problems will be analyzed in forthcoming works.

Acknowledgments

Velazquez thanks for financial support CONICyT/Programa Bicentenario de Ciencia y Tecnología PSD 65. He also wishes to express thanks for partial financial support from the VRIDT-UCN research program.

Appendix.: The γ-exponential function

The γ-exponential function E x; γ is defined with the help of the so-calledgamma distribution as follows:

Equation (A.1)

which vanishes for x < 0. This family of functions is directly related to the lower incomplete gamma function:

Equation (A.2)

which only differs from expression (A.1) because of the factor A s, x = ex/Γ s. Moreover, this can also be obtained applying the fractional integral operator [20]:

Equation (A.3)

to the exponential function. Changing the integration variable, txt, in this last definition, one obtains

Equation (A.4)

which reduces to expression (A.1) when f(t) = exp(t).

Performing the integration by parts, one can obtain from definition (A.1) the recurrent relation

Equation (A.5)

which leads to the power expansion in equation (10). This representation allows us to obtain the following particular expressions:

Equation (A.6)

where n is any positive integer number. In general, the γ-exponential E x; γ is a continuous and differentiable function for every x > 0. This function vanishes at x = 0 whenever γ > 0, while it exhibits a discontinuity at x = 0 when the deformation parameter γ→0. It is easy to verify that the function E x; γ with γ > 0 exhibits an exponential behavior for sufficiently large x and a power-law dependence for small x:

Equation (A.7)

According to the recurrence relation (A.5), the γ-exponential function E(x, γ) always diverges at the origin x = 0 if the deformation parameter γ is negative but not an integer. The behavior of this family of functions is shown in panel (a) of figure A.1 for some values of the deformation parameter γ. Moreover, we have also considered the deviation function:

Equation (A.8)

that characterizes the relative deviation of the γ-exponential E(x, γ) from the ordinary exponential. The behavior of this latter function is illustrated in panel (b) of figure A.1.

Figure A.1.

Figure A.1. Panel (a): some particular examples of the truncated γ-exponential family in a log–log graph. One can note the asymptotic dependences: a power-law behavior for small x and an exponential growth for x large enough. Inset panel: the same dependences in a linear–linear graph for x near the origin, which allows us to appreciate the power-law character of the truncation. Panel (b): dependence of the deviation function (A.8), using a linear–linear scale. According to this figure, the relative deviation from the exponential decreases with increase of the independent variable x, while it grows with increase of the deformation parameter γ.

Standard image High-resolution image

Using the power-expansion expression (10), one can obtain the integration–differentiation rules:

Equation (A.9)

Equation (A.10)

and the convolution formula:

Equation (A.11)

This last relation is obtained using the Beta function:

Equation (A.12)

Notethat definition (A.1) is a particular case of the convolution formula (A.11) for γ = 0. Moreover, this identity accounts for the fractional integration of the γ-exponential function:

Equation (A.13)

Footnotes

  • By itself, a closed system of particles interacting via gravitation or van der Waals forces experiences the influence of evaporation. The key question here is that of the existence of a finite potential barrier to escape from that system. For a system in full equilibrium, a Maxwell distribution of velocities with finite temperature always exhibits a fraction of particles that overcome any finite potential barrier. Evaporation can be avoided by introducing an external field with an infinite potential barrier—for example, a container with impenetrable walls. While the presence of a container with impenetrable walls is a very common assumption for physical laboratories, for confining substances that are found in a gaseous phase, it is a nonphysical notion for real astrophysical situations. Therefore, astrophysical systems can only achieve quasi-stationary evolution because of the presence of evaporation.

Please wait… references are loading.
10.1088/1742-5468/2014/04/P04006