fields and elementary particles

18
FIELDS AND ELEMENTARY PARTICLES ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 863 https://doi.org/10.15407/ujpe63.10.863 K.A. BUGAEV, 1 A.I. IVANYTSKYI, 2, 1 V.V. SAGUN, 1, 3 E.G. NIKONOV, 4 G.M. ZINOVJEV 1 1 Bogolyubov Institute for Theoretical Physics, Nat. Acad. of Sci. of Ukraine (14b, Metrolohichna Str., Kyiv 03680, Ukraine; e-mail: [email protected], [email protected]) 2 Department of Fundamental Physics, University of Salamanca (Plaza de la Merced s/n 37008, Spain; e-mail: [email protected]) 3 CENTRA, Instituto Superior Tยด ecnico, Universidade de Lisboa (Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal; e-mail: [email protected]) 4 Laboratory for Information Technologies, JINR (6, Joliot-Curie Str., Dubna 141980, Russia; e-mail: [email protected]) EQUATION OF STATE OF QUANTUM GASES BEYOND THE VAN DER WAALS APPROXIMATION A recently suggested equation of state with the induced surface tension is generalized to the case of quantum gases with mean-๏ฌeld interaction. The self-consistency conditions of such a model and the conditions necessary for the Third Law of thermodynamics to be satis๏ฌed are found. The quantum virial expansion of the van der Waals models of such a type is analyzed, and its virial coe๏ฌƒcients are given. In contrast to traditional beliefs, it is shown that an in- clusion of the third and higher virial coe๏ฌƒcients of a gas of hard spheres into the interaction pressure of the van der Waals models either breaks down the Third Law of thermodynamics or does not allow one to go beyond the van der Waals approximation at low temperatures. It is demonstrated that the generalized equation of state with the induced surface tension allows one to avoid such problems and to safely go beyond the van der Waals approximation. In addition, the e๏ฌ€ective virial expansion for the quantum version of the induced surface tension equation of state is established, and all corresponding virial coe๏ฌƒcients are found exactly. The explicit expressions for the true quantum virial coe๏ฌƒcients of an arbitrary order of this equation of state are given in the low-density approximation. A few basic constraints on such models which are necessary to describe the nuclear and hadronic matter properties are discussed. Keywords: nuclear matter, hadron resonance gas, induced surface tension, quantum gases, virial coe๏ฌƒcients. 1. Introduction Investigation of the equation of state (EoS) of strongly interacting particles at low temperatures is important for studies of the nuclear liquid-gas phase transition and for properties of neutron stars [1โ€“ 3]. To have a realistic EoS, one has to simultane- c โ—‹ K.A. BUGAEV, A.I. IVANYTSKYI, V.V. SAGUN, E.G. NIKONOV, G.M. ZINOVJEV, 2018 ously account for a short-range repulsive interac- tion, a medium-range attraction, and the quantum properties of particles. Unfortunately, it is not much known about the in-medium quantum distribution functions of particles which experience a strong inter- action. Therefore, a working compromise to account for all these features is to introduce the quasiparticles with quantum properties which interact via the mean ๏ฌeld. One of the ๏ฌrst successful models of such a type

Upload: others

Post on 23-Mar-2022

14 views

Category:

Documents


0 download

TRANSCRIPT

FIELDS AND ELEMENTARY PARTICLES

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 863

https://doi.org/10.15407/ujpe63.10.863

K.A. BUGAEV,1 A.I. IVANYTSKYI,2, 1 V.V. SAGUN,1, 3 E.G. NIKONOV,4

G.M. ZINOVJEV 1

1 Bogolyubov Institute for Theoretical Physics, Nat. Acad. of Sci. of Ukraine(14b, Metrolohichna Str., Kyiv 03680, Ukraine; e-mail: [email protected],[email protected])

2 Department of Fundamental Physics, University of Salamanca(Plaza de la Merced s/n 37008, Spain; e-mail: [email protected])

3 CENTRA, Instituto Superior Tecnico, Universidade de Lisboa(Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal; e-mail: [email protected])

4 Laboratory for Information Technologies, JINR(6, Joliot-Curie Str., Dubna 141980, Russia; e-mail: [email protected])

EQUATION OF STATE OF QUANTUM GASESBEYOND THE VAN DER WAALS APPROXIMATION

A recently suggested equation of state with the induced surface tension is generalized to thecase of quantum gases with mean-field interaction. The self-consistency conditions of such amodel and the conditions necessary for the Third Law of thermodynamics to be satisfied arefound. The quantum virial expansion of the van der Waals models of such a type is analyzed,and its virial coefficients are given. In contrast to traditional beliefs, it is shown that an in-clusion of the third and higher virial coefficients of a gas of hard spheres into the interactionpressure of the van der Waals models either breaks down the Third Law of thermodynamics ordoes not allow one to go beyond the van der Waals approximation at low temperatures. It isdemonstrated that the generalized equation of state with the induced surface tension allows oneto avoid such problems and to safely go beyond the van der Waals approximation. In addition,the effective virial expansion for the quantum version of the induced surface tension equationof state is established, and all corresponding virial coefficients are found exactly. The explicitexpressions for the true quantum virial coefficients of an arbitrary order of this equation ofstate are given in the low-density approximation. A few basic constraints on such models whichare necessary to describe the nuclear and hadronic matter properties are discussed.K e yw o r d s: nuclear matter, hadron resonance gas, induced surface tension, quantum gases,virial coefficients.

1. Introduction

Investigation of the equation of state (EoS) ofstrongly interacting particles at low temperatures isimportant for studies of the nuclear liquid-gas phasetransition and for properties of neutron stars [1โ€“3]. To have a realistic EoS, one has to simultane-

cโ—‹ K.A. BUGAEV, A.I. IVANYTSKYI, V.V. SAGUN,E.G. NIKONOV, G.M. ZINOVJEV, 2018

ously account for a short-range repulsive interac-tion, a medium-range attraction, and the quantumproperties of particles. Unfortunately, it is not muchknown about the in-medium quantum distributionfunctions of particles which experience a strong inter-action. Therefore, a working compromise to accountfor all these features is to introduce the quasiparticleswith quantum properties which interact via the meanfield. One of the first successful models of such a type

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

was a Walecka model [4]. However, the strong de-mands to consider a more realistic interaction whichis not restricted by some kind of effective Lagrangianled to formulating a few phenomenological general-izations of the relativistic mean-field model [5โ€“7]. Al-though a true breakthrough among them was madein work [7] in which the hard-core repulsion was sug-gested for fermions, the introduction of a phenomeno-logical attraction in the spirit of the Skyrmeโ€“Hartreeโ€“Fock approach [8, 9] which depends not on the scalarfield, but on the baryonic charge density, was alsoimportant, since such a dependence of the attractivemean-field is typical of the EoS of real gases [10].

However, in addition to the usual defect of the rel-ativistic mean-field models breaking down the firstand second Van Hove axioms of statistical mechanics[11, 12], the usage of a non-native variable, namely aparticle number density, in the grand canonical en-semble led to the formulation of self-consistency con-ditions [5, 6]. In contrast to the Walecka model [4]and its followers for which the structure of a La-grangian and the extremum condition of the sys-tem pressure with respect to each mean-field au-tomatically provide the fulfillment of the thermo-dynamic identities, the phenomenological mean-fieldEoS of hadronic matter had to be supplemented bythe self-consistency conditions [5,6,13]. The latter al-lows one to, formally, recover the first axiom of sta-tistical mechanics [11, 12] (for the more recent dis-cussion of the self-consistency conditions, see [14โ€“16]). An exception is given by the van der Waals(VdW) hard-core repulsion [7], since such an in-teraction in the grand canonical ensemble dependson the system pressure which is the native variablefor it.

Due to its simplicity, the VdW repulsion is verypopular in various branches of modern physics. Buteven in case of Boltzmann statistics, it is valid onlyat low particle densities for which an inclusion of thesecond virial coefficient is sufficient. For the classicalgases, the realistic EoSs which are able to accountfor several virial coefficients are well-known [10, 17],while a complete quantum mechanical treatment ofthe third and higher virial coefficients is rather hard[18]. Hence, the quantum EoSs with realistic inter-action allowing one to go beyond the second virialcoefficient are of great interest not only for the densehadronic and nuclear/neutron systems, but also forquantum and classical liquids. It is widely believed

that one possible way to go beyond the VdW approx-imation, i.e. beyond the second virial coefficient, is toinclude a sophisticated interaction known from theclassical models [10, 17] into the relativistic mean-field models with the quantum distribution functionsfor quasiparticles [14, 15].

On the other hand, a great success in gettinga high quality description of experimental hadronicmultiplicities measured in the central nuclear colli-sions from AGS (BNL) to LHC (CERN) energiesis achieved recently with the hadron resonance gasmodel which employs both the traditional VdW re-pulsion [19โ€“24] and the induced surface tension (IST)concept for the hard-core repulsion [25โ€“27] motivatesus to formulate and to throughly inspect the quantumversion of this novel class of the IST EoSs in order toapply it in the future to the description of the prop-erties of dense hadronic, nuclear, neutron matter anddense quantum liquids on the same footing. This isa natural choice, since the Boltzmann version of theIST EoS [25, 26] for a single sort of particles simul-taneously accounts for the second, third, and fourthvirial coefficients of the classical gas of hard spheresand, thus, allows one to go beyond the VdW approx-imation. The multicomponent formulation of such anEoS applied to a mixture of nuclear fragments withall possible sizes [28] not only allows one to introducethe compressibility of atomic nuclei into an exactlysolvable version [29] of the statistical nuclear multi-fragmentation model [30], but also it sheds light onthe reason of why this model employing the propervolume approximation for the hard-core repulsion isable to correctly reproduce the low-density virial ex-pansion for all atomic nuclei.

Therefore, the present work has two aims. First,we would like to analyze the popular quantum VdWmodels [14โ€“16] at high and low temperatures in or-der to verify whether a tuning of the interaction al-lows one to go beyond the VdW treatment. In addi-tion, we calculate all virial coefficients for the pres-sure of point-like particles of the quantum VdWEoS. Second, we generalize the recently suggestedIST EoS [25,26] to the quantum case, obtain its effec-tive virial expansion, and calculate all quantum virialcoefficients, including the true virial coefficients forthe low-density limit. Using these results, we discussa few basic constraints on the quantum EoS which arenecessary to model the properties of nuclear/neutronand hadronic matter.

864 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

The work is organized as follows. In Sect. 2 we an-alyze, the quantum VdW EoS and its virial expan-sion and discuss the pitfalls of this EoS. The quan-tum version of the IST EoS is suggested and ana-lyzed in Sect. 3. In Sect. 4, we obtain several virialexpansions of this model and discuss the Third Lawof thermodynamics for the IST EoS. Some simplestapplications to nuclear and hadronic matter EoS arediscussed in Sect. 5, while our conclusions are formu-lated in Sect. 6.

2. Quantum VirialExpansion for the VdW Quasiparticles

Similarly to the ordinary gases, the source of hard-core repulsion in the hadronic or nuclear systems isrelated to the Pauli blocking effect between the in-teracting fermionic constituents existing inside of thecomposite particles (see, e.g., [2]). This effect appearsdue to the requirement of antisymmetrization of thewave function of all fermionic constituents existingin the system. At very high densities, it may lead tothe Mott effect, i.e., to the dissociation of compositeparticles or even the clusters of particles into theirconstituents [2]. Therefore, it is evident that, at suffi-ciently high densities, one cannot ignore the hard-core repulsion or the finite (effective) size of com-posite particles. The success of traditional EoSs usedin the theory of real gases [10] based on the hard-core repulsion approach tells us that this is a fruitfulframework also for quantum systems. Hence, we startfrom the simplest case, i.e., the quantum VdW EoS[15, 16]. The typical form of EoS for quantum quasi-particles of mass ๐‘š๐‘ and degeneracy factor ๐‘‘๐‘ is asfollows:

๐‘(๐‘‡, ๐œ‡, ๐‘›๐‘–๐‘‘) = ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ(๐œ‡, ๐‘›๐‘–๐‘‘))โˆ’ ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘), (1)

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) = ๐‘‘๐‘

โˆซ๐‘‘k

(2๐œ‹3)

๐‘˜2

3๐ธ(๐‘˜)

1

๐‘’(๐ธ(๐‘˜)โˆ’๐œˆ

๐‘‡ ) + ๐œ, (2)

๐œˆ(๐œ‡, ๐‘›๐‘–๐‘‘) = ๐œ‡โˆ’ ๐‘ ๐‘+ ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘), (3)

where the constant ๐‘ โ‰ก 4๐‘‰0 = 16๐œ‹3 ๐‘…3

๐‘ is the ex-cluded volume of particles with the hard-core radius๐‘…๐‘ (here, ๐‘‰0 is their proper volume), the relativis-tic energy of particle with momentum k is ๐ธ(๐‘˜) โ‰กโ‰กโˆšk2 +m2

๐‘, and the density of point-like particles

is defined as ๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ) โ‰ก ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ)๐œ• ๐œˆ . The parameter ๐œ

switches between the Fermi (๐œ = 1), Bose (๐œ = โˆ’1),

and Boltzmann (๐œ = 0) statistics. The interactionpart of the pressure ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘) and the mean-fieldpotential ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘) will be specified later.

Note that, similarly to the Skyrme-like EoS and theEoS of real gases, it is assumed that the interactionbetween quasiparticles described by system (1)โ€“(3) iscompletely accounted by the excluded volume (hard-core repulsion), by the mean-field potential ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘),and by the pressure ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘). This is in contrastto the relativistic mean-field models of the Waleckatype in which the mass shift of quasiparticles is takeninto account. Since such an effect may be importantfor the modeling of the chiral symmetry restorationin hadronic matter (the strongest arguments of itsexistence are recently given in [27]), we leave it for afuture exploration and concentrate here on a simplerEoS defined by Eqs. (1)โ€“(3).

The functions ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘) and ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘) are notindependent, due to the thermodynamic identity๐‘›(๐‘‡, ๐œˆ(๐œ‡, ๐‘›๐‘–๐‘‘)) โ‰ก ๐œ•๐‘(๐‘‡,๐œˆ(๐œ‡,๐‘›๐‘–๐‘‘))

๐œ•๐œ‡ . Therefore, the mean-field terms ๐‘ˆ and ๐‘ƒint should obey the self-consistency condition

๐‘›๐‘–๐‘‘๐œ•๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘)

๐œ•๐‘›๐‘–๐‘‘=

๐œ•๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘)

๐œ•๐‘›๐‘–๐‘‘โ‡’ (4)

โ‡’ ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘) = ๐‘›๐‘–๐‘‘ ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘)โˆ’๐‘›๐‘–๐‘‘โˆซ0

๐‘‘๐‘›๐‘ˆ(๐‘‡, ๐‘›). (5)

After integrating by parts Eq. (4), we used the obvi-ous condition ๐‘ˆ(๐‘‡, 0) < โˆž in (5). If condition (5) isobeyed, then the direct calculation of the ๐œ‡-derivativeof the pressure (1) gives the usual expression for theparticle number density in terms of the density ofpoint-like particles

๐‘› =๐‘›๐‘–๐‘‘

1 + ๐‘ ๐‘›๐‘–๐‘‘, (6)

๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ) = ๐‘‘๐‘

โˆซ๐‘‘k

(2๐œ‹3)

1

๐‘’(๐ธ(๐‘˜)โˆ’๐œˆ

๐‘‡ ) + ๐œ. (7)

From these equations, one finds that ๐‘› โ†’ ๐‘โˆ’1 for๐‘›๐‘–๐‘‘ โ†’ โˆž. The limit ๐‘›๐‘–๐‘‘ โ†’ โˆž is provided by thecondition ๐œˆ โ†’ โˆž or ๐‘‡ โ†’ โˆž for ๐œ = {0; 1}, while, for๐œ = โˆ’1, it is provided by the condition ๐œˆ โ†’ ๐‘š๐‘ โˆ’ 0or ๐‘‡ โ†’ โˆž.

Note that, in contrast to other works discussingEqs. (4) and (5), we will use the density of point-like particles ๐‘›๐‘–๐‘‘ through this paper as an argument

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 865

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

of the functions ๐‘ˆ(๐‘‡, ๐‘›๐‘–๐‘‘) and ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘) instead ofthe physical density of particles ๐‘›, because, for moresophisticated EoSs, their relation will be more com-plicated than (6). In addition, such a representationis convenient for a subsequent analysis, because thevirial expansion of ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) in terms of ๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ) looksextremely simple [18]:

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) = ๐‘‡

โˆžโˆ‘๐‘™=1

๐‘Ž(0)๐‘™ [๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ)]

๐‘™, (8)

where

๐‘Ž(0)1 = 1, (9)

๐‘Ž(0)2 = โˆ’๐‘

(0)2 , (10)

๐‘Ž(0)3 = 4

[๐‘(0)2

]2โˆ’ 2 ๐‘

(0)3 , (11)

๐‘Ž(0)4 = โˆ’20

[๐‘(0)2

]3+ 18 ๐‘

(0)2 ๐‘

(0)3 โˆ’ 3 ๐‘

(0)4 , (12)

.................. (13)

Here, the first several virial coefficients ๐‘Ž(0)๐‘™ of an

ideal quantum gas are expressed in terms of the corre-sponding cluster integrals ๐‘(0)๐‘™>1 which depend only onthe temperature. The latter can be expressed via thethermal density of the auxiliary Boltzmann system๐‘›(0)๐‘–๐‘‘ (๐‘‡, ๐œˆ) โ‰ก ๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ)|๐œ=0 of Eq. (7) [18, 31]

๐‘(0)๐‘™ =

(โˆ“1)๐‘™+1

๐‘™๐‘›(0)๐‘–๐‘‘ (๐‘‡/๐‘™, ๐œˆ)

[๐‘›(0)๐‘–๐‘‘ (๐‘‡, ๐œˆ)

]โˆ’๐‘™

, (14)

where the upper (lower) sign corresponds to Fermi(Bose) statistics. For the non-relativistic case, expres-sion (14) can be further simplified [18]. For an arbi-trary degeneracy factor ๐‘‘๐‘, it acquires the form [31]

๐‘(0)๐‘™

nonrel

โ‰ƒ (โˆ“1)๐‘™+1

๐‘™52

(1

๐‘‘๐‘

[2๐œ‹

๐‘‡ ๐‘š๐‘

] 32

)๐‘™โˆ’1

. (15)

For high temperatures, one can write an ultra-rela-tivistic analog of Eq. (15) for a few values of ๐‘™ == 2, 3, ... โ‰ช ๐‘‡/๐‘š๐‘

๐‘(0)๐‘™

urel

โ‰ƒ (โˆ“1)๐‘™+1

๐‘™4

[๐œ‹2

๐‘‘๐‘ ๐‘‡ 3

]๐‘™โˆ’1

. (16)

Suppose that the coefficients ๐‘Ž(0)๐‘™ from Eq. (8) are

known and that the virial expansion is convergent

for the considered ๐‘‡ . Then, using Eq. (6), we find๐‘›๐‘–๐‘‘ = ๐‘›/(1โˆ’ ๐‘ ๐‘›). Hence, we can rewrite Eq. (8) as

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ)

๐‘‡ ๐‘›=

1

1โˆ’ ๐‘ ๐‘›+

โˆžโˆ‘๐‘™=2

๐‘Ž(0)๐‘™

[๐‘›]๐‘™โˆ’1

[1โˆ’ ๐‘ ๐‘›]๐‘™. (17)

Note that the expansions of such a type for a systempressure which use the variable ๐‘›/(1 โˆ’ ๐‘ ๐‘›) insteadof ๐‘› are well-known for the EoSs of hard discs [32]and hard spheres [33], since they provide a very fastconvergence of the series due to a very fast decreaseof their coefficients.

As one can see from Eqs. (15) and (16), at hightemperatures, all cluster integrals and virial coeffi-cients of the ideal quantum gas strongly decrease withthe temperature ๐‘‡ and, hence, at high temperatures,the virial expansion of ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) is defined by the first(classical) term on the right-hand side of (17). In thiscase, one gets

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ)

๐‘‡ ๐‘›โ‰ƒ 1+4๐‘‰0 ๐‘›+(4๐‘‰0 ๐‘›)

2+(4๐‘‰0 ๐‘›)3+ ... . (18)

Here, after expanding the first term on the right-hand side of (17), we used the relation between ๐‘and ๐‘‰0. From this equation, one sees that only thesecond virial coefficient, 4๐‘‰0, coincides with the onefor the gas of hard spheres, while the third, 16๐‘‰ 2

0 ,and the fourth, 64๐‘‰ 3

0 virial coefficients are essen-tially larger than their counterparts ๐ต3 = 10๐‘‰ 2

0 and๐ต4 = 18.36๐‘‰ 3

0 of the gas of hard spheres. In addi-tion, Eq. (17) can naturally explain why the authorsof work [14] insisted on the interaction pressure ๐‘ƒint

to be a linear function of ๐‘‡ (see a statement afterEq. (62) in [14]): if one chooses the interaction pres-sure in the form

๐‘ƒint(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) = ๐‘‡๐น (๐‘›(๐‘›๐‘–๐‘‘)) = ๐‘‡๐‘›ร—ร—[(๐‘2โˆ’๐ต3)๐‘›

2+ (๐‘3โˆ’๐ต4)๐‘›3+ (๐‘4โˆ’๐ต5)๐‘›

4+ ...], (19)

then, at high temperatures, the quantum correc-tions are negligible. Hence, for such a choice of๐‘ƒint(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) with the corresponding value for themean-field potential ๐‘ˆ(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) obeying the self-consistency condition (4), one can improve the totalpressure of the mean-field model by matching its re-pulsive part to the pressure of hard spheres.

The problem, however, arises at low temperatures,while calculating the entropy density for the modelwith ๐‘ƒint(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) (19). Indeed, for any choice of

866 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

the mean-field potential of the form ๐‘ˆ(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) =๐‘”(๐‘‡ )๐‘“(๐‘›(๐‘›๐‘–๐‘‘)) (note that Eq. (19) has such a form)from the thermodynamic identities ๐‘  = ๐œ•๐‘(๐‘‡,๐œ‡)

๐œ•๐‘‡ and๐‘ ๐‘–๐‘‘ = ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ)

๐œ•๐‘‡ , one finds [14]

๐‘ (๐‘‡, ๐œ‡) =

[๐‘ ๐‘–๐‘‘ +

[๐‘›๐‘–๐‘‘

๐œ•๐‘ˆ

๐œ•๐‘‡โˆ’ ๐œ•๐‘ƒint

๐œ•๐‘‡

]][1 + ๐‘ ๐‘›๐‘–๐‘‘]

โˆ’1=

(20)

=

[๐‘ ๐‘–๐‘‘ +

๐‘‘๐‘”(๐‘‡ )

๐‘‘ ๐‘‡

๐‘›๐‘–๐‘‘โˆซ0

๐‘‘๏ฟฝ๏ฟฝ ๐‘“(๐‘›(๏ฟฝ๏ฟฝ))

][1 + ๐‘ ๐‘›๐‘–๐‘‘]

โˆ’1, (21)

where, in deriving Eq. (21) from Eq. (20), we usedEq. (5) to express the interaction pressure ๐‘ƒint interms of the potential ๐‘ˆ(๐‘‡, ๐‘›(๐‘›๐‘–๐‘‘)) = ๐‘”(๐‘‡ )๐‘“(๐‘›(๐‘›๐‘–๐‘‘)).Using such an expression, one finds the followingderivative:

๐œ•๐‘ƒint

๐œ•๐‘‡= ๐‘›๐‘–๐‘‘๐‘“(๐‘›(๐‘›๐‘–๐‘‘))

๐‘‘๐‘”(๐‘‡ )

๐‘‘ ๐‘‡โˆ’ ๐‘‘๐‘”(๐‘‡ )

๐‘‘๐‘‡

๐‘›๐‘–๐‘‘โˆซ0

๐‘‘๏ฟฝ๏ฟฝ๐‘“(๐‘›(๏ฟฝ๏ฟฝ)).

(22)

Substituting this expression into (20), one getsEq. (21).

As one can see now from Eq. (21), the mean-fieldmodel with the linear ๐‘‡ dependence of ๐‘ˆ or, equiv-alently, of ๐‘ƒint, i.e., ๐‘”(๐‘‡ ) = ๐‘‡ โ‡’ ๐‘‘๐‘”(๐‘‡ )

๐‘‘ ๐‘‡ = 1, breaksdown the Third Law of thermodynamics. Indeed, at๐‘‡ = 0, one finds ๐‘ ๐‘–๐‘‘(๐‘‡ = 0, ๐œˆ) = 0 by construction,whereas, for the full entropy density, one gets

๐‘ (๐‘‡ = 0, ๐œ‡) = [1 + ๐‘ ๐‘›๐‘–๐‘‘]โˆ’1 ๐‘‘๐‘”(๐‘‡ )

๐‘‘ ๐‘‡

๐‘›๐‘–๐‘‘โˆซ0

๐‘‘๏ฟฝ๏ฟฝ ๐‘“(๐‘›(๏ฟฝ๏ฟฝ)) = 0,

unless ๐‘“ โ‰ก 0. Hence, the mean-field model with thelinear ๐‘‡ dependence of ๐‘ƒint suggested in [14] may bevery good at high temperatures, for which the Boltz-mann statistics is valid, but it is unphysical at ๐‘‡ = 0.

Of course, one can repair this defect by choos-ing a more complicated function ๐‘”(๐‘‡ ), which be-haves at high ๐‘‡ as ๐‘”(๐‘‡ ) โˆผ ๐‘‡ . But its derivative๐‘”โ€ฒ(๐‘‡ ) vanishes at ๐‘‡ = 0, providing the fulfillmentof the Third Law of thermodynamics (see an exam-ple in Sect. 5 for which ๐‘”(๐‘‡ ) โˆผ ๐‘‡ 2 at low tempera-tures). However, in this case, the whole idea to com-pensate the defects of the VdW EoS by tuning theinteracting part of the pressure does not work at low๐‘‡ , since, in this case, ๐‘ƒint = ๐‘”(๐‘‡ )๐น (๐‘›๐‘–๐‘‘) would vanish

faster than the first term staying on the right-handside of Eq. (17), i.e., the classical part of the pres-sure ๐‘‡๐‘›๐‘–๐‘‘ = ๐‘‡๐‘›/(1โˆ’ ๐‘๐‘›). Thus, we explicitly showedhere that, at low ๐‘‡, the mean-field models defined byEqs. (1)โ€“(5) either are unphysical, if ๐‘ƒint = ๐‘‡๐น (๐‘›๐‘–๐‘‘),or they cannot go beyond the VdW approximation byadjusting their interaction pressure ๐‘ƒint.

Such a conclusion can be also applied to the one oftwo ways to introduce the excluded volume correctioninto the quantum second virial coefficients discussedin Ref. [34]. Although the model of Ref. [34] containsthe scalar mean-fields which modify the masses ofparticles, the effective potential approach to treatthe excluded volume correction of Ref. [34] with thelinear ๐‘‡ dependence of the repulsive effective poten-tial ๐‘Š๐‘– (equivalent to the mean-field potential โˆ’๐‘ˆ inour notations) of the ๐‘–-th particle sort [see Eqs. (20)and (46) and (47) in [34]] should unavoidably leadto a breakdown of the Third Law of thermodynam-ics. Therefore, we conclude that such a way to intro-duce the excluded volume correction into the quan-tum second virial coefficients discussed in [34] is un-physical. Thus, despite the claims of the author ofRef. [34], such a generalization of the approach [7] toinclude the hard-core repulsion in quantum systemsleads to a problem with the Third Law of thermo-dynamics. To end this section, we express the tradi-tional virial coefficients ๐‘Ž๐‘„๐‘˜ of the quantum VdW gasof Eq. (17) in terms of the classical excluded volume ๐‘and the quantum virial coefficients of point-like par-ticles ๐‘Ž

(0)๐‘˜ . Expanding each denominator in Eq. (17)

into a series in powers of ๐‘›, one can easily find

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) = ๐‘‡

[๐‘›+

โˆžโˆ‘๐‘˜=2

๐‘Ž๐‘„๐‘˜ ๐‘›๐‘˜

], (23)

where

๐‘Ž๐‘„2 = ๐‘+ ๐‘Ž(0)2 , (24)

๐‘Ž๐‘„3 = ๐‘2 + 2 ๐‘ ๐‘Ž(0)2 + ๐‘Ž

(0)3 , (25)

๐‘Ž๐‘„4 = ๐‘3 + 3 ๐‘2 ๐‘Ž(0)2 + 3 ๐‘1 ๐‘Ž

(0)3 + ๐‘Ž

(0)4 , (26)

๐‘Ž๐‘„๐‘˜ = ๐‘๐‘˜โˆ’1 +

๐‘˜โˆ‘๐‘™=2

(๐‘˜ โˆ’ 1)!

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ ๐‘™)!๐‘๐‘˜โˆ’๐‘™๐‘Ž

(0)๐‘™ . (27)

If the interaction pressure ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘(๐‘›)) of model (1)can be expanded into the Taylor series of the particlenumber density ๐‘› at ๐‘› = 0, then one can obtain the

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 867

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

full quantum virial expansion of this EoS. Note thatthe coefficients ๐‘Ž(0)๐‘˜ for model (1) depend on the tem-perature only, while specific features of the EoS arestored in ๐‘ and in ๐‘ƒint(๐‘‡, ๐‘›๐‘–๐‘‘(๐‘›)). For example, usingthe coefficients ๐‘ = 3.42 fm3 and ๐‘ƒint(๐‘‡, ๐‘›) = ๐‘Žattr๐‘›

2

(๐‘Žattr = 329 MeV ยท fm3) found in [15] for the quantumVdW EoS of nuclear matter, one can calculate thefull quantum second virial coefficient of the model as

๐‘Ž๐‘„,tot2 = ๐‘+๐‘Ž

(0)2 โˆ’ ๐‘Žattr

๐‘‡โ‰ƒ ๐‘+

1

252 ๐‘‘๐‘

[2๐œ‹

๐‘‡ ๐‘š๐‘

]32

โˆ’ ๐‘Žattr๐‘‡

,

(28)

where, on the second step of the derivation, we usedthe non-relativistic expression for the cluster integral๐‘(0)2 (15). Taking results from [15], one can find that,

for nucleons (๐‘‘๐‘ = 4,๐‘š๐‘ = 939MeV), the coefficient๐‘Ž๐‘„,tot2 (๐‘‡ ) is zero at ๐‘‡ โ‰ƒ 0.32 MeV and ๐‘‡ โ‰ƒ 90.5 MeV,

is negative between these temperatures. Then, above๐‘‡ โ‰ƒ 90.5 MeV, it grows almost linearly with ๐‘‡ to๐‘Ž๐‘„,tot2 (๐‘‡ = 150MeV) โ‰ƒ (3.42 + 0.101 โˆ’ 2.19) fm3 โ‰ƒ

โ‰ƒ 1.33 fm3 which corresponds to the equivalent hard-core radius ๐‘…eq โ‰ƒ 0.46 fm at ๐‘‡ = 150 MeV. Fromthis estimate, it is evident that the large value of theequivalent hard-core radius ๐‘…eq for model [15] is aconsequence of the unrealistically large hard-core ra-dius of nucleons ๐‘…๐‘› โ‰ƒ 0.59 fm obtained in [15] (seealso a discussion later). In the most advanced versionof the hadron resonance gas model, the hard-core ra-dius of nucleons is 0.365 fm [25โ€“27], and, in the ISTEoS of the nuclear matter, this radius is below 0.4 fm[35]. It is obvious that a more realistic attraction thanthe one used in [15] would decrease the values of ๐‘…eq

and ๐‘…๐‘› to physically more adequate ones. Althoughthe explicit quantum virial expansion (23)โ€“(28) canbe used to find the appropriate attraction in order tocure the problems of the VdW EoS and to extend itto higher particle number densities and high/low ๐‘‡values, the true solution of this problem is suggestedbelow.

3. EoS with Induced Surface Tension

In order to overcome the difficulties of the quantumVdW EoS at high particle number densities, we sug-gest the following EoS

๐‘ = ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)โˆ’ ๐‘ƒint 1(๐‘‡, ๐‘›๐‘–๐‘‘ 1), (29)

ฮฃ = ๐‘…๐‘ [๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ2)โˆ’ ๐‘ƒint 2(๐‘‡, ๐‘›๐‘–๐‘‘ 2)], (30)

๐œˆ1 = ๐œ‡โˆ’ ๐‘‰0 ๐‘โˆ’ ๐‘†0 ฮฃ+ ๐‘ˆ1(๐‘‡, ๐‘›๐‘–๐‘‘ 1), (31)

๐œˆ2 = ๐œ‡โˆ’ ๐‘‰0 ๐‘โˆ’ ๐›ผ๐‘†0 ฮฃ+ ๐‘ˆ2(๐‘‡, ๐‘›๐‘–๐‘‘ 2), (32)

where ๐‘›๐‘–๐‘‘๐ด โ‰ก ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ๐ด)๐œ• ๐œˆ๐ด

with ๐ด = {1; 2}, ๐‘†0 = 4๐œ‹๐‘…2๐‘

denotes the proper surface of the hard-core volume๐‘‰0. Equation (29) is an analog of Eq. (1), while theequation for the induced surface tension coefficient ฮฃ(30) was first introduced for the Boltzmann statisticsin [28]. System (29)โ€“(32) is a quantum generalizationof the Boltzmann EoS in the spirit of work [7]. Asit was argued above, the temperature-dependent ef-fective potentials considered in [34] may lead to anunphysical behavior at low temperatures. Hence, wewould like to study this problem in detail. Below, wewill show what is a principal difference of EoS (29)โ€“(32) with the second way to include the hard-corerepulsion in quantum systems discussed in Ref. [34].

The quantity ฮฃ defined by (30) is the surface partof the hard-core repulsion [26]. As it will be shownlater, the representation of the hard-core repulsionin pressure (29) in two terms, namely via โˆ’๐‘‰0๐‘ andโˆ’๐‘†0ฮฃ, instead of a single term โˆ’4๐‘‰0๐‘, as it is donein the quantum VdW EoS, has great advantages andallows one to go beyond the VdW approximation.

Evidently, the self-consistency conditions for theIST EoS are similar to Eqs. (4) and (5) (๐ด = {1; 2})

๐‘›๐‘–๐‘‘๐ด๐œ•๐‘ˆ๐ด(๐‘‡, ๐‘›๐‘–๐‘‘๐ด)

๐œ• ๐‘›๐‘–๐‘‘๐ด=

๐œ•๐‘ƒintA(๐‘‡, ๐‘›๐‘–๐‘‘๐ด)

๐œ• ๐‘›๐‘–๐‘‘๐ด. (33)

The model parameter ๐›ผ > 1 is a switch between theexcluded and proper volume regimes. To demonstratethis property, let us consider the quantum distribu-tion function

๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) โ‰ก1

๐‘’๐ธ(๐‘˜)โˆ’๐œˆ2

๐‘‡ + ๐œ=

=๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡

๐‘’๐ธ(๐‘˜)โˆ’๐œˆ1

๐‘‡ + ๐œ โˆ’ ๐œ[1โˆ’ ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡

] =

= ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ ร—

ร—{1 +

โˆžโˆ‘๐‘™=2

[๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) ๐œ

(1โˆ’ ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡

)]๐‘™}, (34)

where, in the last step of the derivation, we have ex-panded the longest denominator above into a series of๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) ๐œ

(1โˆ’ ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡

)powers. Consider two lim-

its of (34), namely ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ‰ƒ 1 and ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ†’ 0 for

868 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

๐œ = 0. Then the distribution function (34) can becast as:

๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) โ†’

โ†’ ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡

{for ๐œ = 0, if ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 1,

for โˆ€ ๐œ, if ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ†’ 0.(35)

Further on, we assume that the inequality

(๐›ผโˆ’ 1)๐‘†0 ฮฃ/๐‘›๐‘–๐‘‘ 2 โ‰ซ (๐‘ˆ2 โˆ’ ๐‘ˆ1)/๐‘›๐‘–๐‘‘ 2, (36)

holds in either of the considered limits for๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ . Note that, in the case ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 1, condition

(36) is a natural one, because, at low particle den-sities, it means that the difference of two mean-fieldpotentials (๐‘ˆ2 โˆ’๐‘ˆ1) is weaker than the hard-core re-pulsion term (๐›ผโˆ’ 1)๐‘†0 ฮฃ; whereas, for ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ†’ 0, it

means that such a difference is simply restricted fromabove for large values of ฮฃ, i.e., max{|๐‘ˆ1|; |๐‘ˆ2|} << Const < โˆž. Evidently, in this limit, the mean-fieldpressures should be also finite, i.e. |๐‘ƒintA| < โˆž.

In the case ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ‰ƒ 1, one immediately recoversthe relation

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ2) โ‰ƒ ๐‘’(1โˆ’๐›ผ)๐‘†0 ฮฃ

๐‘‡ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)

for ๐œ = 0, which exactly corresponds to theBoltzmann statistics version [26] of system (29)โ€“(32). Hence, one recovers the virial expansion of๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1) [26] in terms of the particle number den-sity ๐‘›1 = ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ1)

๐œ• ๐œ‡ |๐‘ˆ1, which is calculated under the

condition ๐‘ˆ1 = const:

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)

๐‘‡๐‘›1โ‰ƒ 1 + 4๐‘‰0๐‘›1 + [16โˆ’ 18(๐›ผโˆ’ 1)] ๐‘‰ 2

0 ๐‘›21 +

+

[64 +

243

2(๐›ผโˆ’ 1)2 โˆ’ 216(๐›ผโˆ’ 1)

]๐‘‰ 30 ๐‘›3

1 + ... . (37)

Note that, due to the self-consistency condition (33),one finds ๐œ•๐‘(๐‘‡,๐œˆ1)

๐œ• ๐œ‡ = ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ1)๐œ• ๐œ‡ |๐‘ˆ1 , and, therefore, ๐‘›1

is the physical particle number density.As it was revealed in [26] for ๐›ผ = ๐›ผ๐ต โ‰ก 1.245,

one can reproduce the fourth virial coefficient of thegas of hard spheres exactly, while the value of thethird virial coefficient of such a gas is recovered withthe relative error about 16% only. Therefore, for lowdensities, i.e., for ๐‘‰0๐‘›1 โ‰ช 1, the IST EoS (29)โ€“(32)reproduces the results obtained for ๐œ = 0, if condition(36) is fulfilled.

On the other hand, from Eqs. (34) and (35), onesees that, in the limit ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ†’ 0, the distribu-

tion function ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) with ๐œ = 0 acquires theBoltzmann form. In this limit, we find ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ2) โ‰ƒโ‰ƒ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1) ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ and ๐‘›

(0)๐‘–๐‘‘ 2 โ‰ƒ ๐‘›

(0)๐‘–๐‘‘ 1 ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ . Using

these results and Eq. (36), we can rewrite (30) as

ฮฃ โ‰ƒ ๐‘…๐‘

[๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1) ๐‘’

(1โˆ’๐›ผ)๐‘†0 ฮฃ๐‘‡ โˆ’ ๐‘ƒint 2(๐‘‡, ๐‘›

(0)๐‘–๐‘‘ 2)

]. (38)

Here, we use the same notation as in the previoussection (see a paragraph before Eq. (14)). FromEq. (38), one can see that, for ๐‘‰0 ๐‘๐‘–๐‘‘(๐‘‡,๐œˆ1)

๐‘‡ โ‰ซ 1, thesurface tension coefficient ฮฃ is strongly suppressedcompared to ๐‘…๐‘ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1), i.e., one finds

ฮฃ โ‰ƒ ๐‘‡

๐‘†0 (๐›ผโˆ’ 1)ln

[๐‘…๐‘ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)

ฮฃ

]โ‰ช ๐‘…๐‘ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1).

Note that, for ๐›ผ > 1, the condition ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ†’ 0 canbe provided by ๐‘†0ฮฃ/๐‘‡ โ‰ซ 1 only. Thus, the secondterm on the right-hand side of Eq. (38) cannot domi-nate, since it is finite. It is evident that the inequality๐‘‰0 ๐‘๐‘–๐‘‘(๐‘‡,๐œˆ1)

๐‘‡ โ‰ซ 1 also means that ๐‘›(0)๐‘–๐‘‘ 1๐‘‰0 โ‰ซ 1. There-

fore, in this limit, the effective chemical potential (31)can be approximated as

๐œˆ1 โ‰ƒ ๐œ‡โˆ’ ๐‘‰0 ๐‘+ ๐‘ˆ1(๐‘‡, ๐‘›(0)๐‘–๐‘‘ 1), (39)

i.e., the contribution of the induced surface tensionis negligible compared to the pressure. This resultmeans that, for ๐‘›

(0)๐‘–๐‘‘ 1๐‘‰0 โ‰ซ 1, i.e., at high particle

densities or for ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ†’ 0, the IST EoS correspondsto the proper volume approximation.

On the other hand, Eq. (37) exhibits that, at lowdensities, i.e., for ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 1, the IST EoS recovers

the virial expansion of the gas of hard spheres up tothe fourth power of the particle density ๐‘›1. Therefore,it is natural to expect that, for intermediate values ofthe parameter ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โˆˆ [0; 1], the IST EoS will gradu-

ally evolve from the low-density approximation to thehigh-density one, if condition (36) is obeyed. This isa generalization of the previously obtained result [26]onto the quantum statistics case.

Already from the virial expansion (37), one can seethat the case ๐›ผ = 1 recovers the VdW EoS withthe hard-core repulsion. If, in addition, the mean-field potentials are the same, i.e., ๐‘ˆ2 = ๐‘ˆ1 and,consequently, ๐‘ƒint 2 = ๐‘ƒint 1, then one finds that

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 869

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

๐œˆ2 = ๐œˆ1 and ฮฃ = ๐‘…๐‘ ๐‘(๐‘‡, ๐œˆ1). In this case, the term๐‘‰0 ๐‘ + ๐‘†0 ฮฃ โ‰ก 4๐‘‰0 ๐‘ exactly corresponds to the VdWhard-core repulsion. If, however, ๐‘ˆ2 = ๐‘ˆ1, but bothmean-field potentials are restricted from above, thenthe model can deviate from the VdW EoS at low tem-peratures only, while, at high temperatures, it againcorresponds to the VdW EoS. In the case ๐‘ˆ2 < ๐‘ˆ1,this can be easily seen from Eqs. (34) and (35) in thecase ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 0, if one sets ๐›ผ = 1. Then, using the

same logic as in deriving Eq. (38), one can find thatฮฃ โ‰ช ๐‘…๐‘ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1). Hence, the effective chemical po-tential ๐œˆ1 acquires the form (39). In other words, atlow ๐‘‡ , the surface tension effect becomes negligible,and the IST EoS corresponds to the proper volumeapproximation, if ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 0.

Finally, if the inequality ๐‘ˆ2 > ๐‘ˆ1 is valid, then,at low ๐‘‡, expansion (34) has to be applied to the dis-tribution function ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) instead of ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2)and then one arrives at the unrealistic case, since ฮฃ โ‰ซโ‰ซ ๐‘…๐‘ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1). In this case, the hard-core repulsionwould be completely dominated by the induced sur-face tension term. Hence, even the second virial coef-ficient would not correspond to the excluded volumeof particles.

4. Going Beyond VdW Approximation

Let us closely inspect the IST EoS and show explicitlyits major differences from the VdW one. For such apurpose in this section, we analyze its effective andtrue virial expansions and discuss somewhat unusualproperties of the entropy density.

4.1. Effective virial expansion

First, we analyze the particle densities ๐‘›1(๐‘‡, ๐œˆ1) โ‰กโ‰ก ๐œ•๐‘(๐‘‡,๐œˆ1)

๐œ• ๐œ‡ and ๏ฟฝ๏ฟฝ2(๐‘‡, ๐œˆ2) โ‰ก ๐‘…โˆ’1๐‘

๐œ•ฮฃ(๐‘‡,๐œˆ2)๐œ• ๐œ‡ . For this

purpose, we differentiate Eqs. (29) and (30) withrespect to ๐œ‡ and apply the self-consistency condi-tions (33)๐‘›1 = ๐‘›๐‘–๐‘‘ 1

[1โˆ’ ๐‘‰0๐‘›1 โˆ’ ๐‘†0

๐œ•ฮฃ

๐œ•๐œ‡

], (40)

๐œ•ฮฃ

๐œ•๐œ‡= ๐‘…๐‘ ๐‘›๐‘–๐‘‘ 2

[1โˆ’ ๐‘‰0๐‘›1 โˆ’ ๐›ผ๐‘†0

๐œ•ฮฃ

๐œ•๐œ‡

]. (41)

Expressing ๐œ•ฮฃ๐œ•๐œ‡ from Eq. (41) and substituting it

into (40), one finds the particle number densities(๏ฟฝ๏ฟฝ2(๐‘‡, ๐œˆ2) โ‰ก ๐‘›2(1โˆ’ ๐‘‰0๐‘›1))

๐‘›1 =๐‘›๐‘–๐‘‘ 1 (1โˆ’ 3๐‘‰0 ๐‘›2)

1 + ๐‘‰0 ๐‘›๐‘–๐‘‘ 1 (1โˆ’ 3๐‘‰0 ๐‘›2), (42)

๐‘›2 =๐‘›๐‘–๐‘‘ 2

1 + ๐›ผ 3๐‘‰0 ๐‘›๐‘–๐‘‘ 2, (43)

where we used the relation ๐‘…๐‘๐‘†0 = 3๐‘‰0 for hardspheres. From Eq. (43) for ๐‘›2, one finds that, for๐›ผ > 1, the term (1 โˆ’ 3๐‘‰0 ๐‘›2) staying above is al-ways positive, since, taking the limit ๐‘›๐‘–๐‘‘ 2 โ†’ โˆž inEq. (43), one finds the limiting density of max{๐‘›2} =

= [3๐›ผ๐‘‰0]โˆ’1. Therefore, irrespective of the value of

๐‘›๐‘–๐‘‘ 2 โ‰ฅ 0, one finds in the limit ๐‘›๐‘–๐‘‘ 1๐‘‰0 โ‰ซ 1 thatmax{๐‘›1} = ๐‘‰ โˆ’1

0 . This is another way to prove thatthe limiting density of the IST EoS corresponds tothe proper volume limit, since, at high densities, it isfour times higher than the one of the VdW EoS. Wri-ting the particle number density ๐‘›๐‘–๐‘‘ 1 from Eq. (42)as๐‘›๐‘–๐‘‘ 1 =

๐‘›1

(1โˆ’ ๐‘‰0 ๐‘›1) (1โˆ’ 3๐‘‰0 ๐‘›2), (44)

one can get the formal virial-like expansion for theIST pressure ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1) (29)

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)

๐‘‡=

โˆžโˆ‘๐‘˜=1

๐‘Ž(0)๐‘˜

[1โˆ’ 3๐‘‰0 ๐‘›2]๐‘˜[๐‘›1]

๐‘˜

[1โˆ’ ๐‘‰0 ๐‘›1]๐‘˜, (45)

where the expressions for the coefficients ๐‘Ž(0)๐‘˜ are

given by Eqs. (9)โ€“(16). This result allows us to for-mally write the expansion

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)

๐‘‡โ‰ก

โˆžโˆ‘๐‘˜=1

๐‘Ž(0),๐ผ๐‘†๐‘‡๐‘˜

[๐‘›1]๐‘˜

[1โˆ’ ๐‘‰0 ๐‘›1]๐‘˜

(46)

with the coefficients ๐‘Ž(0),IST๐‘˜ =

๐‘Ž(0)๐‘˜

[1โˆ’3๐‘‰0 ๐‘›2]๐‘˜which de-

pend not only on ๐‘‡ , but also on ๐‘›2. Expansions (45)and (46) are the generalizations of the ones used forthe EoSs of hard discs [32] and hard spheres [33].

Similarly to deriving Eq. (27), one can get thequantum virial expansion for IST pressure ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1)from (46):

๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1) = ๐‘‡

โˆžโˆ‘๐‘˜=1

๐‘Ž๐‘„,IST๐‘˜ ๐‘›๐‘˜

1 , (47)

๐‘Ž๐‘„,IST๐‘˜ =

๐‘˜โˆ‘๐‘™=1

๐ถ(๐‘˜)๐‘™

[1โˆ’ 3๐‘‰0 ๐‘›2]๐‘™, (48)

๐ถ(๐‘˜)๐‘™ =

(๐‘˜ โˆ’ 1)!

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ ๐‘™)!๐‘‰ ๐‘˜โˆ’๐‘™0 ๐‘Ž

(0)๐‘™ , (49)

with the coefficients ๐‘Ž๐‘„,IST๐‘˜ which are ๐‘‡ - and ๐‘›2-de-

pendent. For the interaction pressure ๐‘ƒint 1(๐‘‡, ๐‘›๐‘–๐‘‘ 1)

870 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

which is expandable in terms of the density ๐‘›1,Eq. (48) can be used to estimate the full quantumvirial coefficients of higher orders. Of course, Eq. (47)is not the traditional virial expansion. But the factthat it can be exactly obtained from the grand canon-ical ensemble formulation of the quantum version ofthe IST EoS for the third, fourth, and higher ordervirial coefficients is still remarkable.

4.2. True quantum virial coefficients

Now, we consider an example on how to employ re-sults (47)โ€“(49) to estimate the true virial coefficientsat low densities and at sufficiently high temperatureswhich provide the convergence of the virial expansion(47). Apparently, in this case, one can expand thedensity ๐‘›2 โ‰ƒ ๐ต1๐‘›1(1 + ๐ต2๐‘›1) in powers of the den-sity ๐‘›1. From our above treatment of the low-densitylimit ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ โ‰ƒ 1, it is clear that ๐ต1 = 1. Substitut-

ing this expansion for ๐‘›2 into Eqs. (47) and (48) andkeeping only the terms up to ๐‘›2

1, one can get the truequantum virial coefficients ๐‘Ž๐‘„,tot

๐‘˜ as

๐‘Ž๐‘„,tot2 = ๐‘‰0 + ๐‘Ž

(0)2 + 3๐‘‰0๐ต1 = 4๐‘‰0 + ๐‘Ž

(0)2 , (50)

๐‘Ž๐‘„,tot3 โ‰ƒ 13๐‘‰ 2

0 + 3๐‘‰0๐ต2 + 5๐‘‰0๐‘Ž(0)2 + ๐‘Ž

(0)3 , (51)

๐‘Ž๐‘„,tot๐‘˜โ‰ฅ3 โ‰ƒ

๐‘˜โˆ‘๐‘™=1

๐ถ(๐‘˜)๐‘™ + 3๐‘‰0๐ต1

๐‘˜โˆ’1โˆ‘๐‘™=1

๐ถ(๐‘˜โˆ’1)๐‘™ ๐‘™+

+3๐‘‰0๐ต1

๐‘˜โˆ’2โˆ‘๐‘™=1

๐ถ(๐‘˜โˆ’2)๐‘™

[3

2๐‘™(๐‘™ + 1)๐‘‰0๐ต1 +๐ต2

], (52)

and replace the coefficients ๐‘Ž๐‘„,IST๐‘˜ in Eq. (47) with

the true quantum virial coefficients ๐‘Ž๐‘„,tot๐‘˜ which de-

pend on ๐‘‡ only. Note that the expression for the sec-ond virial coefficient ๐‘Ž๐‘„,tot

2 is exact, while the ex-pressions for the higher order virial coefficients arethe approximate ones, which, nevertheless, at highvalues of temperature are rather accurate. Conside-ring the limit of high temperatures which allows oneto ignore the quantum corrections in Eqs. (50) and(51), one can find the coefficients ๐ต1 = 1 exactlyand ๐ต2 โ‰ƒ [7โˆ’6๐›ผ]๐‘‰0 approximately by comparing ex-pressions (50) and (51) with the corresponding virialcoefficients of the Boltzmann gas in Eq. (37). Substi-tuting the obtained expressions for ๐ต1 and ๐ต2 coeffi-cients into Eq. (52), one gets the approximate formula

for higher-order virial coefficients ๐‘Ž๐‘„,tot๐‘˜โ‰ฅ3 :

๐‘Ž๐‘„,tot๐‘˜โ‰ฅ3 โ‰ƒ

๐‘˜โˆ‘๐‘™=1

๐ถ(๐‘˜)๐‘™ + 3๐‘‰0

๐‘˜โˆ’1โˆ‘๐‘™=1

๐ถ(๐‘˜โˆ’1)๐‘™ ๐‘™+

+3๐‘‰ 20

๐‘˜โˆ’2โˆ‘๐‘™=1

๐ถ(๐‘˜โˆ’2)๐‘™

[3

2๐‘™(๐‘™ + 1) + (7โˆ’ 6๐›ผ)

]=

=

๐‘˜โˆ‘๐‘™=1

(๐‘˜ โˆ’ 1)!๐‘‰ ๐‘˜โˆ’๐‘™0 ๐‘Ž

(0)๐‘™

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ ๐‘™)!+ 3

๐‘˜โˆ’1โˆ‘๐‘™=1

(๐‘˜ โˆ’ 2)!๐‘‰ ๐‘˜โˆ’๐‘™0 ๐‘Ž

(0)๐‘™

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ 1โˆ’ ๐‘™)!๐‘™+

+3

๐‘˜โˆ’2โˆ‘๐‘™=1

(๐‘˜ โˆ’ 3)!๐‘‰ ๐‘˜โˆ’๐‘™0 ๐‘Ž

(0)๐‘™

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ 2โˆ’ ๐‘™)!

[3

2๐‘™(๐‘™ + 1) + (7โˆ’ 6๐›ผ)

],

(53)

where the second equality above is obtained by sub-stituting Eq. (49) for the coefficients ๐ถ

(๐‘˜)๐‘™ into the

first one.Comparing now Eq. (53) for the IST EoS and

Eq. (27) for the VdW EoS, one can see that the firstsum on the right-hand side of (53) is identical to theexpression for the VdW quantum virial coefficientswith the excluded volume ๐‘ = 4๐‘‰0 replaced by theproper volume ๐‘‰0. Apparently, the other two sumson the right-hand side of (53) are the corrections dueto the induced surface tension coefficient.

Note that it is not difficult to get the exact expres-sions for the third or fourth virial coefficients ๐‘Ž๐‘„,tot

๐‘˜

by inserting the higher order terms of the expansion๐‘›2(๐‘›1) in powers of the density ๐‘›1 into Eqs. (47) and(48). Although, comparing the coefficients in front of๐ต1 and ๐ต2 in the last sum of Eq. (52), one can seethat, even for ๐‘™ = 1, the coefficient staying before๐ต1 is essentially larger than the one staying before๐ต2. This means that, at low densities, the role of ๐ต2

is an auxiliary one, if ๐›ผ is between 1 and 1.5.

4.3. Virial expansionfor compressible spheres

It is of interest that the ๐‘˜-th term

1

[1โˆ’ 3๐‘‰0 ๐‘›2]๐‘˜[๐‘›1]

๐‘˜

[1โˆ’ ๐‘‰0 ๐‘›1]๐‘˜,

staying in sum (45) allows for a non-trivial interpreta-tion. Comparing Eq. (17) and Eq. (45) and recallingthe fact that the particle number density ๐‘›1 is pro-portional to the number of spin-isospin configurations

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 871

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

๐‘‘๐‘, one can introduce an effective number of such con-figurations as ๐‘‘eff๐‘ =

๐‘‘๐‘

1โˆ’3๐‘‰0๐‘›2with the simultaneous

replacement of ๐‘‰0 by the effective proper volume

๐‘‰ eff0 = ๐‘‰0 (1โˆ’ 3๐‘‰0๐‘›2)

in all terms which contain the powers of [1 โˆ’ ๐‘‰0๐‘›1]on the right-hand side of (45). Then, at high densi-ties, the effective number of spin-isospin configura-tions ๐‘‘eff๐‘ โ‰ค ๐›ผ๐‘‘๐‘

๐›ผโˆ’1 can be sizably larger than ๐‘‘๐‘, whilethe effective proper volume ๐‘‰ eff

0 can be essentiallysmaller than ๐‘‰0 (i.e., such effective particles are com-pressible), if the coefficient ๐›ผ > 1 is close to 1. More-over, one can also establish the equivalent virial ex-pansion of pressure (45) in terms of ๐‘›1

(1โˆ’3๐‘‰0๐‘›2)pow-

ers. Then, instead of the coefficients ๐‘Ž๐‘„,IST๐‘˜ (48), one

would get

๏ฟฝ๏ฟฝ๐‘„,IST๐‘˜ =

๐‘˜โˆ‘๐‘™=1

(๐‘˜ โˆ’ 1)!

(๐‘™ โˆ’ 1)!(๐‘˜ โˆ’ ๐‘™)!

[๐‘‰ eff0

]๐‘˜โˆ’๐‘™๐‘Ž(0)๐‘™ , (54)

which shows that, at high densities, the contribu-tions of low-order virial coefficients ๐‘Ž

(0)๐‘™ into the co-

efficient ๏ฟฝ๏ฟฝ๐‘„,IST๐‘˜>1 are suppressed due to a decrease of

๐‘‰ eff0 . Eq. (54) quantifies the source of softness of the

IST EoS compared to VdW one at high densities. Itis also interesting that the monotonic decrease of ๐‘‰ eff

0

at high densities is qualitatively similar to the effectof the Lorentz contraction of a proper volume for rel-ativistic particles [36].

Although the present model does not know any-thing about the internal structure of considered par-ticles, but the fact that ๐‘‘eff๐‘ increases with the par-ticle number density ๐‘›2 can be an illustration of thein-medium effect that the IST hard-core interactionโ€œproducesโ€ the additional (or โ€œenhancesโ€ the numberof existing) spin-isospin states which are well knownin quantum physics as excited states, but with anexcitation energy being essentially smaller than themean value of the particle free energy. In this way,one can see that, at high densities, the IST effectivelyincreases the degeneracy factor of particles. This find-ing is a good illustration that the claim of Ref. [34]that accounting for the excluded volume correction inthe quantum case via the effective degeneracy leads toa reduction of the latter (see a discussion of Eqs. (18)and (19) in [34]) is not a general one. On contrary, amore advanced EoS developed above requires not a

reduction of the effective number of degrees of free-dom as it is suggested in [34], but their enhancement.

It is apparent that, for ๐›ผ โ‰ซ 1, the quantities ๐‘‰ eff0

and ๐‘‘eff๐‘ are practically independent of ๐‘›2, i.e., in thiscase, the coefficients ๐‘Ž๐‘„,IST

๐‘˜ and ๏ฟฝ๏ฟฝ๐‘„,IST๐‘˜ are the true

quantum virial coefficients of the VdW EoS, but withthe excluded volume ๐‘ = 4๐‘‰0 replaced by ๐‘‰0.

4.4. Properties of entropy density

Next, we study the entropy density of the ISTEoS. Similarly to finding the derivatives of Eqs. (29)and (30) with respect to ๐œ‡, one has to find theirderivatives with respect to ๐‘‡ in order to get the en-tropy per particle

๐‘ 1๐‘›1

=

[๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1โˆ’ 3๐‘‰0 ๐‘›2

๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2

][1โˆ’ 3๐‘‰0 ๐‘›2]

, (55)

๐‘ ๐‘–๐‘‘๐ด โ‰ก ๐‘ ๐‘–๐‘‘๐ด + ๐‘›๐‘–๐‘‘๐ด๐œ•๐‘ˆ๐ด

๐œ• ๐‘‡โˆ’ ๐œ•๐‘ƒint๐ด

๐œ• ๐‘‡, (56)

where the entropy density of point-like particles isdefined as ๐‘ ๐‘–๐‘‘๐ด โ‰ก ๐œ•๐‘๐‘–๐‘‘(๐‘‡,๐œˆ๐ด)

๐œ• ๐‘‡ and ๐ด โˆˆ {1; 2}. If themean-field potentials of the model have the form

๐‘ˆ๐ด =โˆ‘๐œ†

๐‘”๐œ†๐ด(๐‘‡ )๐‘“๐œ†๐ด(๐‘›๐‘–๐‘‘๐ด) (57)

and, for ๐‘‡ = 0, their derivatives obey the set of con-ditions ๐‘‘๐‘”๐œ†

๐ด(๐‘‡ )๐‘‘ ๐‘‡ = 0, then it is easy to see that the en-

tropy per particle ๐‘ 1๐‘›1

also vanishes at ๐‘‡ = 0, i.e. theThird Law of thermodynamics is obeyed under theseconditions. In a special case where the interactionmean-field potentials do not explicitly depend on thetemperature ๐‘‡ , the expression for the entropy den-sities (56) gets simpler, i.e., ๐‘ ๐‘–๐‘‘๐ด = ๐‘ ๐‘–๐‘‘๐ด. This caseis important for the hadron resonance model and isdiscussed in the Appendix in some details.

Apparently, to provide a positive value of the en-tropy per particle ๐‘ 1

๐‘›1, one has to properly choose

the interaction terms in Eqs. (29) and (30). In otherwords, the Third Law of thermodynamics providesone of the basic constraints to the considered EoS. Itis clear that the corresponding necessary conditionsshould not be very restrictive, because, at low den-sities, i.e. for 3๐‘‰0 ๐‘›2 โ‰ช 1, the coefficient staying infront of the term ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2is very small, while, at high

densities, it is ๐›ผโˆ’1 < 1 for ๐›ผ > 1. Although a discus-sion of such conditions is far beyond the scope of thiswork, we consider two important cases below.

872 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

In the case ๐‘ˆ2(๐‘‡, ๐œŒ) โ‰ก ๐‘ˆ1(๐‘‡, ๐œŒ), condition (36) isvalid for any choice of parameters. Then one can showa validity of the inequality ๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1โ‰ฅ ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2, since, for

๐›ผ > 1, one finds ๐œˆ1 > ๐œˆ2. To prove this inequality,one has to take into account that ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ๐ด), and allits derivatives are monotonically increasing functionsof ๐‘‡ and ๐œˆ๐ด. Then, using relations (34) and (35) be-tween the quantum distribution functions, one canshow the validity of the inequality ๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1โ‰ฅ ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2for

two limits ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ‰ƒ 1 and ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ†’ 0. Similarly, onecan introduce an effective parameter of statistics

๐œeff โ‰ก ๐œ โˆ’ ๐œ[1โˆ’ ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡

]and study the quantities for the distribution func-tion ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) with an effective parameter of statis-tics ๐œeff . However, one can easily understand that theinequality ๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1โ‰ฅ ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2cannot be broken down for

any value of the exponential ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ obeying the in-equalities 0 < ๐‘’

๐œˆ2โˆ’๐œˆ1๐‘‡ < 1. This is so, since the pres-

sure of point-like particles and its partial derivativesare monotonic functions of the parameters ๐‘‡ and ๐œˆ1(or ๐œˆ2), and the non-monotonic behavior of the en-tropy per particle can be caused by the phase tran-sition, which does not exists for an ideal gas. Notethat we do not consider a possible effect of the Boseโ€“Einstein condensation. Using the above inequality be-tween the entropies per particle and requiring that๐‘ˆ1 โ‰ฅ 0 and the inequalities ๐‘‘๐‘”๐œ†

๐ด(๐‘‡ )๐‘‘ ๐‘‡ > 0 for ๐‘‡ > 0 and

๐‘‘๐‘”๐œ†๐ด(๐‘‡=0)๐‘‘ ๐‘‡ = 0 hold, one can show that ๐‘ 1

๐‘›1โ‰ฅ ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2โ‰ฅ 0,

using identity (5).Another important case corresponds to the choice

๐‘ˆ1 > 0 and ๐‘ˆ2 < 0 in Eq. (57), i.e., the mean-field๐‘ˆ1 describes an attraction, while ๐‘ˆ2 represents a re-pulsion. Clearly, condition (36) in this case is alsofulfilled for any choice of parameters. Using the self-consistency relation (33) or its more convenient form(5), one can find that the term describing the mean-field entropy in ๐‘ ๐‘–๐‘‘ 2 can be negative, i.e.,

๐‘›๐‘–๐‘‘ 2๐œ•๐‘ˆ2

๐œ• ๐‘‡โˆ’ ๐œ•๐‘ƒint 2

๐œ• ๐‘‡=โˆ‘๐œ†

๐‘‘๐‘”๐œ†2 (๐‘‡ )

๐‘‘ ๐‘‡

๐‘›๐‘–๐‘‘ 2โˆซ0

๐‘‘๐‘› ๐‘“๐œ†2 (๐‘›) < 0,

(58)

if ๐‘”๐œ†2 (๐‘‡ ) > 0, ๐‘‘๐‘”๐œ†2 (๐‘‡ )๐‘‘ ๐‘‡ > 0, but ๐‘ˆ2 < 0 for ๐‘‡ โ‰ฅ 0 due

to the inequalities ๐‘“๐œ†2 (๐‘›) < 0. Such a choice of the in-

teraction allows one to decrease the effective entropy

density ๐‘ ๐‘–๐‘‘ 2 or even to make it negative by tuning themean-field ๐‘ˆ2 related to the IST coefficient. As a re-sult, this would increase the physical entropy density๐‘ 1. Note that, for the VdW EoS, this is impossible.

5. Application to Nuclearand Hadronic Matter

5.1. Some important examples

As a pedagogical example to our discussion, we con-sider the IST EoS for the nuclear matter and compareit with the VdW EoS (1) having the interaction

๐‘ƒVdWint (๐‘‡, ๐‘›๐‘–๐‘‘) = ๐‘Ž

[๐‘›๐‘–๐‘‘

1 + ๐‘ ๐‘›๐‘–๐‘‘

]2+ ๐‘‡๐‘›๐‘–๐‘‘ โˆ’

๐‘”(๐‘‡ )๐‘›๐‘–๐‘‘

1 + ๐‘ ๐‘›๐‘–๐‘‘โˆ’

โˆ’ ๐‘”(๐‘‡ )๐‘ ๐‘›2๐‘–๐‘‘

[1 + ๐‘ ๐‘›๐‘–๐‘‘]2 โˆ’ ๐‘”(๐‘‡ )๐ต3 ๐‘›

3๐‘–๐‘‘

[1 + ๐‘ ๐‘›๐‘–๐‘‘]3 โˆ’ ๐‘”(๐‘‡ )๐ต4 ๐‘›

4๐‘–๐‘‘

[1 + ๐‘ ๐‘›๐‘–๐‘‘]4 , (59)

where the virial coefficients ๐‘, ๐ต3, and ๐ต4 are intro-duced above, and the function ๐‘”(๐‘‡ ) โ‰ก ๐‘‡ 2

๐‘‡+๐‘‡SWwith

๐‘‡SW = 1 MeV provides the fulfillment of the ThirdLaw of thermodynamics. Note that the term ๐‘‡๐‘›๐‘–๐‘‘

cancels exactly the first term of the quantum virialexpansion for ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ) (see Eq. (17)), while the term๐‘Ž[

๐‘›๐‘–๐‘‘

1+๐‘ ๐‘›๐‘–๐‘‘

]2 in Eq. (59) accounts for an attraction andthe other terms proportional to ๐‘”(๐‘‡ ) are the low-est four powers of the virial expansion for the gasof classical hard spheres for ๐‘‡ โ‰ซ ๐‘‡SW. By construc-tion, such an EoS reproduces, apparently, the fourfirst virial coefficients of the gas of hard spheres at๐‘‡ โ‰ซ ๐‘‡SW. Simultaneously, it obeys the Third Law ofthermodynamics at ๐‘‡ = 0.

For the IST EoS, we choose ๐›ผ = 1.245 [26], ๐‘ƒ ISTint 1 =

= ๐‘Ž[

๐‘›๐‘–๐‘‘ 1

1+๐‘ ๐‘›๐‘–๐‘‘ 1

]2 and ๐‘ƒ IST๐‘–๐‘›๐‘ก 2 = 0 with the same con-

stants ๐‘Ž โ‰ƒ 329 MeV fm3 and ๐‘ = 4๐‘‰0 โ‰ƒ 3.42 fm3

which were found in [15] for the VdW EoS of nuclearmatter (๐‘‘๐‘ = 4,๐‘š๐‘ = 939MeV), i.e., we took just theparameters of Ref. [15] for a proper comparison. Byconstruction, the IST EoS and EoS (59) agree verywell (within one percent) for ๐‘‡ > 120 MeV and par-ticle number densities ๐‘› โ‰ค 0.25 fmโˆ’3. In Fig. 1, wecompare three isotherms at ๐‘‡ = 19, 10, and 0 MeV ofthese two EoS. For ๐‘‡ = 10 MeV, their isotherms agreeup to the packing fraction ๐œ‚ = ๐‘‰0๐‘› โ‰ƒ 0.09 (for thenuclear density ๐‘› โ‰ค 0.11 fmโˆ’3), i.e., within the usualrange of the VdW EoS applicability [25,26]. However,for ๐‘‡ = 0 and ๐‘‡ = 19 MeV isotherms, the both mod-els agree up to the packing fraction ๐œ‚ = ๐‘‰0๐‘› โ‰ƒ 0.03only (for ๐‘› โ‰ค 0.035 fmโˆ’3), i.e., far below the usual

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 873

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

Fig. 1. Behavior of the pressure as a function of the particlenumber density for isotherms of nuclear matter (see the textfor details)

Fig. 2. Packing fraction dependence of the quantum com-pressibility factors ฮ”๐‘๐‘„ of the GVdW EoS and IST EoS (seethe text)

range of the VdW EoS applicability due to the im-portant role of the second and higher order quantumvirial coefficients ๐‘Ž(0)๐‘˜โ‰ฅ2 defined by Eqs. (10)โ€“(15). Thepresent example clearly shows that providing the fourvirial coefficients of the gas of hard spheres for thequantum VdW EoS of Ref. [15] at high temperatures,one can, at most, get a good agreement with the ISTEoS for a single value of the temperature, namely for๐‘‡ = 10 MeV. Figure 1 also shows that, for the sameparameters, the IST EoS is essentially softer that theimproved VdW one, hence, it does not require sostrong attraction and so strong repulsion to reproduce

the properties of normal nuclear matter. This conclu-sion is supported by the results obtained recently forthe nuclear-matter EoS within the IST concept [35].

Recently, an interesting generalization of the quan-tum VdW EoS (GVdW hereafter) was suggested in[37]. This EoS allows one to go beyond the VdW ap-proximation, but, formally, it is similar to the VdWmodels discussed above. In terms of the ideal gaspressure (2), the GVdW pressure can be written as[37] (๐œ‚ = ๐‘‰0๐‘› is the packing fraction):

๐‘G(๐‘‡, ๐œ‡) = ๐‘ค(๐œ‚) ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆG)โˆ’ ๐‘ƒintG(๐‘›), (60)๐œˆG(๐œ‡, ๐‘›) = ๐œ‡+ ๐‘‰0 ๐‘“

โ€ฒ(๐œ‚) ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆG) + ๐‘ˆG(๐‘›), (61)

where ๐‘› is the particle density, and the multiplier๐‘ค(๐œ‚) โ‰ก (๐‘“(๐œ‚)โˆ’ ๐œ‚๐‘“ โ€ฒ(๐œ‚)) is given in terms of the func-tion ๐‘“(๐œ‚) which is defined as

๐‘“(๐œ‚) =

โŽงโŽจโŽฉ๐‘“VdW(๐œ‚) = 1โˆ’ 4๐œ‚, for VdW EoS,

๐‘“CS(๐œ‚) = exp

[โˆ’ (4โˆ’ 3๐œ‚)๐œ‚

(1โˆ’ ๐œ‚)2

], for CS EoS,

(62)

where the function ๐‘“VdW(๐œ‚) corresponds to the VdWcase, whereas the function ๐‘“CS(๐œ‚) is given for the fa-mous Carnahanโ€“Starling (CS) EoS [17]. The interac-tion terms of the GVdW EoS are given in terms ofa function ๐‘ข(๐‘›): ๐‘ˆG = ๐‘ข(๐‘›) + ๐‘›๐‘ขโ€ฒ(๐‘›) and ๐‘ƒintG == โˆ’๐‘›2๐‘ขโ€ฒ(๐‘›). This choice automatically provides theself-consistency condition fulfillment. Since the po-tentials ๐‘ˆG and ๐‘ƒintG are temperature-independent,the Third Law of thermodynamics is obeyed.

The presence of the function ๐‘ค(๐œ‚) in front of theideal gas pressure in (60) allows one to reproducethe famous CS EoS [17] at high temperatures, whileit creates the problems with formulating the GVdWmodel for several hard-core radii, since the pressuresof point-like particles of kinds 1 and 2 cannot beadded to each other, if their functions ๐‘ค(๐œ‚1) and๐‘ค(๐œ‚2) are not the same.

Using the quantum virial expansion (8) and theparticle number density expression ๐‘› = ๐‘“(๐œ‚)ร—ร—๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆG) [37], for ๐‘ƒIG โ‰ก ๐‘ค(๐œ‚) ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆG), one ob-tains๐‘ƒIG = ๐‘ค(๐œ‚)๐‘‡

[๐‘›

๐‘“(๐œ‚)+

โˆžโˆ‘๐‘™=2

๐‘Ž(0)๐‘™

[๐‘›

๐‘“(๐œ‚)

]๐‘™], (63)

๐‘ค(๐œ‚)

๐‘“(๐œ‚)=

โŽงโŽชโŽชโŽจโŽชโŽชโŽฉ1

1โˆ’ 4๐œ‚โ‰ก 1

๐‘“VdW(๐œ‚), for VdW EoS,

1 + ๐œ‚ + ๐œ‚2 โˆ’ ๐œ‚3

(1โˆ’ ๐œ‚)3, for CS EoS.

(64)

874 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

Although this effective expansion can be used to de-rive the true virial expansion for the CS parametriza-tion of the GVdW EoS (for the VdW one, it is givenabove), the result is cumbersome. Nevertheless, theseequations show that, due to the multiplier ๐‘ค(๐œ‚),the first term of the quantum virial expansion inEqs. (63), (8), (17), and (47), i.e., the classical term,exactly reproduces the pressure by the correspond-ing classical EoS. Hence, all other terms in Eqs. (8),(17), (47), and (63) are the quantum ones. A directcomparison of the IST with ๐›ผ = 1.245 and CS EoSfor classical gases shows that, for packing fractions๐œ‚ > 0.22, the IST EoS is softer than the CS one[25,26]. Figure 2 depicts the quantum compressibilityfactors

ฮ”๐‘CS๐‘„ (๐œ‚) =

๐‘ƒIG โˆ’ ๐‘ค(๐œ‚)๐‘‡๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆG)

๐‘‡ ๐‘›

for the CS EoS of the GVdW model and the one forthe IST EoS defined similarly

ฮ”๐‘IST๐‘„ (๐œ‚) =

๐‘๐‘–๐‘‘ 1 โˆ’ ๐‘‡๐‘›๐‘–๐‘‘ 1(๐‘‡, ๐œˆ1)

๐‘‡ ๐‘›1

taken both for the same parameters ๐‘ = 3.42 fm3,๐‘ƒintG(๐‘‡, ๐‘›) = ๐‘Žattr๐‘›

2 with ๐‘Žattr = 329 MeV ยท fm3 (see[37] for more details). As one can see from Fig. 2,the quantum compressibility factors of these EoS dif-fer essentially for ๐œ‚ โ‰ฅ 0.05. Therefore, for ๐œ‚ โ‰ฅ 0.1,both the classical and quantum parts of the ISTpressure with ๐›ผ = 1.245 [26] are essentially softerthan the corresponding terms of the CS version ofthe GVdW model of Ref. [37]. One can easily un-derstand such a conclusion comparing expansions(63) and (45). Since, for the same packing fraction๐œ‚ โ‰ฅ 0.1, the function ๐‘“CS(๐œ‚) of the CS version ofthe GVdW EoS vanishes essentially faster than theterm [1โˆ’3๐‘‰0๐‘›2][1โˆ’๐‘‰0๐‘›1] of the IST EoS, each termproportional to ๐‘›๐‘˜ in (63) with ๐‘˜ > 1 is larger thanthe corresponding term proportional to ๐‘›๐‘˜

1 = ๐‘›๐‘˜ in(45). It is necessary to note that such a property isvery important, because the softer EoS provides awider range of thermodynamic parameters for whichthe EoS is causal, i.e. its speed of sound is smallerthan the speed of light.

5.2. Constraints on nuclearmatter properties

It is appropriate to discuss the most important con-straints on the considered mean-field models which

are necessary to describe the strongly interactingmatter properties. According to Eqs. (17), (47), and(63), the fermionic pressure for the considered EoSconsists of three contributions: the classical pressure(the first term on the right-hand side of (17), (47),and (63)), the quantum part of the pressure and themean-field ๐‘ƒint. At temperatures below 1 MeV, theclassical part is negligible, but the usage of virial ex-pansions discussed above is troublesome due to theconvergency problem. Since the exact parametriza-tion of the function ๐‘ƒint on the particle number den-sity of nucleons is not known, it is evident that allconsidered models are effective by construction. Tofix their parameters, one has to reproduce the usualproperties of normal nuclear matter, i.e. to get azero value for the total pressure at the normal nu-clear density ๐‘›0 โ‰ƒ 0.16 fmโˆ’3 and the binding en-ergy ๐‘Š = โˆ’16 MeV at this density [1]. Similarlyto the high-temperature case discussed at the endof Section 2, there exists a freedom of parametriz-ing the hard-core radius of nucleons, since the attrac-tion pressure can be always adjusted to reproducethe properties of normal nuclear matter and, there-fore, all the model parameters are also effective byconstruction.

However, in addition to the properties of normalnuclear matter, there is the so-called flow constraintat nuclear densities ๐‘› = (2โ€“5)๐‘›0 [38], which setsstrong restrictions on the model pressure dependenceon the nuclear particle density and requires a rathersoft EoS at these densities. Hence, it can be used todetermine the parameters of a realistic EoS at highnuclear densities and ๐‘‡ = 0. Traditionally, such aconstraint creates troubles for the relativistic mean-field EoS based on the Walecka model [4, 39, 40].

The validity of this statement can be seen fromRef. [39] in which it is shown that only 104 of suchEoSs out of 263 analyzed in [39] are able to obeythe flow constraint despite the fact that they have10 or even more adjustable parameters. At the sametime, as one can see from the simplest realization ofthe IST EoS suggested in Ref. [35], the 4-parameterEoS is able to simultaneously reproduce all prop-erties of normal nuclear matter and the flow con-straint. Furthermore, the IST EoS is able not only toreproduce the flow constraint, but, simultaneously, itis able to successfully describe the neutron star prop-erties with the masses more than two Solar ones [41],which sets another strong constraint on the stiffness

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 875

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

of the realistic EoS at high particle densities and thezero temperature. On the other hand, Fig. 2 showsthat the existing CS version of the GVdW EoS ofRef. [37] is very stiff, and, hence, it will also havetroubles to obey the flow constraint [38].

5.3. Constraints on hadronicmatter properties

From the virial expansions of all models discussedhere, one sees that the EoS calibration on the prop-erties of nuclear matter at low ๐‘‡ and at high densi-ties involves mainly the quantum and mean-field pres-sures. But, unfortunately, it also fixes the parametersof the classical pressure at higher temperatures. Itis, however, clear that the one-component mean-fieldmodels of nuclear matter cannot be applied at tem-peratures above 50 MeV, since one has to include themesons, other baryons, and their resonances [31, 42].

Moreover, at high temperatures, the mean-fieldsand the parameters of interaction should be re-ca-librated because the very fact of resonance existencealready corresponds to a partial account for the in-teraction [42]. For many years, it is well known that,for temperatures below 170 MeV and densities be-low ๐‘›0, the mixture of stable hadrons and their res-onances whose interaction is taken into account bythe quantum second virial coefficients behaves as amixture of nearly ideal gases of stable particles. Thelatter, in this case, includes both the hadrons andthe resonances, but taken with their averaged masses[42]. The main reason for such a behavior is rooted ina nearly complete cancellation between the attractionand repulsion contributions. The resulting deviationfrom the ideal gas (a weak repulsion) is usually de-scribed in the hadron resonance gas model (HRGM)[19โ€“27] by the classical second virial coefficients.

Nevertheless, such a repulsion is of principal impor-tance for the HRGM. Otherwise, if one considers amixture of ideal gases of all known hadrons and theirresonances, then, at high temperatures, the pressureof such a system will exceed the one of the ideal gas ofmassless quarks and gluons [43]. Since such a behav-ior contradicts the lattice version of quantum chro-modynamics, the (weak) hard-core repulsion in theHRGM is absolutely necessary. Moreover, to our bestknowledge, there is no other approach which is able toinclude all known hadronic states into considerationand to be consistent with the thermodynamics of lat-

tice quantum chromodynamics at low-energy densi-ties and which, simultaneously, would not contradictit at the higher ones.

Therefore, it seems that the necessity of a weakrepulsion between the hadrons is naturally encodedin the smaller values of their hard-core radii (๐‘…๐‘ << 0.4 fm) obtained within the HRGM compared tothe larger hard-core radius of nucleons in nuclearmatter ๐‘…๐‘› โ‰ฅ 0.52 fm found in [37]. This conclu-sion is well supported by the recent simulations ofthe neutron star properties with masses more thantwo Solar ones [41] which also favors the nucleonhard-core radii below than 0.52 fm. Furthermore, thesmall values of the hard-core radii provide the fulfill-ment of the causality condition in the hadronic phase[25, 26, 41, 46], while a possible break of causalityoccurs in the region, where the hadronic degrees offreedom are not relevant [46]. Hence, in contrast toRef. [37], we do not see any reason to believe that themean-field model describing the nuclear matter prop-erties may set any strict conditions on the hadronichard-core radii of the HRGM.

Moreover, we would like to point out that a greatsuccess achieved recently by the HRGM [19โ€“27] setsa strong restriction on any model of hadronic phasewhich is claimed to be realistic. The point is that,at the chemical freeze-out curve ๐œ‡ = ๐œ‡CFO(๐‘‡ ), themean-field interaction term of pressure (1) or (29)must vanish. Otherwise, one would need a special pro-cedure to transform the mean-field potential energyinto the masses and kinetic energy of non-interactinghadrons (the kinetic freeze-out problem [44,45]). Theexisting versions of the HRGM do not face such aproblem, since this model has the hard-core repul-sion only, while the mean-field interaction in it is setto zero [19โ€“27]. Due to such a choice of the interac-tion, the HRGM has the same energy per particle asan ideal gas. Hence, it can be tuned to describe theexisting experimental hadronic multiplicities in cen-tral nuclear collisions from the lower AGS collisionenergy

โˆš๐‘ ๐‘๐‘ = 2.76 GeV to the ALICE center of

mass energyโˆš๐‘ ๐‘๐‘ = 2.76 TeV with the total quality

of fit ๐œ’2/dof โ‰ƒ 1.04 [25, 26].Therefore, any realistic hadronic EoS of hadronic

matter should be able to reproduce the pressure, en-tropy, and all charge densities obtained by the HRGMat the chemical freeze-out curve ๐œ‡ = ๐œ‡CFO(๐‘‡ ). Inparticular, for the mean-field models discussed here,it means that they should be extended in order to

876 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

include all other hadrons and, at the curve ๐œ‡ == ๐œ‡CFO(๐‘‡ ), the total interaction pressure must van-ish, i.e., ๐‘ƒint = 0, since it does not exist in the HRGM.

In other words, if, at the chemical freeze-out curve,such a model EoS has a non-vanishing attraction,then it must have an additional repulsion to provide๐‘ƒint = 0. Only this condition will help one to avoid ahard mathematical problem of kinetic freeze-out toconvert the interacting particles into a gas of freestreaming particles [44,45], since the HRGM with thehard-core repulsion and with vanishing mean-field in-teraction has the same energy per particle as an idealgas. Due to its importance, we analyzed the IST EoSin Appendix and show that this EoS also possessessuch a property. The condition ๐‘ƒint = 0 at the chemi-cal freeze-out curve will make a direct connection be-tween the realistic EoS and the hadron multiplicitiesmeasured in heavy ion collisions. It is clear that, with-out ๐‘‡ -dependent mean-field interaction ๐‘ƒint, such acondition cannot be fulfilled.

Despite many valuable results obtained with theHRGM, the hard-core radii are presently well estab-lished for the most abundant hadrons only, namely,for pions (๐‘…๐œ‹ โ‰ƒ 0.15 fm), the lightest Kยฑ-mesons(๐‘…๐พ โ‰ƒ 0.395 fm), nucleons (๐‘…๐‘ โ‰ƒ 0.365 fm),and the lightest (anti)ฮ›-hyperons (๐‘…ฮ› โ‰ƒ 0.085 fm)[25, 26]. Nevertheless, we hope for that the new high-quality data on the yields of many strange hadronsrecently measured by the ALICE Collaboration atCERN [47] at the center of mass energy

โˆš๐‘ ๐‘๐‘ =

= 2.76 TeV and the ones which are expected to bemeasured during the Beam Energy Scan II at RHICBNL (Brookhaven) [48], and at the accelerators ofnew generations, i.e., at NICA JINR (Dubna) [49,50]and FAIR GSI (Darmstadt) [51,52] will help us to de-termine their hard-core radii with high accuracy. Wehave to add only that the IST EoS for quantum gasesis well suited for such a task due to the additive pres-sure ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1,2), whereas the generalization of theCS EoS of Ref. [37] to a multicomponent case looksrather problematic, since the CS EoS [17] is the one-component EoS by construction.

6. Conclusions

The self-consistent generalization of the IST EoS forquantum gases is worked out. It is shown that, withthis EoS, one can go beyond the VdW approximationat any temperature. The restrictions on the tempera-

ture dependence of the mean-field potentials are dis-cussed. It is found that, at low temperatures, thesepotentials either should be ๐‘‡ -independent or shouldvanish faster than the first power of the temperatureproviding the fulfillment of the Third Law of thermo-dynamics. The same is true for the quantum VdWEoS. Hence, the idea to improve the quantum VdWEoS by tuning the interaction part of the pressure[14, 15] is disproved for low temperatures ๐‘‡ : if thispart of the pressure is linear in ๐‘‡ , then the VdW EoSbreaks down the Third Law of thermodynamics; ifit vanishes faster than the first power of ๐‘‡ , then theinteraction part of the pressure is useless, since it van-ishes faster than the first term of the quantum virialexpansion. An alternative EoS [37] allowing one toabandon the VdW approximation for nuclear matteris analyzed here, and it is shown that, for the sameparameters, the IST EoS is softer at low temperaturesat packing fractions ๐œ‚ โ‰ฅ 0.05.

The virial expansions for the quantum VdW andIST EoS are established, and the explicit expressionsfor all quantum virial coefficients, exact for VdW andapproximative ones for the IST EoS, are given. The-refore, for the first time, the analytical expressionsfor the third and fourth quantum virial coefficientsare derived for the EoS which is more realistic thanthe VdW one. The source of softness of the IST EoS isdemonstrated, by using the effective virial expansionfor the effective proper volume which turns out tobe compressible. The generalization of the traditionalvirial expansions for the mixtures of particles withdifferent hard-core radii is straightforward.

The general constraints on the realistic EoS fornuclear and hadronic matter are discussed. We hopefor that, by using the revealed properties of the ISTEoS for quantum gases, it will be possible to gofar beyond the traditional VdW approximation, and,due to its advantages, this EoS will become a use-ful tool for heavy ion physics and for nuclear astro-physics. Furthermore, we hope for that the developedEoS will help us to determine the hard-core radii ofhadrons from the new high-quality ALICE data andthe ones which will be measured at RHIC, NICA, andFAIR.

The authors appreciate the valuable comments ofD.B.Blaschke, R. Emaus, B.E.Grinyuk, D.R.Oliiny-chenko, and D.H. Rischke. K.A.B., A.I.I., V.V.S.and G.M.Z. acknowledge a partial support of the

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 877

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

National Academy of Sciences of Ukraine (projectNo. 0118U003197). V.V.S. thanks the Fundacao paraa Ciencia e Tecnologia (FCT), Portugal, the Mul-tidisciplinary Center for Astrophysics (CENTRA),Instituto Superior Tecnico, Universidade de Lisboa,for the partial financial support through the GrantNo. UID/FIS/00099/2013. The work of A.I. wasperformed within the project SA083P17 of Universi-dad de Salamanca launched by the Regional Govern-ment of Castilla y Leon and the European RegionalDevelopment Fund.

APPENDIX

Here, we consider a special choice of the mean-field potentialswhich are temperature-independent, i.e., ๐‘ˆ๐ด = ๐‘ˆ๐ด(๐‘›๐‘–๐‘‘๐ด) andshow that, at low particle densities, the energy per particle ofsuch an EoS coincides with the one of the ideal gas. The analy-sis is made for a single sort of particles, but it is evident that ageneralization to the multicomponent case is straightforward.

For the considered choice of the mean-field potentials,Eq. (55) for the entropy per particle becomes

๐‘ 1

๐‘›1=

[๐‘ ๐‘–๐‘‘ 1๐‘›๐‘–๐‘‘ 1

โˆ’ 3๐‘‰0 ๐‘›2๐‘ ๐‘–๐‘‘ 2๐‘›๐‘–๐‘‘ 2

][1โˆ’ 3๐‘‰0 ๐‘›2]

โ‰ƒ๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1, (65)

where, in the first step, we applied the relation ๐‘ ๐‘–๐‘‘๐ด = ๐‘ ๐‘–๐‘‘๐ด

with ๐ด โˆˆ {1; 2} to Eq. (55), while, in the second step, weused an approximation ๐‘ ๐‘–๐‘‘ 2

๐‘›๐‘–๐‘‘ 2โ‰ƒ ๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1. The latter result fol-

lows from condition (36). Then, in the low-density limit, i.e.,

for ๐‘’๐œˆ2โˆ’๐œˆ1

๐‘‡ โ‰ƒ 1, one gets relation (35) for the distributionfunctions ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) and ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1) which can be approx-imated further on as ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ2) โ‰ƒ ๐œ‘๐‘–๐‘‘(๐‘˜, ๐‘‡, ๐œˆ1). Therefore,one finds ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ2) โ‰ƒ ๐‘๐‘–๐‘‘(๐‘‡, ๐œˆ1), ๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ2) โ‰ƒ ๐‘›๐‘–๐‘‘(๐‘‡, ๐œˆ1) and๐‘ ๐‘–๐‘‘(๐‘‡, ๐œˆ2) โ‰ƒ ๐‘ ๐‘–๐‘‘(๐‘‡, ๐œˆ1).

The energy per particle for EoS (29) can be found from thethermodynamic identity

๐œ–1

๐‘›1= ๐‘‡

๐‘ 1

๐‘›1+ ๐œ‡โˆ’

๐‘(๐‘‡, ๐œ‡)

๐‘›1. (66)

Expressing the chemical potential ๐œ‡ via an effective one ๐œˆ1 fromEq. (31), one can write ๐œ‡ = ๐œˆ1+๐‘‰0๐‘๐‘–๐‘‘ 1โˆ’๐‘‰0๐‘ƒint 1+3๐‘‰0๐‘๐‘–๐‘‘ 2 โˆ’โˆ’ 3๐‘‰0๐‘ƒint 2 โˆ’ ๐‘ˆ1. Substituting this result into Eq. (66), onefinds๐œ–1

๐‘›1โ‰ƒ ๐‘‡

๐‘ ๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1+ ๐œˆ1 โˆ’ ๐‘ˆ1 +

[๐‘‰0 โˆ’

1

๐‘›1

](๐‘๐‘–๐‘‘ 1 โˆ’ ๐‘ƒint 1)+

+3๐‘‰0(๐‘๐‘–๐‘‘ 2 โˆ’ ๐‘ƒint 2), (67)

where Eq. (65) was also used. Approximating the particle num-ber density ๐‘›1 in Eq. (42) as

๐‘›1 โ‰ƒ๐‘›๐‘–๐‘‘ 1

1 + ๐‘‰0 ๐‘›๐‘–๐‘‘ 1 + 3๐‘‰0 ๐‘›2, (68)

and substituting it into Eq. (67), one obtains

๐œ–1

๐‘›1โ‰ƒ

๐œ–๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1+ 3๐‘‰0๐‘›2

[๐‘๐‘–๐‘‘ 2

๐‘›2โˆ’

๐‘๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1

]โˆ’ ๐‘ˆ1 โˆ’

โˆ’[๐‘‰0 โˆ’

1

๐‘›1

]๐‘ƒint 1 โˆ’ 3๐‘‰0๐‘ƒint 2, (69)

where we applied the thermodynamic identity (66) to the en-ergy per particle for a gas of point-like particles with the chem-ical potential ๐œˆ1. To simplify the evaluation, we assume for themoment that all mean-field interaction terms obey the equality

(1โˆ’ ๐‘‰0๐‘›1)

๐‘›1๐‘ƒint 1(๐‘›๐‘–๐‘‘ 1)โˆ’ 3๐‘‰0๐‘ƒint 2(๐‘›๐‘–๐‘‘ 2) = ๐‘ˆ1(๐‘›๐‘–๐‘‘ 1). (70)

Using the first two terms of the virial expansion (8) in Eq. (69)for the pressures ๐‘๐‘–๐‘‘ 1 and ๐‘๐‘–๐‘‘ 2 and Eq. (43) for ๐‘›2, one finds๐‘๐‘–๐‘‘ 2

๐‘›2โˆ’

๐‘๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1โ‰ƒ ๐‘‡

[(1 + ๐‘Ž

(0)2 ๐‘›๐‘–๐‘‘ 2)(1 + 3๐›ผ๐‘‰0๐‘›๐‘–๐‘‘ 2) โˆ’

โˆ’ (1 + ๐‘Ž(0)2 ๐‘›๐‘–๐‘‘ 1)

]โ‰ƒ ๐‘‡ (1 + ๐‘Ž

(0)2 ๐‘›๐‘–๐‘‘ 1)3๐›ผ๐‘‰0๐‘›๐‘–๐‘‘ 1, (71)

where, in the last step of the derivation, we used the low-density approximation ๐‘›๐‘–๐‘‘ 2 โ‰ƒ ๐‘›๐‘–๐‘‘ 1. Finally, under condition(70), Eq. (69) acquires the form๐œ–1

๐‘›1โ‰ƒ

๐œ–๐‘–๐‘‘ 1

๐‘›๐‘–๐‘‘ 1+ 9๐›ผ๐‘‰ 2

0 ๐‘›2๐‘›๐‘–๐‘‘ 1 ๐‘‡ (1 + ๐‘Ž(0)2 ๐‘›๐‘–๐‘‘ 1). (72)

Since the typical packing fractions ๐œ‚ = ๐‘‰0๐‘›1 โ‰ƒ ๐‘‰0๐‘›2 โ‰ƒ ๐‘‰0๐‘›๐‘–๐‘‘ 1

of the hadron resonance gas model at the chemical freeze-outdo not exceed the value 0.05 [25], the second term on the right-hand side of Eq. (72) is not larger than

0.025๐›ผ๐‘‡ (1 + ๐‘Ž(0)2 ๐‘›๐‘–๐‘‘ 1). (73)

Comparing this estimate with the energy per particle for thelightest hadrons, i.e., for pions, in the non-relativistic limit๐œ–๐‘–๐‘‘ 1๐‘›๐‘–๐‘‘ 1

๐œ‹โ‰ƒ ๐‘š๐œ‹ + 3

2๐‘‡ (here, ๐‘š๐œ‹ โ‰ƒ 140 MeV), one can be sure

that, for temperatures at which the hadron gas exists, i.e.,for ๐‘‡ < 160 MeV, term (73) is negligible. Hence, one finds๐œ–1๐‘›1

โ‰ƒ ๐œ–๐‘–๐‘‘ 1๐‘›๐‘–๐‘‘ 1

with high accuracy.Now, we discuss condition (70). It is apparent that, in the

general case, it can hold, if the mean-field interaction is absent,i.e., ๐‘ˆ1 = ๐‘ˆ2 = 0 and ๐‘ƒint 1 = ๐‘ƒint 2 = 0. This is exactly thecase of the hadron resonance gas model. However, one mightthink that there exists a special case for which Eq. (70) is thesimple differential equation for two independent variables ๐‘›๐‘–๐‘‘ 1

and ๐‘›๐‘–๐‘‘ 2. Let us show that this is impossible. First, with thehelp of Eq. (42), we rewrite the term (1โˆ’๐‘‰0๐‘›1)

๐‘›1= [๐‘›๐‘–๐‘‘ 1(1โˆ’

โˆ’ 3๐‘‰0๐‘›2)]โˆ’1. Then Eq. (70) can be cast as

๐‘ƒint 1(๐‘›๐‘–๐‘‘ 1)/๐‘›๐‘–๐‘‘ 1

(1โˆ’ 3๐‘‰0๐‘›2(๐‘›๐‘–๐‘‘ 2))โˆ’ 3๐‘‰0๐‘ƒint 2(๐‘›๐‘–๐‘‘ 2) = ๐‘ˆ1(๐‘›๐‘–๐‘‘ 1). (74)

From this equation, one sees that the only possibility to de-couple the dependences on ๐‘›๐‘–๐‘‘ 1 and ๐‘›2 in the first termabove is to assume that ๐‘ƒint 1 = ๐ถ๐‘›๐‘–๐‘‘ 1 where ๐ถ is someconstant. However, in this case, one finds that the ๐‘›๐‘–๐‘‘ 1-dependence of the right-hand side of Eq. (74) remains, since๐‘ˆ1 = ๐ถ ln(๐‘›๐‘–๐‘‘ 1). Therefore, there is a single possibility to de-couple the functional dependence of ๐‘›๐‘–๐‘‘ 1 from ๐‘›2, namely,๐ถ = 0 which means that ๐‘ƒint 2 = 0.

One can, however, consider Eq. (74) under the low-densityapproximation, by assuming that ๐‘›๐‘–๐‘‘ 2 = ๐‘›๐‘–๐‘‘ 1. In this case,Eq. (74) defines the functional dependence of ๐‘ƒint 2(๐‘›๐‘–๐‘‘ 1) for

878 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10

Equation of State of Quantum Gases Beyond the Van der Waals Approximation

any reasonable choice of the potential ๐‘ˆ1(๐‘›๐‘–๐‘‘ 1). Note that, inthis case, the function ๐‘ƒint 2(๐‘›๐‘–๐‘‘ 1) can be rather complicatedeven for the simplest choice of ๐‘ˆ1(๐‘›๐‘–๐‘‘ 1). Hence, the practicalrealization of dependence (74) seems to be problematic. There-fore, the most direct way to avoid the problem to convert theinteracting particles into the free streaming ones [44, 45] is touse only the hard-core repulsion between hadrons and set allother interactions at the chemical freeze-out to zero.

1. N.K. Glendenning. Compact Stars (Springer, 2000).2. S. Typel, G. Ropke, T. Klahn, D. Blaschke, H.H. Wolter.

Composition and thermodynamics of nuclear matter withlight clusters. Phys. Rev. C 81, 015803 (2010) and refer-ences therein.

3. S. Benic, D. Blaschke, D.E. Alvarez-Castillo, T. Fischer,S. Typel. A new quark-hadron hybrid equation of statefor astrophysics I. High-mass twin compact stars. Astron.Astrophys. 577, A40 (2015).

4. J.D. Walecka. A theory of highly condensed matter. AnnalsPhys. 83, 491 (1974) .

5. J. Zimanyi et al. An interpretable family of equations ofstate for dense hadronic matter. Nucl. Phys. A 484, 647(1988).

6. K.A. Bugaev, M.I. Gorenstein. Thermodynamically self-consistent class of nuclear matter EOS and compressionshocks in relativistic nuclear collisions. Z. Phys. C 43, 261(1989).

7. D.H. Rischke, M.I. Gorenstein, H. Stocker, W. Greiner.Excluded volume effect for the nuclear matter equation ofstate. Z. Phys. C 51, 485 (1991).

8. T.H.R. Skyrme. CVII. The nuclear surface. Phil. Mag. 1,1043 (1956).

9. T.H.R. Skyrme. The effective nuclear potential. Nucl.Phys. 9, 615 (1959).

10. J.P. Hansen, I.R. McDonald. Theory of Simple Liquids(Academic, 2006).

11. L. Van Hove. Quelques proprietes generales de lโ€™integralede configuration Dโ€™un systeme de particules avec interac-tion. Physica 15, 951 (1949).

12. L. Van Hove. Sur lโ€™integrale de configuration pour les sys-temes de particules a une dimension. Physica 16, 137(1950).

13. M.I. Gorenstein, D.H. Rischke, H. Stocker, W. Greiner,K.A. Bugaev. A self-consistent equation of state for nuclearmatter. J. Phys. G 19, L69 (1993).

14. D. Anchishkin, V. Vovchenko. Mean-field approach in themulti-component gas of interacting particles applied to rel-ativistic heavy-ion collisions. arXiv:1411.1444 [nucl-th] andreferences therein.

15. V. Vovchenko, D.V. Anchishkin, M.I. Gorenstein. Van derWaals equation of state with Fermi statistics for nuclearmatter. Phys. Rev. C 91, 064314 (2015).

16. K. Redlich, K. Zalewski. Thermodynamics of van derWaals fluids with quantum statistics. Acta Phys. Polon.B 47, 1943 (2016).

17. N.F. Carnahan, K.E. Starling. Equation of state for nonat-tracting rigid spheres. J. Chem. Phys. 51, 635 (1969).

18. K. Huang. Statistical Mechanics (Wiley, 1963).19. A. Andronic, P. Braun-Munzinger, J. Stachel. Hadron pro-

duction in central nucleusโ€“nucleus collisions at chemicalfreeze-out. Nucl. Phys. A 772, 167 (2006) and referencestherein.

20. K.A. Bugaev, D.R. Oliinychenko, A.S. Sorin, G.M. Zinov-jev. Simple solution to the strangeness horn descriptionpuzzle. Eur. Phys. J. A 49, 30 (2013).

21. D.R. Oliinychenko, K.A. Bugaev, A.S. Sorin. Investigationof hadron multiplicities and hadron yield ratios in heavyion collisions. Ukr. J. Phys. 58, 211 (2013).

22. K.A. Bugaev et al. Chemical freeze-out of strange parti-cles and possible root of strangeness suppression. Euro-phys. Lett. 104, 22002 (2013).

23. J. Stachel, A. Andronic, P. Braun-Munzinger, K. Redlich.Confronting LHC data with the statistical hadronizationmodel. J. Phys. Conf. Ser. 509, 012019 (2014) and refer-ences therein.

24. V.V. Sagun. ฮ›-anomaly in the hadron chemical freeze-out.Ukr. J. Phys. 59, 755 (2014).

25. K.A. Bugaev et al. Going beyond the second virial coef-ficient in the hadron resonance gas mode. Nucl. Phys. A970, 133 (2018).

26. V.V. Sagun et al. Hadron resonance gas model with in-duced surface tension. Eur. Phys. J. A 54, 100 (2018).

27. K.A. Bugaev et al. Threshold collision energy of the QCDphase diagram tricritical endpoint. Phys. Part. Nucl. Lett.15, 210 (2018).

28. V.V. Sagun, K.A. Bugaev, A.I. Ivanytskyi, I.N. Mishustin.The statistical multifragmentation model for liquidโ€“gasphase transition with a compressible nuclear liquid. Nucl.Phys. A 924, 24 (2014).

29. K.A. Bugaev, M.I. Gorenstein, I.N. Mishustin, W. Greiner.Exactly soluble model for nuclear liquid-gas phase transi-tion. Phys. Rev. C 62, 044320 (2000).

30. J.P. Bondorf et al. Statistical multifragmentation of nuclei.Phys. Rep. 257, 131 (1995).

31. A. Kostyuk, M.I. Gorenstein, H. Stocker, W. Greiner. Sec-ond cluster integral and excluded volume effects for thepion gas. Phys. Rev. C 63, 044901 (2001).

32. J. Kolafa, M. Rottner. Simulation-based equation of stateof the hard disk fluid and prediction of higher-order virialcoefficients. Mol. Phys. 104, 3435 (2006).

33. B. Barboy, W.M. Gelbart. Series representation of theequation of state for hard particle fluids. J. Chem. Phys.71, 3053 (1979).

34. S. Typel. Variations on the excluded-volume mechanism.Eur. Phys. J. A 52, 16 (2016).

35. A.I. Ivanytskyi, K.A. Bugaev, V.V. Sagun, L.V. Bravina,E.E. Zabrodin. Influence of flow constraint on the proper-ties of nuclear matter critical endpoint. Phys. Rev. C 97,064905 (2018).

36. K.A. Bugaev. The Van-der-Waals gas EOS for the Lorentzcontracted spheres. Nucl. Phys. A 807, 251 (2008) andreferences therein.

ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10 879

K.A. Bugaev, A.I. Ivanytskyi, V.V. Sagun et al.

37. V. Vovchenko. Equations of state for real gases on the nu-clear scale. Phys. Rev. C 96, 015206 (2017).

38. P. Danielewicz, R. Lacey, W.G. Lynch. Determination ofthe equation of state of dense matter. Science 298, 1592(2002).

39. M. Dutra et al. Relativistic mean-field hadronic modelsunder nuclear matter constraints. Phys. Rev. C 90, 055203(2014) and references therein.

40. O. Lourenco, M. Dutra, D.P. Menezes. Critical parametersof consistent relativistic mean-field models. Phys. Rev. C95, 065212 (2017).

41. V.V. Sagun, I. Lopes. Neutron stars: A novel equation ofstate with induced surface tension. Astrophys. J 850, 75(2017).

42. R. Venugopalan, M. Prakash. Thermal properties of inter-acting hadrons. Nucl. Phys. A 546, 718 (1992).

43. L.M. Satarov, M.N. Dmitriev, I.N. Mishustin. Equation ofstate of hadron resonance gas and the phase diagram ofstrongly interacting matter. Phys. Atom. Nucl. 72, 1390(2009).

44. K.A. Bugaev. Shock-like freeze-out in relativistic hydrody-namics. Nucl. Phys. A 606, 559 (1996).

45. K.A. Bugaev. Relativistic kinetic equations for finite do-mains and freeze-out problem. Phys. Rev. Lett. 90, 252301(2003) and references therein.

46. L.M. Satarov, K.A. Bugaev, I.N. Mishustin. Equation ofstate and sound velocity of a hadronic gas with a hard-core interaction. Phys. Rev. C 91, 055203 (2015).

47. N. Agrawal et. al. Probing the hadronic phase with reso-nances of different life-times in Pb-Pb collisions with AL-ICE. [ALICE Collaboration] arXiv:1711.02408v1 [hep-ex].

48. L. Adamczyk et. al. [STAR Collaboration] Probing partondynamics of QCD matter with ฮฉ and ๐œ‘ production. Phys.Rev. C 93, 021903 (2016).

49. D. Rischke. Exploring strongly interacting matter at highdensities โ€“ NICA white paper. Eur. Phys. J. A 52, 267(2016).

50. P. Senger. Nuclear matter physics at NICA. Eur. Phys. J.A 52, 217 (2016).

51. P. Senger. The compressed baryon matter experiment atFAIR. Nucl. Phys. A 862โ€“863, 139 (2011).

52. T. Ablyazimov et al. Challenges in QCD matter physics โ€“the scientific programme of the Compressed Baryon Matterexperiment at FAIR. Eur. Phys. J. A 53, 60 (2017).

Received 11.04.18

ะš.ะž.ะ‘ัƒะณะฐั”ะฒ, ะž.I. Iะฒะฐะฝะธั†ัŒะบะธะน,ะ’.ะ’.ะกะฐะณัƒะฝ, ะ•.ะ“.ะiะบะพะฝะพะฒ, ะ“.ะœ. ะ—iะฝะพะฒโ€™ั”ะฒ

ะ Iะ’ะะฏะะะฏ ะกะขะะะฃ ะšะ’ะะะขะžะ’ะ˜ะฅ ะ“ะะ—Iะ’ ะŸะžะ—ะะ ะะœะšะะœะ˜ ะะะ‘ะ›ะ˜ะ–ะ•ะะะฏ ะ’ะะ-ะ”ะ•ะ -ะ’ะะะ›ะฌะกะ

ะ  ะต ะท ัŽ ะผ ะต

ะะตั‰ะพะดะฐะฒะฝะพ ะทะฐะฟั€ะพะฟะพะฝะพะฒะฐะฝะต ั€iะฒะฝัะฝะฝั ัั‚ะฐะฝัƒ ะท iะฝะดัƒะบะพะฒะฐะฝะธะผะฟะพะฒะตั€ั…ะฝะตะฒะธะผ ะฝะฐั‚ัะณะพะผ ัƒะทะฐะณะฐะปัŒะฝะตะฝะพ ะฝะฐ ะฒะธะฟะฐะดะพะบ ะบะฒะฐะฝั‚ะพะฒะธั… ะณะฐ-ะทiะฒ iะท ะฒะทะฐั”ะผะพะดiั”ัŽ ัะตั€ะตะดะฝัŒะพะณะพ ะฟะพะปั. ะ”ะปั ั‚ะฐะบะพั— ะผะพะดะตะปi ะทะฝะฐ-ะนะดะตะฝะพ ัƒะผะพะฒะธ ัะฐะผะพัƒะทะณะพะดะถะตะฝะพัั‚i i ัƒะผะพะฒะธ, ะฝะตะพะฑั…iะดะฝi ะดะปั ะฒะธ-ะบะพะฝะฐะฝะฝั ะขั€ะตั‚ัŒะพะณะพ ะŸะพั‡ะฐั‚ะบัƒ ั‚ะตั€ะผะพะดะธะฝะฐะผiะบะธ. ะะฐ ะฒiะดะผiะฝัƒ ะฒiะดั‚ั€ะฐะดะธั†iะนะฝะธั… ัะฟะพะดiะฒะฐะฝัŒ ะฟะพะบะฐะทะฐะฝะพ, ั‰ะพ ะฒะฝะตัะตะฝะฝั ะฒ ั‚ะธัะบ ะผะพ-ะดะตะปi ะ’ะฐะฝ-ะดะตั€-ะ’ะฐะฐะปัŒัะฐ ั‚ั€ะตั‚ัŒะพะณะพ i ะฑiะปัŒัˆ ะฒะธัะพะบะธั… ะฒiั€iะฐะปัŒะฝะธั…ะบะพะตั„iั†iั”ะฝั‚iะฒ ะณะฐะทัƒ ั‚ะฒะตั€ะดะธั… ัั„ะตั€ ะทะฐ ะฝะธะทัŒะบะธั… ั‚ะตะผะฟะตั€ะฐั‚ัƒั€ ะฐะฑะพะฟะพั€ัƒัˆัƒั” ะขั€ะตั‚iะน ะŸะพั‡ะฐั‚ะพะบ ั‚ะตั€ะผะพะดะธะฝะฐะผiะบะธ, ะฐะฑะพ ะฝะต ะดะพะทะฒะพะปัั”ะฒะธะนั‚ะธ ะทะฐ ั€ะฐะผะบะธ ะฝะฐะฑะปะธะถะตะฝะฝั ะ’ะฐะฝ-ะดะตั€-ะ’ะฐะฐะปัŒัะฐ. ะŸั€ะพะดะตะผะพะฝ-ัั‚ั€ะพะฒะฐะฝะพ, ั‰ะพ ัƒะทะฐะณะฐะปัŒะฝะตะฝะต ั€iะฒะฝัะฝะฝั ัั‚ะฐะฝัƒ ะท iะฝะดัƒะบะพะฒะฐะฝะธะผ ะฟะพ-ะฒะตั€ั…ะฝะตะฒะธะผ ะฝะฐั‚ัะณะพะผ ะดะพะทะฒะพะปัั” ัƒะฝะธะบะฝัƒั‚ะธ ั†ะธั… ะฟั€ะพะฑะปะตะผ i ะฒะธะนั‚ะธะทะฐ ั€ะฐะผะบะธ ะฝะฐะฑะปะธะถะตะฝะฝั ะ’ะฐะฝ-ะดะตั€-ะ’ะฐะฐะปัŒัะฐ. ะšั€iะผ ั†ัŒะพะณะพ ะพั‚ั€ะธะผะฐ-ะฝะพ ะตั„ะตะบั‚ะธะฒะฝะต ะฒiั€iะฐะปัŒะฝะต ั€ะพะทะบะปะฐะดะฐะฝะฝั ะบะฒะฐะฝั‚ะพะฒะพั— ะฒะตั€ัiั— ั€iะฒะฝั-ะฝะฝั ัั‚ะฐะฝัƒ ะท iะฝะดัƒะบะพะฒะฐะฝะธะผ ะฝะฐั‚ัะณะพะผ i ะนะพะณะพ ะฒiั€iะฐะปัŒะฝi ะบะพะตั„iั†iั”ะฝ-ั‚ะธ ะทะฝะฐะนะดะตะฝะพ ั‚ะพั‡ะฝะพ. ะฏะฒะฝi ะฒะธั€ะฐะทะธ ะดะปั ัะฟั€ะฐะฒะถะฝiั… ะบะฒะฐะฝั‚ะพะฒะธั…ะฒiั€iะฐะปัŒะฝะธั… ะบะพะตั„iั†iั”ะฝั‚iะฒ ะฑัƒะดัŒ-ัะบะพะณะพ ะฟะพั€ัะดะบัƒ ั†ัŒะพะณะพ ั€iะฒะฝัะฝะฝััั‚ะฐะฝัƒ ะฟะพะดะฐะฝะพ ะฒ ะฝะฐะฑะปะธะถะตะฝะฝi ะฝะธะทัŒะบะพั— ะณัƒัั‚ะธะฝะธ. ะžะฑะผiั€ะบะพะฒะฐะฝะพะดะตัะบi ะฑะฐะทะพะฒi ัƒะผะพะฒะธ ะฝะฐ ั‚ะฐะบi ะผะพะดะตะปi, ัะบi ะฝะตะพะฑั…iะดะฝi ะดะปั ะพะฟะธััƒะฒะปะฐัั‚ะธะฒะพัั‚ะตะน ัะดะตั€ะฝะพั— i ะฐะดั€ะพะฝะฝะพั— ะผะฐั‚ะตั€iะน.

880 ISSN 2071-0194. Ukr. J. Phys. 2018. Vol. 63, No. 10