The self-consistent field model for Fermi systems with account of three-body interactions
On the basis of a microscopic model of self-consistent field, the thermodynamics of the many-particle Fermi system at finite temperatures with account of three-body interactions is built and the quasiparticle equations of motion are obtained. It is shown that the delta-like three-body interaction gi...
Saved in:
Date: | 2015 |
---|---|
Main Authors: | , , |
Format: | Article |
Language: | English |
Published: |
Інститут фізики конденсованих систем НАН України
2015
|
Series: | Condensed Matter Physics |
Online Access: | http://dspace.nbuv.gov.ua/handle/123456789/155261 |
Tags: |
Add Tag
No Tags, Be the first to tag this record!
|
Journal Title: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Cite this: | The self-consistent field model for Fermi systems with account of three-body interactions / Yu.M. Poluektov, A.A. Soroka, S.N. Shulga // Condensed Matter Physics. — 2015. — Т. 18, № 4. — С. 43005: 1–21. — Бібліогр.: 41 назв. — англ. |
Institution
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-155261 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1552612019-06-17T01:28:22Z The self-consistent field model for Fermi systems with account of three-body interactions Poluektov, Yu.M. Soroka, A.A. Shulga, S.N. On the basis of a microscopic model of self-consistent field, the thermodynamics of the many-particle Fermi system at finite temperatures with account of three-body interactions is built and the quasiparticle equations of motion are obtained. It is shown that the delta-like three-body interaction gives no contribution into the self-consistent field, and the description of three-body forces requires their nonlocality to be taken into account. The spatially uniform system is considered in detail, and on the basis of the developed microscopic approach general formulas are derived for the fermion's effective mass and the system's equation of state with account of contribution from three-body forces. The effective mass and pressure are numerically calculated for the potential of ``semi-transparent sphere'' type at zero temperature. Expansions of the effective mass and pressure in powers of density are obtained. It is shown that, with account of only pair forces, the interaction of repulsive character reduces the quasiparticle effective mass relative to the mass of a free particle, and the attractive interaction raises the effective mass. The question of thermodynamic stability of the Fermi system is considered and the three-body repulsive interaction is shown to extend the region of stability of the system with the interparticle pair attraction. The quasiparticle energy spectrum is calculated with account of three-body forces. На основi мiкроскопiчної моделi самоузгодженого поля побудовано термодинамiку системи багатьох фермi-частинок при скiнчених температурах з врахуванням тричастинкових взаємодiй i отримано рiвняння руху квазiчастинок. Показано, що дельтаподiбна тричастинкова взаємодiя не дає внеску в самоузгоджене поле, i для опису тричастинкових сил слiд враховувати їх нелокальнiсть. Детально розглянуто просторовооднорiдну систему i в рамках розвиненого мiкроскопiчного пiдходу отримано загальнi формули для ефективної маси фермiону i рiвняння стану системи з врахуванням внеску тричастинкових взаємодiй. Для потенцiалу типу “напiвпрозорої сфери” при нулi температур чисельно розраховано ефективну масу i тиск. Знайдено розвинення ефективної маси i тиску за ступенями щiльностi. Показано, що при врахуваннi тiльки парних сил, взаємодiя, що має характер вiдштовхування, зменшує ефективну масу квазiчастинки порiвняно з масою вiльної частинки, а у разi притягання — збiльшує. Розглянуто питання термодинамiчної стiйкостi фермi-системи i показано, що тричастинкова взаємодiя, яка має характер вiдштовхування, розширює область стiйкостi системи з мiжчастинковим парним притяганням. Розраховано енергетичний спектр квазiчастинки з врахуванням тричастинкових сил. 2015 Article The self-consistent field model for Fermi systems with account of three-body interactions / Yu.M. Poluektov, A.A. Soroka, S.N. Shulga // Condensed Matter Physics. — 2015. — Т. 18, № 4. — С. 43005: 1–21. — Бібліогр.: 41 назв. — англ. 1607-324X PACS: 05.30.Ch, 05.30.Fk, 05.70.-a DOI:10.5488/CMP.18.43005 arXiv:1503.02428 http://dspace.nbuv.gov.ua/handle/123456789/155261 en Condensed Matter Physics Інститут фізики конденсованих систем НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
description |
On the basis of a microscopic model of self-consistent field, the thermodynamics of the many-particle Fermi system at finite temperatures with account of three-body interactions is built and the quasiparticle equations of motion are obtained. It is shown that the delta-like three-body interaction gives no contribution into the self-consistent field, and the description of three-body forces requires their nonlocality to be taken into account. The spatially uniform system is considered in detail, and on the basis of the developed microscopic approach general formulas are derived for the fermion's effective mass and the system's equation of state with account of contribution from three-body forces. The effective mass and pressure are numerically calculated for the potential of ``semi-transparent sphere'' type at zero temperature. Expansions of the effective mass and pressure in powers of density are obtained. It is shown that, with account of only pair forces, the interaction of repulsive character reduces the quasiparticle effective mass relative to the mass of a free particle, and the attractive interaction raises the effective mass. The question of thermodynamic stability of the Fermi system is considered and the three-body repulsive interaction is shown to extend the region of stability of the system with the interparticle pair attraction. The quasiparticle energy spectrum is calculated with account of three-body forces. |
format |
Article |
author |
Poluektov, Yu.M. Soroka, A.A. Shulga, S.N. |
spellingShingle |
Poluektov, Yu.M. Soroka, A.A. Shulga, S.N. The self-consistent field model for Fermi systems with account of three-body interactions Condensed Matter Physics |
author_facet |
Poluektov, Yu.M. Soroka, A.A. Shulga, S.N. |
author_sort |
Poluektov, Yu.M. |
title |
The self-consistent field model for Fermi systems with account of three-body interactions |
title_short |
The self-consistent field model for Fermi systems with account of three-body interactions |
title_full |
The self-consistent field model for Fermi systems with account of three-body interactions |
title_fullStr |
The self-consistent field model for Fermi systems with account of three-body interactions |
title_full_unstemmed |
The self-consistent field model for Fermi systems with account of three-body interactions |
title_sort |
self-consistent field model for fermi systems with account of three-body interactions |
publisher |
Інститут фізики конденсованих систем НАН України |
publishDate |
2015 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/155261 |
citation_txt |
The self-consistent field model for Fermi systems with account of three-body interactions / Yu.M. Poluektov, A.A. Soroka, S.N. Shulga // Condensed Matter Physics. — 2015. — Т. 18, № 4. — С. 43005: 1–21. — Бібліогр.: 41 назв. — англ. |
series |
Condensed Matter Physics |
work_keys_str_mv |
AT poluektovyum theselfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions AT sorokaaa theselfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions AT shulgasn theselfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions AT poluektovyum selfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions AT sorokaaa selfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions AT shulgasn selfconsistentfieldmodelforfermisystemswithaccountofthreebodyinteractions |
first_indexed |
2025-07-14T07:19:17Z |
last_indexed |
2025-07-14T07:19:17Z |
_version_ |
1837606629231558656 |
fulltext |
Condensed Matter Physics, 2015, Vol. 18, No 4, 43005: 1–21
DOI: 10.5488/CMP.18.43005
http://www.icmp.lviv.ua/journal
The self-consistent field model for Fermi systems
with account of three-body interactions
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
Akhiezer Institute for Theoretical Physics, National Science Center “Kharkiv Institute of Physics and
Technology”, 1 Akademichna St., 61108 Kharkiv, Ukraine
Received August 3, 2015, in final form October 7, 2015
On the basis of a microscopic model of self-consistent field, the thermodynamics of the many-particle Fermi
system at finite temperatures with account of three-body interactions is built and the quasiparticle equations
of motion are obtained. It is shown that the delta-like three-body interaction gives no contribution into the self-
consistent field, and the description of three-body forces requires their nonlocality to be taken into account.
The spatially uniform system is considered in detail, and on the basis of the developed microscopic approach
general formulas are derived for the fermion’s effective mass and the system’s equation of state with account
of contribution from three-body forces. The effective mass and pressure are numerically calculated for the po-
tential of “semi-transparent sphere” type at zero temperature. Expansions of the effective mass and pressure
in powers of density are obtained. It is shown that, with account of only pair forces, the interaction of repul-
sive character reduces the quasiparticle effective mass relative to the mass of a free particle, and the attractive
interaction raises the effective mass. The question of thermodynamic stability of the Fermi system is consid-
ered and the three-body repulsive interaction is shown to extend the region of stability of the system with the
interparticle pair attraction. The quasiparticle energy spectrum is calculated with account of three-body forces.
Key words: self-consistent field, three-body interactions, effective mass, fermion, equation of state
PACS: 05.30.Ch, 05.30.Fk, 05.70.-a
1. Introduction
The self-consistent field model is an effective approach for describing systems of a large number
of particles, even in the case when the interaction between particles cannot be considered to be weak
and the density of a system to be low. This model is applicable both for systems with a finite number of
particles, such as atomic nuclei [1–3] or electronic shells of atoms andmolecules [4, 5], as well as formany-
particle systems, when describing them by methods of statistical physics. The method of self-consistent
field is especially useful in studying spatially non-uniform systems and phase transitions [6, 7]. The self-
consistent field model serves as an efficient main approximation in constructing a perturbation theory,
including such a theory for spatially non-uniform systems [1] and many-particle systems with broken
symmetries [8–11].
Systems of a large number of particles, obeying Fermi statistics, within a phenomenological approach
at large enough densities are described in the language of the Fermi liquid theory [12, 13], which origi-
nally was developed for normal limitless systems. Afterwards, the Fermi liquid theory was generalized
both for systems of finite dimensions [14, 15] and for many-particle Fermi systems with broken phase
symmetry possessing superfluid and superconducting properties [16]. The Fermi-liquid approach is, in
essence, a phenomenological variant of the self-consistent field theory [17, 18].
In contrast to the Fermi-liquid approach, in which the interaction of quasiparticles is described phe-
nomenologically by means of interaction amplitudes, the account for the interaction between real parti-
cles is laid in the basis of the developed microscopic approach. Usually the interaction between particles
is described by means of pair potentials, on the assumption that the presence of other particles does
not influence the interaction between the two selected particles. Meanwhile, for particles possessing an
© Yu.M. Poluektov, A.A. Soroka, S.N. Shulga, 2015 43005-1
http://dx.doi.org/10.5488/CMP.18.43005
http://www.icmp.lviv.ua/journal
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
internal structure, the interaction between a pair of particles is changed due to the presence of a third
particle, that can be taken into account by introducing potentials depending on the coordinates of three
particles. Such a representation follows from the consideration of exchange and multipole interactions
of more than two particles in different orders of the perturbation theory [19–21]. Essential is the fact that
contribution from three-body forces not only gives quantitative corrections to characteristics of a system
calculated with account of only pair interactions, but can be necessary for qualitative understanding of
some effects.
Accounting for three-body interactions is also important in the theory of nuclear forces [3, 15, 22],
because they model the dependence of the nucleon-nucleon interaction potential on density. In particu-
lar, in the case of the interaction proposed by Skyrme [23, 24], three-body interactions are described by
a simple delta-like potential. The role of three-body forces within the framework of the quantum field
approach is discussed in reference [25].
The effects of three-body interactions should also manifest themselves in the interaction of structure-
less particles located in a polarizable medium, for example electrons in a lattice. However, this issue is
completely unexplored so far.
In this paper, self-consistent field equations are obtained for normal (non-superfluid) Fermi systems
at finite temperatures with account of three-body forces within the approach developed earlier for Fermi
systems with pair interactions [9, 10, 17, 18]. In general, the mean (or self-consistent) field approxima-
tion can be formulated in different ways and, accordingly, different variants of this approximation are
possible. In most cases this approximation is introduced at the level of equations of motion. In the sta-
tistical description of many-particle systems it is convenient to introduce the self-consistent field model,
most consistently and in most general form, on the level of Hamiltonian rather than on the level of equa-
tions of motion. This permits in a natural way not only to derive the self-consistent equations of motion,
but also to build the thermodynamics of a many-particle system already within the scope of a particular
model.
If the self-consistent field model is formulated in a way such that all thermodynamic relations hold
exactly in it, as it takes place for the ideal quantum gases [26], then the structure of the model and all
its parameters will be determined uniquely. The method of constructing a model within the Hamiltonian
approach consists in splitting the exact initial Hamiltonian into a sum of an approximating Hamiltonian
and a residual correlation Hamiltonian. The approximating (or self-consistent) Hamiltonian is chosen as
a quadratic operator form of most general kind, that contains indeterminate fields at the initial stage.
Such a Hamiltonian, describing a many-particle system in the self-consistent field approximation, is re-
duced to a diagonal form and the fields entering it are determined from the variational principle. Within
this description, the concept of quasiparticles naturally arises in the microscopic approach. The effects
associated with the interaction of quasiparticles are accounted for by the correlation Hamiltonian and
can be calculated using the perturbation theory. It is worth noting that this method is so general that it
can be also used for a consistent description of normal and superfluid Bose systems [27, 28], in particular
phonons [29], as well as relativistic quantized fields with broken symmetries [30, 31].
From the derived self-consistent equations accounting for both pair and three-body interactions, it fol-
lows that the delta-like three-body interaction gives no contribution into the self-consistent field. There-
fore, a description of three-body interactions requires accounting for their nonlocality. Thermodynamic
properties of the spatially uniform system are studied more in detail in the paper, and general formulae
are derived for the quasiparticle effective mass and the system’s equation of state at finite temperatures
with account of both pair and three-body interactions. It is shown that, with account of only pair forces,
the repulsive interaction reduces the quasiparticle effective mass relative to the mass of a free particle,
and the attractive interaction raises it.
The effective mass of fermions and the pressure of system are numerically calculated for the “semi-
transparent sphere” potential at zero temperature, and expansions of these quantities in powers of den-
sity with account of three-body interactions are obtained. The influence of three-body forces on the stabil-
ity of the many-particle Fermi system is considered and it is shown that accounting for three-body forces
of repulsive character extends the region of stability and can lead to stabilization of the system with pair
attraction between particles. The quasiparticle energy spectrum is calculated with account of three-body
forces.
43005-2
The self-consistent field model for Fermi systems
2. Hamiltonian of the Fermi system with account of three-body
interactions
Potential energy of a system of N particles possessing an internal structure can be represented as a
sum of pair, three-body, etc. interactions
U (r1,r2, . . . ,rN ) = ∑
i< j
U (ri ,r j )+ ∑
i< j<k
U (ri ,r j ,rk )+ . . . ,
(2.1)
whereU (ri ,r j ) =U (r j ,ri ),U (ri ,r j ,rk ) is a symmetric function in all permutations of its coordinates. In
the second quantization representation, the Hamiltonian of the many-particle system with account of
pair and tree-body interactions has the form
H = T +V2 +V3 . (2.2)
Here,
T =
∫
dqdq ′Ψ+(q)H0(q, q ′)Ψ(q ′) (2.3)
is the kinetic energy and the energy in external fieldU0(q), and
H0(q, q ′) =− ħ2
2m
∆δ(q −q ′)+U0(q)δ(q −q ′). (2.4)
The energies of pair V2 and three-body V3 interactions can be written in the form
V2 = 1
2!
∫
dqdq ′Ψ+(q)Ψ+(q ′)U (q, q ′)Ψ(q ′)Ψ(q), (2.5)
V3 = 1
3!
∫
dqdq ′dq ′′Ψ+(q)Ψ+(q ′)Ψ+(q ′′)U (q, q ′, q ′′)Ψ(q ′′)Ψ(q ′)Ψ(q). (2.6)
In the formulae (2.3)–(2.6) and below, the symbol q ≡ (r,σ) designates the space coordinate r and the
spin projection σ. We assume s = 1/2. The field operators obey the known anticommutation relations
[26]. In what follows, we consider the case when all the potentials depend only on the space coordinates:
U (q, q ′) =U (r,r′),U (q, q ′, q ′′) =U (r,r′,r′′). Besides that, we assume that external potential does not de-
pend on the spin projection, so that U0(q) = U0(r). In order not to take into account the condition of
conservation of the total number of particles in all computations, we assume that the considered sys-
tem can exchange particles with a thermostat. To account for it, a term with the chemical potential µ is
introduced into Hamiltonian, so that H0(q, q ′) in (2.3) is replaced by one-particle Hamiltonian
H(q, q ′) = H0(q, q ′)−µδ(q −q ′). (2.7)
In what follows, when talking about Hamiltonian, we imply that it includes the term with the chemical
potential.
3. The self-consistent field model with account of three-body
interactions
In order to proceed to the description of a many-particle Fermi system within the self-consistent field
model, let us represent the total Hamiltonian (2.2) as a sum of two terms, i.e., the self-consistent Hamilto-
nian H0 and the correlation Hamiltonian HC:
H = H0 +HC. (3.1)
The self-consistent Hamiltonian is defined by the relation
H0 =
∫
dqdq ′Ψ+(q)
[
H(q, q ′)+W (q, q ′)
]
Ψ(q ′)+E0, (3.2)
43005-3
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
which is quadratic in the field operators of creation and annihilation. Equation (3.2) includes the self-
consistent potentialW (q, q ′), representing the mean field acting on a single particle, as well as the non-
operator term E0, both still indeterminate. Owing to hermiticity of the Hamiltonian, the next property
holds W (q, q ′) = W ∗(q ′, q). Note that taking account of the non-operator term in (3.2) is essential for a
consistent description of the thermodynamics of a system within the considered approach. The correla-
tion Hamiltonian accounts for all the effects, not accounted for in the self-consistent field model:
HC = 1
2!
∫
dqdq ′Ψ+(q)Ψ+(q ′)U (r,r′)Ψ(q ′)Ψ(q)
+ 1
3!
∫
dqdq ′dq ′′Ψ+(q)Ψ+(q ′)Ψ+(q ′′)U (r,r′,r′′)Ψ(q ′′)Ψ(q ′)Ψ(q)
−
∫
dqdq ′Ψ+(q)W (q, q ′)Ψ(q ′)−E0. (3.3)
Obviously, the total Hamiltonian H has not changed in consequence of the performed decomposition.
Transition to the self-consistent field model consists in describing the system using the approximate
quadratic Hamiltonian (3.2) instead of the exact Hamiltonian (3.1). Entering (3.2), so far indeterminate,
quantitiesW (q, q ′) and E0 should be chosen in an optimal manner, as it will be done below. The effects
conditioned by the correlation Hamiltonian can be accounted for by the perturbation theory [1, 9–11]. In
this paper we confine ourselves to consideration of only the main approximation. Note also that in con-
structing the self-consistent field theory, we do not take into account the effects connected with breaking
the phase symmetry and those leading to the properties of superfluidity and superconductivity [9, 10].
Since Hamiltonian (3.2) is quadratic in the field operators, it can be represented in the form of Hamil-
tonian of an ideal gas of quasiparticles. The field operators and the operators of creation and annihilation
of free particles a+
j , a j are connected by the relations
Ψ(q) =∑
j
φ(0)
j (q) a j , Ψ+(q) =∑
j
φ(0)∗
j (q) a+
j ,
(3.4)
where the functions φ(0)
j (q) are solutions of the Schrödinger equation for free particles∫
dq ′H0(q, q ′)φ(0)
j (q ′) = ε(0)
j φ(0)
j (q). (3.5)
Here, j ≡ (ν,σ), where ν is a full set of quantum numbers describing the state of a particle except the
spin projection σ. To represent Hamiltonian (3.2) in the form analogous to that for free particles, let us
introduce the quasiparticle operators γ+j , γ j connected with the field operators (3.4) by the relations
Ψ(q) =∑
j
φ j (q)γ j , Ψ+(q) =∑
j
φ∗
j (q)γ+j ,
(3.6)
where the functions φ j (q) are now solutions of the self-consistent equation∫
dq ′[H(q, q ′)+W (q, q ′)
]
φ j (q ′) = ε jφ j (q). (3.7)
Note that for φ j (q) functions, the same as for φ(0)
j (q), the conditions of orthonormality and completeness
hold ∫
dqφ∗
j (q)φ j ′ (q) = δ j j ′ ,
∑
j
φ∗
j (q)φ j (q ′) = δ(q −q ′). (3.8)
As a result, the self-consistent Hamiltonian (3.2) acquires the form of the Hamiltonian of an ideal gas of
quasiparticles
H0 =
∑
j
ε j γ
+
j γ j +E0,
(3.9)
where ε j means the quasiparticle energy and E0 means the energy of the background on which the
quasiparticles exist. In the self-consistent field model, the reference of the energy cannot be arbitrary
43005-4
The self-consistent field model for Fermi systems
but, as will be shown below, should be chosen in a way such that the thermodynamic relations are satis-
fied. From (3.4) and (3.6) it follows that the quasiparticle operators are explicitly expressed through the
operators of free particles
γ j =
∑
j ′
a j ′
∫
dqφ∗
j (q)φ(0)
j ′ (q), γ+j =∑
j ′
a+
j ′
∫
dqφ j (q)φ(0)∗
j ′ (q). (3.10)
In a spatially uniform non-magnetic state, the operators of particles and quasiparticles coincide, which
corresponds to the known reasoning in the Fermi liquid theory regarding the invariance of classification
of states during adiabatic “switching-on” of the interaction [12, 13].
4. Derivation of the self-consistent potential
Let us define the statistical operator
ρ̂0 = expβ(Ω−H0), (4.1)
where β= 1/T is the inverse temperature, the constantΩ=−T ln
[
Spe−βH0
]
is determined from the nor-
mality condition Spρ̂0 = 1 and, as will be seen below, has the meaning of the thermodynamic potential of
the system in the self-consistent fieldmodel. The average of an arbitrary operator A in this approximation
is defined by the relation
〈A〉 ≡ Sp(ρ̂0 A). (4.2)
Since Hamiltonian (3.9) is quadratic, then for the averages with the statistical operator (4.1) the Bloch-de
Dominicis (Wick) theorem [32] holds.
The ground state energy in the self-consistent field model is determined from the requirement of
equality of the averages calculated with the statistical operator (4.1) for the exact (3.1) and for the ap-
proximating (3.2) Hamiltonians:
〈H〉 = 〈H0〉. (4.3)
Hence, we have
E0 = 1
2!
∫
dqdq ′U (r,r′)
〈
Ψ+(q)Ψ+(q ′)Ψ(q ′)Ψ(q)
〉
+ 1
3!
∫
dqdq ′dq ′′U (r,r′,r′′)
〈
Ψ+(q)Ψ+(q ′)Ψ+(q ′′)Ψ(q ′′)Ψ(q ′)Ψ(q)
〉
−
∫
dqdq ′W (q, q ′)
〈
Ψ+(q)Ψ(q ′)
〉
. (4.4)
Let us define the one-particle density matrix by the relation
ρ(q, q ′) = 〈
Ψ+(q ′)Ψ(q)
〉=∑
i
φi (q)φ∗
i (q ′) fi .
(4.5)
Here, the quasiparticle distribution function is defined by the expression
〈
γ+i γ j
〉= fi δi j . Based on the
form of Hamiltonian (3.9), a straightforward calculation gives the Fermi-type distribution function
fi = f (εi ) = [
exp(βεi )+1
]−1. (4.6)
Since the quasiparticle energy εi is the functional of fi , the formula (4.6) represents a complicated non-
linear equation for the distribution function, being similar to that which takes place in the Landau phe-
nomenological theory of a Fermi liquid [12]. The energy E0 expressed through ρ(q, q ′) has the form:
E0 = 1
2!
∫
dqdq ′U (r,r′)
[
ρ(q, q)ρ(q ′, q ′)−ρ(q ′, q)ρ(q, q ′)
]
+ 1
3!
∫
dqdq ′dq ′′U (r,r′,r′′)
[
2ρ(q, q ′)ρ(q ′, q ′′)ρ(q ′′, q)
−3ρ(q, q)ρ(q ′, q ′′)ρ(q ′′, q ′)+ρ(q, q)ρ(q ′, q ′)ρ(q ′′, q ′′)
]
−
∫
dqdq ′W (q, q ′)ρ(q ′, q). (4.7)
43005-5
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
The variation of the thermodynamic potential Ω = −T ln
[∑
n
〈
n
∣∣e−βH0
∣∣n〉]
is equal to the averaged
variation of Hamiltonian (3.2):
δΩ=
∑
n
〈
n
∣∣e−βH0δH0
∣∣n〉∑
n
〈
n
∣∣e−βH0
∣∣n〉 = 〈
δH0
〉
. (4.8)
From the requirement that the variation of the thermodynamic potential with respect to the density
matrix (4.5) vanishes δΩ
/
δρ(q, q ′) = 0, we obtain the expression for the self-consistent potential which
consists of the contributions from pair and three-body interactions W (q, q ′) = W (2)(q, q ′)+W (3)(q, q ′),
where
W (2)(q, q ′) =−U (r,r′)ρ(q, q ′)+δ(q −q ′)
∫
dq ′′U (r,r′′)ρ(q ′′, q ′′), (4.9)
W (3)(q, q ′) =
∫
dq ′′U (r,r′,r′′)
[
ρ(q, q ′′)ρ(q ′′, q ′)−ρ(q, q ′)ρ(q ′′, q ′′)
]
+1
2
δ(q −q ′)
∫
dq ′′dq ′′′U (r,r′′,r′′′)
[
ρ(q ′′, q ′′)ρ(q ′′′, q ′′′)−ρ(q ′′, q ′′′)ρ(q ′′′, q ′′)
]
. (4.10)
The self-consistent field can be also derived by an equivalent method from the requirement that the
variation of the functional I ≡ 〈
H − H0
〉
with respect to the density matrix vanishes, where W (q, q ′)
and the parameter E0 are not varied [9, 10]. Such variation rule is conditioned by the fact that, as may be
checked, the variations of the functional δI with respect to δW (q, q ′) and δE0 aremutually compensated.
Equation (3.7), together with the derived potentials (4.9), (4.10), enables us to obtain the quasiparticle
wave functions φi (q) and the quasiparticle energies εi :
εi =− ħ2
2m
∫
dqφ∗
i (q)∆φi (q)+
∫
dq U0(r)
∣∣φi (q)
∣∣2 +
∫
dqdq ′W (q, q ′)φ∗
i (q)φi (q ′)−µ. (4.11)
Substitution of the potentials (4.9), (4.10) into equation (3.7) leads to the integro-differential equation
− ħ2
2m
∆φi (q)+ [
U0(r)−µ−εi
]
φi (q)
+
{∫
dq ′U (r,r′)ρ(q ′, q ′)+ 1
2
∫
dq ′dq ′′U (r,r′,r′′)
[
ρ(q ′, q ′)ρ(q ′′, q ′′)−ρ(q ′, q ′′)ρ(q ′′, q ′)
]}
φi (q)
+
∫
dq ′dq ′′U (r,r′,r′′)
[
ρ(q, q ′′)ρ(q ′′, q ′)−ρ(q, q ′)ρ(q ′′, q ′′)
]
φi (q ′)
−
∫
dq ′U (r,r′)ρ(q, q ′)φi (q ′) = 0. (4.12)
The chemical potential µ is associated with the average number of particles N by the relation
N =
∫
drn(r), n(r) =∑
σ
ρ(q, q). (4.13)
In many cases, finding the equilibrium characteristics of the researched system does not require the
calculation of the quasiparticle wave functions, but it is sufficient to know the one-particle densitymatrix.
From equations (4.5) and (4.12), the equation for the one-particle density matrix follows
ħ2
2m
[
∆ρ(q, q ′)−∆′ρ(q, q ′)
]−[
U0(r)−U0(r′)
]
ρ(q, q ′)
+
∫
dq ′′[U (r,r′′)−U (r′,r′′)
][
ρ(q, q ′′)ρ(q ′′, q ′)−ρ(q ′′, q ′′)ρ(q, q ′)
]
+
∫
dq ′′dq ′′′[U (r,r′′,r′′′)−U (r′,r′′,r′′′)
][
ρ(q, q ′′′)ρ(q ′′′, q ′′)−ρ(q, q ′′)ρ(q ′′′, q ′′′)
]
ρ(q ′′, q ′)
+1
2
ρ(q, q ′)
∫
dq ′′dq ′′′[U (r,r′′,r′′′)−U (r′,r′′,r′′′)
][
ρ(q ′′, q ′′)ρ(q ′′′, q ′′′)−ρ(q ′′, q ′′′)ρ(q ′′′, q ′′)
]= 0. (4.14)
43005-6
The self-consistent field model for Fermi systems
In the absence of magnetic effects ρ(q, q ′) = ρ(r,r′)δσσ′ , and the self-consistent potential is diagonal
in spin indices as wellW (q, q ′) = [W (2)(r,r′)+W (3)(r,r′)]δσσ′ , where
W (2)(r,r′) = −U (r,r′)ρ(r,r′)+2δ(r− r′)
∫
dr′′U (r,r′′)ρ(r′′,r′′), (4.15)
W (3)(r,r′) =
∫
dr′′U (r,r′,r′′)ρ(r,r′′)ρ(r′′,r′)−2ρ(r,r′)
∫
dr′′U (r,r′,r′′)ρ(r′′,r′′)
+δ(r− r′)
∫
dr′′dr′′′U (r,r′′,r′′′)
[
2ρ(r′′,r′′)ρ(r′′′,r′′′)−ρ(r′′,r′′′)ρ(r′′′,r′′)
]
. (4.16)
In what follows we consider the system without account of magnetic effects.
5. Thermodynamic relations
Let us formulate the self-consistent field model at finite temperatures in a way such that all the ther-
modynamic relations would hold already in this approximation. This requirement leads to a unique for-
mulation of the model. Entering the definition of the statistical operator quantity Ω has the meaning of
the grand thermodynamic potential in the self-consistent field model. The entropy is defined through the
statistical operator (4.1) by the known expression
S =−Sp(ρ̂0 ln ρ̂0
)
. (5.1)
With the help of (5.1) it is easy to verify that the usual thermodynamic definition of the grand thermo-
dynamic potential holds Ω = E −T S −µN , where E is the total energy of the system. This potential is a
function of the temperature T and of the chemical potential µ, as well as a functional of the one-particle
density matrix ρ(q, q ′) = ρ(q, q ′;T,µ), which also depends on these quantities. However, by virtue of the
fact that the self-consistent potential was derived from the condition δΩ
/
δρ(q, q ′) = 0, when finding the
derivatives of Ω with respect to temperature and chemical potential, one should account for only the
explicit dependence of the thermodynamic potential on these quantities. As a consequence, at a fixed
volume of the system, the usual thermodynamic relation proves to be fulfilled in the self-consistent field
model:
dΩ=−SdT −N dµ. (5.2)
Calculation of the entropy and the number of particles either by means of the averaging with the density
matrix (4.1) or with the help of the thermodynamic relations S = −(∂Ω/∂T )µ, N = −(∂Ω/∂µ)T gives the
same result. With the correct choice of the energy E0, any kind of inconsistency in statistical description
of systems within the self-consistent field model, mentioned in [1], does not appear.
The energy (4.7), with account of the obtained self-consistent potential (4.9), (4.10), acquires the form
E0 = 1
2
∫
dqdq ′U (r,r′)
[
ρ(q, q ′)ρ(q ′, q)−ρ(q, q)ρ(q ′, q ′)
]
+
∫
dqdq ′dq ′′U (r,r′,r′′)
[
ρ(q, q)ρ(q ′, q ′′)ρ(q ′′, q ′)−1
3
ρ(q, q)ρ(q ′, q ′)ρ(q ′′, q ′′)
−2
3
ρ(q, q ′)ρ(q ′, q ′′)ρ(q ′′, q)
]
, (5.3)
and the thermodynamic potential is determined by the formula
Ω= E0 −T
∑
i
ln
(
1+e−βεi
)
. (5.4)
The total energy can be calculated either by averaging the Hamiltonian operator or by the use of the
thermodynamic potential (5.4)
E =Ω−µ∂Ω
∂µ
−T
∂Ω
∂T
(5.5)
43005-7
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
and has the form
E = ∑
i
εi fi +µN + 1
2
∫
dqdq ′U (r,r′)
[
ρ(q, q ′)ρ(q ′, q)−ρ(q, q)ρ(q ′, q ′)
]
+
∫
dqdq ′dq ′′U (r,r′,r′′)
[
ρ(q, q)ρ(q ′, q ′′)ρ(q ′′, q ′)−1
3
ρ(q, q)ρ(q ′, q ′)ρ(q ′′, q ′′)
−2
3
ρ(q, q ′)ρ(q ′, q ′′)ρ(q ′′, q)
]
. (5.6)
The entropy in this model is expressed through the distribution function formally in the same way as
in the model of an ideal gas
S =−∑
i
[
fi ln fi + (1− fi ) ln(1− fi )
]
,
(5.7)
but, as remarked previously, the distribution function itself is derived from the complicated nonlinear
equation (4.6) and includes the effects conditioned by both pair and three-body interactions.
The total number of particles (4.13) is expressed through the distribution function by the formula
N =∑
i
fi .
(5.8)
Since, owing to interaction, the energy of a single particle (4.11) differs from the energy of a free particle
and includes collective effects, then in this case a particle should be treated as a quasiparticle and the
function fi as the distribution function of quasiparticles. According to (4.13), (5.8), the number of initial
free particles and the number of quasiparticles coincide, as it takes place in the Landau theory of a normal
Fermi liquid [12]. In the presence of pair correlations, leading to the property of superfluidity, the number
of quasiparticles is less than the number of initial free particles [9, 10], because some fraction of particles
participates in the formation of the pair condensate.
Within the framework of the microscopic model of self-consistent field, formulae (5.3)–(5.8) describe
in general form the thermodynamics of the many-particle Fermi system at finite temperatures with ac-
count of pair and three-body interactions.
6. Interaction potentials in the self-consistent field model
In calculations within the self-consistent field approximation, instead of realistic potentials by which
individual particles interact in the vacuum and that usually model a strong repulsion at short distances,
the effective potentials with a set of adjustable parameters are used [23, 24]. A delta-like potential is
often chosen as the simplest model potential of the interparticle interaction. Such a choice of the pair
interaction potentialU (r′,r′′) = t2δ(r′− r′′) results in the following form of the self-consistent field
W (2)(r,r′) = t2
2
δ(r− r′)n(r), (6.1)
where n(r) = 2ρ(r,r) is the particle number density. In this case, the contribution into the self-consistent
field (4.15) of the first (exchange) term proves to be two times less by absolute value than the contribution
of the second term, describing the direct interaction, and has an opposite sign.
As can be immediately verified, when choosing three-body forces in the delta-like form
U (r′,r′′,r′′′) = t3δ(r′− r′′)δ(r′− r′′′), (6.2)
which is used, for example, in [23], the self-consistent potential W (3)(r,r′) (4.16) vanishes. Therefore, in
order to obtain a nonvanishing contribution of three-body forces into the self-consistent field, we have to
take into account their nonlocality, that is their finite radius of action. Physically this property is condi-
tioned by the Fermi statistics, not allowing for the presence of more than two particles with opposite spins
in the same space point. In the case of an arbitrary three-body potential of the interparticle interaction,
its contribution into the self-consistent potential (4.16) can be represented in the from
W (3)(r,r′) = a1(r,r′)−2ρ(r,r′)a2(r,r′)+δ(r− r′)
[
2b1(r)−b2(r)
]
, (6.3)
43005-8
The self-consistent field model for Fermi systems
where
a1(r,r′) ≡
∫
dr′′U (r,r′,r′′)ρ(r,r′′)ρ(r′′,r′), a2(r,r′) ≡
∫
dr′′U (r,r′,r′′)ρ(r′′,r′′),
b1(r) ≡
∫
dr′dr′′U (r,r′,r′′)ρ(r′,r′)ρ(r′′,r′′), b2(r) ≡
∫
dr′dr′′U (r,r′,r′′)ρ(r′,r′′)ρ(r′′,r′). (6.4)
Let us discuss the issue of choosing the potentials of interparticle interactions, which are suitable for
the self-consistent field model. In a spatially uniform state, the interaction potentials should satisfy two
conditions. They must not change under the replacement r → r+ a (a is an arbitrary vector) of all the
vectors they depend on, and must be symmetric relative to any permutation of their arguments. The pair
potential satisfying these conditions has the form
U (r,r′) = 1
2
[
U2(r− r′)+U2(r′− r)
]
, (6.5)
where U2(r) is the function of a vector argument. The three-body potential, depending on the pairs of
vector differences and satisfying the similar conditions, is as follows:
U (r,r′,r′′) = 1
6
[
U3(r− r′,r− r′′)+U3(r− r′′,r− r′)+U3(r′− r,r′− r′′)
+U3(r′− r′′,r′− r)+U3(r′′− r,r′′− r′)+U3(r′′− r′,r′′− r)
]
. (6.6)
Here, U3(r,r′) is a function of two vector arguments. If there is a dependence only on the distances be-
tween particles, and the function in (6.6) is symmetricU3(r,r′) =U3(r′,r), thenU (r,r′) =U2(|r− r′|) and
U (r,r′,r′′) = 1
3
[
U3(|r− r′|, |r− r′′|)+U3(|r′− r|, |r′− r′′)|+U3(|r′′− r|, |r′′− r′|)]. (6.7)
In particular, the three-body potential can be chosen in the form proposed in reference [21]:
U (|r− r′|, |r− r′′|) = u0 exp
(
− |r− r′|+ |r− r′′|
r0
)
. (6.8)
There is also a possibility of choosing the potential in the form of the Gauss function:
U (|r− r′|, |r− r′′|) = u0
π3/2r 3
0
exp
[
− (r− r′)2 + (r− r′′)2
r 2
0
]
. (6.9)
Such a choice is characteristic because in the limit r0 → 0, the potential (6.9) turns into the potential
of zero radius (6.2), and its contribution into the self-consistent field vanishes. In principle, the model
three-body potential can be chosen to depend on the three distances between three particles
U (r,r′,r′′) =U3(|r− r′|, |r− r′′|, |r′− r′′|). (6.10)
Note that the derivation from first principles of the potential of interaction of three atoms as structureless
entities presents a complex quantum mechanical problem [33].
The presently developed programs use different types of themodel interatomic three-body potentials.
These potentials are often of a complex form and in most cases include the angular dependence [34],
essential in the formation of the crystalline phase. In our paper, which can be used, for example, for the
description of the system of
3He atoms in the gaseous or liquid phases, we restrict ourselves to the use of
potentials of the simplest form, not depending on angles.
The potentials which rapidly tend to infinity at small distances are often used for describing the in-
teraction between particles. An example of such a potential is given by the Lennard-Jones potential
ULJ(r ) = 4ε
[(σ
r
)12
−
(σ
r
)6
]
, (6.11)
containing two parameters: the distance σ and the energy ε. The use of the potentials with the hard core
leads to considerable difficulties, especially in the quantum mechanical description [35]. Such potentials
43005-9
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
do not have the Fourier image, and the self-consistent field becomes infinity when using them. Some-
times, this fact is used as justification of non-applicability of the self-consistent field model in this or that
case, for example for describing the liquid [36]. Thus, the issue of the choice of the interparticle inter-
action potentials is of principal significance for problems of statistical physics. The use of the potential
which rapidly tends to infinity at small distances means that an atom or another compound particle re-
tains its identity at arbitrary high pressures. Meanwhile, it is clear that the critical pressure should exist
at which the atoms approach each other so closely that they should be “crushed” and loose their identity.
Therefore, the requirement of absolute impermeability of particles at arbitrary high pressures is exces-
sively rigorous and unphysical and, in our opinion, it is more reasonable to use the potentials which
tend to a finite value at small distances. An example of such a potential is given by the known Morse
potentialUM(r ) = ε{exp[−2(r −r0)/a]−2exp[−(r −r0)/a]
}
, where ε is the energy parameter and r0, a are
specific distances. Notice that, from the quantummechanical point of view, the use of bounded potentials
means the possibility for particles to tunnel through each other with some probability. The problem of
calculating the integral
∫
U (r )dr diverging for the potentials of type (6.11) is encountered, for example,
when deriving the known Gross-Pitaevskii equation, which is presently widely used in describing atomic
Bose-Einstein condensates [37]. Here, the diverging integral is replaced by a finite value coinciding with
the scattering length under the fulfilment of the Born approximation, which in essence means the use
of the potential not diverging rapidly to infinity at small distances. It should also be noted that quantum
chemical calculations lead to the potentials having a finite value of energy at zero [38, 39].
The model of “semi-transparent sphere” potential is quite often used as the simplest form of the pair
potential where the noted problems with divergences are absent:
U2(r ) =
{
U2m, r < r2,
0, r > r2.
(6.12)
This potential is determined by two parameters, one of whichU2m has the energy dimension and defines
the interaction strength, and the second one r2 of the length dimension defines the radius of interaction.
Such an observable quantity as the scattering length in the Born approximation (which, as should be
noted, is far from being always valid for realistic potentials) can be expressed through the parameters
of the potential (6.12): a0 = (m/4πħ2)
∫
U2(r )dr = mU2mr 3
2
/
3ħ2
. A similar model can be also used in the
case of three-body forces. For the potential (6.7) depending on pairs of distances between particles, we
have
U3(r,r ′) =
{
U3m, r < r3, r ′ < r3,
0, else.
(6.13)
Here, there are also two parameters: the strengthU3m and the radius r3. For this choice, the total potential
U (r,r′,r′′) =U3m, if the distances between each pair of three particles are less than r3. However, if two
distances between particles are less than r3, and the third one is greater than r3 and less than 2r3, then
U (r,r′,r′′) =U3m/3. In other cases, the potential vanishes.
For the potential (6.10), depending on three distances between particles, the model of “semi-trans-
parent sphere” potential is defined by the formula
U3(r,r ′,r ′′) =
{
U3m, r < r3, r ′ < r3, r ′′ < r3,
0, else.
(6.14)
In this case, the potential is nonzero only under the condition that the distances between each pair of
three particles are less than r3. Let us consider more in detail the spatially uniform system of Fermi
particles in the absence of an external field and with account of three-body forces.
7. The spatially uniform system
In the spatially uniform state, the one-particle density matrix is a function of absolute value of the
coordinate difference ρ(r,r′) = ρ(|r− r′)|, and the particle number density n = 2ρ(0) is constant. For the
potential (6.7) depending on pairs of distances between particles, the coefficients (6.4) which determine
43005-10
The self-consistent field model for Fermi systems
the contribution of three-body forces into the self-consistent potential acquire the form
a1(r− r′) = 2
3
∫
dr′′U3(|r− r′|,r ′′)ρ(r ′′)ρ(|r− r′+ r′′|)+ 1
3
∫
dr′′U3(|r− r′+ r′′|,r ′′)ρ(r ′′)ρ(|r− r′+ r′′|),
a2(r− r′) = 2ρ(0)
3
∫
dr′′U3(|r− r′|,r ′′)+ ρ(0)
3
∫
dr′′U3(|r− r′+ r′′|,r ′′),
b1 = ρ2(0)
∫
drdr′U3(r,r ′),
b2 = 1
3
∫
drdr′U3(|r− r′|,r ′)ρ2(r )+ 2
3
∫
drdr′U3(r,r ′)ρ2(r ′). (7.1)
Expansions of the three-body potential and the one-particle density matrix in Legendre polynomials can
be used:
U3(|r− r′|,r ′′) =
∞∑
l=0
U3l (r,r ′;r ′′)Pl (cosθ),
ρ(|r− r′|) =
∞∑
l=0
ρl (r,r ′)Pl (cosθ), (7.2)
where
U3l (r,r ′;r ′′) = 2l +1
2
1∫
−1
U3
(√
r 2 + r ′2 −2r r ′x,r ′′
)
Pl (x)dx,
ρl (r,r ′) = 2l +1
2
1∫
−1
ρ
(√
r 2 + r ′2 −2r r ′x
)
Pl (x)dx = 2l +1
2π2
∞∫
0
f (εk ) jl (kr ) jl (kr ′)k2dk, (7.3)
jl (x) is the spherical Bessel function. With these expansions taken into account we have:
a1(r ) = 8π
3
∞∫
0
U3(r,r ′)ρ(r ′)ρ0(r,r ′)r ′2dr ′+ 4π
3
∞∑
l=0
1
2l +1
∞∫
0
U3l (r,r ′;r ′)ρ(r ′)ρl (r,r ′)r ′2dr ′,
a2(r ) = 8πρ(0)
3
∞∫
0
U3(r,r ′)r ′2dr ′+ 4πρ(0)
3
∞∫
0
U30(r,r ′;r ′)r ′2dr ′,
b1 = 16π2ρ2(0)
∞∫
0
r 2dr
∞∫
0
U3(r,r ′)r ′2dr ′,
b2 = 32π2
3
∞∫
0
ρ2(r )r 2dr
∞∫
0
U3(r,r ′)r ′2dr ′+ 16π2
3
∞∫
0
ρ2(r )r 2dr
∞∫
0
U30(r,r ′;r ′)r ′2dr ′. (7.4)
In the spatially uniform case, the state of a single particle can be characterized by its momentum, and
equation (3.7) admits solutions in the form of plane waves:
φ j (q) = δσσ′p
V
eikr, (7.5)
where q ≡ (r,σ), j ≡ (k,σ′). As was noted, in the absence of magnetic effects ρ(q, q ′) = δσσ′ρ(r − r′),
W (q, q ′) = δσσ′W (r−r′). Equation (4.11) gives the quasiparticle energy not depending on the spin projec-
tion:
εk = ħ2k2
2m
−µ+Wk, (7.6)
where Wk = ∫
drW (r)eikr
. The self-consistent potential can be represented as a sum of direct and ex-
change terms
W (r) =W0δ(r)+WC(r), (7.7)
43005-11
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
and, if the interaction depends only on a distance between particles, then alsoWC(r) =WC(r ). In this case,
εk = ħ2k2
2m
−µ+W0 + 4π
k
∞∫
0
dr r WC(r )sin(kr ). (7.8)
Considering that both pair and three-body interactions give contribution into the self-consistent potential
W0 =W (2)
0 +W (3)
0 ,WC(r ) =W (2)
C
(r )+W (3)
C (r ), where
W (2)
0 = nU20, W (2)
C
(r ) =−U2(r )ρ(r ),
W (3)
0 = 2b1 −b2, W (3)
C
(r ) = a1(r )−2ρ(r )a2(r ). (7.9)
Here, U2(r ) is the potential of the pair interaction, U20 = ∫
U2(r )dr, n = 2ρ(0) is the particle number
density, and quantities a1(r ), a2(r ), b1, b2 for the potential (6.7) are defined by the formulae (7.4). The
one-particle density matrix has the form
ρ(r ) = 1
2π2r
∞∫
0
dk k sin(kr ) f (εk ), (7.10)
where f (εk ) = [
exp(βεk )+1
]−1
.
At low temperatures in a degenerate system, its properties are determined by quasiparticles located
near the Fermi surface. In this case, the notion of the effective mass of a quasiparticle m∗ can be intro-
duced. We define the Fermi wave number at finite temperatures by the relation
ħ2k2
F
2m
−µ+W0 + 4π
kF
∞∫
0
dr r WC(r )sin(kFr ) = 0. (7.11)
Then, near the Fermi surface k = kF+∆k , so that the dispersion law of a quasiparticle can be represented
in the form εk = (ħ2kF/m∗)∆k , where the effective mass is defined by the formula
1
m∗
= 1
m
− 4π
kFħ2
∞∫
0
dr r 3WC(r ) j1(kFr ), (7.12)
and j1(x) = (sin x−x cos x)/x2
is the first order spherical Bessel function. It is seen that the effective mass
is determined by the exchange part of the self-consistent potential. It should be stressed that within the
self-consistent field model, the formula (7.12) is exact, and it is fair for any densities at which the use
of the self-consistent field approximation is still permissible. Within the scope of the model itself, the
constraint on density cannot be derived. For estimation of the limiting value of density at which the self-
consistent field model remains true, one should calculate a correction due to the correlation Hamiltonian
(3.3) by the perturbation theory and requires it to be small relative to the main approximation [11]. The
contribution of three-body interactions into the effective mass of a quasiparticle with different choices
of the model potential will be studied in detail in a separate paper. In the approximation of the effective
mass, the distribution function acquires the form
fk = 1
exp
[
β
ħ2kF
m∗
(k −kF)
]
+1
≈ 1
exp
[
β
ħ2
2m∗
(
k2 −k2
F
)]+1
. (7.13)
For degenerate systems, both representations of the distribution function (7.13) are equivalent. The only
difference between the functions (7.13) and the distribution function of an ideal gas is the dependence of
the effective mass entering (7.13) on temperature and density.
The effective mass can be introduced for the non-degenerate system as well, when the main contri-
bution into the thermodynamics is given by particles with small momenta. Having used the expansion
sinkr ≈ kr − (kr )3/3!, from (7.8), we find the dispersion law of quasiparticles in the non-degenerate case:
43005-12
The self-consistent field model for Fermi systems
εk = ħ2k2/2m∗c−µ∗, where the effective mass m∗c and the chemical potential µ∗ are now determined
by the formulae:
1
m∗c
= 1
m
− 4π
3ħ2
∞∫
0
dr r 4WC(r ),
µ∗ = µ−W0 −4π
∞∫
0
dr r 2WC(r ). (7.14)
Note that the formula (7.12) for the effective mass turns into the formula (7.14), if j1(kFr ) ≈ kFr /3 is put
in (7.12).
The thermodynamic potential per unit volume, with account of three-body interactions depending on
pairs of distances between particles and in the approximation of the effective mass, according to (5.3),
(5.4) has the form
Ω
V
= −2T
Λ3Φ5/2
(
β
ħ2k2
F
2m∗
)
+4π
∞∫
0
U2(r )ρ2(r )r 2dr −2ρ2(0)
∞∫
0
U2(r )r 2dr
+16π2
8
3
ρ(0)
∞∫
0
r 2dr
∞∫
0
U3(r,r ′)ρ2(r ′)r ′2dr ′+ 4
3
ρ(0)
∞∫
0
ρ2(r )r 2dr
∞∫
0
U30(r,r ′;r ′)r ′2dr ′
− 8
3
ρ3(0)
∞∫
0
r 2dr
∞∫
0
U3(r,r ′)r ′2dr ′− 4
3
∞∫
0
ρ(r )r 2dr
∞∫
0
U3(r,r ′)ρ(r ′)ρ0(r,r ′)r ′2dr ′
. (7.15)
Here, Λ≡ (
2πħ2/m∗T
)1/2
is the thermal wavelength. The first term contains one of the integrals of the
Fermi–Dirac function tabulated in the McDougall and Stoner paper [40]
Φs (t ) = 1
Γ(s)
∞∫
0
zs−1 dz
ez−t +1
(7.16)
[Γ(s) is the gamma function], by which all the thermodynamical quantities of an ideal Fermi gas can be
expressed. The first term in (7.15) gives the contribution of a gas of non-interacting quasiparticles. Its
difference from the thermodynamical potential of an ideal Fermi gas consists only in the replacement of
mass by the effective mass. The rest terms in (7.15) are specific for the self-consistent field model and are
determined by pair and three-body interactions. Since the pressure is associated with the thermodynam-
ical potential by the known relation p =−Ω/
V , then the formula (7.15) determines, except for sign, also
the pressure. The formula for the particle number density
n = 2
Λ3 Φ3/2
(
β
ħ2k2
F
2m∗
)
, (7.17)
together with (7.15), determines the pressure as a function of density, that is the equation of state.
8. Fermi system at zero temperaturewith three-body interactions in the
model of “semi-transparent sphere”
The system of Fermi particles at zero temperature can be investigated in detail if the three-body inter-
action is chosen in the form (6.10) which leads to somewhat more simple formulae than for the potential
(6.7). For the potential (6.10), the parameters determining the self-consistent potential (6.4) are given by
43005-13
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
the formulae
a1(r ) = 4π
∞∑
l=0
1
2l +1
∞∫
0
U3l (r,r ′)ρl (r,r ′)ρ(r ′)r ′2dr ′, a2(r ) = 4πρ(0)
∞∫
0
U30(r,r ′)r ′2dr ′,
b1 = 16π2ρ2(0)
∞∫
0
r 2dr
∞∫
0
U30(r,r ′)r ′2dr ′, b2 = 16π2
∞∫
0
r 2dr
∞∫
0
U30(r,r ′)ρ2(r ′)r ′2dr ′. (8.1)
Here,
U3(r,r ′, |r− r′|) =
∞∑
l=0
U3l (r,r ′)Pl (cosθ),
U3l (r,r ′) = 2l +1
2
1∫
−1
U3
(
r,r ′,
√
r 2 + r ′2 −2r r ′x
)
Pl (x)dx. (8.2)
In particular for the interaction potential of “semi-transparent sphere” type (6.14), we have:
U3l (r,r ′) =U3mθ(r3 − r )θ(r3 − r ′)(2l +1)
[
θ(r3 − r − r ′)δl0 +θ(r + r ′− r3)
1
2
1∫
x0
Pl (x)dx
]
, (8.3)
where x0 = (r 2 + r ′2 − r 2
3 )
/
2r r ′
. At zero temperature fk = θ(kF −k) and
ρ(r ) = k2
F
2π2r
j1(kFr ), ρ(0) = n
2
= k3
F
6π2 , ρl (r,r ′) = 2l +1
2π2
kF∫
0
dk k2 jl (kr ) jl (kr ′). (8.4)
Taking into account the latter relations, we find
a1(r ) = U3mθ(r3 − r )
k3
F
2π3
1
kFr
j 2
0 (kFr3)− j0(kFr3) j0[kF(r3 − r )]+
kFr3∫
−kF(r3−r )
dy j0(kFr − y) j1(y)
,
a2(r ) = U3m
72π
(kFr3)3θ(r3 − r )
[(
r
r3
)3
−12
(
r
r3
)
+16
]
,
b1 = 5π2
6
ρ2(0)U3mr 6
3 = 5
216π2 U3m(kFr3)6 = 0.0023U3m(kFr3)6,
b2 = U3m
12π2
[
B3(kFr3)−12(kFr3)2B1(kFr3)+16(kFr3)3B0(kFr3)
]
, (8.5)
where Bn(z) ≡ ∫ z
0 yn j 2
1 (y)dy . When calculating a1(r ), the summation formula is used (see [41], p. 133)
∞∑
l=0
(2l +1) jl (u) jl (v)Pl (x) = j0
(√
u2 + v2 −2uv x
)
. (8.6)
In the limit kFr3 ¿ 1, we have more simple formulae
b2 ≈ 5
216π2 U3m(kFr3)6, a1(r ) ≈ U3m
432π3r 3
3
(kFr3)6θ(r3 − r )
[(
r
r3
)3
−12
(
r
r3
)
+16
]
. (8.7)
Note that even at kFr3 = 1, the formulae (8.7) give amistake of the order and less than 10%. For the follow-
ing, it is convenient to define a characteristic value of density through the radius of the pair interaction:
1
n∗
≡ 4π
3
r 3
2 . (8.8)
43005-14
The self-consistent field model for Fermi systems
If, for example, r2 = 3×10−8
cm is taken, then n∗ = 0.88×1022
cm
−3
. This density is close to the particle
number density in liquid. Note that, for example, the particle number density in liquid helium-3 at near-
zero temperature and vapor pressure: n3He ≈ 1.6×1022
cm
−3
. Since kFr3 = (9π/4)1/3(r3/r2)(n/n∗)1/3
, then
the condition kFr3 ¿ 1 can be written in the form
(r3/r2)(n/n∗)1/3 ¿ 1. (8.9)
Since it should be considered likely that r3 ∼ r2, the noted condition is satisfied in the limit of low den-
sities. However, as was noted, even at n ∼ n∗ the formulae (8.7) are valid with good accuracy. Let us
introduce the designation that will be used below for the dimensionless density
ñ ≡ n
n∗
=
(
r2
rS
)3
, (8.10)
where rS = [3/(4πn)]1/3
is the radius of a sphere whose volume equals the volume per particle. It is im-
portant to note that, even in the case when the dimensionless density (8.10) is considerably less than
unity, the value of the dimensional density proves to be much greater than that of low-density gases. In
the following formulae we will also use the next designations:
λ≡ r3
r2
, α≡ a0
r2
, u ≡ U3m
U2m
, (8.11)
a0 = mU2mr 3
2
/
3ħ2
. With account of (7.9), the relations (8.4), (8.5) determine the contribution of three-
body interactions into the direct and exchange parts of the self-consistent potential
W (3)
0 = 2b1 −b2, W (3)
C
(r ) = a1(r )−3n
j1(kFr )
kFr
a2(r ). (8.12)
The contribution of pair forces into these quantities for the potential of “semi-transparent sphere” type
(6.12) is given by the formulas
W (2)
0 =U2mñ, W (2)
C
(r ) =
−U2m
3
2
n
j1(kFr )
kFr
, r < r2,
0, r > r2.
(8.13)
The obtained relations (8.12), (8.13) for the self-consistent potential enable us to determine by the
formula (7.12) the contribution of pair and three-body interactions into the effective mass. With account
of only pair forces, the effective mass is determined by the expression
m
m∗2
= 1+ 6
π
α
(kFr2)2 B2(kFr2). (8.14)
Hence, it follows that for the repulsive pair interaction U2m > 0, the effective mass of a quasiparticle
proves to be less than the mass of a free particle, and for the attractionU2m < 0, it is greater. The effective
mass with account of three-body interactions in this case has the form
m
m∗
= m
m∗2
− 4πm
kFħ2
r3∫
0
a1(r ) j1(kFr )r 3dr + 12πmn
k2
Fħ2
r3∫
0
j 2
1 (kFr )a2(r )r 2dr. (8.15)
It should be stressed that, for all densities at which the self-consistent field approximation holds, the
formula (8.15) is exact. At low relative densities, when the condition (8.9) is satisfied, the formula for the
total effective mass (8.15) can be represented in the form of expansion in powers of density
m
m∗
= 1+ 3
10
α ñ − 3
70
(
9π
4
)2/3
α ñ5/3 + 159
2560
αuλ8 ñ2
= 1+0.3α ñ −0.158α ñ5/3 +0.0621αuλ8 ñ2. (8.16)
43005-15
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
Figure 1. (Color online) Dependencies of the effective mass on density: (1) α = −2.5, u = 0; (2) α = −2.5,
u =−0.3; (3) α= 2.5, u = 0; (4) α= 2.5, u = 1. It is everywhere assumed that λ= 1.
In (8.14), (8.16) and below we use the designations (8.11). The second and the third terms in (8.16) are
determined by pair forces, and the fourth by three-body forces. As it is seen, with increasing density the
contribution of three-body forces into the effective mass rises faster than that of pair forces, and the
relative role of three-body forces rises. Furthermore, attention is drawn to a strong dependence (as λ8
)
of the contribution of three-body forces on the ratio of the radii of three-body and pair interactions.
Dependencies of the effective mass on density, calculated by the exact formula (8.15), are shown in
figure 1. For the attractive interaction and with neglect of three-body forces, the effective mass proves
to be greater than the mass of a free particle and monotonously rises with an increasing density, reach-
ing its maximum. With a further increase of density, the rise of the effective mass changes into its fall,
although, as before for all physically reasonable densities, it remains greater than the mass of a free
particle (curve 1 in figure 1). Accounting for the three-body interaction with a positive constant leads to a
reduction of the region of the rise of the effective mass and to its faster decrease at high densities (curve 2
in figure 1). For the repulsive interaction and in the absence of three-body forces, the effective mass for
all reasonable densities is less than the mass of a free particle (curve 3 in figure 1). Accounting for the
three-body interaction with a positive constant leads to a faster monotonous decrease of the effective
mass (curve 4 in figure 1).
The total pressure p = −Ω/
V of the Fermi system with the three-body interaction of the form (6.10)
at finite temperatures, according to the general formulae (5.3), (5.4), is as follows
p = p0 +p2 +p3,
p0 = 2T
Λ3Φ5/2
(
β
ħ2k2
F
2m∗
)
, p2 = 4π
− ∞∫
0
U2(r )ρ2(r )r 2dr +2ρ2(0)
∞∫
0
U2(r )r 2dr
,
p3 =
∫
drdr′U3(r,r ′, |r− r′|)
[
−4ρ(0)ρ2(r ′)+ 8
3
ρ3(0)+ 4
3
ρ(r )ρ(r ′)ρ(|r− r′|)
]
, (8.17)
where the term p0 is a contribution of a gas of fermions with the effective massm∗, and the terms p2 and
p3 give, respectively, contributions of the pair and three-body interactions. Together with the formula
(7.17) for the particle number density, formulae (8.17) define in a general form the system’s equation of
state for the considered form (6.10) of three-body interactions.
At zero temperature, the pressure of the fermion gas
p0 = (3π2)2/3
5
ħ2
m∗
n5/3, (8.18)
and contributions of the pair and three-body forces into the total pressure (8.17) in the model of “semi-
transparent sphere” potentials for them [formulae (6.12), (6.14)], with the expansion formulae (8.1)–(8.4)
43005-16
The self-consistent field model for Fermi systems
taken into account, acquire the form
p2 = −U2mn∗
2
[ 6
π
ñ B0(kFr2)− ñ2
]
, (8.19)
p3 = 4ρ(0)
(2
3
b1 −b2
)
+ 16π
3
∞∫
0
ρ(r )a1(r )r 2dr, (8.20)
where b1, b2, a1(r ) are defined by the formulae (8.5). In the case of densities n É n∗, using the expansion
Bn(z) ≈ zn+3
9(n +3)
− zn+5
45(n +5)
+ zn+7
525(n +7)
we find the dependence of the pressure due to the pair interaction on density
p2
p0∗
= 5α
(12π2)1/3
[
ñ2 + 3
25
(
9π
4
)2/3
ñ8/3 − 81π
4900
(
9π
4
)1/3
ñ10/3
]
= 5α
(12π2)1/3
[
ñ2 +0.442 ñ8/3 −0.0997 ñ10/3
]
. (8.21)
Here the pressure is related to that of a gas of particles at the characteristic density (8.8): p0∗ =
1
5 (3π2)2/3(ħ2/m)n5/3∗ . The sign of pressure (8.21) is determined by the sign of the pair interaction con-
stant and, at negative value of this constant, the interaction contribution into the pressure is negative.
The main term in the expansion of pressure due to the three-body interaction (8.20) has the form
p3
p0∗
= 477
2560
αuλ8 ñ11/3 = 0.186αuλ8 ñ11/3. (8.22)
Attention should be paid to the fact that the expansion of p2 holds in even powers of the quantity ñ1/3
and the expansion of p3 in odd powers, and the latter begins with a high power.
Since the effective mass depends on density, then the pressure of a gas of quasiparticles (8.18), with
account of (8.16), can also be represented in the form of expansion in powers of density:
p0
p0∗
= ñ5/3 + 3
10
α ñ8/3 − 3
70
(
9π
4
)2/3
α ñ10/3 + 159
2560
αuλ8 ñ11/3
= ñ5/3 +0.3α ñ8/3 −0.158α ñ10/3 +0.0621αuλ8 ñ11/3. (8.23)
Taking into account (8.21)–(8.23), we find the expansion of the total pressure in powers of density:
p
p0∗
= ñ5/3 + 5
(12π2)1/3
α ñ2 + 3
4
α ñ8/3 − 207
1960
(
3π2
2
)1/3
α ñ10/3 + 159
640
αuλ8 ñ11/3
= ñ5/3 +1.018α ñ2 +0.75α ñ8/3 −0.259α ñ10/3 +0.248αuλ8 ñ11/3. (8.24)
Comparison of the accurate dependence of the total pressure on density, calculated by the formulae
(8.18)–(8.20) and the approximate dependence, calculated by the formula (8.24), is given in figure 2. It
is seen that even at ñ ∼ 1, the pressure, calculated by the approximate formula, differs weakly from its
accurate value.
Let us discuss the issue of the thermodynamic stability of the Fermi system, assuming ñ < 1. For a sys-
tem to be stable, it is necessary that its compressibility or (which is the same) the squared speed of sound
should be positive, so that ∂p/∂n > 0. Since in this case the two first terms give the main contribution into
the total pressure (8.24) and qualitatively taking into account the new effect of three-body interactions,
we find the condition of stability in the form
1+ 9
2
(
4
9π
)2/3
α ñ1/3 + 1749
3200
αuλ8 ñ2 > 0. (8.25)
43005-17
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
Figure 2. (Color online) Comparison of the accurate dependence of the pressure (1), calculated by the
formulae (8.18)–(8.20) and the approximate dependence (2), calculated by the formula (8.24), at α=−0.3,
u =−1, λ= 1. (3) Ideal Fermi gas.
For positive constants of pair and three-body interactions, the system is always stable. More interesting
is the case when the constant of the pair interaction is negative. Then, without account of three-body
interactions, the condition of stability should be satisfied which can be represented in equivalent forms
ñ1/3 <
(
π2
18
)1/3
|α|−1 ≈ 0.82 |α|−1, rS >
(
18
π2
)1/3
|a0| ≈ 1.22 |a0|. (8.26)
Accounting for the three-body interaction, if it is of a repulsion character, extends the region of stability
and can lead to stabilization of the system with the negative pair potential at arbitrary densities in case
of fulfilment of the following (8.25) condition
|u| > 200
583π4
55
33
|α|5
λ8 ≈ 0.408
|α|5
λ8 . (8.27)
Some dependencies of the pressure on density for different signs of the interparticle interaction are
shown in figure 3. In the case of attraction and with neglect of three-body interactions, the spatially
Figure 3. (Color online) Dependencies of the pressure on density: (1) α = −0.3, u = 0; (2) α = −0.3, u =
−0.5; (3) an ideal Fermi gas; (4) α= 0.3, u = 1. It is everywhere assumed that λ= 1.
43005-18
The self-consistent field model for Fermi systems
Figure 4. (Color online) The effect of three-body repulsive forces (U3m > 0) on stability of the Fermi
system with pair attractive forces: (1) u = 0; (2) u =−0.085; (3) u =−0.098; (4) u =−0.11; (5) u =−0.13. It
is everywhere assumed that α=−0.3 and λ= 1.
uniform state is stable only at low densities for which the condition (8.26) is satisfied. At high densities,
the pressure decreases with an increase of density (curve 1 in figure 3) and the spatially uniform state
ceases to be stable. Sufficiently intensive three-body repulsive forces lead to stabilization of the system
with pair attractive forces (curve 2 in figure 3).
The effect of three-body repulsive forces on the stability of the Fermi system is illustrated more in
detail in figure 4. With an increasing strength of the three-body repulsive interaction, the regions of
stability of such a system extend (curve 2, 3 in figure 4), and a further growth of the intensity of the
three-body interaction leads to stabilization of the system for all physically reasonable values of density.
The form of the quasiparticle energy spectrumwith account of the interparticle interactions is shown
in figure 5. Accounting for only the pair attraction leads to a slowing down of the energy increase with
increasing momentum relative to an ideal Fermi gas (curve 2 in figure 5). Accounting for the repulsive
three-body forces, against the background of the attractive pair forces, additionally yields a small increase
Figure 5. (Color online) The quasiparticle energy spectrum ε̃(κ) (in the variables ε̃ ≡ εk /εF, κ ≡ k/kF):
(1) Ideal Fermi gas [ε̃(κ) = κ2 −1, m̃∗ ≡ m∗/m = 1]; (2) α = −2.5, u = 0, ñ = 1 (m̃∗ = 1.78); (3) α = −2.5,
u =−1, λ= 1, ñ = 1 (m̃∗ = 1.50); (4) α= 2.5, u =−1, λ= 1, ñ = 1 (m̃∗ = 0.75).
43005-19
Yu.M. Poluektov, A.A. Soroka, S.N. Shulga
of the quasiparticle energy (curve 3 in figure 5). The effect of three-body forces on the spectrum is appre-
ciably weaker than the role of pair forces, so that accounting for the attractive three-body forces in the
presence of the pair repulsion weakly prevents the increase of the quasiparticle energy as momentum
increases (curve 4 in figure 5).
9. Conclusion
The self-consistent field equations are obtained, and, within this model, thermodynamic relations are
derived for a normal system of Fermi particles with account of both pair and three-body forces. The
satisfaction of the essential requirement of fulfilment of all thermodynamic relations already in the self-
consistent approximation leads to a unique formulation of the self-consistent field model. Emphasized
are the novelty and universality of the proposed approach to the formulation of this model and the use-
fulness of the method for describing not only Fermi, but also Bose systems, in particular phonons, and
relativistic quantized fields. The approach being developed allows us also, as shown in the paper, to ac-
count for many-body interactions naturally.
It is shown that three-body interactions of zero radius give no contribution into the self-consistent
field and, in order to account for the effects due to such interactions, it is necessary to account for their
nonlocality. The case of the spatially uniform system is considered in detail. General formulae are derived
for the system’s equation of state and the effective mass of quasiparticles with account of three-body
forces. Dependencies of the quasiparticle effective mass and the system’s pressure on density at zero
temperature are obtained, with pair and three-body forces accounted for in the model of interaction
potentials of “semi-transparent sphere” type.
It is shown that pair interactions of repulsive character reduce the quasiparticle effective mass rela-
tive to the mass of a free particle while attractive pair interactions, on the contrary, raise it. The effective
mass and pressure are numerically calculated at zero temperature and expansions of these quantities are
derived in powers of the relative density with account of three-body forces. It is shown that the relative
contribution of three-body interactions into thermodynamic quantities rises with an increasing density.
The effect of three-body forces on the stability of the Fermi system is considered, and it is shown that in
the case of repulsion, their being taken into account extends the region of stability and can lead to stabi-
lization of the system with pair attraction. The quasiparticle energy spectrum is calculated with account
of the interparticle interactions.
References
1. Kirzhnits D.A., Field Theoretical Methods in Many-Body systems, Pergamon, Oxford, 1967.
2. Eisenberg J., Greiner W., Microscopic Theory of the Nucleus, North-Holland, London, 1972.
3. Barts B.I., Bolotin Yu.L., Inopin E.V., Gonchar V.Yu., The Hartree-Fock Method in Nuclear Theory, Naukova
Dumka, Kiev, 1982 (in Russian).
4. Hartree D.R., The Calculation of Atomic Structures, Wiley, New York, 1957.
5. Slater J.C., The Self-Consistent Field for Molecules and Solids, McGraw-Hill, New York, 1974.
6. Braut R., Phase Transitions, Benjamin, New York, 1965.
7. Vaks V.G., Larkin A.I., Pikin S.A., Sov. Phys. JETP, 1967, 24, 240.
8. Poluektov Yu.M., Low Temp. Phys., 1997, 23, 685; doi:10.1063/1.593364.
9. Poluektov Yu.M., Bulletin of Kharkiv National University, Phys. Ser. “Nuclei, particles, fields”, 2001, 522, 3.
10. Poluektov Yu.M., Ukr. J. Phys., 2005, 50, 1303; Preprint arXiv:1303.4913, 2013.
11. Poluektov Yu.M., Ukr. J. Phys., 2015, 60, 553.
12. Landau L.D., Sov. Phys. JETP, 1957, 3, 920.
13. Pines D., Nozières P., The Theory of Quantum Liquids, Benjamin, New York, 1966.
14. Migdal A.B., Theory of Finite Fermi Systems and Applications to Atomic Nuclei, Nauka,Moscow, 1983 (in Russian).
15. Kolomiets V.M., Nuclear Fermi-Liquid, Naukova Dumka, Kiev, 2009 (in Russian).
16. Akhiezer A.I., Krasil’nnikov V.V., Peletminskii S.V., Yatsenko A.A., Phys. Rep., 1994, 245, 1;
doi:10.1016/0370-1573(94)90060-4.
17. Poluektov Yu.M., Fiz. Nizk. Temp., 1995, 21, 300 (in Russian).
18. Poluektov Yu.M., Fiz. Nizk. Temp., 1996, 22, 402 (in Russian).
43005-20
http://dx.doi.org/10.1063/1.593364
http://arxiv.org/abs/1303.4913
http://dx.doi.org/10.1016/0370-1573(94)90060-4
The self-consistent field model for Fermi systems
19. Axilrod B.M., Teller E., J. Chem. Phys., 1943, 11, 299; doi:10.1063/1.1723844.
20. Basu A.N., Sengupta S., Phys. Status Solidi B., 1968, 29, 367; doi:10.1002/pssb.19680290137.
21. Sarkar A.K., Sengupta S., Solid State Commun., 1969, 7, 135; doi:10.1016/0038-1098(69)90710-8.
22. Bethe H.A., Ann. Rev. Nucl. Sci., 1971, 21, 93; doi:10.1146/annurev.ns.21.120171.000521.
23. Skyrme T.H.R., Nucl. Phys., 1959, 9, 615; doi:10.1016/0029-5582(58)90345-6.
24. Vautherin D., Brink D.M., Phys. Rev. C, 1972, 5, 626; doi:10.1103/PhysRevC.5.626.
25. Hammer H.-W., Nogga A., Schwenk A., Rev. Mod. Phys., 2013, 85, 197; doi:10.1103/RevModPhys.85.197.
26. Landau L.D., Lifshitz E.M., Quantum Mechanics, Pergamon Press, 1977.
27. Poluektov Yu.M., Low Temp. Phys., 2002, 28, 429; doi:10.1063/1.1491184.
28. Poluektov Yu.M., Ukr. J. Phys., 2007, 52, 579; Preprint arXiv:1306.2103, 2013.
29. Poluektov Yu.M., Fiz. Nizk. Temp., 2015, 41, 1181 (in Russian) [Low Temp. Phys., 2015, 41, 922;
doi:10.1063/1.4936228].
30. Poluektov Yu.M., Bulletin of Kharkiv National University, Phys. Ser. “Nuclei, particles, fields”, 2009, 859, 9;
Preprint arXiv:1507.00246v2, 2015.
31. Poluektov Yu.M., Russ. Phys. J., 2010, 53, 163; doi:10.1007/s11182-010-9401-6.
32. Bogoliubov N.N., Bogoliubov N.N. (jr), Introduction to Quantum Statistical Mechanics, World Scientific PC, Singa-
pore, 2009.
33. Sarry A.M., Sarry M.F., Tech. Phys., 2014, 59, 474; doi:10.1134/S1063784214040215.
34. Gale J.D., J. Chem. Soc., Faraday Trans., 1997, 93, 629; doi:10.1039/A606455H.
35. Thouless D.J., The Quantum Mechanics of Many-Body Systems, Academic Press, New York, 1972.
36. Bazarov I.P., Statistical Theory of the Crystalline State, Moscow University Press, Moscow, 1972 (in Russian).
37. Pitaevskii L.P., Stringari S., Bose-Einstein Condensation, Oxford University Press, Oxford, 2003.
38. Aziz R.A., Slaman M.J., J. Chem. Phys., 1991, 94, 8047; doi:10.1063/1.460139.
39. Anderson J.B., Traynor C.A., Boghosian B.M., J. Chem. Phys., 1993, 99, 345; doi:10.1063/1.465812.
40. McDougall J., Stoner E.C., Phil. Trans. Roy. Soc. A, 1938, 237, 67; doi:10.1098/rsta.1938.0004.
41. Varshalovich D.A., Moskalev A.N., Khersonskii V.K., Quantum Theory of Angular Momentum, Nauka, Leningrad,
1975 (in Russian).
Модель самоузгодженого поля для фермi-систем
з врахуванням тричастинкових взаємодiй
Ю.М. Полуектов, О.О. Сорока, С.М.Шульга
Iнститут теоретичної фiзики iм. О.I. Ахiєзера, ННЦ ХФТI, вул. Академiчна, 1, 61108 Харкiв, Україна
На основi мiкроскопiчної моделi самоузгодженого поля побудовано термодинамiку системи багатьох
фермi-частинок при скiнчених температурах з врахуванням тричастинкових взаємодiй i отримано рiвня-
ння руху квазiчастинок. Показано, що дельтаподiбна тричастинкова взаємодiя не дає внеску в самоузго-
джене поле, i для опису тричастинкових сил слiд враховувати їх нелокальнiсть. Детально розглянуто про-
сторовооднорiдну систему i в рамках розвиненого мiкроскопiчного пiдходу отримано загальнi формули
для ефективної маси фермiону i рiвняння стану системи з врахуванням внеску тричастинкових взаємодiй.
Для потенцiалу типу “напiвпрозорої сфери” при нулi температур чисельно розраховано ефективну масу i
тиск. Знайдено розвинення ефективної маси i тиску за ступенямищiльностi.Показано,що при врахуваннi
тiльки парних сил, взаємодiя, що має характер вiдштовхування, зменшує ефективну масу квазiчастинки
порiвняно з масою вiльної частинки, а у разi притягання— збiльшує. Розглянуто питання термодинамi-
чної стiйкостi фермi-системи i показано, що тричастинкова взаємодiя, яка має характер вiдштовхування,
розширює область стiйкостi системи з мiжчастинковим парним притяганням. Розраховано енергетичний
спектр квазiчастинки з врахуванням тричастинкових сил.
Ключовi слова: самоузгоджене поле, тричастинковi взаємодiї, ефективна маса, фермiон,
рiвняння стану
43005-21
http://dx.doi.org/10.1063/1.1723844
http://dx.doi.org/10.1002/pssb.19680290137
http://dx.doi.org/10.1016/0038-1098(69)90710-8
http://dx.doi.org/10.1146/annurev.ns.21.120171.000521
http://dx.doi.org/10.1016/0029-5582(58)90345-6
http://dx.doi.org/10.1103/PhysRevC.5.626
http://dx.doi.org/10.1103/RevModPhys.85.197
http://dx.doi.org/10.1063/1.1491184
http://arxiv.org/abs/1306.2103
http://dx.doi.org/10.1063/1.4936228
http://arxiv.org/abs/1507.00246v2
http://dx.doi.org/10.1007/s11182-010-9401-6
http://dx.doi.org/10.1134/S1063784214040215
http://dx.doi.org/10.1039/A606455H
http://dx.doi.org/10.1063/1.460139
http://dx.doi.org/10.1063/1.465812
http://dx.doi.org/10.1098/rsta.1938.0004
Introduction
Hamiltonian of the Fermi system with account of three-body interactions
The self-consistent field model with account of three-body interactions
Derivation of the self-consistent potential
Thermodynamic relations
Interaction potentials in the self-consistent field model
The spatially uniform system
Fermi system at zero temperature with three-body interactions in the model of ``semi-transparent sphere''
Conclusion
|