Quantum kinetic theory of metal clusters in an intense electromagnetic field
A quantum kinetic theory for weakly inhomogeneous charged particle systems is derived within the framework of nonequilibrium Green’s functions. The results are of relevance for valence electrons of metal clusters as well as for confined Coulomb systems, such as electrons in quantum dots or ultracold...
Збережено в:
Дата: | 2004 |
---|---|
Автори: | , |
Формат: | Стаття |
Мова: | English |
Опубліковано: |
Інститут фізики конденсованих систем НАН України
2004
|
Назва видання: | Condensed Matter Physics |
Онлайн доступ: | http://dspace.nbuv.gov.ua/handle/123456789/119037 |
Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Цитувати: | Quantum kinetic theory of metal clusters in an intense electromagnetic field / M. Bonitz, J.W. Dufty // Condensed Matter Physics. — 2004. — Т. 7, № 3(39). — С. 483–525. — Бібліогр.: 35 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-119037 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1190372017-06-04T03:04:28Z Quantum kinetic theory of metal clusters in an intense electromagnetic field Bonitz, M. Dufty, J.W. A quantum kinetic theory for weakly inhomogeneous charged particle systems is derived within the framework of nonequilibrium Green’s functions. The results are of relevance for valence electrons of metal clusters as well as for confined Coulomb systems, such as electrons in quantum dots or ultracold ions in traps and similar systems. To be specific, here we concentrate on the application to metal clusters, but the results are straightforwardly generalized. Therefore, we first give an introduction to the physics of correlated valence electrons of metal clusters in strong electromagnetic fields. After a brief overview on the jellium model and the standard density functional approach to the ground state properties, we focus on the extension of the theory to nonequilibrium. To this end a general gauge-invariant kinetic theory is developed. The results include the equations of motion of the two-time correlation functions, the equation for the Wigner function and an analysis of the spectral function. Here, the concept of an effective quantum potential is introduced which retains the convenient local form of the propagators. This allows us to derive explicit results for the spectral function of electrons in a combined strong electromagnetic field and a weakly inhomogeneous confinement potential. На основі нерівноважних функцій Гріна розроблено квантову теорію слабонеоднорідної системи заряджених частинок. Отримані результати мають суттєве значення для опису валентних електронів металічних кластерів, а також для замкнутих кулонівських систем, таких як електрони в квантових точках чи ультрахолодні іони в пастках. Зокрема, в даній роботі ми зупиняємося на розгляді металічних кластерів, хоча результати можна безпосередньо узагальнити на інші випадки. Ми розпочинаємо з введення в фізику скорельованих валентних електронів у металічних кластерах в сильних електромагнітних полях. Після короткого огляду моделі “желе” та стандартного методу функціоналу густини для опису основних станів ми зосереджуємося на узагальненні теорії на випадок нерівноважних процесів. З цією метою розроблено узагальнену калібрувально-інваріантну кінетичну теорію. Ці результати включають рівняння руху для двочасових кореляційних функцій, рівняння для функцій Вігнера та аналіз спектральної функції. Вводиться поняття ефективного квантового потенціалу, який залишає незмінною локальну форму пропагаторів. Це дає нам змогу отримати явні вирази для спектральної функції електронів при суперпозиції сильного електромагнітного поля та слабонеоднорідного утримуючого потенціалу. 2004 Article Quantum kinetic theory of metal clusters in an intense electromagnetic field / M. Bonitz, J.W. Dufty // Condensed Matter Physics. — 2004. — Т. 7, № 3(39). — С. 483–525. — Бібліогр.: 35 назв. — англ. 1607-324X DOI:10.5488/CMP.7.3.483 PACS: 05.30.-d,36.40.Wa,52.38.-r http://dspace.nbuv.gov.ua/handle/123456789/119037 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 |
A quantum kinetic theory for weakly inhomogeneous charged particle systems is derived within the framework of nonequilibrium Green’s functions. The results are of relevance for valence electrons of metal clusters as well as for confined Coulomb systems, such as electrons in quantum dots or ultracold ions in traps and similar systems. To be specific, here we concentrate on the application to metal clusters, but the results are straightforwardly generalized. Therefore, we first give an introduction to the physics of correlated valence electrons of metal clusters in strong electromagnetic fields. After a brief overview on the jellium model and the standard density functional approach to the ground state properties, we focus on the extension of the theory to nonequilibrium. To this end a general gauge-invariant kinetic theory is developed. The results include the equations of motion of the two-time correlation functions, the equation for the Wigner function and an analysis of the spectral function. Here, the concept of an effective quantum potential is introduced which retains the convenient local form of the propagators. This allows us to derive explicit results for the spectral function of electrons in a combined strong electromagnetic field and a weakly inhomogeneous confinement potential. |
format |
Article |
author |
Bonitz, M. Dufty, J.W. |
spellingShingle |
Bonitz, M. Dufty, J.W. Quantum kinetic theory of metal clusters in an intense electromagnetic field Condensed Matter Physics |
author_facet |
Bonitz, M. Dufty, J.W. |
author_sort |
Bonitz, M. |
title |
Quantum kinetic theory of metal clusters in an intense electromagnetic field |
title_short |
Quantum kinetic theory of metal clusters in an intense electromagnetic field |
title_full |
Quantum kinetic theory of metal clusters in an intense electromagnetic field |
title_fullStr |
Quantum kinetic theory of metal clusters in an intense electromagnetic field |
title_full_unstemmed |
Quantum kinetic theory of metal clusters in an intense electromagnetic field |
title_sort |
quantum kinetic theory of metal clusters in an intense electromagnetic field |
publisher |
Інститут фізики конденсованих систем НАН України |
publishDate |
2004 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/119037 |
citation_txt |
Quantum kinetic theory of metal clusters in an intense electromagnetic field / M. Bonitz, J.W. Dufty // Condensed Matter Physics. — 2004. — Т. 7, № 3(39). — С. 483–525. — Бібліогр.: 35 назв. — англ. |
series |
Condensed Matter Physics |
work_keys_str_mv |
AT bonitzm quantumkinetictheoryofmetalclustersinanintenseelectromagneticfield AT duftyjw quantumkinetictheoryofmetalclustersinanintenseelectromagneticfield |
first_indexed |
2025-07-08T15:07:47Z |
last_indexed |
2025-07-08T15:07:47Z |
_version_ |
1837091804132933632 |
fulltext |
Condensed Matter Physics, 2004, Vol. 7, No. 3(39), pp. 483–525
Quantum kinetic theory of metal
clusters in an intense electromagnetic
field I ∗
M.Bonitz 1,2 , J.W.Dufty 2
1 Institut für Theoretische Physik und Astrophysik,
Christian-Albrechts-Universität zu Kiel,
Leibnizstr. 15, 24098 Kiel, Germany
2 Department of Physics, University of Florida,
Gainesville, FL 32611–8440
Received March 30, 2004
A quantum kinetic theory for weakly inhomogeneous charged particle sys-
tems is derived within the framework of nonequilibrium Green’s functions.
The results are of relevance for valence electrons of metal clusters as well
as for confined Coulomb systems, such as electrons in quantum dots or
ultracold ions in traps and similar systems. To be specific, here we con-
centrate on the application to metal clusters, but the results are straightfor-
wardly generalized. Therefore, we first give an introduction to the physics
of correlated valence electrons of metal clusters in strong electromagnetic
fields. After a brief overview on the jellium model and the standard density
functional approach to the ground state properties, we focus on the exten-
sion of the theory to nonequilibrium. To this end a general gauge-invariant
kinetic theory is developed. The results include the equations of motion of
the two-time correlation functions, the equation for the Wigner function and
an analysis of the spectral function. Here, the concept of an effective quan-
tum potential is introduced which retains the convenient local form of the
propagators. This allows us to derive explicit results for the spectral func-
tion of electrons in a combined strong electromagnetic field and a weakly
inhomogeneous confinement potential.
Key words: quantum statistical mechanics, charged clusters,
laser-plasma interactions
PACS: 05.30.-d,36.40.Wa,52.38.-r
∗This paper is devoted to the memory of Yuri L’vovich Klimontovich
c© M.Bonitz, J.W.Dufty 483
M.Bonitz, J.W.Dufty
1. Introduction
Correlated charged particles are of growing interest in many fields of physics.
Here the activities date back to the work on electrolytes and plasmas in the first
half of the 20th century, for an overview see [1] and references therein. A systematic
kinetic theory for correlated many-body systems was derived from the basic equa-
tions of quantum mechanics (von Neumann equation) by Bogolyubov, Born, Green,
Kirkwood, Yvon and others, resulting in the famous BBGKY-hierarchy of equations
for the nonequilibrium reduced density operators, see [2] for an overview.
A major breakthrough in the theory of nonideal gases and plasmas was achieved
by the development of field theoretical methods (second quantization). One branch,
represented by Feynman, Schwinger, Martin, Keldysh, Baym, Kadanoff and many
others, led to a systematic quantum field theory of plasmas – formulated in terms
of Green’s functions, e.g. [1,3]. An independent line, pioneered by Klimontovich,
focussed on a physically very similar, but formally very different method – Klimon-
tovich’s microscopic phase space density, e.g. [4,5]. This approach turned out to be
very fruitful for classical nonideal gases and plasmas, whereas for quantum many-
body systems, the Green’s functions approach appears to be more efficient. The
latter, therefore, will be the method of choice in the present paper.
Strong field effects in quantum many-body systems have been successfully in-
corporated into the Green’s functions method by many authors, for an overview
and references, cf. [6]. Among the problems of this theory (and similar methods as
well) is that the results of many-body approximations (such as gradient expansions)
are known to be dependent on the chosen gauge. Therefore, an explicitly gauge-
invariant theory [7] provides a convenient starting point from which approximations
can be systematically derived. This method turned out to be fruitful in many fields
including semiconductor quantum transport, e.g. [6] and dense plasmas [8,9].
The latter results, however, were limited to spatially homogeneous systems. It
is, therefore, of great interest to extend them to the inhomogeneous case. This will
allow us to extend the theory to a variety of new systems, including the valence
electrons of metal clusters, charged particles in traps, electrons in quantum dots [10]
and so on. In this paper, we extend previous work [8,9] to these weakly inhomoge-
neous systems, concentrating on the additional modifications arising in the quantum
kinetic equations from an external confinement field which is the source of the inho-
mogeneity. For definiteness, we wil consider the potential VI confining the valence
electrons in metal clusters, but the situation in quantum dots or particle traps is
analogous.
The basic physics of these systems is best illustrated on the jellium approximation
(i.e. homogeneous ion charge density within a sphere), as it directly generalizes the
very successful one-component plasma model to confined systems. This is done in
the reminder of this introductory section and in section 2. As a result of mean-field
effects (Hartree-Fock selfenergy effects) this potential is renormalized and becomes
density (and possibly time) dependent. The ground state configuration of the valence
electrons in this effective potential V eff
I is usually obtained within density functional
484
Quantum kinetic theory of metal clusters
theory. To go beyond the ground state properties and to analyze the nonequilibrium
behavior of the electrons in a strong electromagnetic field requires development of
a quantum kinetic theory which is outlined in section 3. There, a gauge-invariant
derivation is presented which selfconsistently includes the external field and the
weakly inhomogeneous confinement potential. The kinetic equations for the two-
time correlation functions and for the Wigner distribution are derived.
Special attention is given to an analysis of the spectral properties of the elec-
trons, cf. section 3.9. Introducing the concept of an effective quantum potential
which replaces the potential V eff
I , the spectral function of the valence electrons in
a combined electromagnetic field and thus the confinement potential is derived. Its
main characteristic feature is that, even without correlations, the spectral function
is broadened and shifted (towards higher energy) as a results of the external field
and of inhomogeneity effects. This has direct consequences for various scattering
processes and the collision integrals. Further applications of the results to solutions
of the quantum kinetic equation in linear response, to dielectric properties and to
explicit solutions in a strong field go beyond this paper and will be presented in a
forthcoming publication [11].
1.1. Basic equations and simplifications
We consider a finite system of N nuclei with atomic mass A and charge Z.
Thus, the total number of elementary charged particles – nuclei and electrons – is
N(Z+1), each interacting by the bare Coulomb potential. The particles are subject
to an electromagnetic field given by the vector and scalar potentials A(r, t) and
φ(r, t). The hamiltonian of the full system consists of a field part and a coupled
field-matter part,
Ĥ = ĤF + ĤFM,
ĤF =
∑
k,λ
~ωkb
†
kλbkλ,
ĤFM = Ĥ0 + Ĥint,
Ĥ0 =
N
∑
α=1
{
(
Pα − e0
c
A(rα, t)
)2
2M
+ φ(rα, t) +
Z
∑
i=1
(
pαi + e0
c
A(rα, t)
)2
2me
+ φ(rαi, t)
}
,
Ĥint =
N
∑
α=1
N
∑
β=1
{
1
2
(Ze0)
2
|Rα − Rβ|
+
Z
∑
i=1
[
− Ze20
|Rα − rβi|
+
Z
∑
j=1
e20
|rαi − rβj|
]}
,
where p and P denote the electron and ion momentum, respectively, r and R the
corresponding positions and me and M the corresponding masses. e0 is the free
electron charge times minus one, and in Hint selfinteraction terms are not included
(first and third terms). bkλ and b†kλ are annihilation and creation operators of photons
of mode k (normally k denotes the momentum) with polarization λ which obey
bosonic commutation rules, [bkλ, b
†
k′λ′ ] = δkk′δλλ′ , all other commutators vanish. The
485
M.Bonitz, J.W.Dufty
transverse vector potential of the electromagnetic field is expressed in terms of these
operators by [12]
Â(r) =
∑
k,λ
(
2π~c2
ωkV
)1/2
(
bkλε̂kλe
ikr + b†kλε̂
∗
kλe
−ikr
)
, (1)
where the expansion is in a complete set of plane waves1, for convenience, and V
denotes the volume. The transversality condition puts on the constraint k · ε̂kλ = 0.
For sake of generality, we also include a longitudinal electric field which is given
by the scalar potential φ. Some issues related to various presentations of the field
(gauge problem) will be discussed below.
1.1.1. Limit of a classical electromagnetic field
In this paper we are interested specifically in intense electromagnetic field pro-
duced by lasers or free electron lasers. These electromagnetic fields are characterized
by high coherence and a large number of photons in each mode which justifies a clas-
sical treatment. For example, the number of photons in a laser field with intensity
I = 1 W/cm2 and photon energy of ~ω = 1 eV in a typical coherence volume of
V = 1 cm−3 can be estimated by
Nkλ =
I
(~ω)
V
c
≈ 2 · 108.
This large number assures that fluctuations of the eigenvalues of the photon number
operator nkλ = b†kλbkλ around its mean value which is given by a classical treatment
〈nkλ〉 ≈ Nkλ will be small. The resulting classical result for the vector potential is
Â(r, t) → A(r, t) =
∑
kλ
c
2ωk
{
Ekλe
i(kr−ωkt−φkλ) + E∗
kλe
−i(kr−ωkt−φkλ)
}
, (3)
where the electric field amplitude is defined by the average (classical) number of
photons in the mode, Ekλ =
√
(8π~ωkλNkλ) /V · ε̂kλ, and the φkλ are operators de-
scribing the relative phases of different modes. For a single-mode laser which we will
consider below, φkλ leads just to a time shift and does not enter any physical result
and may thus be dropped. For a strict justification of the classical result, see [12].
1.1.2. Separation of valence electrons
A first step to simplify the above system is to subdivide the electrons into valence
(weakly bound or quasi-free) and core (deeply bound) ones. This is reasonable for
1This is an operator in the Schrödinger picture. It can be made time-dependent by transforming
to the Heisenberg representation by
e
i
~
ĤFtÂ(r)e−
i
~
ĤFt =
∑
k,λ
(
2π~c2
ωkV
)1/2
(
bkλε̂kλei(kr−ωkt) + b†kλε̂∗kλe−i(kr−ωkt)
)
. (2)
486
Quantum kinetic theory of metal clusters
metals and metal clusters and has been found to well reproduce dominant qualita-
tive features, e.g. [13,14]. On the other hand, it has to be kept in mind that this
subdivision is never strict and, in situations of strong excitation, it has to be justi-
fied in each case. Moreover, in nonequilibrium situations inner core electrons may
be ionized or become valence electrons.
Using this idea, the system is reconsidered as the one composed of N ions, each
of which is w−fold charged2, and w · N valence electrons. This means that the
field-matter hamiltonian ĤFM is now replaced by the sum of two terms – an ion
hamiltonian and an effective electron hamiltonian, ĤFM = ĤI + Ĥe, given by
ĤI =
N
∑
α=1
{
(
Pα − e0
c
A(rα, t)
)2
2M
+ φ(rα, t) +
N
∑
β<α
(we0)
2
|Rα − Rβ|
}
, (4)
Ĥe =
wN
∑
i=1
{
(
pi + e0
c
A(ri, t)
)2
2me
+ φ(ri, t) + VI(ri, t) +
wN
∑
j<i
e20
|ri − rj|
}
, (5)
VI(ri, t) =
N
∑
α=1
Vps(|Rα(t) − ri|). (6)
The major simplification is that the electron-ion interactions are replaced by an
effective single-particle external potential in which the electrons move.
1.1.3. Adiabatic approximation
Due to the large ion mass, often the ionic motion can be neglected (it is no
principal problem to avoid this approximation and perform a selfconstistent numer-
ical treatment where the ion coordinates are updated on comparatively large time
intervals). Since we are interested in fast processes related to the electron dynam-
ics prior to the Coulomb explosion we will use this assumption. At the same time
this allows us to neglect the external field effect on the ions, so we will neglect the
potentials A and φ in the ionic hamiltonian ĤI and, moreover, consider the ions as
“frozen”. Their only effect on the electrons is then to provide a neutralizing back-
ground (which is inhomogeneous though, see section 1.2 below) and the effective
potential VI . There still remains a time dependence in the potential VI(t) which is
due to ionization processes changing the charge of the ions.
1.2. Spherical jellium model
In the following we will use atomic units, i.e. lengths will be in units of the
hydrogen Bohr radius, aB = ~
2/(e20me) and energies in units of Hartree, 1Ha =
2Ry = e20/aB. Then, the interparticle distance a is measured by the Wigner-Seitz
radius rs = a/aB.
2this is analogous to switching from a physical to a chemical picture in partially ionized plasmas,
e.g. [1]
487
M.Bonitz, J.W.Dufty
The simplest approximation to treat the influence of the ions on the electrons is
the spherical jellium model3. It is essentially a generalization of the one-component
plasma model to the case of finite systems: the ion charge density ρI is assumed to
be homogeneous, inside a sphere of radius RI and zero outside,
ρI(r) = ρI0Θ(RI − r), RI = rsN
1/3. (7)
Since the number density is assumed to be constant, given by the bulk value rs, the
radius of the sphere is determined by the particle number N , and the volume of the
ion sphere is VI = 4πr3
sN/3. Thus the charge density of the sphere, ρI0, is given by
the charge of all ions inside the sphere,
QI(RI) = wNe0 : ρI0 =
QI(RI)
VI
=
3
4π
we0
r3
s
.
We first compute the electric field of the charged sphere. Outside the sphere, it
is just the field of a point charge equal to the total charge, while inside it is the field
of a point charge confined in the sphere of radius r,
EI(r) =
wNe0
r2
, r > RI ,
QI(r)
r2
=
we0
r3
s
r, r 6 RI ,
(8)
since QI(r) = ρ0I4πr
3/3 = we0r
3/r3
s . From the electric field we immediately obtain
the electrostatic potential, given by
φI(r) = −
∫ r
dr̄EI(r̄) + C.
Outside the ion sphere, the potential is wNe0/r and, determining the constant from
the continuity of the potential at r = RI , we obtain
C = we0
[
N
RI
+
R2
I
2r3
s
]
and
φI(r) =
wNe0
r
, r > RI ,
we0
2r3
s
(
3R2
I − r2
)
, r 6 RI .
(9)
Thus, the jellium potential (energy) VI which one electron will feel will be given by
VI(r) = −e0φI(r),
VI(r) =
−wNe
2
0
r
= −R
3
I
r3
s
we20
r
, r > RI ,
−wNe
2
0
2RI
[
3 −
(
r
RI
)2
]
= −we
2
0
2r3
s
(
3R2
I − r2
)
, r 6 RI .
(10)
3This model is analogous to the ion sphere model in strongly correlated plasmas which is due
to Salpeter [16] and the Wigner-Seitz sphere concept in solid state physics, e.g.[15,17].
488
Quantum kinetic theory of metal clusters
Let us briefly discuss the basic properties of the jellium model.
i) By construction, VI is continuous for all r including the sphere boundary r = RI .
Also, its first derivative, i.e. the force acting on the electrons toward the cluster
center, FI = −∇VI ,
FI(r) =
−wNe
2
0
r3
r, r > RI ,
−wNe
2
0
R3
I
r, r 6 RI ,
(11)
is continuous at r = RI .
ii) The derivative of FI is not continuous. At the point r = RI , F
′
I(r) jumps from
(
2wNe20
)
/R3
I to −
(
wNe20
)
/R3
I . This discontinuity is obvious since φI obeys
Poisson’s equation with the source ρI . In terms of VI , the Poisson equation
reads V ′′
I (r) = −4π(−e0)ρI(r) with a discontinuity of 4πe0ρI0 =
(
3wNe20
)
/R3
I
at r = RI , in agreement with the above result.
iii) The electrostatic energy contained in the ion sphere is readily computed,
WI =
1
2
∫
d3rρI(r)φI(r) =
4π
2
3
4π
we0
r3
s
∫ RI
0
drr2we0
2r3
s
(3R2
I − r2)
=
3
5
(wNe0)
2
RI
. (12)
iv) Electron binding: the Nve valence electrons are confined in the potential VI . The
deepest value of the potential is at the center of the ion sphere,
VI(0) = −3
2
we20N
2/3
rs
.
We note that this binding energy would refer to a single electron. For many
electrons, the Pauli principle leads to occupation of higher energy states be-
tween the bottom of the potential, VI(0) and the Fermi energy, see section 2.
The spatial distribution of the valence electrons in the potential well is gov-
erned by the shape of VI and by quantum and interaction effects among the
electrons – i.e. by mean-field, exchange and correlation effects. A simple ap-
proach is provided by Thomas–Fermi theory, see section 2.
As noted above, the potential VI may change in time. In particular, each act
of ionization increases the ion charge number wN → wN + 1, thus deepening
the potential and increasing the binding.
The potential (10) enters the electron hamiltonian Ĥe (6), replacing the sum over
the pseudopotentials Vps.
489
M.Bonitz, J.W.Dufty
2. Valence electron ground state configuration
To obtain the ground state of the ensemble of valence electrons one can use the
concepts of density functional theory. There one minimizes the total energy with
respect to the valence electron number density profile n(r),
δEve[n(r)]
δn(r)
= 0, with
∫
d3r n(r) = Nve = const. (13)
If the cluster is neutral, the total number of electrons in the cloud equals the number
of valence electrons Nve = w ·N , but for generality we will retain the number Nve.
The total energy can be written as
Eve[n(r)] = Tve[n(r)] + UH
ve[n(r)] + Vve[n(r)] + EXC
ve [n(r)], (14)
consisting of kinetic energy (first term), Hartree mean field (second), external con-
finement (third) and a remainder (fourth) containing all exchange and correlation
contributions. The quality of the result depends on the approximation for the various
energy contributions.
2.1. Thomas–Fermi model
An illustrative nontrivial approximation for the total energy of the valence elec-
trons is to include the electron kinetic energy, the external confinement VI in the
jellium approximation and the electron mean field (Hartree potential)4,
Eve[n(r)] = Tve[n(r)] + UH
ve[n(r)] + Vve[n(r)]. (15)
The kinetic energy may be approximated by the Thomas-Fermi result, i.e. one as-
sumes that the valence electrons are governed by a zero-temperature Fermi dis-
tribution which parametrically depends on a space-dependent density n(r), (local
approximation, LDA). With the kinetic energy density of the ideal electron gas,
ε =
3
10
~
2(3π2)2/3
me
n5/3,
we obtain
Tve [n(r)] =
3
10
~
2(3π2)2/3
me
∫
d3r n5/3(r). (16)
Next, the Hartree mean-field energy is given by
UH
ve =
e20
2
∫
d3rd3r′
n(r)n(r′)
|r− r′| =
1
2
∫
d3r n(r)ΣH
ve[r;n(r)], (17)
ΣH
ve[r;n(r)] = e20
∫
d3r′
n(r′)
|r− r′| , (18)
4thus neglecting in equation (14) the contributions EXC from electron exchange (Fock energy)
and correlations.
490
Quantum kinetic theory of metal clusters
where we introduced the Hartree selfenergy of the valence electrons, and we note
that it is the solution of Poisson’s equation
∇2ΣH
ve[r;n(r)] = −4πe20n(r). (19)
Finally, the confinement energy of the valence electrons in the jellium potential is
given by the integral of the ion electrostatic potential multiplied with the electron
charge density,
Vve[n(r)] = −e0
∫
d3r φI(r)n(r) =
∫
d3r VI(r)n(r). (20)
Collecting the results for the three energy contributions, we can perform the
variation of the total energy (15) with respect to the valence electron density profile
δE =
∫
d3r
{
~
2(3π2)2/3
2me
n2/3(r) + e20
∫
d3r′
n(r′)
|r− r′| + VI(r) − µ
}
δn(r), (21)
where we introduced the chemical potential as Lagrange multiplier. Requiring the
expression in curly brackets to vanish, we obtain
n(r) =
1
3π2~3
{
2me
[
|VI(r)| − ΣH
ve [r;n(r)] + µ
]}3/2
. (22)
This is not an explicit equation for the density because the Hartree term depends on
n too. We can recast it into a closed differential equation for the Hartree selfenergy
by inserting (22) into the r.h.s. of the Poisson equation (19),
∇2ΣH
ve[r;n(r)] = − 4e20
3π~3
{
2me
[
|VI(r)| − ΣH
ve[r;n(r)] + µ
]}3/2
. (23)
We will concentrate below on the jellium potential VI . Due to its isotropy, the
valence electron density will be isotropic as well. Then, from equation (22), there
are two obvious constraints on the density n(r): since VI and ΣH
ve are monotonical-
ly decaying with increasing distance from the center5, n(r) will be monotonically
decaying too, reaching zero at a finite r0 and remain zero for r > r0. The critical
radius r0 follows from the normalization,
4π
∫ r0
0
drr2n(r) = Nve (24)
and defines the chemical potential:
µ = −|VI(r0)| + ΣH
ve[r0;n(r)]. (25)
5the latter statement has to be verified on the final solution for the density
491
M.Bonitz, J.W.Dufty
The angle integration in the selfenergy (18) can be carried out6
ΣH
ve[r;n(r)] = 4πe20
∫ r0
0
dr′r′2n(r′)g(r, r′),
with g(r, r′) =
1
r
, r > r′,
1
r′
, r 6 r′.
(27)
In particular, for r > r0 : ΣH
ve(r) =
Nvee
2
0
r
. Using this result at r = r0 and expression
(10) for the jellium potential7 we obtain from (25)
µ = −(wN −Nve)
e20
r0
. (28)
We see that for a neutral cluster, the chemical potential vanishes, in agreement with
atomic Thomas-Fermi theory, e.g. [18]. For positively charged clusters where some
electrons have been removed from the cloud, Nve < wN , µ is negative.
Equations (22), (24), (28) provide a closed system to compute the valence elec-
tron density profile, the cloud radius r0, the chemical potential and the Fermi energy,
EF = −µ, of the valence electrons surrounding the jellium droplet.
2.2. Free Electron gas
The simplest limiting case is the ideal electron gas where the Hartree mean
field is neglected as well. This case is instructive as all results can be obtained
analytically. Moreover, this limit serves as a generalization of the well-known model
of the (macroscopic) ideal electron gas to finite systems, and it becomes exact in the
high-density limit, rs → 0. Then we obtain from equations (22), (28)
µid = −wNe
2
0
r0
, (29)
nid(r) =
(2me)
3/2
3π2~3
(
wNe20
2RI
)3/2
[
3 −
(
r
RI
)2
− 2
RI
r0
]3/2
, r 6 RI = rsN
1/3;
(2me)
3/2
3π2~3
(
wNe20
r
)3/2 [
1 − r
r0
]3/2
, RI 6 r 6 r0;
0, r > r0.
(30)
6due to isotropy,
∫
d3r′
n(r′)
|r − r′| = 2π
∫ ∞
0
dr′r′2
∫ 1
−1
dz
n(r′)√
r2 + r′2 − 2rr′z
(26)
=
2π
r
∫ ∞
0
dr′r′
{
√
r2 + r′2 + 2rr′ −
√
r2 + r′2 − 2rr′
}
,
leading to the result (2.1).
7The point r0 is expected to be outside the ion sphere, so we use the branch of VI which
corresponds to r > RI .
492
Quantum kinetic theory of metal clusters
Finally, the boundary of the electron cloud, r0, follows from equation (24) where
the integration can be performed yielding a transcendental equation for x0 = r0/RI
which, however, is rather complicated8, so a numerical solution of equation (24) is
advantageous.
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0
0.5
1
1.5
2
2.5
3
x 10
28 Na
20
2+
radius in Angstrom
de
ns
ity
in
m
−
3
electronic density
ionic density
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
−8
−7
−6
−5
−4
−3
−2
−1
Na
20
2+
radius in Angstrom
po
te
nt
ia
l e
ne
rg
y
in
e
V
chem. potential
phi
tot
Figure 1. Jellium model result for a two-fold charged sodium cluster of 20 atoms
in the Hartree approximation. Upper figure: density of ions and valence elec-
trons vanishing at RI and r0, respectively. Lower figure: total potential in-
cluding VI and the Hartree mean field. Also shown is the chemical potential.
Its intersection with the total potential defines the electron cloud boundary r0.
Figure courtesy of Th. Fennel.
8for completeness, we provide the result of the integration using equation (30) for the density
profile. Introducing the definitions n ≡ n
nid
, x0 ≡ r0
RI
, α ≡ Nve
wN
, n0 ≡ (2me)
3/2
3π2~3
(
wNe2
0
2RI
)3/2
, x0
is the solution of the following equation:
Nve
4πR3
In0
= − (2α − x0)
3
16x3
0
arccot
[
√
2 − 2α
x0
]
−
√
2 − 2α
x0
48x2
0
(6α − 7x0)(2α + x0)
+
( x0
2α
)3
{
arctan
[√
α
1 − α
]
− arccot
[
√
x0
α
− 1
]}
− 1
24α2
{
(2α − 3)(4α − 1)x3
0
√
1 − α
α
− (2α − 3x0)(4α − x0)x
3
0
√
x0
α
− 1
}
. (31)
493
M.Bonitz, J.W.Dufty
2.3. Electron ground state including the mean field
Using the result for the ideal electron gas, we can now improve the valence
electron density by including the Hartree mean field. A simple perturbation approach
consists in computing ΣH, equation (2.1), using nid and inserting the result into (22)
to obtain an improved approximation for the density profile. This procedure can
be repeated (at each step, the cut-off radius r0 has to be computed again) until
convergence is reached.
Numerical results for the electron density in the Hartree approximation are shown
in figure 1. One sees that the valence electron density (upper figure) extends beyond
the limits of the core ion distribution up to the maximum radius r0 which is defined
by the intersection of the chemical potential with the effective potential, cf. equa-
tion (25). The effective potential deviates from the quadratic (inside the cluster)
jellium potential, equation (10) – the mean field effects cause a flat potential shape
in the center of the cluster. At the boundary of the ion charge the potential curvature
changes turning to an 1/r decay which is only weakly influenced by many-particle
effects.
2.4. Electron ground state including exchange in local appr oximation
The mean field result shown in figure 1 is readily improved by including exchange
effects. The exchange energy in local approximation was derived by Dirac [19]
EX[n(r)] = −3
4
(
3
π
)1/3
e20
∫
d3r [n(r)]4/3, (32)
and its variation yields the exchange potential
V X[n(r)] = −e
2
0
π
(
3π2
)1/3
∫
d3r [n(r)]1/3, (33)
which has to be added to the density equation (22). For our isotropic jellium problem
this equation then becomes
n(r) =
1
3π2~3
{
2me
[
|VI(r)| − ΣH
ve[r;n(r)] + V X[n(r)] − (wN −Nve)
e20
r0
]}3/2
, (34)
where we used the solution (28) for the chemical potential.
Summarizing this introductory part, we have given a brief overview on the ground
state properties of valence electrons in metal clusters. An illustrative picture is given
by the Thomas-Fermi model which is generalized by the concepts of density func-
tional theory by including mean field and exchange (plus correlations, in some ap-
proximation). The DFT approach is very successful in describing the ground state
of clusters, and it allows to systematically improve the model. However, it is not
able to correctly describe clusters at elevated temperatures and in nonequilibrium,
in particular, under the combined influence of strong fields and nontrivial electronic
correlations.
494
Quantum kinetic theory of metal clusters
We now proceed to a rigorous and consistent theoretical description of nonequi-
librium properties of metal clusters in a strong electromagnetic field, based on quan-
tum kinetic theory.
3. Quantum kinetic description of cluster valence electron s
in a strong electromagnetic field
In the previous section we have established the model for our treatment of the
valence electrons in metal clusters: it is based on the hamiltonian Ĥe, equation (6).
The ions are treated as frozen (adiabatic approximation), they have a homogeneous
density inside the droplet of radius RI , their influence on the electrons is condensed
in the jellium potential VI . In the absence of an external field, the electrons will
form a cloud around the ionic core with a density profile n(r), decreasing away
from the center, cf. figure 1, (except for possible Friedel oscillations). This spatial
density profile has been computed for the zero temperature limit. At the same
time, in momentum space, the electrons will be described by a Fermi distribution
(step function) where all momentum states are occupied up to the Fermi energy of
the electron cloud, given by minus the chemical potential (28), which is essentially
defined by the overall charge of the cluster.
Of course, at finite temperature and, more generally, under nonequilibrium con-
ditions created by a strong field, the momentum distribution of the valence electrons
may be very far from the zero-temperature result. To outline a theory which allows to
systematically calculate this nonequilibrium distribution and the related transport
quantities of the valence electrons is the main subject of this Section.
3.1. Outline of the presented theory
While a quantum kinetic theory can be formulated for arbitrary spatial proper-
ties, in a weakly inhomogeneous system the formulation strongly simplifies. More-
over, it is reasonable to expect that, under a number of relevant conditions, the
spatial shape of the electron cloud will remain close the ground state case, at least
for an intermediate period of time prior to the Coulomb explosion. In particular,
such a situation is expected to exist, even in nonequilibrium, in the case of
a. homogeneous laser excitation, i.e. if the laser focus and the wave length are much
larger than the cluster - which is usually fulfilled, at least in the optical an UV
range
b. collective effects influencing the electron cloud as a whole, which is the case for
plasma oscillations with sufficiently small wave number, k � 2π/r0, where r0
denotes the extension of the valence electron cloud discussed above,
c. correlation and scattering effects being sufficiently weak.
If these conditions are fulfilled we may expect that, on the scale of the cluster, spatial
variations and spatial density gradients will be small (compared to the unperturbed
495
M.Bonitz, J.W.Dufty
density profile), and the main effect of the excitation will be a deformation of the
momentum distribution of the electrons. Under these conditions the quantum kinetic
theory is effectively formulated in the Wigner representation. Assuming throughout
this paper weak spatial inhomogeneity we expect that the local approximation plus
first order gradient corrections will give the dominant contributions to the theory.
Finally, since gradient corrections are gauge dependent, we will develop a theory
which is gauge-invariant from the beginning.
3.2. Valence Electron Density Matrix
The central quantity to describe the statistical properties of the electrons in
nonequilibrium is the single-particle density matrix F1(r1, r
′
1, t). It can be derived
from the von Neumann equation which includes the electron hamiltonian (6)
i~
∂
∂t
ρe − [He, ρe] = 0, (35)
where ρe is the density matrix of all Nve valence electrons. Computing the partial
trace of ρe over particles 2, . . .Nve yields the single-particle density operator F1 which
obeys the first equation of the BBGKY hierarchy, e.g. [2]. In coordinate represen-
tation it becomes the one-particle density matrix F1(r1, r
′
1, t). In the following, it
will be convenient to transform to center of mass (R) and relative (r) coordinates,
and to proceed analogously with time and momentum arguments, according to the
relations
R ≡ (r1 + r′1)/2, r ≡ r1 − r′1 , (36)
Q ≡ (k1 + k′
1)/2, k ≡ k1 − k′
1 , (37)
t ≡ (t1 + t′1)/2, τ ≡ t1 − t′1 . (38)
In the new coordinates, the density matrix becomes F1(R + r/2,R − r/2, t) ≡
F̃1(R, r, t). Fourier transform with respect to r then yields the Wigner distribution
function f(R,p, t), which is analogous to the classical distribution function (how-
ever, in contrast to the latter it has the known problems of possible negative and
complex values resulting from the Heisenberg uncertainty principle).
Yet, there are a number of nontrivial questions concerning the equation of mo-
tion for this function in the special situation of interest which are related to the
appearance of the time-dependent field, the spatial inhomogeneity of the cluster etc.
For this reason, we turn now to a careful derivation of this equation. In particular,
we will use a gauge-invariant approach to exclude any ambiguity of the results. The
most straightforward approach to do this is to use nonequilibrium Green’s functions
as was demonstrated for a homogeneous system in [8,9] which we will generalize to
inhomogeneous systems in section 3.6.
3.3. Local approximation in nonequilibrium and first order g radient correc-
tions
As a first approximation to the rather complicated problem of the valence elec-
tron dynamics we introduce a local approximation, in analogy to the concepts used
496
Quantum kinetic theory of metal clusters
in density functional theory9. In situations where no sharp spatial changes of the
system properties occur (geometrical boundaries etc.) and the number of electrons
involved is not small, it is reasonable to assume that the density matrices depend
on space (center of mass coordinate) only via the density,
F̃1(R, r, t) ≡ F̃1[n(R, t), r, t], (39)
and, therefore, ∇RF̃1(R, r, t) =
δF̃1
δn(R, t)
∇R n(R, t), (40)
and, similarly for the Wigner function,
f(R,k, t) ≡ f [n(R, t),k, t], (41)
and, therefore, ∇Rf(R,k, t) =
δf
δn(R, t)
∇R n(R, t). (42)
Analogous behavior is expected for other dynamical quantities, including two-particle
and higher order distributions and the two-time correlation functions, see below.
The situation is more complex for more general quantities which are, for example,
a product of several distribution functions. Then, applying the local approximation
to each factor does not necessarily result in an adequate approximation for the
product. Consider an arbitrary physical quantity C which is the product of A and
B. In the Wigner representation, we can write (time arguments appear here only as
parameters and are omitted),
C(R,k) =
∫
d3Q1
(2π~)3
∫
d3Q2
(2π~)3
∫
d3r1
∫
d3r2 e
i
~
(k−Q2)r1e−
i
~
(k−Q1)r2
×A
(
R +
r1
2
,Q1
)
B
(
R +
r2
2
,Q2
)
, (43)
which is still general for an arbitrary inhomogeneous case, cf. e.g. [30]. Now, expand-
ing A and B with respect to small deviations of the coordinates with respect to R,
up to first order, yields
A
(
R +
r1
2
)
B
(
R +
r2
2
)
≈ A (R)B (R)
+
1
2
{
[r1 · ∇R1 + r2 · ∇R2 ]A (R1)B (R2)
}∣
∣
∣
R1=R2=R
. (44)
Using in equation (43) the identities r1 = −~/i∇Q2 and r2 = ~/i∇Q1 with the
expansion (44) allows to perform the integrations over r1 and r2 and, with the
help of the resulting delta functions, also the integrations over Q1 and Q1, yielding
Q1 = Q2 = k, with the final result
C(R,k) =
{
[
1 +
i~
2
(∇R1∇Q2 −∇R2∇Q1)
]
A (R1,Q1)B (R2,Q2)
}∣
∣
∣
(R,k)
≡
(
1 +
i~
2
{...}
)
AB, (45)
9and similar (and even more general) concepts in hydrodynamics
497
M.Bonitz, J.W.Dufty
where we introduced the short notation (R,k) to indicate that, after differentiation,
one has to set R1 = R2 = R and Q1 = Q2 = k. The first term in equation (45) gives
the local approximation for C, while the term in curly brackets (Poisson brackets)
contains the first gradient corrections.
With this general result which will be used below, e.g. for the collision integrals,
it is straightforward to test, in each case of interest, the validity of the local approx-
imation and to estimate the magnitude of corrections. On the other hand, it turns
out that several physical contributions to the Wigner function, such as the exchange
self energy, are missing in the local approximation and appear only in first order
gradient corrections, see section 3.7.
3.4. Kadanoff-Baym/Keldysh equations for the valence elec trons
The field theoretical description10 of the cluster valence electrons is based on
the fermionic creation and annihilation operators ψ† and ψ which are defined to
guarantee the spin statistics theorem,
ψ(1)ψ(2) + ψ(2)ψ(1) = ψ†(1)ψ†(2) + ψ†(2)ψ†(1) = 0,
ψ(1)ψ†(2) + ψ†(2)ψ(1) = δ(1 − 2),
where t1 = t2 has been assumed, and 1 ≡ (r1, t1, s
3
1), (below, we will drop the
spin index). The nonequilibrium state of the electrons is described by the two–time
correlation functions which are statistical averages (with the initial density operator
of the system) of field operator products
g>(1, 1′) =
1
i~
〈ψ(1)ψ†(1′)〉 , g<(1, 1′) = − 1
i~
〈ψ†(1′)ψ(1)〉, (46)
where g> and g< are, in nonequilibrium, independent from one another. They contain
the complete dynamical and statistical information. The latter follows from their
elements along the time diagonal: the one-particle density matrix is immediately
obtained from the function g< according to
F1(r1, r
′
1, t) = −i~g<(1, 1′)|t1=t′1
, (47)
whereas the dynamical information (e.g. the single–particle spectrum and the cor-
relations) follows from the function values across the diagonal in the t1 − t′1–plane,
in particular, from the spectral function a(1, 1′),
a(1, 1′) ≡ i~ {g>(1, 1′) − g<(1, 1′)} = i~
{
gR(1, 1′) − gA(1, 1′)
}
, (48)
where gR/A are the retarded and advanced Green’s functions, defined below in equa-
tion (52). In particular, in cases where the microscopic variables vary on much
smaller scales than the macroscopic ones, cf. definition (36, 38), it is advantageous
10In this section we follow [9].
498
Quantum kinetic theory of metal clusters
to perform a Fourier transformation with respect to τ and/or r which leads to the fre-
quency and momentum variables ω and p, respectively. In particular, equation (47)
then yields the familiar Wigner distribution function
f(p,R, t) = −i~g<(p,R; t1, t
′
1)|t1=t′1=t . (49)
The time evolution of the correlation functions in an electromagnetic field is de-
termined by the Kadanoff–Baym/Keldysh equations which follow from the Heisen-
berg equations of motion for the field operators in the presence of the hamiltonian
Ĥe (in the following, the electrostatic potential φ will be neglected which will be
justified in section 3.5)
[
i~
∂
∂t1
− 1
2me
(
~
i
∇1 +
e0
c
A(1)
)2
− VI(1)
]
g≷(1, 1′) =
=
∫
dr̄1 ΣHF(1, r̄1t1)g
≷(r̄1t1, 1
′) +
∫ t1
t0
d1̄ [Σ>(1, 1̄) − Σ<(1, 1̄)] g≷(1̄, 1′)
−
∫ t′1
t0
d1̄Σ≷(1, 1̄) [g>(1̄, 1′) − g<(1̄, 1′)] , (50)
which have to be fulfilled together with the adjoint equations. The l.h.s. of this
equation contains all single-particle terms, whereas the r.h.s. contains all corrections
due to mean field, exchange and correlations. Further, t0 denotes the initial time
where the system is assumed to be uncorrelated (otherwise, the equations have to
be supplemented with an initial correlation contribution to Σ, cf. [20,21]). ΣHF is
the Hartree–Fock selfenergy (Hartree mean–field plus exchange energy),
ΣHF(11′) = −i~δ(t1 − t1′)
{
∫
dr2V (r1 − r2)g
<(22+) − V (r1 − r′1)g
<(11′)
}
, (51)
and Σ≷ are the correlation selfenergies which describe collision processes, ionization
and so on.
The properties of the cluster valence electrons are mainly determined by the
choice of the confinement potential and the collision integrals (selfenergies). In the
derivations below, the potential VI will be kept completely general and it will not
necessarily be identified with the smooth jellium potential of section 2. Further we
may account for varying valence electron number due to ionization of core electrons
(or, vice versa, recombination) as well as due to “evaporation” of electrons. These
processes can be readily incorporated into the model by choosing the appropriate
selfenergy contributions and by properly updating the confinement potential so that
it is consistent with the actual number of core and valence electrons.
For the following derivations, it is useful to introduce, in addition to the two-time
correlation functions, the retarded and advanced Green’s functions
gR/A(1, 1′) = ±Θ[±(t1 − t′1)] {g>(1, 1′) − g<(1, 1′)} , (52)
499
M.Bonitz, J.W.Dufty
which are related by
[gR(1, 1′)]∗ = gA(1′, 1), (53)
and obey the simpler equations
[
i~
∂
∂t1
− 1
2me
(
~
i
∇1 +
e0
c
A(1)
)2
− VI(1)
]
gR/A(1, 1′)
−
∫
d2 ΣR/A(1, 2)gR/A(2, 1′) = δ(1 − 1′), (54)
where the functions ΣR/A are related to Σ≷ in analogy to equation (52). gR/A describe
the single-particle spectrum under the influence of the external field, the jellium
potential and exchange and correlation effects and are directly related to the spectral
function, see definition (48).
3.5. Gauge-invariant Fourier transform
It is well known that the electromagnetic field can be introduced in various ways
(gauges) which may lead to essentially different explicit forms of the resulting ki-
netic equations. Although alternative derivations are successfully applied too, gauge
invariance becomes a particular problem if the resulting kinetic equations are treat-
ed by means of approximations, such as retardation or gradient expansions as in
the present case. A critical issue is that the result of these approximations maybe
essentially different in different gauges, see e.g. [6] for illustrative examples. To avoid
these difficulties, we will formulate the theory in terms of correlation functions which
are made explicitly gauge–invariant.
In this section, we use a co-variant 4-vector notation as it makes the following
transformations more compact and symmetric. The corresponding definitions are
Aµ = (cφ,A), xµ = (cτ, r), Xµ = (ct,R),
and the conventions aµ = (a0, a), aµ = (a0,−a) and aµb
µ = a0b0−ab are being used.
One readily proofs that the Kadanoff–Baym/Keldysh equations (50) remain co-
variant under gauge transformations, i.e. under the following transformations of the
potentials and field operators
A′
µ(x) = Aµ(x) − ∂µχ(x), ψ′
a(x) = e−
i
~
e0
c
χ(x)ψa(x). (55)
From this one immediately obtains the gauge transform of the Green’s functions
g′(x,X) = e−
i
~
e0
c [χ(X+ x
2 )−χ(X−x
2 )]g(x,X) .
Following an idea of Fujita [7], we now introduce a gauge–invariant Green’s function
g(k, X) which is given by the modified Fourier transform
g̃(k, X) =
∫
d4x
(2π)4
exp
{
i
∫ 1
2
− 1
2
dλ xµ
[
kµ − e0
c
Aµ(X + λx)
]
}
g(x,X), (56)
500
Quantum kinetic theory of metal clusters
where use has been made of the identity
χ
(
X +
x
2
)
− χ
(
X − x
2
)
=
∫ 1
2
− 1
2
dλ
d
dλ
χ (X + λx) = xµ∂
µ
∫ 1
2
− 1
2
dλχ (X + λx) .
Indeed, one readily confirms that under any gauge transform (55), the phase factors
cancel, and g′(k, X) ≡ g(k, X).
In the following, we focus on spatially homogeneous electric fields and use the
vector potential gauge
A0 = φ = 0, A = −c
∫ t
−∞
dt̄E(t̄). (57)
In this case, relation (56) simplifies to
g̃(k, ω;R, t) =
∫
dτdr exp
iωτ − i
~
r ·
k − e0
c
t+ τ
2
∫
t− τ
2
dt′
τ
A(t′)
g(r, τ ;R, t), (58)
what means that the gauge–invariant Green’s function g(k) follows from the (con-
ventional) Wigner transformed function ga(p) by replacing the canonical momentum
p by the gauge–invariant kinematic momentum k according to [9]
p(t, τ) = k − e0
c
t+ τ
2
∫
t− τ
2
dt′
A(t′)
τ
. (59)
Consider two important examples. First, for a constant homogeneous electric field,
E = E0, it follows
A(t) = −cE0t, p(t) = k + e0E0t. (60)
Second, for a spatially homogeneous harmonic electric field given by E(t) = E0 cos Ωt,
the vector potential and the momentum relation become, according to equations (57)
and (59),
A(t) = −cE0
Ω
sin Ωt, p(t, τ) = k +
2
τ
e0E0
Ω2
sin Ωt sin
Ωτ
2
. (61)
3.6. Gauge-invariant quantum kinetic equation for the Wign er distribution
We are now ready to derive gauge-invariant quantum kinetic equations for the
correlation functions g≷ and the Wigner distribution. Our starting point is the ki-
netic equation (50) which we transform to the Wigner representation. For this we
assume that macroscopic and microscopic length scales can be separated, where the
first is given by the cluster geometry and described by the center of mass variable
501
M.Bonitz, J.W.Dufty
R. Gauge-invariant Fourier transformation with respect to the microscopic scale r
gives rise to the momentum variable k.
In center of mass and relative variables the gradient in the momentum operator
in equation (50) becomes ∇r1 = 1/2∇R +∇r. Then, applying the spatial part of the
transform (58) to the l.h.s. of equation (50) yields
∫
d3re
− i
~
r
k−
e0
c
t1
∫
t′
1
dt̄
A(t̄)
t1−t′
1
[
i~
∂
∂t1
− 1
2me
(
~
i
∇r +
1
2
~
i
∇R +
e0
c
A(1)
)2
− VI(1)
]
×
∫
d3k1
(2π~)3
e
i
~
r
k1−
e0
c
t1
∫
t′1
dt̄ A(t̄)
t1−t′
1
g̃≷(R,k1, t1, t
′
1). (62)
The t1 and r derivatives on the second exponential are readily performed, and we
notice that the vector potential terms in the exponentials cancel. Further, using
r = ~/i∇k1 , the r−integration can be performed, except for the potential term,
with the result (details of the derivation are given in the appendix)
{[
i~
∂
∂t1
+ i~
KA(t1, t
′
1)
t1 − t′1
∇k
]
− 1
2me
[
1
2
~
i
∇R + k + KA(t1, t
′
1)
]2 }
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
, t1
)
g̃≷(R,k1, t1, t
′
1) =
= Ĩ≷(R,k, t1, t
′
1) + Ĩ≷
F (R,k, t1, t
′
1), (63)
where the field-dependent momentum KA is defined by
KA(t, t′) ≡ −e0
c
∫ t
t′
dt′′
A(t) −A(t′′)
t− t′
. (64)
For the above two examples of a constant and monochromatic field, equations (60,61),
respectively, this has the form
KA(t, t′) = e0E0
t− t′
2
, E ≡ E0 , (65)
KA(t, t′) =
e0E0
Ω2(t− t′)
{Ω(t− t′) sin Ωt+ cos Ωt− cos Ωt′} , E ≡ E0 cos Ωt. (66)
In equation (63) we introduced an effective time-dependent confinement potential
which includes the Hartree mean field,
V eff
I (R, t) ≡ VI(R, t) + ΣH(R, t), (67)
ΣH(R, t) =
∫
dr V (R − r)n(r, t) =
∫
d3r
d3k
(2π~)3
V (R− r)f(r,k, t), (68)
which depends on the nonequilibrium electron density profile. Finally, the r.h.s. of
equation (63) contains all remaining two-particle and higher order contributions
502
Quantum kinetic theory of metal clusters
to the dynamics (i.e. collisions and exchange mean field) the explicit structure of
which will be discussed below in section 3.7. These gauge-invariant equations of
motion for the two-time correlation functions are the basis for all further analysis.
In principle, their direct numerical solution is possible, following previous numerical
work for spatially homogeneous charged particle systems, e.g. [2,22]. Here, however,
we concentrate on the derivation of a closed equation of motion for the (single-time)
Wigner function as it is essentially more simple.
To obtain from equation (63) the gauge-invariant equation for the single-time
distribution function we also need the adjoint equation and compute the difference
of the two, e.g. [2,3]. Taking this difference equation for g< at equal times, t1 = t′1,
allows us to express g< by the Wigner distribution according to relation (47) with
the result (details are given in the appendix)
{
∂
∂t
+
k
me
∇R − e0E(t)∇k
}
f(R,k, t)
+
2
~
∫
d3r
∫
d3k1
(2π~)3
sin
r (k − k1)
~
V eff
I
(
R +
r
2
, t
)
f(R,k1, t) =
= I(R,k, t) + IF(R,k, t), (69)
where I(R,k, t) ≡ −2Re[Ĩ<(R,k, t, t)] and IF(R,k, t) ≡ −2Re[Ĩ<
F (R,k, t, t)]. Here
E is the total electric field (external plus induced) which obeys Maxwell’s equations
which have to be solved selfconsistently with the kinetic equation (69).
Equation (69) is not yet the final equation for the Wigner distribution which
will be presented below, in section 3.8, after evaluation of the collision and exchange
terms in section 3.7. But before doing this, we analyze the term with the external
potential (the integral on the l.h.s.) more in detail. The integral representation of
V eff
I in equation (69) is still completely general and does not use any assumptions on
the space dependence of the confinement potential. The present form applies even
to sharp spatial changes and can be used with any form of pseudo-potentials for the
core ions of clusters and it fully includes quantum effects.
Analytical simplifications are possible in the case of the jellium potential due to
its smooth variation in space. In that case it is possible to eliminate the integrations
over the jellium potential by expanding V eff
I around the center of mass coordinate,
e.g. [2],
V eff
I
(
R ± r
2
)
=
∞
∑
l=0
(±1)l
2ll!
∂lV eff
I (R)
∂Ri∂Rj . . . ∂Rl
· rirj . . . rl , (70)
where rj∂V
eff
I /∂Rj denotes the scalar product ∇Rj
V eff
I · rj and so on. With this, the
integral term in equation (69) becomes
2
~
∫
d3r
∫
d3k1
(2π~)3
sin
r (k − k1)
~
V eff
I
(
R +
r
2
, t
)
f(R,k1, t) =
= −
∞
∑
l=0
(i~)2l+1
22l(2l + 1)!
∂2l+1V eff
I (R, t)
∂Ri∂Rj . . . ∂R2l+1
· ∂2l+1f(R,k, t)
∂ki∂kj . . . ∂k2l+1
503
M.Bonitz, J.W.Dufty
= −
{
∇RV
eff
I (R, t)∇k −
~
2
24
∂3V eff
I (R, t)
∂R1∂R2∂R3
· ∂3
∂k1∂k2∂k3
. . .
}
f(R,k, t). (71)
Applying this expansion to the jellium potential (10) we readily see that the first
term of the expansion of V eff
I is dominating: indeed, inside the cluster where VI is
parabolic, the third derivative vanishes exactly. Outside, where VI is Coulomb-like,
the third derivative is short range, it decays as 1/r4 and gives only a small correction.
We may expect that more realistic pseudopotentials will behave similarly since they
do not exhibit sharp changes or even divergencies [13].
3.7. Gauge-invariant expression for the Hartree-Fock and c ollision terms.
Nonlocal corrections
We now perform the gauge-invariant Fourier transform of the right hand side of
the kinetic equation which contains all two-particle quantities related to mean field
and correlation effects. Let us start with the collision integral. If the characteristic
length scales of the scattering processes are small compared to the cluster radius,
the local approximation can be used. Then the collision integrals retain their form
of the homogeneous case with R being an additional parameter, and our previous
results [8,9] can be applied leading to
I(R,k, t) = −2Re
∫ t
t0
dt̄
{
Σ>[R,k + KA(t, t̄), t, t̄]g<[R,k + KA(t, t̄), t̄, t]
− Σ<[R,k + KA(t, t̄), t, t̄]g>[R,k + KA(t, t̄), t̄, t]
}
. (72)
Note that the momentum arguments in all functions are shifted by the field-dependent
momentum KA which reflects the explicit field-dependence of the two-particle scat-
tering process. This is an important effect leading to the so-called intracollisional
field effect and to nonlinear phenomena including collisional harmonics generation,
(inverse) bremsstrahlung, multiphoton excitation and ionization etc [8,9].
The result (72) gives the local approximation for the collision integral. It is
however, straightforward to write down the first gradient corrections. For this we
recognize that (72) is of the general structure (45), except for the time integral. We
thus can write [using the notation of equation (45)]
Igrad(R,k, t) = ~ Im
∫ t
t0
dt̄ (∇R1∇k2 −∇R2∇k1)
×
{
Σ>[R1,k1 + KA(t, t̄), t, t̄]g<[R2,k2 + KA(t, t̄), t̄, t]
− Σ<[R1,k1 + KA(t, t̄), t, t̄]g>[R2,k2 + KA(t, t̄), t̄, t]
}
∣
∣
∣
(R,k)
. (73)
This result has to be added to the local approximation (72).
Expressions (72,73) are still completely general. A particular collision process is
specified by the appropriate choice of the selfenergies. Further, the two-time correla-
tion functions have to be expressed by the Wigner distributions via the so-called re-
construction ansatz. Its original form due to Baym and Kadanoff [3] was generalized
504
Quantum kinetic theory of metal clusters
by Lipavski, Spicka and Velicky [33] to nonequilibrium systems, properly accounting
for causality and retardation effects. The latter ansatz was further generalized to the
presence of external fields [6,9], and here we account, in addition, for weak spatial
inhomogeneity which gives rise to gradient corrections in the reconstruction ansatz:
±g≷(R,k, t, t′) = gR(R,k, t, t′)f≷[R,k − KA(t′, t), t′]
− f≷[R,k − KA(t, t′), t]gA(R,k, t, t′) +
i~
2
(∇R1∇k2 −∇R2∇k1)
×
{
gR(R1,k1, t, t
′)f≷[R2,k2 −KA(t′, t), t′]
− f≷[R1,k1 − KA(t, t′), t]gA(R2,k2, t, t
′)
}∣
∣
∣
(R,k)
, (74)
where f> ≡ 1 − f and f< ≡ f .
Consider now the Hartree-Fock terms. They have a product form analogous to
the collision term. Furthermore, due to the time-diagonal structure of ΣHF, the field-
induced momentum shift KA vanishes (KA(t, t) = 0). In particular, the Hartree term
can be written as (see appendix)
ΣH(r1, t1; r
′
1, t
′
1) = δ(r1 − r′1)δ(t1 − t′1)
∫
d3r2V (r1 − r2)n(r2, t). (75)
This potential is local in space and is treated exactly like the jellium potential
VI . Both terms can be combined into an effective confinement potential V eff
I , see
equation (82) below.
Let us now turn to the nonequilibrium exchange term which is given by
ΣF(r1, t1; r
′
1, t
′
1) = δ(t1 − t′1)Σ
F(r1, r
′
1, t1),
ΣF(r1, r
′
1, t) = −V (r1 − r′1)f(r1, r
′
1, t), (76)
and has the Fourier representation
ΣF(R,k, t) = −
∫
d3k2
(2π~)3
V (k2)f(R,k − k2, t). (77)
In the kinetic equation it appears in the combination
IF(R,k, t) = −2Re[ΣF(R,k, t)g<(R,k, t, t)], (78)
which vanishes exactly (on the time-diagonal, g< is purely imaginary). This means
that the only nonzero exchange contributions arise from the gradient corrections
which appear due to the product ΣF · g<. Using the general expression (45), we
obtain
IF(R,k, t) = 2Re
(
1 +
i~
2
{. . .}
)
ΣF (R1,Q1, t) g
<(R2,Q2, t, t)
∣
∣
∣
(R,k)
=
=
∫
d3k1
(2π~)3
V (k1)
{
∇R1f(R1,Q1 − k1, t)∇Q2f(R2,Q2, t)
−∇Q1f(R1,Q1 − k1, t)∇R2f(R2,Q2, t)
}
∣
∣
∣
(R,k)
. (79)
505
M.Bonitz, J.W.Dufty
Of the two terms in curly brackets, the first has the form of the gradient of a lo-
cal exchange potential, ∇RΣF (R,k) · ∇kf(R,k). This term may be included into
the gradient of a local total potential together with the confinement and Hartree
field, V tot
I (R,k, t) = VI(R) + ΣH(R, t) + ΣF (R,k, t). However, the second term ap-
pears to be of the same order and cannot be neglected if the first is retained11.
This second contribution has the form ∇kΣ
F (R,k) · ∇Rf(R,k) and effectively
renormalizes the velocity in the drift term of the kinetic equation, k/me∇Rf →
[k/me +∇kΣF (R,k)]∇Rf . We may further generalize the drift term by permitting
a general single-particle energy dispersion (or band structure) k2/2me → ε(k), as
our derivation did not depend on the explicit form. Finally, we can write the drift
term as ∇kε̄(k) with the effective single-particle energy being renormalized by the
Hartree-Fock and confinement energy (using the fact that ΣH and VI are momentum-
independent), ε̄(R,k, t) ≡ ε(k)+ΣH(R, t)+ΣF (R,k, t)+VI(R). Finally, the gradient
of the effective potential can be written in terms of the same effective single-particle
energy, since the kinetic energy is R-independent, ∇RV
eff
I = ∇Rε̄, allowing for a
highly symmetric form of the kinetic equation 12.
3.8. Resulting quantum kinetic equation
Collecting all terms together, the resulting kinetic equation which includes terms
up to first order gradients reads
{
∂
∂t
+ ∇k ε̄(k,R, t)∇R − [e0E(t) + ∇R ε̄(R,k, t)]∇k
}
f(R,k, t) = Ī(R,k, t),
(80)
ε̄(R,k, t) ≡ ε(k) + V eff
I (R, t) + ΣF(R,k, t), (81)
V eff
I (R, t) ≡ VI(R) + ΣH(R, t), (82)
Ī(R,k, t) ≡ I(R,k, t) + Igrad(R,k, t), (83)
where ΣH,ΣF , I and Igrad have been defined in equations (72), (73), (75), (77).
We underline that this equation is very general. It applies to arbitrary nonequi-
librium situations and strong electromagnetic fields with arbitrary amplitude and
time-dependence (the only assumption is that the field does not vary significantly
on the space scale RI of the cluster). Due to the gauge-invariant derivation, the
resulting kinetic equation with all gradient correcations is gauge-invariant as well.
This equation is internally consistent on the level of first gradients. The neglect of
higher order spatial derivatives assumes that RI |∇2n(R)| � |∇n(R)|.
Equation (80) directly generalizes the density functional results considered above
to nonequilibrium situations and the presence of a strong field: in particular, the to-
tal energy Eve[n(r)] appearing in equation (14) is generalized to the nonequilibrium
11In particular, the use – in a classical Vlasov equation – of an effective potential which includes
exchange, as done by several authors, e.g. [27,28], while neglecting the second gradient contribution
in equation (79) can not be justified from the perspective of quantum kinetic theory.
12The general structure is, of course, the same as in the quasiparticle equations of Landau Fermi
liquid theory.
506
Quantum kinetic theory of metal clusters
expression ε̄[f ] depending on the nonequilibrium momentum-dependent Wigner dis-
tribution instead of the local density:
E(R, t) =
∫
d3k
(2π~)3
ε̄ (R, t, [f(R,k, t)]) f(R,k, t) + Ucor(R, t), (84)
where the correlation energy contribution Ucor arises from the collision integral.
Finally, we discuss possible generalizations of the above kinetic equation. The
main simplification involved in the derivation of equation (80) is the restriction to
first order gradient corrections. On the other hand, we can use the full quantum
result for the local potential and the Hartree term. The exchange contribution,
however, cannot be included since it is momentum dependent and has a gradient
expansion completely different from that of V eff
I
{
∂
∂t
+ ∇k ε(k)∇R − e0E(t)∇k
}
f(R,k, t) + ÎF
(
[ΣF ], [f ]
)
+
2
~
∫
d3r
∫
d3k1
(2π~)3
sin
r (k − k1)
~
V eff
I
(
R +
r
2
, t
)
f(R,k1, t) = Î(R,k, t), (85)
V eff
I (R,k, t) ≡ VI(R) + ΣH(R, t). (86)
Here the integral term with V eff
I contains gradient corrections of all orders, the first
two were given in equation (71), and ÎF and Î denote the gradient expansions of
the exchange term and the collision integral, respectively, the first orders of which
were given above by equations (73), (79). To go beyond the kinetic equation (80)
in a strict manner would require to retain in all terms all contributions up to a
given derivative ∂kf/∂Rk which can be expected to be total energy and sum rule
conserving and fully internally consistent. On the other hand, simpler schemes seem
possible. First of all a simplified treatment of the collision integral compared to the
remaining terms, (e.g. by retaining only the 0th or 0th plus 1st gradient term) seems
reasonable for most applications. Secondly, a treatment of the Fock term on the level
of first gradients (cf. equation (80) should capture the dominant exchange effects.
At the same time, a full quantum treatment of V eff
I is possible in order to analyze
the role of quantum effects such as tunneling etc.
3.9. Spectral properties of the valence electrons.
Propagator in a strong field
The strength of the Green’s functions approach is that statistical properties – giv-
en by the Wigner distribution – and dynamical information – related to the spectral
function or density of states – are treated fully selfconsistently. This is of particular
importance if approximations are being developed. How the spectral information
enters the Green’s functions is seen explicitly in the reconstruction relation (74),
[although it is an approximation]. Here the spectral information is contained in the
retarded and advanced Green’s functions which, consequently, have to be determined
together with the distribution function.
507
M.Bonitz, J.W.Dufty
Indeed, both the equations of motion for f , equation (80), and gR/A are derived
from the more general equations of motion for the two-time correlations functions
g≷(t, t′), the Kadanoff-Baym/Keldysh equation (63) and its adjoint, equation (131).
So far we considered only one possible combination of the two – its difference, which
resulted in the equation of motion for the Wigner distribution (taking the difference,
yields essentially an equation containing a commutator of f and the hamiltonian).
Also, in taking this equation only on the time diagonal, t = t′ or, equivalently, τ = 0
by no means exhausted the full information contained in the two-time funtions.
3.9.1. Gauge-invariant equation of motion for the retarded Green’s function
The additional spectral information is recovered by considering the Green’s func-
tions away from the time diagonal, i.e. as a function of the relative time τ . This is
most easily analyzed by computing the sum of equation (63) and its adjoint. The
result is (details are given in the appendix)
{
i~
∂
∂τ
+ i~
KA
+(t1, t
′
1)
t1 − t′1
∇k −
1
2me
[(
~
2i
∇R
)2
+ k2 +
(KA)2(t1, t
′
1) + (KA)2(t′1, t1)
2
+
~
i
KA
−(t1, t
′
1)∇R + 2KA
+(t1, t
′
1) k
]}
g̃R/A(R,k, t1, t
′
1)
− 1
2
∫
d3r
∫
d3k1
(2π~)3
{
e−i
r(k−k1)
~ V eff
I
(
R +
r
2
, t1
)
+ ei
r(k−k1)
~ V eff
I
(
R − r
2
, t′1
)}
× g̃R/A(R,k1, t1, t
′
1) = I
R/A
− (R,k, t1, t
′
1) + δ(t1 − t′1), (87)
where the definition of I
R/A
− is analogous to that of I≷
− and we introduced
KA
±(t, t′) ≡
[
KA(t, t′) ± KA(t′, t)
]
. (88)
From the retarded and advanced Green’s functions (in fact, due to the symmetry
(53) knowledge of one function in one half-plane, e.g. knowledge of gR for τ > 0
is sufficient) the complete spectral information is known. In particular, the spectral
function follows from the definition (48). The result (87) is still completely general.
It includes strong field effects, spatial inhomogeneities and correlations. Analytical
expressions are of special interest, and we now consider some important cases.
3.9.2. Spectral function for a homogeneous system in an exte rnal field
Consider first the simplest case of a spatially homogenous ideal electron gas (no
mean field and collisions) in the absence of confinement and external fields, i.e.
KA = ∇R = VI = 0. The resulting equation is readily solved
{
i~
∂
∂τ
− k2
2me
}
gR(k, τ, t) = δ(τ), gR(k, τ) = − i
~
Θ(τ) e−
i
~
k2
2me
τ . (89)
With the definition (48) this yields the spectral function
a(k, τ) = e−
i
~
k2
2me
τ , (90)
508
Quantum kinetic theory of metal clusters
corresponding, in frequency space, to a sharp energy spectrum,
a(k, ω) = 2π~δ(~ω − k2/(2me)), which is peaked at the free single-particle energy.
Next, we include mean field effects (and possible band structure effects) which does
not change the structure, except for a possible slow (macroscopic) time-dependence
of the effective single-particle energy (which includes the Hartree-Fock energy) via
the Wigner distribution,
{
i~
∂
∂τ
− ε(k, t)
}
gR(k, τ, t) = δ(τ), a(k, τ, t) = e
− i
~
t+τ/2
∫
t−τ/2
dt̄ ε(k,t̄)
. (91)
The corresponding frequency spectrum again consists of a single sharp line, now
shifted from k2/2m to ε(k, t).
Further, correlation effects (related to the collision integral) cause a damping of
the spectral function (91), so the oscillations in τ are modulated with an overall
decay with increasing |τ | which is naturally interpreted as a finite life time of the
(quasi-)particle. Correspondingly, in frequency space, a broadening of the peak is
observed. A detailed analysis has been given, e.g. in [2,31]. Here we only recall
the main result: an intuitive inclusion of damping by a constant exponential factor
e−γ|τ | leading to a replacement of the energy delta function by a Lorentzian, gives
rise to unphysical results which are due to the slow decay at large frequencies of
that spectral function. Therefore, improved analytical results have been derived [31]
which have a zero slope at the time diagonal τ = 0 and cure this defect.
For our present analysis, where we are interested in the effect of intense electro-
magnetic fields on the spectrum, we expect that a detailed selfconsistent treatment
of the correlation effects on the spectral function can be avoided. As was found in
many investigations, in strong fields inclusion of all external fields into the spec-
tral function together with Hartree-Fock effects, is crucial for a correct modeling of
the many-particle behavior, see e.g. [32]. Correlation effects lead to an energy shift
and broadening which is generally well understood. Therefore, these effects can be
added to the collisionless result in perturbation theory. We will thus concentrate in
the following on the first step of perturbation theory where collision effects on the
spectral function are neglected.
The next step is to include, in addition to the single-particle energy, an external
spatially homogeneous field. Then, the equation reads
{
i~
∂
∂τ
+
i~
τ
KA
+(t1, t
′
1)∇k
− 1
4me
[
(
k + KA(t1, t
′
1)
)2
+
(
k + KA(t′1, t1)
)2
]}
gR(R,k, t1, t
′
1) = δ(τ), (92)
and again a solution is possible, e.g. [9],
a(k, τ, t) = exp
− i
4me~
t+τ/2
∫
t−τ/2
dt̄
[
2k2 +
(
KA(t̄, t′)
)2
+
(
KA(t′, t̄)
)2
]
. (93)
509
M.Bonitz, J.W.Dufty
One readily obtains the results for limiting cases: for a constant electric field, cf.
equation (65), it follows
a(k, τ ;E0) = exp
{
− i
~
[
k2
2me
τ +
e20E
2
0
24me
τ 3
]}
. (94)
The field dependence give rise to a nonharmonic dependence on the difference time
τ which transforms to a modified nonharmonic energy spectrum. Calculating the
Fourier transform of (94) yields, e.g. [6],
a(k, ω;E0) =
2π~
α(E0)
Ai
(
k2/2me − ~ω
α(E0)
)
, (95)
where Ai is the Airy function13 and α = (~2e20E
2
0/8me)
1/3. This is an exact result
valid for arbitrary field strength. Its main feature is that, due to the action of the
field, the sharp energy spectrum of the field-free case, equation (90), is replaced by
a broadened peak, together with additional lower side peaks.
Consider now the case of a monochromatic time-dependent field, equation (66),
then the solution (94) becomes
a(k, τ ;E0,Ω) = exp
{
− i
~
[(
k2
2me
+ Upond
)
τ + ∆E(t, τ ;E0,Ω)
]}
, (97)
where Upond ≡ (e2E2
0)/(4meΩ
2) is the mean kinetic energy of a free particle in the
field (ponderomotive energy) and the field-dependent energy correction is
∆E(t, τ ;E0,Ω) ≡ −Upond
(
sin Ωτ cos 2Ωt
Ω
− sin2 Ωt sin2 Ωτ/2
Ω2τ
)
.
As discussed e.g. in [6,9], this spectral function describes an “electron-photon-quasiparticle”.
For example, in a harmonic field with frequency Ω and amplitude E0, the peak in the
corresponding energy spectrum is shifted from k2/2me to k2/2me +Upond, and there
appear additional peaks at a distance of integer multiples of ~Ω from this peak.
3.9.3. Spectral function for a weakly inhomogeneous system .
Effective quantum potential
Let us now discuss the influence of a space-dependent confinement potential VI ,
first without external field A. Then, the equation for gR reads
{
i~
∂
∂τ
− 1
2me
[ (
~
2i
∇R
)2
+ k2
]}
gR(R,k, t1, t
′
1)
− 1
2
∫
d3r
∫
d3k1
(2π~)3
{
e−i
r(k−k1)
~ V eff
I
(
R +
r
2
, t1
)
+ ei
r(k−k1)
~ V eff
I
(
R − r
2
, t′1
)}
× g̃R/A(R,k1, t1, t
′
1) = δ(τ). (98)
13We make use of the integral representation of the Airy function
∫ ∞
0
cos (at3 ± xt)dt =
π
(3a)1/3
Ai
[
± x
(3a)1/3
]
. (96)
510
Quantum kinetic theory of metal clusters
If the potential is weakly inhomogeneous, i.e. ∇RV
eff
I is small and V eff
I (R + r/2) ≈
V eff
I (R), then the solution of equation (98) is given by (for generality, we include the
time-dependent Hartree mean field into the potential)
a(R, k, τ, t) = e
− i
~
[
k2
2me
τ+
t+τ/2
∫
t−τ/2
dt̄ V eff
I (R,t̄)
]
. (99)
It is obvious that this “local” approximation for the propagator where the electrons
at each space point are assigned a definite single-particle energy is valid only in the
classical limit. In contrast, for quantum particles, this is prevented by the Heisenberg
uncertainty principle which has the consequence that kinetic and potential energy
operators do not commute. In the following we will define an effective quantum
potential V eff
Q [23] which allows us to retain the classical local form of the propagator
but properly takes into account quantum diffraction effects,
a(R,k, τ, t) = e
− i
~
[
k2
2me
τ+
t+τ/2
∫
t−τ/2
dt̄ V eff
Q (R,k,t̄)
]
. (100)
This effective potential is momentum dependent and will be analyzed in section 3.9.4.
For completeness, we mention that similar concepts of an effective quantum poten-
tial have been used in a variety of contexts. We mention the work of Feynman and
Kleinert on the application of path integrals, e.g. [24], Kelbg and others on quan-
tum pair potentials, cf. [34,35] and references therein and Ferry and co-workers on
effective potentials for quantum transport [25,26]. Here, for the first time this idea
is applied to the nonequilibrium spectral properties of quantum particles in a strong
field. For the present application, the advantage of this definition of the quantum
potential is the formal decoupling of the spectral problem from the cluster dynamics.
Finally, we restore the external electric field and assume that the same represen-
tation in form of a local spectral function is possible. In that case we have to allow
for a field dependent quantum potential,
a(R,k, τ, t) = e
− i
~
k2
2me
τ+
t+τ/2
∫
t−τ/2
dt̄
(K
A(t̄,t′))
2
+(KA(t′,t̄))
2
4me
+V eff
Q (R,k,t̄;A)
. (101)
Naturally, equations (100) and (101) are only an ansatz. But we will demonstrate
below in section (3.9.4) that this ansatz yields a closed equation for the quantum
potential which can be solved in a number of important cases.
The spectral function (101) represents a far-reaching generalization of all the
above special cases. Even without knowing yet the explicit analytical structure of the
quantum potential, with its help all properties which are due to space-dependence
of the spectral function can be analyzed straightforwardly. These properties will
show up in a variety of places, including the polarization function and the local ap-
proximation to the collision integral. The most important occurences of the spectral
function are combinations such as the product
gR(R,k1, τ, t)g
A(R,k2,−τ, t) = gR(R,k1, τ, t)[g
R(R,k2, τ, t)]
∗ =
= − i
~
Θ(τ)e
− i
~
{
(
k2
1
2me
−
k2
2
2me
)
τ+
t+τ/2
∫
t−τ/2
dt̄ [V eff
Q (R,k1,t̄;A)−V eff
Q (R,k2,t̄;A)]
}
, (102)
511
M.Bonitz, J.W.Dufty
which enters the collision integrals, see section 3.10, and the RPA polarization [9,11].
Further, we can compute the gradient of the propagator,
∇Rg
R(R,k, τ, t) = − i
~
t+τ/2
∫
t−τ/2
dt̄∇RV
eff
Q (R,k, t̄;A) · gR(R,k, τ, t), (103)
which appears in the local approximation of the collision integral etc.
3.9.4. Calculation of the Quantum potential V
eff
Q
We now discuss the quantum potential more in detail, starting with the field
free case. To simplify the notation we consider a potential which is only weakly
time-dependent, i.e. τ∂V eff
I /∂t � V eff
I (t), allowing to approximate V eff
I (t ± τ/2) ≈
V eff
I (t)14. Then, the ansatz (100) simplifies to
gR(R,k, τ, t) = − i
~
Θ(τ)e
− i
~
{
k2
2m
+V eff
Q (R,k,t)
}
τ
. (104)
Here, the quantum potential still contains a possible weak (parametric) dependence
on the macroscopic time which will be suppressed below to shorten the notation.
Inserting this expression into equation (98), we readily obtain the equation for the
quantum potential [23]
∫
d3k1
(2π~)3
{
δ(k − k1)
[
V eff
Q (R,k) − 1
8me
(
i~τ∇2
RV
eff
Q (R,k) + [∇RV
eff
Q (R,k)]2τ 2
)
]
−
∫
d3r cos
[
r
k − k1
~
]
V eff
I
(
R +
r
2
, t
)}
e
− i
~
[
k2
1
2me
+V eff
Q (R,k1)
]
τ
. (105)
This is an exact equation and various analytical or numerical solution schemes
are possible. Since we are interested in weakly inhomogeneous systems it is nat-
ural to develop a perturbation solution in terms of gradients of the (effective)
confinement potential. The zeroth order approximation is given by the neglect of
all terms involving ∇RV
eff
I , thus V eff
I (R + r/2) ≈ V eff
I (R), and the r−integral
yields −V eff
I (R)(2π~)3δ(k − k1). In the curly brackets in (105) we are left with
δ(k−k1)[V
eff
Q (R,k)−V eff
I (R)], so the zeroth order result for the quantum potential
is
V
eff (0)
Q (R,k, t) = V eff
I (R, t), (106)
and we essentially recover the previous result of the local spectral function (99).
14This is usually fulfilled for time dependencies arising from the Hartree-Fock selfenergy, i.e. via
the distribution function. A possible exception are ionization/recombination processes where the
carrier density, and thus, the effective potential V eff
I may change rapidly during a short period of
time.
512
Quantum kinetic theory of metal clusters
The next approximation is obtained by solving equation (105) by replacing, in
all gradient terms, V eff
Q (R,k) → V eff
I (R) and calculating the r−integral in first order
in gradient terms, given by
−(2π~)3δ(k − k1)
[
V eff
I (R) − ~
2
8
∇2
RV
eff
I
∂2
∂k2
1
]
.
Carrying out the momentum differentiation on the zeroth order Green’s function,
we obtain the first iteration for the quantum potential
V
eff (1)
Q (R,k, t) = V
eff (0)
Q (R, t) + i
~τ
4m
∇2
RV
eff
I (R, t)
+
τ 2
8me
[
[∇RV
eff
I (R, t)]2 +
k2
me
∇2
RV
eff
I (R, t)
]
, (107)
which contains, in addition to the zeroth approximation, all contributions up to
second order in the gradients. The ratio of the imaginary part (second term on the
r.h.s.) to the last term is of the order of 1/τcor where τcor is the correlation time (or
quasi-particle life time) and can be neglected. Using (107) we obtain for the spectral
function
a(R,k, τ, t) = e
− i
~
{[
k2
2m
+V eff
I (R,t)
]
τ+δVQ(R,k,t)τ3
}
, (108)
with δVQ(R,k, t) =
1
8me
[
[∇RV
eff
I (R, t)]2 +
k2
me
∇2
RV
eff
I (R, t)
]
, (109)
where, as expected, the quantum potential becomes momentum dependent. Interest-
ingly, the inhomogeneity of the confinement field causes an anharmonic correction
to the spectral function which is similar to that of a constant electric field, cf. equa-
tion (94). Again we readily obtain the energy spectrum by a Fourier transform,
a(R,k, ω, t) =
2π~
[3~2δVQ]1/3
Ai
(
k2/2me + V eff
I (R, t) − ~ω
[3~2δVQ]1/3
)
, (110)
which is shown in figure 2. As in the case of a homogeneous field, cf. equation (95),
the peak of the spectral function around ~ω = k2/2me +V eff
I (R, t) broadens, i.e. the
sharp single-particle energy is smeared out (in the figure, the parameter α denotes
α ≡ (3~
2δVQ)1/3). But here this is due to spatial inhomogeneity – which directly re-
flects the coordinate-momentum uncertainty of quantum mechanics15. Correspond-
ingly, with increasing inhomogeneity, the peak position shifts to higher energies,
i.e. the effective local single-particle energy increases. This is readily understood:
in an inhomogeneous confinement field, a quantum particle aquires an additional
kinetic energy which arises from spatial compression of its wave function which is
proportional to the local curvature of the field.
15In contrast to the field case, this is a perturbation result in terms of potential gradients.
513
M.Bonitz, J.W.Dufty
Figure 2. Spectral function a(R,k, ω, t;α), equation (110), at fixed time t and
a fixed phase space point (R,k) such that the dimensionless local single-particle
energy ε(R,k) = k2/2m+V eff
I (R, t) = 1. Thus, the ideal spectral function would
have a singularity at ~ω = 1 which is observed in the limit of vanishing inho-
mogeneity α ≡ (3~
2δVQ)1/3 → 0. With increasing inhomogeneity, due to the
Heisenberg uncertainty, the main peak broadens and shifts towards higher ener-
gies.
3.9.5. Quantum potential for a system in an external field
We now restore the electromagnetic field in our equations and generalize the
result for the quantum potential. The calculations follow exactly the same lines as
before, so we limit ourselves to presenting the equation for the effective quantum
potential in lowest order gradient approximation
{
i~
∂
∂τ
+
i~
τ
KA
+(t1, t
′
1)∇k −
1
2me
(
~
2i
∇R
)2
+
i~
4me
KA
−(t1, t
′
1)∇R
− 1
4me
[
(
k + KA(t1, t
′
1)
)2
+
(
k + KA(t′1, t1)
)2
]
−
(
V eff
I (R, t) − ~
2
8
∇2
RV
eff
I (R, t)
∂2
∂k2
) }
a(R,k, t1, t
′
1) = 0, (111)
514
Quantum kinetic theory of metal clusters
with the ansatz for the spectral function
a(R, k, τ, t;A) = e
− i
~
[
k2
2me
+V eff
Q (R,k,t;A)
]
τ− i
4me~
t+τ/2
∫
t−τ/2
dt̄
[
2k2+(KA(t̄,t′))
2
+(KA(t′,t̄))
2
]
. (112)
Compared to the field-free equation for the quantum potential, equation (111) con-
tains a homogeneous electromagnetic field of arbitrary strength and time-dependence.
The field couples to the space dependence of the quantum potential via the term
KA
−(t1, t
′
1)∇R giving rise to an additional contribution to the quantum potential
(second term on the r.h.s.):
V
eff (1)
Q (R,k, t;A) = V
eff (0)
Q (R, t) +
τ
4me
KA
−(t1, t
′
1)∇RV
eff
I (R, t)
+
τ 2
8me
[
[∇RV
eff
I (R, t)]2 +
k2
me
∇2
RV
eff
I (R, t)
]
, (113)
which is proportional to the scalar product of the external field and the force acting
on the electrons from the effective confinement potential. Inserting this result in
equation (112) we obtain a very general result for the spectral function of electrons
in a combined external electromagnetic field and weakly inhomogeneous confinement
potential.
While this result represents the lowest order approximation in terms of potential
gradients, it is valid for an arbitrary electromagnetic field. For special time depen-
dencies of the field further analytical progress can be made. As an example, consider
again a time-independent field. Then equation (88) yields KA
−(t1, t
′
1) = e0E0τ/2, and
the quantum potential becomes
V
eff (1)
Q (R,k, t;A)τ = V
eff (0)
Q (R, t)τ + δVQ(R,k, t;A)τ 3 , (114)
δVQ(R,k, t;A) =
1
8me
[
[∇RV
eff
I ]2 +
k2
me
∇2
RV
eff
I + e0E0 · ∇RV
eff
I
]
.
The resulting spectral function in frequency space has again the form (110) and
is represented by figure 2 with α denoting (3~
2δVQ)1/3, where now δVQ has to be
replaced by the generalized field dependent result (114).
3.10. Collision integral
In this section we derive the collision integral in the kinetic equation of the
valence electrons. The main goal here is to see both the effect of the external field
and the confinement potential on the collision process. An important issue is the
choice of the asymptotic states of the electrons long before and after the collision
process. In contrast to plasmas, here the electrons are bound to the cluster core,
i.e. their wave functions are not free plane waves. The modification can be two-fold:
first, the single-particle spectrum is modified and second, the asymptotic states of
the scattering electron pair may be correlated. In the following, we will neglect the
latter aspect, improvements can be made by adopting results from metal theory
515
M.Bonitz, J.W.Dufty
and compute T-matrix cross sections if necessary, e.g. [27]. The first aspect, on the
other hand, is expected to be captured correctly by the electron spectral functions
(or, equivalently, the propagators gR/A), cf. section 3.9. Finally, the use of a static
potential should be criticized as well. We expect essential modifications to arise
from collective excitations (plasmons) and efficient scattering processes which involve
emission and absorption of Mie plasmons by the electrons. The generalization of the
collision integral to the dynamically screened Born approximation is straightforward
and can follow earlier work for homogeneous plasmas in strong field [9].
Having these remarks in mind, it is reasonable to start with the simplest approx-
imation – the static second Born approximation. The corresponding gauge-invariant
expression for the selfenergy [8] in local approximation is given by
Σ≷(R,k1, t, t
′) = ~
2
∫
dk2dk̄1dk̄2
(2π~)9
V 2(k1 − k̄2) δ(k1 + k2 − k̄1 − k̄2)
× g≷(R, k̄1, t, t
′)g≷(R, k̄2, t, t
′)g≶(R,k2, t
′, t), (115)
which, using (72), leads to the collision integral
I(R,k1, t) = −2~
2Re
∫
dk2dk̄1dk̄2
(2π~)9
V 2(k1 − k̄2) δ(k1 + k2 − k̄1 − k̄2)
∫ t
t0
dt̄
{
g>(R, k̄A
1 , t, t̄)g
>(R, k̄A
2 , t, t̄)g
<(R,kA
2 , t̄, t)g
<(R,kA
1 , t̄, t) − (>↔<)
}
,(116)
where the superscript “A” denotes that all momenta are shifted, according to kA
1 ≡
k1 + KA(t, t̄) etc. Now we express the two-time correlation functions by the Wigner
distributions using the generalized Kadanoff-Baym ansatz (74),
I(R,k1, t) = −2~
2Re
∫
dk2dk̄1dk̄2
(2π~)9
V 2(k1 − k̄2) δ(k1 + k2 − k̄1 − k̄2)
×
∫ t
t0
dt̄ gR(R, k̄A
1 , t, t̄)g
R(R, k̄A
2 , t, t̄)[g
R(R,kA
2 , t, t̄)]
∗[gR(R,kA
1 , t, t̄)]
∗
×
{
f>(R, k̄Q
1 , t̄)f
>(R, k̄Q
2 , t̄)f
<(R,kQ
2 , t̄)f
<(R,kQ
1 , t̄) − (>↔<)
}
, (117)
where the shift of the momentum arguments in the distribution functions is now
given by QA, equation (127) in the Appendix, i.e. kQ
1 ≡ k1 + QA(t, t̄) and so on.
What is left now is to evaluate the spectral information of the four propagators
which determines the energy balance of the scattering event in the combined ex-
ternal and confinement fields. Using the result for the product of two propagators,
equation (102), we immediately obtain
I(R,k1, t) = −2
∫
dk2dk̄1dk̄2
(2π~)9
V 2(k1 − k̄2) δ(k1 + k2 − k̄1 − k̄2)
×
∫ t
t0
dt̄ cos
{
1
~
∫ t−t̄
t0
dt1
[
ε(R, k̄1, t1) + ε(R, k̄2, t1) − ε(R,k1, t1) − ε(R,k2, t1)
]
}
×
{
f>(R, k̄Q
1 , t̄)f
>(R, k̄Q
2 , t̄)f
<(R,kQ
2 , t̄)f
<(R,kQ
1 , t̄) − (>↔<)
}
, (118)
516
Quantum kinetic theory of metal clusters
where the effective single-particle energy which enters the energy balance is given
by
ε(R,k, t) ≡ k2
2me
+ V eff
Q (R,k, t;A). (119)
As in the homogeneous case, the field drops out of the energy balance of scattering
of particles with same charge to mass ratio since the field does not change their
distance. In contrast, in the case of electron ion scattering the field changes the
energy balance by [8] (k̄1−k1)RA(t, t̄), where RA here is the distance change of two
particles in the electromagnetic field which is defined in analogy to definition (128).
If the ions are treated as a fixed background (as it is the case with the jellium model),
no electron-ion collision integral appears; electron-ion scattering then appears via
the jellium potential on the left hand side of the kinetic equation. The corresponding
nonlinear effects in a strong field are analyzed in [11].
We see from equation (118) that the confinement field does in fact have an
influence on the scattering process which arises from the momentum dependence of
the quantum potential. This is directly seen from the explicit result for the quantum
potential, equation (113). Using this result, we can rewrite the energy difference ∆ε
in equation (118) and see that the last term in the quantum potential (113) modifies
the energy balance to
∆ε12(R,k1,k2, k̄1, k̄2, t;A) ≡
≡ ε(R,k1, t;A) + ε(R,k2, t;A) − ε(R, k̄1, t;A) − ε(R, k̄2, t;A) =
= ∆E12
[
1 +
τ 2
4me
∇2V eff
I (R, t)
]
, (120)
where ∆E12 = (k2
1 + k2
2 − k̄2
1 − k̄2
2)/2me. We see that the difference of quasiparticle
energies (120) contains, in addition to the difference of kinetic energies of the particle
pair, a term proportional to the local curvature of the effective confinement potential.
To verify if this has a consequence on the energy balance in a two-particle collision
we consider the Markov limit of the collision integral in Born approximation, equa-
tion (118): neglecting the time dependence of the distribution functions compared
to the correlation time and extending the t̄−integration to infinity the integration
can be performed using (96). As a result the energy kernel of the collision integral
becomes
π(4me)
1/3
[
∆E12∇2V eff
I
]Ai
[
(
4me(∆E12)
2
∇2V eff
I
)1/3
]
. (121)
This function is singular at ∆E12 = 0, i.e. the dominant spectral weight falls on
scattering processes which conserve the kinetic energy of the particle pair, as in the
case of a homogeneous system (in the Markov limit). The latter case is recovered
by the limit ∇V eff
I → 0 and leads to the familiar result πδ(∆E12). A larger effect
of the inhomogeneity on the scattering process occurs on short time scales of the
517
M.Bonitz, J.W.Dufty
order of the correlation time where the Markov limit fails, e.g. [2]. Then the energy
broadening arising from the finite collision duration is additionally increased due to
the inhomogeneity of the confinement field.
Aside from electron-electron scattering, there are numerous physical situations
where the presence of an inhomogeneous field will have an even more pronounced
effect on the microscopic scattering probability. The most important one is inelastic
scattering. Indeed if particles, after the collision appear in a different quantum state
(energy level or band) with a different dispersion (effective mass), the quantum
potential will be different before and after the collision, even if the external field
is independent of the quantum state. This effect should be directly observable in
confined quantum systems undergoing e.g. collisional excitation or ionization.
4. Discussion
In this paper a gauge-invariant nonequilibrium Green’s functions theory for weak-
ly inhomogeneous systems has been developed. Weak inhomogeneity covers a broad
class of many-particle systems of current interest, including electrons in quantum
dots, ultracold ions in traps, valence electrons in metal clusters and so on. These sys-
tems are conveniently treated within the Wigner representation by a direct extension
of the quantum kinetic theory for spatially homogeneous systems by including spatial
gradient corrections. Here, we derived the corresponding Kadanoff-Baym/Keldysh
equations for the two-time correlation functions including an arbitrary (homoge-
neous) strong time-dependent electromagnetic field and a weakly inhomogeneous
confinement potential. From the KBE the gauge-invariant equations for the Wigner
distribution and for the retarded Green’s function (propagator) have been obtained.
Special attention has been devoted to an analysis of the spectral properties of
the electrons. Introducing the concept of an effective quantum potential which re-
places the effective (mean-field) potential V eff
I , the spectral function of electrons in
a combined electromagnetic field and confinement field has been derived. Its main
characteristics is that, even without correlations, the spectral function is broadened
and blue-shifted as a result of kinetic energy gain of the electrons from the electro-
magnetic field as well as from spatial localization (wave function compression) in the
confinement field. This has direct consequences for scattering processes and modifies
the energy balance and thus, the collision cross section and scattering rates.
Further applications of the results of the present theory to solutions of the quan-
tum kinetic equation in linear response, to the dielectric properties and to solutions
in a strong field are straightforward and will be presented in a separate publica-
tion [11].
518
Quantum kinetic theory of metal clusters
Appendix: Gauge invariant Kadanoff-Baym/Keldysh equatio ns
for correlated electrons in a strong electromagnetic field a nd
inhomogeneous external potential
In this appendix we present details of the derivation of the gauge-invariant
Kadanoff-Baym/Keldysh equations (KBE) in Wigner representation, the equations
for the Wigner function f(R,k, t) and for the retarded and advanced Green’s func-
tion (propagators) gR/A(R,k, t, t′). The derivation is kept as general as possible. In
particular, we allow for an arbitrary single-particle energy dispersion ε(k), thus, mak-
ing the results applicable to electrons in condensed matter systems as well. The re-
sults of the main text follow by using the limit of parabolic dispersion ε(k) = k2/2me.
The KBE in the presence of a strong homogeneous electric field and an inhomo-
geneous confinement potential VI read, in coordinate space,
[
i~
∂
∂t1
− ε
(
~
i
∇1 +
e0
c
A(1)
)
− VI(1)
]
g≷(1, 1′) =
∫
dr̄1 ΣHF(1, r̄1t1)g
≷(r̄1t1, 1
′)
+
∫ t1
t0
d1̄ [Σ>(1, 1̄) − Σ<(1, 1̄)] g≷(1̄, 1′) −
∫ t′1
t0
d1̄ Σ≷(1, 1̄) [g>(1̄, 1′) − g<(1̄, 1′)] ,
(122)
where we denoted 1 ≡ (r1, t1, s
3
1). We now derive the gauge-invariant Wigner repre-
sentation of this equation. For this we introduce macroscopic and microscopic length
scales, R ≡ (r1+r1)/2 and r ≡ r1−r2, where the first is determined by the geometry
of the confinement potential VI . Then, the gradient in the momentum operator in
equation (50) becomes ∇r1 = 1/2∇R + ∇r. We now apply the spatial part of the
transform (58) to the l.h.s. of equation (122),
∫
d3re
− i
~
r
k−
e0
c
t1
∫
t′1
dt̄
A(t̄)
t1−t′
1
[
i~
∂
∂t1
− ε
(
~
i
∇r +
1
2
~
i
∇R +
e0
c
A(1)
)
− VI(1)
]
×
∫
d3k1
(2π~)3
e
i
~
r
k1−
e0
c
t1
∫
t′1
dt̄
A(t̄)
t1−t′
1
g̃≷(R,k1, t1, t
′
1), (123)
where the tilde denotes the Wigner transformed functions. Performing the t1 and r
derivatives on the second exponential and noticing that the vector potential terms
in the exponentials cancel, we obtain from equation (123)
∫
d3r
∫
d3k1
(2π~)3
e−
i
~
r(k−k1)
{[
i~
∂
∂t1
− r
KA(t1, t
′
1)
t1 − t′1
− VI(1)
]
− ε
[
1
2
~
i
∇R + k1 −KA(t1, t
′
1)
]}
g̃≷(R,k1, t1, t
′
1), (124)
where we introduced the definition
KA(t, t′) ≡ −e0
c
∫ t
t′
dt′′
A(t) −A(t′′)
t− t′
(125)
519
M.Bonitz, J.W.Dufty
with the properties KA(t, t) = 0 and
KA(t, t′) − KA(t′, t) = −e0
c
{A(t) − A(t′)} = e0
∫ t
t′
dt′′ E(t′′). (126)
Further, we note the relation of KA to the momentum gain and displacement of a
free particle in the electromagnetic field, within the time interval [t′, t],
QA(t, t′) = KA(t, t′) − KA(t′, t), (127)
RA(t, t′) = − 1
me
KA(t, t′)(t− t′). (128)
Using r = ~/i∇k1 , the r−integration can be performed16, giving k1 = k, ex-
cept for the potential term where the center of mass and relative coordinates, in
general, do not separate. We obtain the gauge-invariant kinetic equation in Wigner
representation
{ [
i~
∂
∂t1
+ i~
KA(t1, t
′
1)
t1 − t′1
∇k
]
− ε
[
1
2
~
i
∇R + k + KA(t1, t
′
1)
]}
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
)
g̃≷(R,k1, t1, t
′
1) =
= Ĩ≷
F (R,k, t1, t
′
1) + Ĩ≷(R,k, t1, t
′
1), (129)
where the Hartree mean field is local (momentum independent) and, therefore, can
be included into the effective confinement potential,
V eff
I (R, t) ≡ VI(R, t) + ΣH(R, t), (130)
and ΣH(R, t) is related to the definition (51) by
ΣH(r1, t1; r
′
1, t
′
1) = δ(r1 − r′1)δ(t1 − t′1)Σ
H(r1, t1),
ΣH(r1, t1) =
∫
d3r2V (r1 − r2)f(r2, r2, t) =
∫
d3r2V (r1 − r2)n(r2, t).
The r.h.s. of equation (129) comprises the remaining two-particle contributions and
all higher order correlations – the exchange (“F”) and correlation (collision, denoted
by ”≷”) terms.
To obtain the equation for the single-time distribution function and for the prop-
agator, we also need the adjoint equation,
{ [
−i~
∂
∂t′1
+ i~
KA(t′1, t1)
t1 − t′1
∇k
]
− ε
[
−1
2
~
i
∇R + k + KA(t′1, t1)
]}
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
e+ i
~
r(k−k1) VI
(
R +
r
2
)
g̃≷(R,k1, t1, t
′
1) =
= −
[
Ĩ≷(R,k, t′1, t1)
]∗
−
[
Ĩ≷
F (R,k, t′1, t1)
]∗
, (131)
16The r−integration can be extended to infinity as the electron density outside the cluster is
zero.
520
Quantum kinetic theory of metal clusters
where in the final expression we exchanged t1 ↔ t′1 and used the property [g≷(t, t′)]∗ =
−g≷(t′, t).
Consider now the exchange term Ĩ≷
F . The gauge-invariant Fourier transform of
the coordinate space definition (51) is readily computed. Due to the locality in time
of ΣF
ΣF(r1, t1; r
′
1, t
′
1) = δ(t1 − t′1)Σ
F(r1, r
′
1, t1),
ΣF(r1, r
′
1, t) = −V (r1 − r′1)f(r1, r
′
1, t), (132)
the external field drops out of the Fourier transform (KA(t, t) = 0),
ΣF(R,k, t) = −
∫
d3k2
(2π~)3
V (k2)f(R,k − k2, t). (133)
Thus the result for Ĩ≷
F , up to first order gradient corrections, is given by
Ĩ≷
F (R,k, t, t′) =
[
1 +
i~
2
(∇R1∇Q2 −∇R2∇Q1)
]
ΣF (R1,Q1, t) g
≷(R2,Q2, t, t
′)
∣
∣
∣
(R,k)
=
∫
d3k1
(2π~)3
V (k1)
{
∇R1f(R1,Q1 − k1, t)∇Q2f(R2,Q2, t)
−∇Q1f(R1,Q1 − k1, t)∇R2f(R2,Q2, t)
}
∣
∣
∣
(R,k)
, (134)
where we introduced the short notation (R,k) to indicate that, after differentiation,
one has to set R1 = R2 = R and Q1 = Q2 = k.
The corresponding term in the adjoint equation (131) is
−
[
Ĩ≷
F (R,k, t′, t)
]∗
= −
[
1 − i~
2
(∇R1∇Q2 −∇R2∇Q1)
]
ΣF (R1,Q1, t
′)
×
[
g≷(R2,Q2, t
′, t)
]∗
∣
∣
∣
(R,k)
. (135)
In the kinetic equation for the Wigner function the Fock terms appear in the com-
bination
IF(R,k, t) = −2Re[ΣF(R,k, t)g<(R,k, t, t)], (136)
which vanishes exactly (on the time-diagonal, g< is purely imaginary). This means
that the only nonzero exchange contributions arise from the gradient corrections
which appear due to the product ΣF · g<. Using the general expression (45), we
obtain
IF(R,k, t) = 2Re
(
1 +
i~
2
{. . .}
)
ΣF (R1,Q1, t) g
<(R2,Q2, t, t)
∣
∣
∣
(R,k)
= − (∇R1∇Q2 −∇R2∇Q1) ΣF (R1,Q1, t) f(R2,Q2, t)
∣
∣
∣
(R,k)
=
∫
d3k1
(2π~)3
V (k1)
{
∇R1f(R1,Q1 − k1, t)∇Q2f(R2,Q2, t)
−∇Q1f(R1,Q1 − k1, t)∇R2f(R2,Q2, t)
}
∣
∣
∣
(R,k)
. (137)
521
M.Bonitz, J.W.Dufty
Now we calculate the sum and difference of the two Kadanoff-Baym/Keldysh
equations in Wigner representation, equations (129), (131)
{
i~
[
∂
∂t1
∓ ∂
∂t′1
]
+ 2i~
KA
±(t1, t
′
1)
t1 − t′1
∇k − ε
[
1
2
~
i
∇R + k + KA(t1, t
′
1)
]
∓ ε
[
−1
2
~
i
∇R + k + KA(t′1, t1)
]}
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
{
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
, t1
)
± e
i
~
r(k−k1) V eff
I
(
R − r
2
, t′1
)}
× g̃≷(R,k1, t1, t
′
1) = I≷
∓ (R,k, t1, t
′
1), (138)
where I≷
± (R,k, t1, t
′
1) ≡ Ĩ≷(R,k, t1, t
′
1)±
[
Ĩ≷(R,k, t′1, t1)
]∗
, and KA
± are short nota-
tions for the sum and difference of the field-induced momenta
KA
±(t1, t
′
1) ≡
1
2
[
KA(t1, t
′
1) ±KA(t′1, t1)
]
. (139)
To obtain the equation for the Wigner distribution, we need the difference (lower
signs) and, for simplification of the notation, we, from now on, specify to a parabolic
energy dispersion, ε(k) → k2/2m:
{
i~
∂
∂t
+ 2i~
KA
−(t1, t
′
1)
t1 − t′1
∇k −
~
i
k
me
∇R − 1
2me
[
(KA)2(t1, t
′
1) − (KA)2(t′1, t1)
+ 4kKA
−(t1, t
′
1) −
2~
i
KA
+(t1, t
′
1)∇R
]
}
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
{
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
, t1
)
− e
i
~
r(k−k1) V eff
I
(
R − r
2
, t′1
)}
× g̃≷(R,k1, t1, t
′
1) = I≷
+ (R,k, t1, t
′
1). (140)
Finally, the kinetic equation for the Wigner function f is obtained by taking equation
(140) for the function g< at equal times, t1 = t′1 = t, and further expressing g< by
f , using relation (47)
{
∂
∂t
+
k
me
∇R − e0E(t)∇k
}
f(R,k, t) (141)
+
2
~
∫
d3r
∫
d3k1
(2π~)3
sin
r (k − k1)
~
V eff
I
(
R +
r
2
, t
)
f(R,k1, t) = I(R,k, t),
where I(R,k, t) ≡ −2Re[Ĩ<(R,k, t, t)] = −I<
+ (R,k, t, t).
Further, to compute the propagators (retarded and advanced Green’s functions)
and the spectral function, we consider the sum of the two equations (129,131), i.e.
the upper sign in equation (138). Using again a parabolic dispersion, we obtain
{
i~
∂
∂τ
+ i~
KA
+(t1, t
′
1)
t1 − t′1
∇k −
1
2me
[(
~
2i
∇R
)2
+ k2 +
(KA)2(t1, t
′
1) + (KA)2(t′1, t1)
2
522
Quantum kinetic theory of metal clusters
+ 2kKA
+(t1, t
′
1) +
~
i
KA
−(t1, t
′
1)∇R
]}
g̃≷(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
{
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
, t1
)
+ e
i
~
r(k−k1) V eff
I
(
R − r
2
, t′1
)}
× g̃≷(R,k1, t1, t
′
1) = I≷
− (R,k, t1, t
′
1). (142)
What is left in order to obtain the equation for gR/A is to subtract the two
equations (142) for g> and g<, and to add the singular term δ(t1 − t′1) on the right
and hand side, cf. definition (54),
{
i~
∂
∂τ
+ i~
KA
+(t1, t
′
1)
t1 − t′1
∇k − 1
2me
[(
~
2i
∇R
)2
+ k2 +
(KA)2(t1, t
′
1) + (KA)2(t′1, t1)
2
+ 2kKA
+(t1, t
′
1) +
~
i
KA
−(t1, t
′
1)∇R
]}
g̃R/A(R,k, t1, t
′
1)
−
∫
d3r
∫
d3k1
(2π~)3
{
e−
i
~
r(k−k1) V eff
I
(
R +
r
2
, t1
)
+ e
i
~
r(k−k1) V eff
I
(
R − r
2
, t′1
)}
× g̃R/A(R,k1, t1, t
′
1) = I
R/A
− (R,k, t1, t
′
1), (143)
where the definition of I
R/A
− is analogous to that of I≷
− .
5. Acknowledgements
MB is grateful to the Physics Department of the University of Florida, Gainesville,
where the bulk of this work has been performed, for its hospitality. We acknowl-
edge discussions with Thomas Fennel (Rostock). This work has been supported
by the Deutsche Forschungsgemeinschaft via SFB 198/B10, the National Science
Foundation and the Department of Energy (grants DE FG03–98DP00218 and DE
FG02ER54677).
523
M.Bonitz, J.W.Dufty
References
1. Kraeft W.D., Kremp D., Ebeling W., Röpke G. Quantum Statistics of Charged Particle
Systems. Berlin, Akademieverlag, 1986.
2. Bonitz M. Quantum Kinetic Theory. Stuttgart/Leipzig, Teubner, 1998.
3. Kadanoff L.P., Baym G. Quantum Statistical Mechanics, 2nd ed. Addison-Wesley,
1989.
4. Klimontovich Yu.L. Kinetic Theory of Nonideal Gases and Nonideal Plasmas. Oxford,
Pergamon Press, 1982, (Russian edition: Moscow, Nauka, 1975).
5. Klimontovich Yu.L. Statistical Physics. Chur, Harwood Acad. Publ., 1986, (Russian
edition: Moscow, Nauka, 1982).
6. Haug H., Jauho A.P. Quantum Kinetics in Transport and Optics of Semiconductors.
New York, Springer-Verlag, Heidelberg, 1996.
7. S. Fujita Introduction to Nonequilibrium Quantum Statistical Mechanics Philadel-
phia, London, W.B. Saunders Company, 1966.
8. Kremp D., Bornath Th., Bonitz M., Schlanges M. // Phys. Rev. E, 1999, vol. 60,
p. 4725.
9. Bonitz M., Bornath Th., Kremp D., Schlanges M., Kraeft W.D. // Contrib. Plasma
Phys., 1999, vol. 39, p. 329.
10. Bonitz M., Dufty J.W. // Superlattices and Microstructures, 2004, vol. 34, p. 225.
11. Bonitz M., Dufty J.W., to be published
12. Mittelman M.H. Introduction to the theory of laser-atom interactions. New York,
Plenum Press, 1982.
13. Brack M. // Rev. Mod. Phys., 1993, vol. 65, p. 677.
14. de Heer W.A. // Rev. Mod. Phys., 1993, vol. 65, p. 611.
15. Ichimaru S., Plasma Physics, Benjamin/Cummings Publ. Co. Inc. 1986
16. Salpeter E.E. // Aust. J. Phys., 1954, vol. 7, p. 373.
17. Pines D. Elementary Excitations in Solids. New York, Benjamin, 1963.
18. Bertsch G.F., Yabana K. Density Functional Theory. Introduction to computational
methods for many-body systems, (ed. Bonitz M.). Princeton, Rinton Press, to be
published.
19. Dirac P.A.M. // Proc. Cambridge Philos. Soc., 1930, vol. 26, p. 376.
20. Semkat D., Kremp D., Bonitz M. // Phys. Rev. E, 1999, vol. 59, p. 1557.
21. Semkat D., Kremp D., Bonitz M. // J. Math. Phys., 2000, vol. 41, p. 7458.
22. Bonitz M. et al. // J. Phys. Cond. Matt., 1996, vol. 8, p. 6075.
23. Bonitz M., Dufty J.W., submitted for publication
24. Kleinert H., Path Integrals in Quantum Mechanics, Statistics and Polymer Physics.
World Scientific, Second edition, 1995.
25. Ferry D.K., Zhou J.-R. // Phys. Rev. B, 1993, vol. 48, p. 7944.
26. Vasileska D., Akis R., Knezevic I., Milicic S.N., Ahmed S.S., Ferry D.K. // Microlec-
tronic Engineering, 2002, vol. 63, p. 233.
27. Domps A., Reinhard P.-G., Suraud E. // Ann. Phys. (N.Y.), 2000, vol. 280, p. 211.
28. Calvayrac F., Reinhard P.-G., Suraud E., Ullrich C.A. // Phys. Rep., 2000, vol. 337,
p. 493.
29. Kwong N.H., Bonitz M. // Phys. Rev. Lett., 2000, vol. 84, p. 1768.
30. Schäfer W., M. Wegener, Semiconductor Optics and Transport Phenomena, Springer
2002
524
Quantum kinetic theory of metal clusters
31. Bonitz M., Semkat D., Haug H. // Europ. Phys. J. B, 1999, vol. 9, p. 309.
32. Kwong N.H., Bonitz M., Binder R., Köhler H.S. // Phys. Stat. Sol. B, 1998, vol. 206,
p. 197.
33. Lipavsky P., Spicka V., Velicky B. // Phys. Rev. B, 1986, vol. 34, p. 6933.
34. Filinov A., Bonitz M., Ebeling W. // J. Phys. A: Math. Gen., 2003, vol. 36, p. 5957.
35. Filinov A., Golubnychiy V., Bonitz M., Ebeling W., Dufty J.W., accepted for publi-
cation in Phys. Rev. E, 2004.
Квантова кінетична теорія металічних кластерів в
потужному електромагнітному полі
М.Боніц 1,2 , Д.В.Дафті 2
1 Інститут теоретичної фізики та астрофізики
Кілльського університету Крістіана Альбрехта,
Лейбніцштрассе 15, 24098 Кілль, Німеччина
2 Фізичний факультет університету Флориди,
Гейнсвілл, FL 32611-8440
Отримано 30 березня 2004 р.
На основі нерівноважних функцій Гріна розроблено квантову теорію
слабонеоднорідної системи заряджених частинок. Отримані резуль-
тати мають суттєве значення для опису валентних електронів мета-
лічних кластерів, а також для замкнутих кулонівських систем, таких
як електрони в квантових точках чи ультрахолодні іони в пастках. Зо-
крема, в даній роботі ми зупиняємося на розгляді металічних клас-
терів, хоча результати можна безпосередньо узагальнити на інші ви-
падки. Ми розпочинаємо з введення в фізику скорельованих вален-
тних електронів у металічних кластерах в сильних електромагнітних
полях. Після короткого огляду моделі “желе” та стандартного методу
функціоналу густини для опису основних станів ми зосереджуємо-
ся на узагальненні теорії на випадок нерівноважних процесів. З цією
метою розроблено узагальнену калібрувально-інваріантну кінетичну
теорію. Ці результати включають рівняння руху для двочасових ко-
реляційних функцій, рівняння для функцій Вігнера та аналіз спект-
ральної функції. Вводиться поняття ефективного квантового потен-
ціалу, який залишає незмінною локальну форму пропагаторів. Це
дає нам змогу отримати явні вирази для спектральної функції елек-
тронів при суперпозиції сильного електромагнітного поля та слабо-
неоднорідного утримуючого потенціалу.
Ключові слова: квантова статистична механіка, заряджені
кластери, взаємодія лазер-плазма
PACS: 05.30.-d,36.40.Wa,52.38.-r
525
526
|