ArticlePDF Available

Hund's Rules and Spin Density Waves in Quantum Dots

Authors:

Abstract and Figures

Spin density functional theory is used to calculate the ground state electronic structures of circular parabolic quantum dots. We find that such dots either have a spin configuration determined by Hund's rule or make a spin-density-wave-like state with zero total spin. The dependence of the spin-density-wave amplitudes on the density of the two-dimensional electron gas is studied.
Content may be subject to copyright.
VOLUME 79, NUMBER 7 PHYSICAL REVIEW LETTERS 18AUGUST 1997
Hund’s Rules and Spin Density Waves in Quantum Dots
M. Koskinen and M. Manninen
Department of Physics, University of Jyväskylä, P.O. Box 35, FIN-40351 Jyväskylä, Finland
S.M. Reimann
Niels Bohr Institute, DK-2100 Copenhagen, Denmark
(Received 17 March 1997)
Spin density functional theory is used to calculate the ground state electronic structures of circular
parabolic quantum dots. We find that such dots either have a spin configuration determined by
Hund’s rule or make a spin-density-wave-like state with zero total spin. The dependence of the
spin-density-wave amplitudes on the density of the two-dimensional electron gas is studied. [S0031-
9007(97)03740-X]
PACS numbers: 75.30.Fv, 71.10.Ca, 73.20.Dx
Semiconductor technology now allows the fabrication
of quantum dots being so tiny that they contain only a
few electrons. Usually, such dots are formed by lateral
confinement of a high-mobility two-dimensional electron
gas (2DEG) in a semiconductor heterostructure. Their
electronic properties are determined by the interplay of
the external confinement and the electron-electron inter-
actions, manifesting a quantum-mechanical many-particle
problem (see, e.g., Ref. [1] for exact diagonalization stud-
ies and Ref. [2] for a mean-field approach, to mention
only a few from a broad field of research). The proper-
ties of such small dots strongly depend on the number of
confined electrons, and the situation is quite similar to the
different properties of the first elements of the periodic
table, why quantum dots now often are called “artificial
atoms” [3].
Recently, Tarucha et al. [4] have developed a vertically
confined quantum dot, where they could experimentally
show that for weak or zero magnetic fields the electronic
structure of a small circular dot containing up to 20
electrons is mostly determined by the subsequent filling
of shells, obeying Hund’s rules as in atoms.
Motivated by their experimental study, we performed
spin-density functional calculations for such circular, par-
abolic quantum dots containing up to N 46 electrons.
Complicated magnetic structures and excited states of
the 2DEG have been obtained earlier in the presence of
an external magnetic field [57].
To our own surprise we found that quantum dots have
a rich variety of different magnetic structures in the
ground state, even without an external magnetic field.
As one would expect from the knowledge of atomic
physics, Hund’s first rule dominates for the smallest sizes.
However, some dots have zero total spin, but exhibit
a space-dependent spin polarization, a so-called spin-
density wave (SDW) [8].
At very low temperatures the electrons in the 2DEG
are confined to the lowest subband, and it is thus suf-
ficient to consider them as being bound laterally in the
x-y plane. For such small dots as studied here, we then
make the frequently used [1,2] approximation that the
external potential is harmonic, V m
p
v
2
sx
2
1 y
2
dy2.
In such a parabolic dot the single-particle electron lev-
els form a pronounced gross-shell structure, the magic
shells corresponding to electron numbers 2, 6, 12, 20,
30,.... The exact degeneracy of the shells of the 2D
oscillator, however, is reduced by the electron-electron
interactions. The situation in the case of open shells
is analogous to that in atoms. For example, if a shell
is half filled, the spins align according to Hund’s rules.
This causes the empty states with opposite spin to be
higher in energy and removes the degeneracy of the Fermi
level. In the case of atoms the ionization potential as a
function of the atomic number has maxima at half-filled
p shells. The experiments of Tarucha et al. [4] for para-
bolic quantum dots show a quite similar behavior: A half-
filled shell shows a maximum in the addition energies as a
manifestation of Hund’s first rule.
For the electronic structure calculations for N electrons
in the parabolic dot, we apply density functional theory
[9,10] and treat the exchange-correlation part of the
electron-electron interactions in the local spin-density [11]
approximation. To be more specific, we solve the single-
particle Kohn-Sham [10] equations
2
¯h
2
2m
p
=
2
x
1 V
s
eff
sxd
c
i,s
sxd e
i,s
c
i,s
sxd , (1)
where x sx, yd and the index s accounts for the spin
(" or #). The effective Kohn-Sham potential consists of
the external harmonic confinement, the Hartree potential
of the electrons, and the functional derivative of the local
exchange-correlation energy
E
xc
Z
dx nsxde
xc
sssnsxd, z sxdddd , (2)
where n is the electron density and z sn
"
2 n
#
dyn the
spin polarization. For the exchange-correlation energy of
the homogeneous 2D electron gas, we use the parame-
trized form of Tanatar and Ceperley [12] for nonpolarized
0031-9007y97y79(7)y1389(4)$10.00 © 1997 The American Physical Society 1389
VOLUME 79, NUMBER 7 PHYSICAL REVIEW LETTERS 18AUGUST 1997
(z 0) and ferromagnetic (z 1) cases. For interme-
diate polarizations, following the work of von Barth and
Hedin [11] as well as Perdew and Zunger [13], which is
frequently used for electronic structure calculations in 3D
systems, one can write
e
xc
sn, z d e
xc
sn,0d 1 fszdfe
xc
sn,1d 2e
xc
sn,0dg .
(3)
The polarization dependence fsz d in 2D [14] is then
fsz d
s1 1zd
3y2
1s12zd
3y2
22
2
3y2
22
. (4)
In order to obtain the electron densities which minimize
the total energy functional Efn
"
, n
#
g, the Kohn-Sham
equations are solved self-consistently. To avoid any
symmetry restrictions for the wave functions, we use a
plane-wave basis. To find the ground state of all the
possible spin configurations, the iterative solution of the
Kohn-Sham equations was started with different forms
and depths of the initial potential for the spin up and down
densities [15]. This assures that one is not trapped in a
local minimum, but with a high probability can separate
the electronic ground-state configuration from the lowest
excited states.
The results will be given in effective atomic units with
Ryp m
p
e
4
yh
2
s4pe
0
ed
2
and a
p
B
¯h
2
s4pe
0
edym
p
e
2
,
where m
p
is the effective mass and e the dielectric
constant. The results can then be scaled to the actual
values for typical semiconductor materials.
The calculations are done for different values of the
density parameter r
s
, which approximately correspond to
the average particle density in the dot, n
0
1yspr
2
s
d.
For the external parabolic confinement, which actually
determines the average particle density, we then use v
2
e
2
ys4pe
0
em
p
r
3
s
p
Nd.
We have first calculated ground-state and isomeric elec-
tronic structures for quantum dots corresponding to the
equilibrium density of the two-dimensional electron gas
with r
s
1.51a
p
B
. The results for dots with even elec-
tron numbers N are summarized [16] in Table I. Dots
with 2, 6, 12, 20, and 30 electrons correspond to “magic”
configurations of a 2D harmonic well and have a par-
ticularly large Fermi gap. Consequently, the system is
fully paramagnetic and the ground-state total spin in all
these cases is zero. For nonclosed shells, however, in
most of the cases the total spin [17] is determined by
Hund’s rule, which maximizes the spin for orbital de-
generacy. From Table I we see that dots with N
4, 8, 10, 14, 18, 22, 28, and 32 electrons have total spin
S 1 in the ground state, whereas for dots with 16 and
26 electrons ground states with a total spin S 2 were
found. For larger dots we found up to three different “spin
isomers,” being little higher in energy than the ground-
state configuration. As an example, we show in the first
row of Fig. 1 the spin down and spin up ground-state
TABLE I. Total spin S of the ground states and some low-
energy spin isomers for r
s
1.51a
p
B
. States with a S 0 spin-
density wave are labeled SDW, whereas nonzero total spins
according to Hund’s rules are labeled with H.
Number of Ground Excited DE
electrons state state(s) (mRy*)
20
41H
60
8 1 H 0 SDW 23.0
10 1 H 0 SDW 4.62
12 0
14 1 H
16 2 H 0 SDW 18.1
1 H 20.2
0 SDW 20.7
18 1 H 0 SDW 11.05
20 0
22 1 H 0 SDW 9.06
24 0 SDW 2 H 0.72
26 2 H 0 SDW 2.75
0 SDW 6.46
28 1 H 0 SDW 1.98
30 0
32 1 H 0 SDW 7.06
34 0 SDW 2 H 0.73
FIG. 1. Spin down and spin up densities n
#
, n
"
and normalized
polarization
˜
z sx, yd for the ground state (first row) and excited
states (lower rows) for a dot with N 16 electrons and
r
s
1.51a
p
B
. The maximum amplitude for the polarization
corresponds to
˜
z sx, y0.23.
1390
VOLUME 79, NUMBER 7 PHYSICAL REVIEW LETTERS 18AUGUST 1997
densities for N 16, together with the normalized po-
larization
˜
z sn
"
2 n
#
dyn
0
. In this case the total spin
is S 2. It is intriguing to see that
˜
z shows a pro-
nounced radial oscillation, which means that the excess
spin is not homogeneously distributed over the whole dot
regime. This effect was also seen for other sizes N when
the ground state had nonzero total spin. It reminds one of
the spin inversion states found by Gudmundsson et al. [6]
for finite magnetic fields.
For many of the excited states at r
s
1.51a
p
B
the total
spin is zero. A priori one would in these cases expect that
the system is fully unpolarized. But now looking again
at Fig. 1, where the three lower rows show the densities
and polarizations of the three different spin isomers for
N 16, we see that the electronic structure is more
complicated. Both the second and fourth rows of Fig. 1
show isomeric S 0 states, being slightly higher in energy
than the ground state (see Table I). Their space-dependent
spin polarization shows apparent spatial oscillations, which
remind one of the phenomenon of spin-density waves [8] in
the bulk. The spin-density-wave-like states inthe finite 2D
system (which in the following are called “SDW states”)
associate a certain preferred spin direction with a given
spatial region in the dot.
Studying these effects further, the next surprise is that
in larger dots this rather peculiar electronic structure
with total spin S 0 gets even lower in energy than
the Hund state. For N 24 and N 34, the ground
state has spin zero, but is associated with a SDW state
which has a lower energy than the uniform state. (Its
shape is similar to the one shown in Fig. 3 below for
r
s
5.0a
p
B
.) The general possibility of such ground-state
configurations with uniform or nearly uniform electron
density, but a nonuniform density of spin magnetization,
was first discussed by Overhauser [8]. He stated that
the nonmagnetic state must become unstable with respect
to SDW formation at low densities, whereas at higher,
metallic densities, it seems unlikely that the SDW state
is of enhanced stability. The energy balance, however, is
very delicate. Overhauser pointed out [8] that when the
SDW states are stable, they are energetically only slightly
lower than the nonmagnetic state.
On the basis of the results discussed above it is now
interesting to study how the SDW states depend on the
density of the two-dimensional electron gas. We thus
calculated for selected electron numbers the ground state
as a function of the strength v of the confining harmonic
potential, corresponding to r
s
values from 0.25a
p
B
to 5.0a
p
B
.
Figure 2 shows the maximum amplitude of the nor-
malized polarization
˜
z as a function of r
s
for dots with
N 24, 34, and 46 electrons, where the SDW was found
to be the ground state.
It can be clearly seen that, depending on the size N,
there is a critical value of r
s
, where the SDW sets in,
and then rises its amplitude with increasing r
s
. Note,
however, that the N dependence of the critical value of
FIG. 2. Maximum amplitudes of
˜
z sx, yd of the ground-state
SDW for N 24, 34, and 46 as a function of r
s
.
r
s
does not generally follow the trend suggested from
this figure: For some N , 46, the SDW sets in at
a higher critical value of r
s
, but with a comparable
amplitude. From our calculations we find that even in
magic configurations, which are fully paramagnetic with
total spin zero at low r
s
values, for sufficiently large r
s
the SDW formation sets in. In the case of the filled shells
N 12, 20, and 30 a SDW state was obtained for very
large r
s
* 5a
p
B
. Increasing r
s
still further, this SDW state
gets more pronounced, in a similar way than the examples
with N 24, 34, and 46 discussed above.
Finally, we note that the occurrence of a SDW is
related to a change in the pattern of the single-particle
levels. As an example, we show in the bottom of Fig. 3
the Kohn-Sham single-particle spectra of N 34 at a
density corresponding to r
s
5.0a
p
B
, both for the ground-
state SDW with S 0 and the excited state with total
spin S 2, being 1.9 mRy
p
higher in energy. The third
spectrum shows the result of a LDA calculation for the
unpolarized case, which gives a 2.13 mRy
p
higher energy
than the SDW state. Note that for a SDW, there is always
one pair of degenerate spin up and spin down orbitals, and
the Fermi gap of the SDW state is much higher than in the
S 2 case or within LDA. In quite a similar way to the
spontaneous shape deformations of nuclei with nonclosed
shells [18], the SDW opens a large energy gap at the
Fermi surface, leading to a more stabilized electronic
structure in the dot.
In conclusion, we have found that in finite quantum
dots a static spin-density-wave-like state occurs even at
rather high densities of the two-dimensional electron gas.
For many open-shell systems we found that the SDW
state with zero spin has a higher energy than the ground
state. In other cases the SDW is the ground state. Also
for nonzero total spin a strong spatial dependence of the
spin polarization was found. The amplitude of the SDW
strongly depends on the size of the Fermi gap: The
1391
VOLUME 79, NUMBER 7 PHYSICAL REVIEW LETTERS 18AUGUST 1997
FIG. 3. Upper part: Polarization
˜
z sx, yd for N 34 at r
s
5a
p
B
. Left: Ground state with S 0. Right: Excited state
with S 2. The scales are the same as in Fig. 1 above.
Lower part: Single-particle spectra for N 34 at the same
density r
s
5a
p
B
. The numbers in the spectra indicate the
degeneracies of the occupied single-particle states. The shorter
lines indicate the lowest unoccupied states. Left: Hund’s case
with S 2. Middle: SDW ground state with S 0. Right:
LDA result, unpolarized. The total width of the spectrum is
about 73.5 mRy
p
.
inset of the SDW-like state for the magic configurations
occurs at much higher r
s
values than for nonclosed
shells. Recent calculations have shown that the SDW-like
states are rather stable against distortions of the external
confinement. In a real quantum dot, the deformations
of the effective confinement (caused, for example, by
quantum point contacts) will thus lead to a pinning of the
polarization. Rather peculiar properties of such quantum
dots could be expected.
We would like to thank J. Helgesson, P. E. Lindelof,
and B. R. Mottelson for helpful discussions. This work
was partially financed by the Studienstiftung des deut-
schen Volkes, the BASF AG, the Academy of Finland,
and CNAST. S. M. R. thanks the University of Jyväskylä
for its hospitality.
[1] D. Pfannkuche, V. Gudmundsson, and P.A. Maksym,
Phys. Rev. B 47, 2244 (1993); P. Hawrylak and
D. Pfannkuche, Phys. Rev. Lett. 70, 485 (1993); P. A.
Maksym, Phys. Rev. B 53, 10871 (1996).
[2] A. Kumar, S.E. Laux, and F. Stern, Phys. Rev. B 42,
5166 (1990); V. Gudmundsson and R.R. Gerhardts, Phys.
Rev. B 43, 12098 (1991); T. Darnhofer, D.A. Broido, and
U. Rössler, Phys. Rev. B 50, 15412 (1994); M. Ferconi
and G. Vignale, Phys. Rev. B 50, 14722 (1994); M. Stopa,
Phys. Rev. B 53, 9595 (1996); H.-M. Mueller and S.E.
Koonin, Phys. Rev. B 54, 14532 (1996).
[3] M. A. Kastner, Rev. Mod. Phys. 64, 849 (1992); Phys.
Today 46, No. 1, 24 (1993); R. C. Ashoori, Nature
(London) 379, 413 (1996).
[4] S. Tarucha, D.G. Austing, T. Honda, R.J. van der Hage,
and L.P. Kouwenhoven, Phys. Rev. Lett. 77, 3613 (1996).
[5] L. Brey and B. I. Halperin, Phys. Rev. B 40, 11634
(1989); C.E. Hembree et al., Phys. Rev. B 48, 9162
(1993).
[6] V. Gudmundsson and G. Palsson, Phys. Scr. T54,92
(1994); V. Gudmundsson and J.J. Palacios, Phys. Rev.
B 52, 11 266 (1995).
[7] S. L. Sondhi, A. Karlhede, S.A. Kivelson, and E. H.
Rezayi, Phys. Rev. B 47, 16419 (1993); H.A. Fertig,
L. Brey, R. Cote, and A.H. McDonald, Phys. Rev. B 50,
11018 (1994).
[8] A. W. Overhauser, Phys. Rev. Lett. 4, 462 (1960); Phys.
Rev. 128, 1437 (1962); 167, 691 (1968).
[9] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).
[10] W. Kohn and L.J. Sham, Phys. Rev. 140, A1133 (1965).
[11] U. von Barth and L. Hedin, J. Phys. C 5, 1629 (1972).
[12] B. Tanatar and D.M. Ceperley, Phys. Rev. B 39, 5005
(1989).
[13] J. P. Perdew and A. Zunger, Phys. Rev. 23, 5048 (1981).
[14] M. I. Lublin, O. Heinonen, and M.D. Johnson, Report
No. Cond-Mat 9701079, 1997.
[15] The initial potentials for both spins were chosen to be of
square well type, with different depths V
",#
0
s1 6 hdV
0
.
Choosing h 0.0 and h 0.3 was sufficient to obtain
the unpolarized and partly polarized (Hund’s rules) cases.
Different radial forms R
",#
m
sfd R
0
f1 1ecossmfdg were
chosen for fm, eg f0, 0g, f3, 0.3g, f4, 0.4g, and f6, 0.6g,
where R
"
m
and R
#
m
are twisted against each other by an
angle u pym.
[16] Odd particle numbers were left out because in most of
the cases, one gets a net spin of 0.5 or 1.5, respectively,
corresponding either to the trivial case of one spin left
over in the highest orbital, or Hund’s rule, which also
occurs for even N.
[17] In this formalism, the wave function is a single Slater
determinant of the Kohn-Sham single-particle wave func-
tions. Only in cases where spin up and down space wave
functions are exactly the same, the Slater determinant is
the eigenstate of the
ˆ
S
2
operator with S S
z
, but gener-
ally this is only approximately true.
[18] Å. Bohr and B.R. Mottelson, Nuclear Structure (Ben-
jamin, New York, 1975), Vol. II.
1392
... Materials discovery is usually an ad hoc process guided by physical intuition, physical rules of thumb, and experimental data, among many other factors. For example, when designing the bandgap of binary semiconducting systems, one can be guided by Vegard's law [38], which says that the bandgap of the binary semiconductor alloy can be linearly interpolated between that of its constituent bandgaps; when designing magnetic materials, the magnetic ground state can be estimated using Hund's rule [39], which states that the valence electrons arrange to maximize the total spin and angular momentum quantum number; when searching for flat band materials, one typically seeks high-symmetric geometric structures, e.g., kagome or Lieb lattice [40][41][42]. ...
Article
Traditionally, materials discovery has been guided by basic physical rules, and such rules embody the basic understanding of the physical characteristics of interest of the material. However, the discovery of physical rules remains a challenging task due to the inherent difficulty in recognizing patterns in the high-dimensional and highly nonuniform distributed materials space. The standard data analytics approach using machine learning (ML) may fall short in producing meaningful results due to fundamental differences between the underlying assumptions and goals of ML vs materials discovery. ML is mainly focused on estimating complex black-box predictive models (that are nonlinear and multivariate), whereas in materials discovery, the goal is to come up with interpretable data-driven physical rules. Here, we attempt to tackle this problem by proposing a robust data analytics framework that allows us to derive basic physical rules from data. We introduce the concept of global and local modeling, utilizing both supervised and unsupervised learning, for highly nonuniformly distributed materials data. To enhance the model interpretation, we also introduce a model-independent interpretation technique to assist human experts in extracting useful physical rules. The proposed framework for extracting data-derived physical rules at the global and local level is illustrated using two case studies: (1) classification of van der Waals (vdW) and non-vdW (nvdW) materials and (2) classification of wide bandgap and non-wide bandgap vdW materials.
... In an external magnetic field it is more natural to replace the electron density with the local filling factor ν(r) = 2πl 2 n e (r) and the spin components thereof [42,43]. The exchange and correlation functional is then [44] B xc (ν, ζ) = ∞ xc (ν)e −f (ν) + 0 xc (ν, ζ)(1 − e −f (ν) ), (A2) where the interpolation between vanishing magnetic field and an infinite one depends on f (ν) = (3ν/2) + 7ν 4 and the high field limit is ∞ xc (ν) = −0.782 √ νe 2 /(κl). ...
Preprint
The orbital and spin magnetization of a cavity-embedded quantum dot array defined in a GaAs heterostructure are calculated within quantum-electrodynamical density-functional theory (QEDFT). To this end a gradient-based exchange-correlation functional recently employed for atomic systems is adapted to the hosting two-dimensional electron gas (2DEG) submitted to an external perpendicular homogeneous magnetic field. Numerical results reveal the polarizing effects of the cavity photon field on the electron charge distribution and nontrivial changes of the orbital magnetization. We discuss its intertwined dependence on the electron number in each dot, and on the electron-photon coupling strength. In particular, the calculated dispersion of the photon-dressed electron states around the Fermi energy as a function of the electron-photon coupling strength indicates the formation of magnetoplasmon-polaritons in the dots.
... However, its general applicability is often hindered by uncontrolled approximations used for constructing the approximate exchange-correlation functional. Notwithstanding its drawbacks, one-electron theories such as DFT in the Kohn-Sham framework of approximate exchange and correlation (XC) energy functionals [37][38][39][40] and the Hartree-Fock (HF) [15,[41][42][43][44][45][46][47][48][49][50] approximation often achieve a qualitatively correct agreement with experiment. In contrast to DFT calculations, Full Configuration Interaction (FCI) investigations of QDs yield exact results for a given basis set and have been applied to QDs in a number of studies [4,49,[51][52][53][54][55][56][57][58][59][60][61][62][63][64][65]. ...
Preprint
We present a study of the two dimensional circular quantum dot model Hamiltonian using a range of quantum chemical ab initio methods. Ground and excited state energies are computed on different levels of perturbation theories including the coupled cluster method. We outline a scheme to compute the required Coulomb integrals in real space and utilize a semi-analytic solution to the integral over the Coulomb kernel in the vicinity of the singularity. Furthermore, we show that the remaining basis set incompleteness error for two dimensional quantum dots scales with the inverse number of virtual orbitals, allowing us to extrapolate to the complete basis set limit energy. By varying the harmonic potential parameter we tune the correlation strength and investigate the predicted ground and excited state energies.
... Density functional theory methods with various flavours (local spin-density, current density functional theory, etc) represent another class of methods used to study semiconductor quantum dots [96][97][98][99]. ...
Chapter
Semiconductor quantum dots represent nanoscale systems with few electrons confined in a semiconductor host crystal. The importance of semiconductor quantum dots lies primarily in their tunability and sensitivity to external parameters as electrons are confined in all dimensions. The bulk of semiconductor quantum dots are fabricated by applying a lateral confinement potential to a two-dimensional electron gas. Quantum confinement profoundly affects the way electrons interact with each other and with external parameters, such as a magnetic field. Quantum confinement of electrons is just one of several ways quantum mechanics reveals itself. Another pure quantum phenomena associated with electrons is their spin. An external magnetic field affects both orbital and spin motion of electrons. External control of the full quantum wave function in a semiconductor quantum dot may lead to novel technological application involving both charge and spin. From a theoretical point of view, semiconductor quantum dots represent a unique opportunity to study fundamental quantum theories in a tunable atomic like setup. In this work, we review some of the theoretical approaches used to study two-dimensional few-electron semiconductor quantum dots. The main emphasis is to clarify the relations between different theories and methods for few-electron semiconductor quantum dots in an external parameter, a perpendicular magnetic field. Properties of few-electron semiconductor quantum dots in the weak magnetic regime are explained well through single-electron theory concepts. However, challenges do exist when considering stronger external magnetic fields. A strong magnetic field, when applied perpendicular to the quantum dot, changes the quantum nature of the electronic correlations and spin-polarizes the electrons. As the strength of the external magnetic field increases, the confined electrons start to manifest collective quantum behavior as seen in the integer and fractional quantum Hall effect regime. Theoretical and computational challenges to studies of semiconductor quantum dots as the magnetic field changes from weak to strong are reviewed. Specific examples are introduced to illustrate the transformation of the quantum wave function into a Laughlin-like one as the magnetic field increases.
Article
Singlet and triplet spin state energies for three-dimensional Hooke atoms, that is, electrons in a quadratic confinement, with even number of electrons (2, 4, 6, 8, 10) is discussed using Full-CI and CASSCF type wavefunctions with a variety of basis sets and considering perturbative corrections up to second order. The effect of the screening of the electron–electron interaction is also discussed by using a Yukawa-type potential with different values of the Yukawa screening parameter (λee = 0.2, 0.4, 0.6, 0.8, 1.0). Our results show that the singlet state is the ground state for two and eight electron Hooke atoms, whereas the triplet is the ground spin state for 4-, 6-, and 10-electron systems. This suggests the following Aufbau structure 1s < 1p < 1d with singlet ground spin states for systems in which the generation of the triplet implies an inter-shell one-electron promotion, and triplet ground states in cases when there is a partial filling of electrons of a given shell. It is also observed that the screening of electron–electron interactions has a sizable quantitative effect on the relative energies of both spin states, specially in the case of two- and eight-electron systems, favoring the singlet state over the triplet. However, the screening of the electron–electron interaction does not provoke a change in the nature of the ground spin state of these systems. By analyzing the different components of the energy, we have gained a deeper understanding of the effects of the kinetic, confinement and electron–electron interaction components of the energy.
Article
The orbital and spin magnetization of a cavity-embedded quantum dot array defined in a GaAs heterostructure are calculated within quantum-electrodynamical density-functional theory. To this end, a gradient-based exchange-correlation functional recently employed for atomic systems is adapted to the hosting two-dimensional electron gas submitted to an external perpendicular homogeneous magnetic field. Numerical results reveal the polarizing effects of the cavity photon field on the electron charge distribution and nontrivial changes of the orbital magnetization. We discuss its intertwined dependence on the electron number in each dot, and on the electron-photon coupling strength. In particular, the calculated dispersion of the photon-dressed electron states around the Fermi energy as a function of the electron-photon coupling strength indicates the formation of magnetoplasmon-polaritons in the dots.
Article
The effects of gadolinium addition on crystallographic characteristics (lattice parameters, c/a ratios, solid solution substitutional position) and physical properties (electrical conductivity, Vickers hardness and mechanical performance) of Mg-Gd alloy were investigated in the present work. The results show that the Gd atoms in Mg-Gd binary alloy prefer to be located on the prismatic plane. The solid solution strengthening and ductilizing effect of Mg-Gd alloy mainly occurs in the composition range of 1∼4 wt.% Gd. With the increase of Gd content, the axial ratio c/a of the extruded alloy first decreases and then increases, which is consistent with the change law of the alloy ductility. With the increase of Gd content in the low addition range, the starting resistance of basal slip increases sharply, while that of prismatic slip increases slowly. The difference between these two kinds of resistance gradually approaches to a certain extent with the increase of Gd content, which is beneficial to the coordination of basal slip and prismatic slip, and plays a plasticizing role. However, when the Gd content exceeds a certain amount, the volume fraction of the second phase in the alloy is too high, which is unfavourable to the elasticity. Under the high addition of Gd, when the starting resistance of prismatic slip and basal slip tend to be close and the difference is relatively constant, the hardening exponent also decreases to a relatively low stable value.
Article
The energy and wave function of a harmonically confined N-electron system coupled to light is calculated by separating the wave functions of the relative and center-of-mass (CM) motions. The light only couples to the CM variable, and the coupled equation can be solved analytically. The relative motion wave function has to be numerically approximated, but the relative Hamiltonian is independent of the coupling strength and it only gives a shift in the energy. The approach works for any coupling strength and the effective coupling strength can be increased by increasing the number of electrons. This gives an extra degree of freedom to fine tune the resonances and other properties of the light-matter coupled systems. Examples of wave functions of light-matter hybrid states are presented.
Article
We present a study of the two-dimensional circular quantum dot model Hamiltonian using a range of quantum chemical ab initio methods. Ground and excited state energies are computed on different levels of perturbation theories, including the coupled cluster method. We outline a scheme to compute the required Coulomb integrals in real space and utilize a semianalytic solution to the integral over the Coulomb kernel in the vicinity of the singularity. Furthermore, we show that the remaining basis set incompleteness error for two-dimensional quantum dots scales with the inverse number of virtual orbitals, allowing us to extrapolate to the complete basis set limit energy. By varying the harmonic potential parameter we tune the correlation strength and investigate the predicted ground and excited state energies.
Article
An explicitly correlated Gaussian basis is used to calculate the energies and wave functions of one-dimensional few-electron systems in confinement potentials created by external potentials or coupling to light in cavity. The appearance and properties of electron density peaks as the function of the relative strength of the confinement and the Coulomb interaction are studied. It is shown that similar Wigner crystal-like structures can be formed by coupling electrons to light due to the dipole self-interaction term in the light-matter Hamiltonian, provided an additional extremely weak confining potential is present. The relation of these systems to Wigner crystals is discussed.
Article
Full-text available
The exact density functional for the ground-state energy is strictly self-interaction-free (i.e., orbitals demonstrably do not self-interact), but many approximations to it, including the local-spin-density (LSD) approximation for exchange and correlation, are not. We present two related methods for the self-interaction correction (SIC) of any density functional for the energy; correction of the self-consistent one-electron potenial follows naturally from the variational principle. Both methods are sanctioned by the Hohenberg-Kohn theorem. Although the first method introduces an orbital-dependent single-particle potential, the second involves a local potential as in the Kohn-Sham scheme. We apply the first method to LSD and show that it properly conserves the number content of the exchange-correlation hole, while substantially improving the description of its shape. We apply this method to a number of physical problems, where the uncorrected LSD approach produces systematic errors. We find systematic improvements, qualitative as well as quantitative, from this simple correction. Benefits of SIC in atomic calculations include (i) improved values for the total energy and for the separate exchange and correlation pieces of it, (ii) accurate binding energies of negative ions, which are wrongly unstable in LSD, (iii) more accurate electron densities, (iv) orbital eigenvalues that closely approximate physical removal energies, including relaxation, and (v) correct longrange behavior of the potential and density. It appears that SIC can also remedy the LSD underestimate of the band gaps in insulators (as shown by numerical calculations for the rare-gas solids and CuCl), and the LSD overestimate of the cohesive energies of transition metals. The LSD spin splitting in atomic Ni and $s$-${}d$ interconfigurational energies of transition elements are almost unchanged by SIC. We also discuss the admissibility of fractional occupation numbers, and present a parametrization of the electron-gas correlation energy at any density, based on the recent results of Ceperley and Alder.
Conference Paper
We investigate the spin splitting of the Landau bands (LB's) in a confined two-dimensional electron gas (2DEG) using the Hartree-Fock approximation (HFA) for the mutual Coulomb interaction of the electrons. The exchange term of the interaction causes a large splitting of the spin levels of a LB whenever the chemical potential lies between them. These oscillations of the splitting with the filling factor of the LB's are conveniently interpreted as an oscillating enhancement of the effective g-factor, g*. The reduction of g* when a LB is becoming completely filled is accompanied by a spontaneous formation of a static spin-density wave state whose details depend on the system size and temperature.
Article
Progress in semiconductor technology has enabled the fabrication of structures so small that they can contain just one mobile electron. By varying controllably the number of electrons in these 'artificial atoms' and measuring the energy required to add successive electrons, one can conduct atomic physics experiments in a regime that is inaccessible to experiments on real atoms.
Article
The discovery of periodic conductance oscillations as a function of charge density in very small transistors has led to a new understanding of the behavior of electrons in such small structures. It has been demonstrated that, whereas a conventional transistor turns on only once as electrons are added to it, submicronsize transistors, isolated from their leads by tunnel junctions, turn on and off again every time an electron is added. This unusual behavior is primarily the result of the quantization of charge and the Coulomb interaction between electrons on the small transistor. However, recent experiments demonstrate that the quantization of energy is important as well.
Article
Hartree–Fock scheme;low-energy states
Article
Modern semiconductor technology makes it possible to fabricate particles of metals or pools of electrons in a semiconductor that are only a few hundred angstroms in size. Artificial atoms have a unique and spectacular property: the current through such an atom or the capacitance between its leads can vary by many orders of magnitude when its charge is changed by a single electron. Why this is so, and how this property can be used to measure the level spectrum of an artificial atom, is the subject of this paper. The various type of atoms are described schematically. 19 refs., 6 figs.
Article
This paper deals with the ground state of an interacting electron gas in an external potential v(r). It is proved that there exists a universal functional of the density, Fn(r), independent of v(r), such that the expression Ev(r)n(r)dr+Fn(r) has as its minimum value the correct ground-state energy associated with v(r). The functional Fn(r) is then discussed for two situations: (1) n(r)=n0+n(r), n/n01, and (2) n(r)= (r/r0) with arbitrary and r0. In both cases F can be expressed entirely in terms of the correlation energy and linear and higher order electronic polarizabilities of a uniform electron gas. This approach also sheds some light on generalized Thomas-Fermi methods and their limitations. Some new extensions of these methods are presented.
Article
The influence of electron-electron correlation on exchange instabilities of a metal is examined. The employment of screened interactions does not constitute a proper treatment. Correlation effects suppress ferromagnetic instabilities, as is well known, but they need not supress instabilities of the spin-density-wave type. On the contrary, it is shown that correlation enhances exchange instability of the charge-density-wave type. For either type, the wave vector of such a state adjusts so that the Fermi surface makes critical contact with the energy gaps introduced by the instability. This circumstance optimizes the correlation energy. The observed conjunction of the long-period-superlattice periodicity with the Fermi surface in order-disorder alloys is probably an example of this phenomenon. It is suggested that charge-density-wave ground states are likely in simple metals having weak Born-Mayer ion-ion interactions, such as the alkali metals. The intensity of Bragg reflection satellites caused by a concomitant positive-ion modulation is computed.
Article
It is shown rigorously that the paramagnetic state of an electron gas is never the Hartree-Fock ground state, even in the high-density—or weak-interaction—limit. The paramagnetic state is always unstable with respect to formation of a static spin density wave. The instability occurs for spin-density waves having a wave vector Q≈2kF, the diameter of the Fermi sphere. It follows that the (Hartree-Fock) spin susceptibility of the paramagnetic state is not a monotonic decreasing function with increasing Q, but rather a function with a singularity near Q=2kF. Rather convincing experimental evidence that the antiferromagnetic ground state of chromium is a large-amplitude spin density wave state is summarized. A number of consequences of such states are discussed, including the problem of detecting them by neutron diffraction.
Article
From a theory of Hohenberg and Kohn, approximation methods for treating an inhomogeneous system of interacting electrons are developed. These methods are exact for systems of slowly varying or high density. For the ground state, they lead to self-consistent equations analogous to the Hartree and Hartree-Fock equations, respectively. In these equations the exchange and correlation portions of the chemical potential of a uniform electron gas appear as additional effective potentials. (The exchange portion of our effective potential differs from that due to Slater by a factor of 23.) Electronic systems at finite temperatures and in magnetic fields are also treated by similar methods. An appendix deals with a further correction for systems with short-wavelength density oscillations.