Sie sind auf Seite 1von 15

The combined effect of pressure and temperature

on the impurity binding energy in a cubic


quantum dot using the FEM simulation
A. Sali
a,
, H. Satori
b,c
a
E.N.S de Fs, Dpartement de Physique, B.P. 5206 Bensouda, Fs, Morocco
b
Department of Mathematics and Computer Science, Mohammed Premier University, Faculty Pluridisciplinary, 300 Selouane, 62700
Nador, Morocco
c
Department of Mathematics and Computer Science, Sidi Mohamed Ben Abdellah University, Faculty of Science, B.P. 1796 Dhar El
Mehraz, Fez, Morocco
a r t i c l e i n f o
Article history:
Received 22 December 2013
Received in revised form 9 January 2014
Accepted 20 January 2014
Available online 6 February 2014
Keywords:
Binding energy
Donor impurity
Quantum dot
Hydrostatic pressure
Temperature
Finite barrier conning potential
Finite element method
a b s t r a c t
Using the nite element method, the binding energies of a hydro-
genic shallow donor impurity are investigated in a cubic GaAs/
Ga
1x
Al
x
A
s
quantum dot structure with realistic potential barrier
height under the combined effect of hydrostatic pressure and
temperature in the framework of the effective mass approxima-
tion. In the calculations, we have taken into account the elec-
tronic effective mass, dielectric constant, and conduction band
offset between the dot and barriers varying with pressure and
temperature. The results we have obtained show that the donor
binding energy varies with the dot size, the conning potential,
the position of impurity, the hydrostatic pressure and tempera-
ture. It is also found that the donor binding energy increases lin-
early with the pressure in direct gap regime and its variation is
larger for narrower dots only and drops slightly with the temper-
ature. A good agreement is obtained with the existing literature
values.
2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.spmi.2014.01.011
0749-6036/ 2014 Elsevier Ltd. All rights reserved.

Corresponding author. Tel.: +212 650775673.


E-mail address: sali_ahm@hotmail.com (A. Sali).
Superlattices and Microstructures 69 (2014) 3852
Contents lists available at ScienceDirect
Superlattices and Microstructures
j our nal homepage: www. el sevi er . com/ l ocat e/ super l at t i ces
1. Introduction
In last past decades, great developments in nano-fabrication technology have made it now pos-
sible to fabricate low-dimensional semiconductor structures whose dimensions are comparable with
inter-atomic distances in solids such as quantum dots (QDs). The movements of charge carriers in
these structures are constrained by potential barriers which conne the carriers in three dimensions.
This results in the restriction of the degrees of freedom for motion to zero. The three-dimensional
quantum connement of carriers in these structures, has led to the formation of atomic-like discrete
energy levels (subbands) which can be tailored to a specic need by changing the size, shape and
materials of the dot. Such systems have a small scale in zero spatial dimensions that their electronic
and optical properties [1] are signicantly different from the same material in bulk form. The quan-
tum dot systems have many applications in transistors, solar cells, LEDs, and diode lasers, medical
imaging etc.
In the last ten years, a profusion investigation of hydrogenic impurities states in quantum conned
zero-dimensional electron systems has been witnessed due to rapid developments in their synthesis.
Shallow hydrogenic impurities increase the conductivity of a semiconductor by several orders of mag-
nitude and play a very important role in its revolution. Without impurities, there would be no elec-
tronic components such as diode, transistor, or any semiconductor science and technology.
Understanding the effects of impurities on electronic states is a fundamental question in semiconduc-
tor physics because their presence can dramatically alter the performance of a quantum device, such
as quantum transport and optical properties. It is has been found that when the dimensions of the sys-
tem are reduced, the quantum size effect becomes prominent. This is due to the fact that an electron
moves only in a smaller space and spends most of its time close to the impurity ion. Thus, the effective
strength of the electronimpurity ion Coulomb interaction increases which lead to an enhancement of
the binding energy in lower dimensions. As a consequence, the electronic and optical properties of
quantum dot structures such as binding energy, photoionization cross-section [2], absorption spectra
and other optical properties [3] can be drastically changed.
The study of shallow hydrogenic impurity states in a semiconductor quantum well was initiated in
the early 1980s through the Bastards pioneering work [4]. Since then, a number of theoretical inves-
tigations with various methods of donor impurity states in low dimensional semiconductors and par-
ticularly in quantum dots have received considerable attention. The rst studies of the connement
effects on the impurity states in these structures have been made by Porras-Montenegro and Perez-
Merchancano [5] and Zhu et al. [6] who calculated the binding energies for the ground and excited
states as a function of dot size and impurity position.
The eigenstates of quantum dots are obtained by solving time-independent Schrdinger equation,
but, in many cases, because of complex geometry, it is very complicated to solve the equation using
analytical method due to the existence of the three-dimensional connement and Coulomb potential.
In order to solve this problem, many authors adopted several numerical methods. The most of them
have employed the variational approach [715], a technique in which the trial wave function takes
into account the connement of the carrier in the dot and the inuence of the electronimpurity
ion Coulomb interaction or alternatively, perturbation methods limited to the strong connement re-
gime [1618], an analytical transfer matrix method [19], a potential morphing method [20], exact
solution which has been obtained only for centered charged impurities [21] or impurities [6], Quan-
tum Genetic Algorithm and HartreeFock Roothaan method [22], exact diagonalization method [23],
the nite difference method [24,25], plane wave method [26] and so on [2729].
In the above works, due to the presence of electronimpurity Coulomb potential and the complex-
ity of the conguration, appropriate basis functions are difcult to obtain or are not available in closed
form for most QDs with complex geometries. We therefore present in this paper a numerical varia-
tional methodology of the nite element method (FEM) which provides a more exible method to
approximate differential equations since it can easily be extended to higher order approximations
and can be used for complex geometries.
The FEM became a dominant method in applied mathematics for numerical modeling of physical
systems in many engineering and scientic disciplines, e.g. electromagnetic, solid or uid dynamics as
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 39
well as civil and aeronautical engineering due to the intensive computationally of modern supercom-
puters with large amounts of memory and higher frequency of CPU processing.
The nite element method is a popular numerical technique and a typical real-space method for
solving problems which are described by partial differential equations (PDE) or can be formulated
as functional minimization and is one such method that transforms the PDE into matrix operations.
Unlike conventional variational schemes, the FEM does not require predetermined globally dened
trial functions. Rather, the entire domain of interest is discretized into smaller regions called elements
and a simpler solution is assumed by incorporating a local polynomial representation within each ele-
ment. This exible approach yields well converged eigenvalues.
In our previous work [30], we have calculated the binding energy of an impurity conned in a
spherical quantum dot by using the FEM choosing a one dimensional physical domain and quadratic
Lagrange as basis functions for each nite element. In this paper, we calculate the binding energy of
a hydrogenic donor impurity in a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot structure by considering a
three dimensional physical region in which elements are tetrahedral Since accuracy of the results
and computational time depend heavily besides the number of elements and the size of the local
element, on the dimension of discretization and on the type of the interpolation functions in each
element.
Understanding the effects of the quantum connement on the impurity states in the low dimen-
sional systems and especially in quantum dot structures is important in physics. In addition, uniaxial
stress and external perturbations such as the hydrostatic pressure and temperature on the electronic
and optical properties of doped nano-structures have been of great interest to researchers from both
experimental and theoretical points of view. Lefebvre et al. [31,32] presented theoretical and exper-
imental calculations on the pressure coefcients of carrier effective mass and the well-width and
barrier-height dependent excitonic transitions in quantum well systems. Their results indicated that
the increase of the effectives masses and the decrease of the barrier heights were the main reason
for the decrease of the pressure coefcients with reduced well widths. An experimental investigation
of the effect of an in-plane uniaxial stress on the characteristic transition energies of both the heavy-
hole and the light-hole excitons conned in GaAsGa
1x
Al
x
A
s
quantum wells have been carried out
[33].
As it is well known, the hydrostatic pressure has an inuence on the semiconductor parameters
such as energy gap, carrier effective mass, dielectric constant, lattice vibration, volume of the hetero-
structure and so on and an increase in the pressure results in a nearly linear increase in donor binding
energy without altering the symmetry of the heterostructure system in direct-gap regime. The effect
of hydrostatic pressure on the optical and electrical properties in quantum dot systems has been
investigated by several authors [3444].
By means of the LuttingerKohn effective mass equation and the direct diagonalization method,
Rezaei and Doostimotlagh [34] have recently studied the combined effects of hydrostatic pressure,
electric eld, and conduction band non-parabolicity on the binding energies and the diamagnetic
susceptibility of an impurity in a typical GaAsGa
1x
Al
x
A
s
spherical QD. Dane et al. [35] have reported
calculations of the normalized ground state binding energy of a hydrogenic donor impurity located at
the center of a GaAs/GaAlAs spherical QD under the inuence of hydrostatic pressure and electric
eld using a variational procedure. The inuence of the hydrostatic pressure on the binding energies
of the ground and the few excited states along with diamagnetic susceptibility of an on-center
hydrogenic impurity conned in typical GaAsGa
1x
Al
x
A
s
spherical QDs is theoretically investigated
using the matrix diagonalization method and the conduction band non-parabolicity effect [36].
Rajashabala and Kannan [37] have investigated the simultaneous effects of hydrostatic pressure and
geometry on the ionization energies of a hydrogenic donor and the metal-insulator transition in a
GaAsGa
1x
Al
x
A
s
cubical quantum dot system. Perez-Merchancano et al. [38] have calculated the
binding energies of shallow donors and acceptors in a spherical GaAsGa
1x
Al
x
A
s
quantum dot under
isotropic hydrostatic pressure for both a nite and an innite high barrier by using a variational
approach within the effective mass approximation. The binding energy of a shallow hydrogenic
impurity in a spherical quantum dot under hydrostatic pressure with square well potential is
calculated using a variational approach and the non-parabolicity effect [39]. Perez-Merchancano
et al. [40] have studied the inuences of hydrostatic-pressure on the donor binding energy in
40 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
GaAs(Ga, Al)As, quantum dots. Within the framework of effective-mass approximation, Xia et al. [41]
have investigated the hydrostatic pressure effects on the donor binding energy of a hydrogenic impu-
rity in InAs/GaAs self-assembled QD by means of a variational method. John Peter [42] has calculated
the binding energy of shallow hydrogenic impurities as a function of the dot size in GaAsGa
1x
Al
x
A
s
spherical QDs in the inuence of pressure using a variational approach within the effective mass
approximation. Correa et al. [43] have calculated the effects of hydrostatic pressure on the binding
energy and photoionization cross-section in spherical QDs for different dimensions of the structure
and radial impurity position. Oyoko et al. [44] have studied the effect of an unaxial stress on the bind-
ing energy of shallow impurities in parallelepiped-shaped GaAs/GaAlAs QDs. In all the above works, it
is have been found that the pressure increases almost linearly the donor binding energy in the regime
of low hydrostatic pressure and diminishes with the size of the structure.
The effects of temperature on the impurity states of conned charge carriers in a nanostructure
quantum dot have been carried out by some authors [4547]. The binding energy of the ground state
and the low-lying excited state of an impurity atom in a GaAs parabolic quantum dot have been re-
ported as a function of the temperature and the effective connement length of the QD by the sec-
ond-order RayleighSchrodinger perturbation theory [45]. Elabsy has studied the effect of
temperature on the binding energy of the donor impurity in an articial semiconductor atom [46]
and spherical quantum dot system [47].
Several works on the combined effects of hydrostatic pressure and temperature on the physical
properties of semiconductor quantum dot structures have drawn the attention of many scientists
recently [4854]. More recently, Sivakami and Gayathri [48] have studied the simultaneous effects
of pressure and temperature with dielectric mismatch effect of an on-center hydrogenic impurity
binding energy in a GaAs spherical QD, using the variational approach within the effective mass
approximation. Also, in a more recent work, Kirak et al. [49] have calculated the effects of the
hydrostatic pressure and temperature on the electronic and the linear and nonlinear optical prop-
erties in a spherical QD in the presence of the electric eld. Vaseghi and Sajadi [50] have examined
the binding energies and diamagnetic susceptibility of an impurity in a spherical GaAs QD under the
simultaneous inuence of static pressure, temperature and laser radiation. Based on the effective-
mass approximation within a matrix diagonalization scheme, simultaneous effects of external elec-
tric eld, hydrostatic pressure and temperature on the binding energy of an off-center hydrogenic
donor conned by a spherical Gaussian potential have been calculated [51]. Liang and Xie [52] have
investigated the combined effects of the hydrostatic pressure and temperature on the binding en-
ergy, the oscillator strength and the third-order susceptibility of third harmonic generation of a
hydrogenic impurity in a spherical QD, in the presence of the external electric eld, by means of
the perturbation approach. The combined effects of hydrostatic pressure and temperature on the
ground state binding energy of two electrons in a GaAs SQD have been studied by Sivakami and
Mahendran [53] using a perturbation approach within the effective-mass approximation. Their re-
sults show that an increment in temperature results in a decrease of the correlation energy while
an increment in the pressure for the same temperature increases the correlation energy at a
particular dot size. Yesilgu et al. [54] have reported a calculation of the binding energy and the pho-
toionization cross-section of a shallow hydrogenic impurity in quantum dots under hydrostatic
pressure and temperature as a function of the dot sizes, for incident light polarized along the axis
of the dots.
The present research is concerned with a theoretical study of the combined effects of hydrostatic
pressure and temperature on the binding energy of the ground state of a hydrogenic shallow donor
impurity in a cubic GaAs/Ga
1x
Al
x
A
s
QD with nite connement potential barriers using the FEM. Re-
sults are calculated for different dot sizes, shallow donor impurity positions, hydrostatic pressure and
temperature. The difference of the electron effective masses and dielectric constants between the
quantum dot region and barriers have been considered within the effective mass and parabolic band
approximations and by restricting ourselves to range of pressure where there is no CX crossover
effect.
The paper is organized as follows: in Section 2 we describe the theoretical framework, Section 3 is
dedicated to the results and discussion, and nally, our conclusions are given in Section 4.
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 41
2. Basic theory
Within the effective mass and parabolic band approximations, the hydrogenic donor impurity lo-
cated at the position (x
0
, y
0
, z
0
) and conned in a cubic QD of width L, embedded in a dielectric matrix,
under the combined effects of hydrostatic pressure (P) and temperature (T) can be described by the
stationary Schrdinger equation,
H Ew 0 1
where
H
m

e
m

d;b
P; T
r
2
2

0

d;b
P; T
1

x x
0

2
y y
0

2
z z
0

2
_
Vx; y; z; P; T
R

_
_

_ 2
E denotes a discrete eigenvalue of the Hamiltonian H and w is the eigenfunction expressed as the sum
over the elemental wave function:
wx; y; z

N
E
a
w
a
x; y; z 3
where N
E
is the number of element and w
a
is the wave function of element(a) which can be expressed
as a linear combination of the interpolation polynomials basis functions:
w
a
x; y; z

n
i1
C
a
i
u
a
i
x; y; z 4
where C
a
i
are unknown coefcients that represent the amplitude of the wave function at a particular
node, u
a
i
x; y; z are the basis functions in each element and the summation upper limit is the number
of basis functions per element. In our case, the elements are three dimensional and have tetrahedron
shape with four interpolation functions (n = 4).
By inserting Eqs. (3) and (4) in Eq. (1), we obtained a general matrix eigenvalue problem:
H ESC 0 5
The elements of the Hamiltonian matrix H and the overlap matrix S are given by:
H
ij
hu
i
; Hu
j
i
_
X
u
i
Hu
j
dX 6
and S hu
i
; u
j
i
_
X
u
i
u
j
dX 7
in which X is the entire physical domain of interest.
Because the wave function vanishes at the innity, w(x, y, z) ?0 as (x, y, z) ?1 and hence there is
no surface contribution, and because the FEM being a numerical approximation, we therefore reduce
the upper limit of integration (1) to a nite value, (x, y, z) = L
c
[55]. As a consequence, the entire phys-
ical domain is segmented in two nite domains X
d
and X
b
where
X
d

jxj LP=2
jyj LP=2
jzj LP=2
_

_
8
is the domain related to the GaAs quantum dot material and,
X
b

LP=2 < jxj 6 L
c
LP=2 < jyj 6 L
c
LP=2 < jzj 6 L
c
_

_
9
42 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
is the Ga
1x
Al
x
A
s
barrier domain as seen in Fig. 1.
Here L(P) is the cubic quantum dot side width depending on the pressure.
The Hamiltonian Eq. (2) is written in reduced atomic units which correspond to a length unit of an
effective Bohr radius, a


0
h
2
=m

e
e
2
, and an effective Rydberg, R

e
e
4
=2
0
h
2
. The two subscripts d
and b refer to the quantum dot and the barrier layer materials, respectively.
r

x x
0

2
y y
0

2
z z
0

2
_
is the distance between the electron and impurity site. m

e
and

0
are, respectively, the static dielectric constant and the electron effective mass of the GaAs quantum
dot at zero pressure. The application of hydrostatic pressure modies the lattice constants, dot size,
barrier height, effective masses and dielectric constants. m

b;d
P; T are the pressure and temperature
dependent effective masses for the electron of the quantum dot and barrier layer, respectively. For
a GaAs quantum dot, the parabolic conduction effective mass is determined from the expression
[5658]
m

d
P; T 1 E
C
p
P; T
2
E
C
g
P; T

1
E
C
g
P; T D
0
_ _ _ _
1
m
0
10
where m
0
is the free electron mass, E
C
p
P; T 7:51 eV is the energy related to the momentum matrix
element, D
0
= 0.341 eV is the spin-orbit splitting. E
C
g
P; T is the pressure and low temperature depen-
dent energy gap for the GaAs QD semiconductor at the C-point in units of eV taken from Ref. [59].
E
C
g
P; T E
C
g
0; 0 5:40510
4
K
1
T
2
T 204K
10710
3
GPa
1
P 3:7710
3
GPa
2
P
2
_ _
eV:
11
E
C
g
0; 0 1:519 eV is the energy gap for GaAs quantum dot at the C-point and at temperature T = 0
and pressure P = 0.
The barrier materials parabolic conduction effective masse as a function of pressure and tempera-
ture is determined from the expression [5658]
m

b
P; T m

d
P; T 0:083xm
0
; 12
where x is the Al mole fraction of Aluminum in the Ga
1x
Al
x
A
s
layer.
In the above expression Eq. (1),
d;b
P; T are the pressure and temperature dependent static dielec-
tric constant of both materials respectively. In the GaAs quantumdot region,
d
P; T is given by [60,61]

b
O
A L/2
x
y
z
L
c
Fig. 1. A 3D-cubic quantum dot with center in origin. OA is the diagonal of the cube.
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 43

d
P; T
12:74exp1:67 10
2
GPa
1
Pexp9:4 10
5
T 75:6T 6 200 K
13:18exp1:73 10
2
GPa
1
Pexp20:4 10
5
T 300T > 200 K
_
13
The static dielectric constant of material barrier is obtained from a linear interpolation of the
dielectric constants of GaAs and Ga
1x
Al
x
A
s
is given by [58]

b
x; P; T
d
P; T 3:12x: 14
V(x, y, z, P, T) in Eq. (2) is the connement potential energy which connes the donor electron and is
given by
Vx; y; z; P; T
0; jxj; jyj; jzj 6 LP=2
V
0
P; T; jxj; jyj; jzj > LP=2;
_
15
where L(P) is the pressure dependent of the cubic quantum dot side length. V
0
(P, T) is the barrier
height expressed as a function of the applied pressure P, at temperature T and is is given by the expres-
sion [60,61]
V
0
P; T 0:658DE
C
g
x; P; T: 16
DE
C
g
x; P; T stands for the difference in the band gap energy of GaAs and Ga
1-x
Al
x
A
s
at the C point as a
function of temperature T and pressure P and is expressed as [5658]
DE
C
g
x; P; T DE
C
g
x CxP DxT: 17
The function DE
C
g
x is the variation of the energy gap difference in the absence of pressure and
temperature and is given by
DE
C
g
x 1:155x 0:37x
2
eV: 18
The quantities C(x) and D(x) are usually dened as the pressure and temperature coefcients,
respectively of the band gap difference and are given by
Cx 1:3 10
2
xGPa
1
eV; 19
Dx 1:15 10
4
xK
1
eV: 20
The hydrostatic pressure effects on the geometric dimensions of the zinc blende low-dimensional
structures are obtained from the Murnaghan relation [62,63]
aP
a
0
1 P
B
0
0
B
0
_ _

1
3B
0
0
21
The change in lattice constant is related to the change in volume by [63]
DV=V
0
3Da=a
0
; 22
where DV VP V
0
and Da aP a
0
23
Here V is the volume of the dot versus the pressure, V
0
is the original volume at atmospheric hydro-
static pressure, a(P) is the lattice constant as a function of pressure and a
0
is the static lattice constant.
B
0
= (C
11
+ 2C
12
)/3 is the bulk modulus [64], B
0
0
is the pressure derivative of the bulk modulus and C
11
and C
12
are the elastic constants.
By developing the term a(P) as a function of P, the variation of the cubic quantum dot side width
L(P) with the pressure is therefore
LP L
0
1 P=B
0
P
2
=6B
2
0
_ _1
3
; 24
which can be reduced by omitting the P
2
term to
LP L
0
1 3S
11
2S
12
P
1
3
; 25
44 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
where L
0
is the cubic GaAs dot size at atmospheric hydrostatic pressure, S
11
and S
12
are the compliance
constants, which can be calculated using elastic constant C
11
and C
12
of GaAs [65]
S
11
C
11
C
12
=C
11
C
12
C
11
2C
12
; 26
S
12
C
12
=C
11
C
12
C
11
2C
12
; 27
where S
11
= 11.6 10
3
GPa
1
and S
12
= 3.7 10
3
GPa
1
and L
0
is the well width at atmospheric
hydrostatic pressure.
In Eq. (1), the hydrostatic pressure and temperature dependence of the ground state binding energy
E
b
(P, T) can be given as follows,
E
b
P; T E
sb
P; T E
i
P; T; 28
where E
sb
(P, T) is the eigenvalue of Hamiltonian in Eq. (2) without the impurity potential term and E
i
(P,
T) is the eigenvalue with the impurity potential term.
3. Results and discussions
We have calculated the donor binding energy by means of the FEM technique using the equidistant
discretization and considering tetrahedral elements and taking the critical cubic size value L
c
to be ten
times the effective Bohr radius (L
c
= 10a

).
In our theoretical study, we have considered a cubic QD made of GaAs and Ga
1x
Al
x
A
s
material
where x is the aluminum concentration since it is the most widely investigated and in which all
the material properties are well known. For numerical calculations, in all what follows, we expressed
the binding energies in effective electron Rydberg units R

= 5.302 m eV and the sizes of the cubic


quantum dot in effective electron Bohr radius a

103; 56 Awhich correspond to an electron effective


mass m

e
0:067m
0
and
0
12; 74 without applied hydrostatic pressure and in the regime of low
temperature [60,61].
We have calculated the donor binding energy of the shallow hydrogenic impurity as functions of
the size of the cubic QD, the Al concentration x, the impurity position, the hydrostatic pressure P
and the temperature T in Zinc Blend GaAs/Ga
1x
Al
x
A
s
QD. We limited our calculations in the regime
of small values of the pressure [04 GPa] where the GaAs QD has a direct band gap since the transition
from direct to indirect band gap for GaAs occurs at higher pressures (larger than 4 GPa) where the
gamma-X crossover effect cannot be neglected [66].
To understand clearly the dot size width effect on the impurity electronic state, we have plotted in
Fig. 2 the binding energy of an on-center shallow donor impurity as a function of the half quantum dot
side width L/2 for an Al concentration x = 0.15 and for ve different hydrostatic pressure values
P = 0, P = 1, P = 2, P = 3 and P = 4 GPa with T = 0 K. The behavior of the binding energies without pres-
sure (P = 0) is similar to the previous results found in Ref. [67] using the plane wave basis method. For
each value of the pressure, we observe that the binding energy increases from its bulk value in GaAs as
the dot side width is reduced, reaches a maximum value, and then drops to the bulk value character-
istic of the barrier material as the cubic dot side length goes to zero. For large dot side widths, the
binding energy is small and nearly independent of dot size. Since, the electron is away from the impu-
rity ion and it behaves as if it is in bulk material, resulting in bulk binding energies. While for very
small dots, the tunneling effects become dominant and most of the wave function of electron pene-
trates into the barrier region, which results in a rapidly decreasing binding energies. For small dots,
the maximum of the binding energy is due to a more localized wave function in the dot region caused
by the combined effect of connement and Coulombic attraction potentials.
Note that the binding energy increases with the hydrostatic pressure for any dot side width, reect-
ing the additional connement due to the pressure; i.e. when the hydrostatic pressure is increased, the
donor impurity becomes more conned and the binding energy increases. Also we observe that the
pressure effect is more noticeable for narrow dots, and the maximum position goes slightly to small
dot sizes when the pressure increases. This is due to the increment of the dot effective masses as well
as to the decrease of dielectric constant and the barrier height with the pressure [39,48]. On the
other hand, the shrinkage of the dot size with the increment of the pressure [39,48] results in
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 45
shortening of the effective electronimpurity distance which also leads to an increase in the
binding energies. This behavior is clearly shown in Fig. 3 for the impurity fractional energy shift
DE
b
= (E
b
(P, L) E
b
(P = 0, L
0
))/E
b
(P = 0, L
0
) as a function of the half quantum dot side width L/2 for
different values of the hydrostatic pressures with an Al concentration x = 0, 15 at T = 0 K.
Fig. 4 shows the dependence of the binding energy of an on-center hydrogenic donor impurity on
the hydrostatic pressure P for two different values of the half quantum dot sizes L/2 = 0.5a

and 1a

with a xed temperature value T = 200 K and an Al concentration x = 0.30. By keeping xe the value
of the temperature and barrier potential connement and for a given value of the dot size, the binding
energy shows an approximately linear increase with the applied pressure. These results are in
accordance with those obtained previously in quantum dots [48,51], quantum wires [68] and quan-
tum wells [69,70]. Additionally, it is clear that as the half size of the cubic QD structure increases,
Fig. 2. Binding energy E
b
of an on-center shallow donor impurity in a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot as a function of the
half quantum dot side width L/2 for ve hydrostatic pressure values P = 0, P = 1, P = 2, P = 3 and P = 4 GPa with an Al
concentration x = 0, 15 and for T = 0 K.
Fig. 3. The impurity fractional energy shift, DE
b
= (E
b
(P, L) E
b
(P = 0, L
0
))/E
b
(P = 0, L
0
) as a function of the half quantum dot side
width L/2 for different values of the hydrostatic pressures with an Al concentration x = 0, 15 and for T = 0 K.
46 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
the binding energy decreases, reecting the lower connement potential. This is because the wave
function does not fell the small compression in the structure when the size of the structure is very
large. Fig. 4 reveals also that the difference in binding energy between the two curves L/2 = 0.5a

and 1a

increases from P = 0 to P = 4 GPa with increasing pressure because of the combined effect of
the QD size and the hydrostatic pressure. As a consequence the application of hydrostatic pressure
is leading to a more connement of the impurity electron for small dot dimensions, in agreement with
an observation made earlier in Fig. 2.
The Variation of the binding energy of an on-center hydrogenic donor impurity in a cubic GaAs/
Ga
1x
Al
x
A
s
quantum dot as a function of the hydrostatic pressure P for three values of the Al concen-
trations x = 0.15, x = 0.30 and x = 0.45 is given in Fig. 5 for a xed temperature value T = 200 K and the
Fig. 4. The dependence of the binding energy of an on-center hydrogenic donor impurity on the hydrostatic pressure P for two
values of the half quantum dot sizes L/2 = 0.5a

and 1a

with a xed temperature value T = 200 K and an Al concentration


x = 0.30.
Fig. 5. Variation of the binding energy of an on-center hydrogenic donor impurity in a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot as a
function of the hydrostatic pressure P for three values of the Al concentrations x = 0.15, x = 0.30 and x = 0.45 for a xed
temperature value T = 200 K and the half quantum dot size L/2 = 1a

.
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 47
half quantum dot size L/2 = 1a

. It can be more clearly seen from this gure that the binding energy is
an increasing function of the Al concentration x for a xed dot size, pressure and temperature. Since as
the Al concentration x increases, the height of the barrier potential increases and the wave function
becomes more localized inside the dot, which leads to more probability of nding the electron inside
the dot and the electron is better pushed toward the impurity center. The effective Coulomb interac-
tion is therefore more enhanced and as a consequence the binding energy increases. In contrast to
Fig. 4, the three curves of the gure relative to x = 0.15, x = 0.30 and x = 0.45 are almost, parallel, in
the calculated pressure range, from 0 to 4 GPa. This means that the difference in the binding energy
between any two values of the Al concentration x remains nearly the same for all the range of the
pressure.
To study the effect of the barrier height on the donor binding energy, we present in Fig. 6, the bind-
ing energy E
b
of a shallow donor impurity located at the center of a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot
as a function of the half dot side width L/2 for two different Al concentrations x = 0.15 (dashed line)
and x = 0.30 (solid line) and three hydrostatic pressure values P = 0, P = 1 and P = 2GPa with T = 0 K.
As seen from this gure, the binding energy of the shallow hydrogenic impurity depends highly on
the height of the connement potential since high (low) Aluminum concentration implies a high
(small) potential barrier. The raise of the aluminum content x in the Ga
1x
Al
x
A
s
material increases
the binding energy for a xed value of the pressure, reecting the higher connement potential. This
increment in the binding energy is due to the fact that the excess of aluminum content expands the
energy gap between the two materials which in turn increases the potential barrier height leading to
more connement of the donor electron inside the QD, and consequently greater binding energy. We
also note that the increasing of mole fractions of aluminum shifts the maximum of the binding energy
to lower dot side width. These results agree with those obtained by Li and Xia [26] who introduced a
uniform method to calculate the binding energy of the ground state as a function of the cubic QD side
width using the plane wave basis in the absence of the applied hydrostatic pressure. For QD with large
dot sizes, the effects of the hydrostatic pressure and the potential barrier height on the binding energy
is negligible, since the impurity wave function is more spread because the potential barriers are far
away and the impurity feels a bulk like environment for these large dimensions of the QD.
In Fig. 7, we display the variation of the binding energy E
b
of a hydrogenic donor impurity in a cubic
GaAs/Ga
1x
Al
x
A
s
QD of half side L/2 = 1a

as a function of the impurity position x


0
inside the dot, along
the cube diagonal (see Fig. 1) with x
0
= y
0
= z
0
where x
0
varying from 0 to L/2 at three hydrostatic pres-
sure values P = 0, P = 1 and P = 2 GPa and for two values of the Al concentrations x = 0.15 (solid line)
Fig. 6. Binding energy E
b
of a shallow donor impurity located at the center of a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot as a function
of the half quantum dot side width L/2 for two different Al concentrations x = 0.15 (dashed line) and x = 0.30 (solid line) and
three hydrostatic pressure values P = 0, P = 1 and P = 2 GPa with T = 0 K .
48 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
and x = 0.30 (dashed line). According to this gure, the maximum binding energy is obtained for impu-
rity located at the center of the cubic QD for the large band offset (x = 0.30) and P = 2 GPa pressure va-
lue. For a specic value of the pressure and the Al concentration x, the binding energy decreases from
its maximum as the impurity moves towards the corner of the cube (L/2). This is in agreement with
the results obtained previously by Mendoza et al. [71] in a cubic QD with innite potential barrier.
We also show that the binding energy increases with increasing applied pressure and Al concen-
tration x, depending on the impurity position inside the dot. We have also observed that the variation
of the binding energy as a function of the pressure and the Al concentration x for the shallow donor
impurity in different positions is not homogeneous. As an example, for impurity positions closed to
the barriers, the binding energy is nearly insensitive to the hydrostatic pressure and the potential bar-
rier height effects due to the small Coulomb interaction between the electron and the impurity and the
potential barrier repulsion. These results are in agreement with those obtained by Perez-Merchancano
et al. [40], who showed that the pressure effects are less pronounced for impurities on the edge.
The binding energy for an on-center shallow donor impurity as a function of the half quantum dot
side width L/2 for an Al concentration x = 0.15 is shown in Fig. 8 for ve different temperature values
T = 0, T = 200 GPa, T = 400 GPa, T = 600 GPa and T = 800 GPa and for P = 0. It is clear from Fig. 8 that, at a
xed temperature the binding energy increases until it reaches a maximum value at a specic QD size
and then decreases as the half quantum dot side width increases. In contrast to the hydrostatic pres-
sure effect shown in Fig. 2, the binding energy decreases with a raise in the temperature for all the dot
sizes with a constant applied pressure. For very large QD side width, all curves approach the bulk limit
value of GaAs. These behaviors are the same as of Refs. [5,46]. The increment of temperature leads to
an increase of the effective Bohr radius a

e
d
h
2
=m

d
e
2
due to the decreasing effective mass and
increasing the dielectric constant [43]. As a result, the value of the effective Rydberg constant,
R

= e
2
/(2e
d
a

) reduces, which in turns decreases the potential barrier height and as a consequence
the binding energy decreases.
The dependencies of the donor binding energy as a function of the temperature in the cubic GaAs/
Ga
1x
Al
x
A
s
quantum dot is shown in Fig. 9 at three different values of the Al concentration x = 0.15,
x = 0.30 and x = 0.45 with a xed value of the QD half width, L/2 = 0.5a

and an applied hydrostatic


pressure P = 2 GPa. The impurity is considered at the center of the dot. As can be seen from this gure,
for a given constant applied pressure and at a specied dot size, as the temperature is increased, the
ground state binding energy of a donor impurity decreases. This effect leads to the weakening of
Fig. 7. Binding energy E
b
of a hydrogenic donor impurity in a cubic GaAs/Ga
1x
Al
x
A
s
QD of half side L/2 = 1a

as a function of the
impurity position x
0
inside the dot along the cube diagonal with x
0
= y
0
= z
0
where x
0
varying from 0 to L/2 for three hydrostatic
pressure values P = 0, P = 1 and P = 2 GPa and for two values of the Al concentrations x = 0.15 (solid line) and x = 0.30 (dashed
line).
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 49
electron localization near the impurity with the enhancement of temperature. We also note that for
T 6 200 K, the binding energy decreases more slowly than for T > 200 K. Similar results have been also
seen in quantum dots [46,48] and quantum wells [72]. This behavior is caused by the difference
between the temperature coefcients in the dielectric constant for the two ranges of temperature that
we have considered here. This result indicates that the temperature effect is quite signicant in small
quantum dots only.
4. Conclusions
In conclusion, by using the FEM, we have studied and computed the binding energy of a donor
impurity in a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot with the realistic potential barrier height under
Fig. 8. The binding energy for an on-center shallow donor impurity as a function of the half quantum dot side width L/2 for Al
concentration x = 0.15 and for ve different temperature values T = 0, T = 200 GPa, T = 400 GPa, T = 600 GPa and T = 800 GPa at
zero pressure.
Fig. 9. Variation of an on-center donor binding energy versus the temperature in a cubic GaAs/Ga
1x
Al
x
A
s
quantum dot at three
different values of the Al concentration x = 0.15, x = 0.30 and x = 0.45 with a xed value of the QD half width, L/2 = 0.5a

and an
applied hydrostatic pressure P = 2 GPa and impurity position (x
0
, y
0
, z
0
) = (0, 0, 0).
50 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852
the simultaneous effects of hydrostatic pressure and temperature. The calculations have been made in
the effective mass and parabolic band approximations. The study considers also variation in the impu-
rity position and in the dimensions of the quantum dot. The main ndings can be summarized as
follows:
- The hydrostatic pressure increases almost linearly the binding energy and its effect is more notice-
able for narrow dots.
- The binding energy is an increasing function of the Al concentration x for a xed dot size, pressure
and temperature and the effect of the potential barrier height is more signicant at smaller dots.
- The binding energy has a maximum at the center of the QD and decreases as the impurity moves
towards the corner of the cube (L/2).
- The increment of temperature leads to a reduction of the donor ground state binding energy and its
effect is quite signicant in small quantum dots only.
- We note also that for T 6 200 K, the binding energy decreases more slowly than for T > 200 K.
- Our results are in a agreement with other previous works.
References
[1] G. Wang, Phys. Rev. B 72 (2005) 155329.
[2] A. Sali, H. Satori, M. Fliyou, H. Loumrhari, Phys. Status Solidi B 232 (2002) 209. and references therein.
[3] M. Sahin, Phys. Rev. B 77 (2008) 045317.
[4] G. Bastard, Surf. Sci. 113 (1982) 165.
[5] N. Porras-Montenegro, S.T. Perez-Merchancano, Phys. Rev. B 46 (1992) 9780.
[6] J.L. Zhu, J.J. Xiong, B.L. Gu, Phys. Rev. B 41 (1990) 6001.
[7] F. Jiang, C. Xia, S. Wei, Phys. B 403 (2008) 165.
[8] A. John Peter, K. Navaneethakrishnan, Physica E 40 (2008) 2747.
[9] S.T.P. Merchancano, H.P. Gutierrez, J.S. Valencia, J. Phys.: Condens. Matter 19 (2007) 026225.
[10] C. Bose, K. Midya, M.K. Bose, Physica E 33 (2006) 116.
[11] R.S.D. Bella, K. Navaneethakrishnan, Solid State Commun. 130 (2004) 773.
[12] H. Satori, A. Sali, K. Satori, Physica E 14 (2002) 184.
[13] H. Satori, M. Fliyou, A. Sali, A. Nougaoui, L. Tayebi, Phys. Low. Dim. Struct. 1 (2) (2001) 73.
[14] Y.P. Varshni, Superlattices Microstruct. 29 (2001) 233.
[15] F.J. Ribeiro, A. Latg, Phys. Rev. B 50 (1994) 4913.
[16] Y. Yakar, B. akr, A. zmen, Superlattices Microstruct. 60 (2013) 389.
[17] J.-H. Yuan, C. Liu, Physica E 41 (2008) 41.
[18] C. Bose, Physica E 4 (1990) 180.
[19] T. Xu, L. Yuan, J. Fang, Physica B 404 (2009) 3445.
[20] A.F. Terzis, S. Baskoutas, J. Phys: Conf. Ser. 10 (2005) 77.
[21] W. Xie, Phys. Lett. A 263 (1999) 127.
[22] Y. Yakar, B. akr, A. zmen, J. Lumin. 134 (2013) 778.
[23] W. Xie, Physica B 403 (2008) 2828.
[24] L. Gong, Y. Shu, J. Xu, Q.Z. Wang, Superlattices Microstruct. 60 (2013) 311.
[25] C.S. Yang, Microelectron. J. 39 (2008) 1469.
[26] S.-S. Li, J.-B. Xia, Phys. Lett. A 366 (2007) 120.
[27] C. Gonzlez-Santander, T. Apostolova, F. Domnguez-Adame, J. Phys.: Condens. Matter 25 (2013) 335802.
[28] E. Sadeghi, A. Avazpour, Physica B 406 (2011) 241.
[29] A. Gharaati, R. Khordad, Superlattices Microstruct. 48 (2010) 276.
[30] H. Satori, A. Sali, Physica E 48 (2013) 171.
[31] P. Lefebvre, B. Gil, H. Mathieu, Phys. Rev. B 35 (1987) 5630.
[32] P. Lefebvre, B. Gil, J. Allegre, H. Mathieu, Y. Chen, C. Raisin, Phys. Rev. B 35 (1987) 1230.
[33] B. Gil, P. Lefebvre, H. Mathieu, G. Platero, M. Altarelli, T. Fukunaga, H. Nakashima, Phys. Rev. B 38 (1988) 1215.
[34] G. Rezaei, N.A. Doostimotlagh, Physica E 44 (2012) 833.
[35] C. Dane, H. Akbas, A. Guleroglu, S. Minez, K. Kasapoglu, Physica E 44 (2011) 186.
[36] G. Rezaei, N.A. Doostimotlagh, B. Vaseghi, Commun. Theor. Phys. 56 (2011) 377.
[37] S. Rajashabala, R. Kannan, J. Nano-Electron. Phys. 3 (2011) 1041.
[38] S.T. Perez-Merchancano, R. Franco, J. Silva-Valencia Microelectron. J. 39 (2008) 383.
[39] A. Sivakami, M. Mahendran, Physica B 405 (2010) 1403.
[40] S.T. Perez-Merchancano, H. Paredes-Gutierrez, J. Silva-Valencia, J. Phys.: Condens. Matter 19 (2007) 026225.
[41] C. Xia, Y. Lui, S. Wei, Appl. Surf. Sci. 254 (2008) 3479.
[42] A. John Peter, Physica E 28 (2005) 225.
[43] J.D. Correa, N. Porras-Montenegro, C.A. Duque, Phys. Status Solidi B 241 (2004) 2440.
[44] H.O. Oyoko, C.A. Duque, N. Porras-Montenegro, J. Appl. Phys. 90 (2001) 819.
[45] S.-H. Chen, J.-L. Xiao, Physica B 393 (2007) 213.
A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852 51
[46] A.M. Elabsy, Egypt. J. Sol. 23 (2000) 267.
[47] A.M. Elabsy, Phys. Scr. 59 (1999) 328.
[48] A. Sivakami, V. Gayathri, Superlatt. Microstruct. 58 (2013) 218.
[49] M. Kirak, Y. Altinok, S. Yilmaz, J. Lumin. 136 (2013) 415.
[50] B. Vaseghi, T. Sajadi, Physica B 407 (2012) 2790.
[51] G. Rezaei, S.F. Taghizadeh, A.A. Enshaeian, Physica E 44 (2012) 1562.
[52] S.J. Liang, W.F. Xie, Eur. Phys. J. B 81 (2011) 79.
[53] A. Sivakami, M. Mahendran, Superlatt. Microstruct. 47 (2010) 530.
[54] U. Yesilgu, E. Kasapoglu, H. Sari, I. Sokmen, Superlatt. Microstruct. 48 (2010) 509.
[55] L.R. Ram-Mohan, Finite Element and Boundary Element Applications in Quantum Mechanics, Oxford University Press,
Oxford, UK, 2002.
[56] B. Welber, M. Cardona, C.K. Kim, S. Rodriquez, Phys. Rev. B 12 (1975) 5729.
[57] D.E. Aspnes, Phys. Rev. B 14 (1976) 5331.
[58] S. Adachi, J. Appl. Phys. 58 (1985) R1.
[59] M.E. Mora-Ramos, S.Y. Lpez, C.A. Duque, Eur. Phys. J. B 62 (2008) 257.
[60] R.F. Kopf, M.H. Herman, M. Lamont Schnoes, A.P. Perley, G. Livescu, M. Ohring, J. Appl. Phys 71 (1992) 5004.
[61] G.A. Samara, Phys. Rev. B 27 (1983) 3494.
[62] F.P. Murnaghan, Proc. Natl. Acad. Sci. USA 30 (1944) 244.
[63] B. Rockwell, H.R. Chandrasekhar, M. Chandrasekhar, A.K. Ramdas, M. Kobayashi, R.L. Gunshor, Phys. Rev. B 44 (1991)
11307.
[64] J.A. Tuchman, I.P. Herman, Phys. Rev. 45 (1992) 11929.
[65] P.Y. Yu, M. Cardona, Fundamentals of Semiconductors, Springer, Berlin, 1998.
[66] D.J. Wolford, J.A. Bradly, Solid State Commun. 53 (1985) 1069.
[67] S.-S. Li, J.-B. Xia, J. Appl. Phys. 101 (2007) 093716.
[68] G. Rezaei, S. Mousavi, E. Sadeghi, Physica B 407 (2012) 2637.
[69] G.J. Zhao, X.X. Liang, S.L. Ban, Phys. Lett. A 319 (2003) 191.
[70] A.L. Morales, A. Montes, S.Y. Lpez, C.A. Duque, J. Phys. Condens. Matter 14 (2002) 987.
[71] C.I. Mendoza, G.J. Vazquez, M. del Castillo-Mussot, H. Spector, Phys. Rev. B 71 (2005) 075330.
[72] A. Hakimyfard, M.G. Barseghyan, C.A. Duque, A.A. Kirakosyan, Physica B 404 (2009) 5159.
52 A. Sali, H. Satori / Superlattices and Microstructures 69 (2014) 3852

Das könnte Ihnen auch gefallen