You are on page 1of 5

The high hydrostatic pressure effect on shallow donor binding energies in

GaAs(Ga, Al)As cylindrical quantum well wires at selected temperatures


H.D. Karki, S. Elagoz
n
, P. Baser
Cumhuriyet University, Department of Physics, 58140, Sivas, Turkey
a r t i c l e i n f o
Article history:
Received 16 July 2010
Received in revised form
1 March 2011
Accepted 3 March 2011
Available online 10 March 2011
Keywords:
Impurity and defect levels
Energy states of adsorbed species
Quantum wells
Quantum wires
a b s t r a c t
We have presented the behavior of a shallow donor impurity with binding energy in cylindrical-shaped
GaAs/Ga
0.7
Al
0.3
As quantum well wires under high hydrostatic pressure values. Our results are obtained
in the effective mass approximation using the variational procedures. In our calculations, we have not
considered the pressure related GX crossover effects. The hydrostatic pressure dependence on the
expectation value of ground state binding energy is calculated as a function of wire radius at selected
temperatures. We have also discussed the effects of high hydrostatic pressure and temperature on
some physical parameters such as effective mass, dielectric constant, and barrier height. A detailed
analysis of these calculations has proved that the effective mass is the most important parameter,
which explains the dependency of donor impurity binding energies on the high hydrostatic pressure
values.
& 2011 Elsevier B.V. All rights reserved.
1. Introduction
A deep understanding of the effects of shallow impurities on
electronic states of semiconductor heterostructures is a fundamental
question in semiconductor physics because their presence alters the
performance of quantum devices dramatically. Due to its impor-
tance for potential applications in electronics and optoelectronics
devices, many experimental and theoretical studies have been
devoted to the understanding of physical properties of impurities
in low dimensional semiconductors such as quantum wells
(QWs) [15], quantum well wires (QWWs) [613] and quantum
dots (QDs) [14,15] over the last decade. The range of hydrostatic
pressure in these works is less than 10 GPa with or without
including the temperature effects. As a general feature of these
experimental and theoretical studies, it is known that the binding
energy for shallow impurities increases with increasing external
hydrostatic pressure for low-dimensional heterostructures.
The effects of high hydrostatic pressure in low-dimensional
structures have proven to be a very interesting study, since
Elabsy [17] has exhibited that with a xed QW thickness, the
calculated donor binding energy is enhanced with increasing pres-
sure, and then decreased when the hydrostatic pressure approaches
the crossover pressure between the quantum well G states and the
barrier X states [16]. Similar studies are done by Zhao et al. They
showed that up to 4 GPa, the binding energy of donors increases
almost linearly with the hydrostatic pressure for quantumwells with
an Al concentration of 0.15 [18]. Including information on the
temperature effects, Peter and Navaneethakrishnan [19] have
demonstrated that for hydrostatic pressure values up to 3.3 GPa
the binding energy increases with pressure but decreases with
increasing temperatures . Furthermore, Panahi et al. demonstrated
that similar to the ground state, excited state donor binding energies
also increase almost linearly with hydrostatic pressure (up to 1 GPa)
in the direct gap regime [20]. In a recent study of Baser et al. [21]
have calculated CQWWs for even higher pressure values up to
25 GPa and in their study and conrmed that binding energy
increases as hydrostatic pressure increases .
It is clear from Elabsys studies that the donor binding energy rst
increases and then decreases with increasing pressure values and he
concludes that it is purely from GX crossover. However, in this
paper, we have calculated donor binding energy under high hydro-
static pressure as a function of different wire radii for on center
shallow-donor impurities in cylindrical shaped GaAs/(Ga
,
Al)As
quantum well wires at selected temperatures without GX cross-
over effects to see if the binding energy increases linearly as opposed
to dropping down for high pressure values. In calculations we have
used effective mass and variational approximations, including the
temperature and the hydrostatic pressure dependency of physical
parameters such as effective mass, dielectric constants, potential
heights, and wire radii.
2. Theory
Using the effective-mass approximation, the Hamiltonian for a
shallow donor impurity in a single GaAs/(Ga
,
Al)As cylindrical
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/physb
Physica B
0921-4526/$ - see front matter & 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.physb.2011.03.004
n
Corresponding author. Tel.: 90 346 219 16 87; fax: 90 346 221 91 59.
E-mail address: elagoz@cumhuriyet.edu.tr (S. Elagoz).
Physica B 406 (2011) 21162120
quantum wire is given by
^
Hc r
!
Ec r
!
1
with
^
H
_
2
2
r
-
1
m

w,b
P,T
r
-
_ _
Vz,P,T
e
2
e
w,b
x,P,T r
!
r
!
o

2
where m

w,b
P,T and e
w,b
P,T are hydrostatic pressure and tem-
perature dependent effective mass and dielectric constant,
respectively in the quantum wire and barrier layers. The sub-
scripts w and b stand for quantum wire layer GaAs and barrier
layer (Ga
,
Al)As materials, respectively. Vr, P, T is spatial
connement potential; in this case it is the cylindrical wire
potential. In the above equation, r
!
and r
!
o
are electron and
impurity position vectors, respectively and r
!
r
!
o

is written as

r
2
r
2
o
2rr
o
cosjj
o
z
2
_
. In this expression, the z coordi-
nate denotes the relative separation of the electronimpurity
along the wire axis. In this work we deal with on-center
impurities so r
!
o
is chosen as zero. Both pressure and temperature
values modify the lattice constants, so wire radius, barrier height,
effective mass and dielectric constant all need to be written as
dependent on P and T. The pressure and temperature dependent
effective masses are written as
m

w
P,T 17:51
2
E
g
P,T

1
E
g
P,T0:341
_ _ _ _
1
m
o
3
and
m

b
P,T m

w
P,T0:0083x m
o
where m
o
is free electron mass [22,23]. In Eq. (3), E
g
P, T is
pressure and temperature dependent energy band gap for GaAs
quantum wire at G-point and is given by
E
g
P,T 1:5195:405 10
4
T
2
T 204
_ _
1:26 10
1
P3:77 10
3
P
2
4
In this equation, energies are taken in eV, P in GPa, and T in
Kelvin [24]. Similarly, dielectric constants in wire and barrier
regions are given by
e
w
P,T
12:74exp1:73 10
2
Pexp9:4 10
5
T75:6, T r200K
13:18exp1:73 10
2
Pexp20:4 10
5
T300, T 4200K
_
e
b
x,P,T e
w
P,T3:12x 5
where the dielectric constant in barrier region, e
b
x, P, T, is
obtained from a linear interpolation of the dielectric constants
of GaAs and Ga
1x
Al
x
As [24]. The spatial connement potential
Vr, P, T is dened as
Vr,P,T
0, roRP=2
V
o
x,P,T, rZRP=2
_
6
where V
o
x, P, T can be written using conduction band offset
parameter Q as
V
o
x,P,T Q DE
G
g
x,P,T
DE
G
g
x,P,T 1:155x0:37x
2
DxPGxT 7
the value of Q is taken as 0.658 [25]. The coefcients of hydro-
static pressure and temperature parameters are Dx 1:3
10
2
xeVGPa
1
and Gx 1:15 10
4
x eVK
1
, respectively
[25,26]. The hydrostatic pressure dependence of the quantum wire
radius RP is obtained by using the fractional changes in volume of
the zinc-blende structure and is given by
RP R0 1S
11
2S
12
P 8
where R0 is the quantum wire radius in the absence of the
hydrostatic pressure; S
11
and S
12
are the elastic constants of GaAs
and they correspond to 1:16 10
2
and 3:7 10
3
GPa
1
,
respectively [27].
It is evident that due to the inclusion of the Coulomb potential
in the Hamiltonian, the Schr odinger equation cannot be analyti-
cally solved in cylindrical coordinates. Therefore, it is necessary to
utilize a variational approach to calculate the eigenvalues and
eigenfunctions of the Hamiltonian. To calculate the impurity
ground state binding energy, following Brown and Spector [6],
we have assumed a suitable trail wave function for the ground
state as the product of appropriate modied Bessel function and a
variational term to include the effects of Coulombic interaction.
Using the cylindrical conning symmetry for the system, angular
part of the equation is trivial and solved easily. Therefore, we deal
only with radial part of the solution and the trial radial solutions
of the wave function for the ground state can be written as
Cr,z,P,T
NP,TJ
o
a
01
re
l

r
2
z
2
p
, 0rrrRP
NP,T
Joa01R
Kob
01
R
K
o
b
01
re
l

r
2
z
2
p
, r4RP
_
_
_
9
where J
o
a
01
r and K
o
b
01
r are the modied Bessel function of
rst and second kind of order (zero), which are the corresponding
solutions inside and outside the quantum wire, respectively.
In Eq. (9), e
l

r
2
z
2
p
is a variational term to include the effect of
Coulombic interaction and l P, T is a variational parameter.
a
01
P, T and b
01
P, T are the parameters of the modied Bessel
functions related to the ground state electronic energy, which are
worked out numerically by using the boundary conditions at
r RP in the absence of the Coulomb term. Finally, NP, T is a
normalization constant derived as
1
N
2
2p
dKM
dl
10
with
K
_
RP
0
rJ
2
o
a
01
rK
o
2lrdr
M
J
2
o
a
01
R
K
2
o
b
01
R
_
1
RP
rK
2
o
b
01
rK
o
2lrdr 11
As usual, the ground state binding energy of the shallow donor
E
b
is dened as the energy difference between the energies of the
system with and without the Coulomb term. In this case, E
b
is
obtained as
E
b
r,P,T
_
2
a
2
01
2m

w,b
p_
2
N
2
a
2
01
l
2
m
w
dK
dl

l
2
b
2
01
m
b
dM
dl
_ _ _

4p e
2
N
2
e
w,b
KM 2pV
o
N
2
dM
dl
_
12
For numerical calculations, it is customary to express the
binding energy in dimensionless variables. By introducing the
effective Rydberg constant R
Y
e
2
=2e
o
a
B
as the unit of energy and
the effective Bohr radius a
B
_
2
e
o
=m

e
2
as the unit of length, for
GaAs these parameters are R
Y
5.8 meV and a
B
98 A
o
[28].
Binding energy can be rewritten, in terms of dimensionless
variables after r R t transformation and some algebra, as
~
E
b
R,P,T l ~ a
B

2
m

1
m

w
dQ
dl

1
m

b
dP
dl
_ _ _
4~ a
B
e
o
Q
e
w

P
e
b
_ __
dQP
dl
_ _
1
13
H.D. Karki et al. / Physica B 406 (2011) 21162120 2117
where
Q R
2
_
1
0
tJ
2
o
a
01
RtK
o
2lRtdt
P
J
2
o
a
01
R
K
2
o
b
0
01
R
R
2
_
1
1
tK
2
o
b
01
RtK
o
2lRtdt 14
To obtain ground state binding energy
~
E
b
, we calculate the
value of l that minimizes /Hr, f, zS. The integration for the
expression Q and P is performed numerically.
3. Results and discussion
In numerical calculations, we have used the hydrostatic pressure
and temperature dependent physical parameters for GaAs/Ga
0.7
A-
l
0.3
As systemwhere P values span from 5 to 35 GPa and the selected
temperature range is from 4 to 300 K. To understand the wire
radius effect clearly, we have plotted the ground state binding
energies of shallow donor impurities located at the center of the
GaAs cylindrical quantum wire as a function of wire radius for
different values of the hydrostatic pressures at three given tem-
peratures 4, 200, and 300 K in Fig. 1(a)(c). As seen in Fig. 1(a)(c),
the ve curves of the donor binding energies vs. wire radius for
different T values display almost similar behavior for all hydrostatic
pressure values. For large wire radii, the binding energy is small and
nearly insensitive to the wire radius. Because, the particle is away
from the impurity ion, it behaves as if it is in bulk material,
resulting in bulk binding energies. On further decreasing the wire
radius, the particle starts to interact with the barrier, which results
in a more localized wave function in the wire region. In this regime,
particle is affected by both spatial connement and Coulombic
attraction, causing its binding energy to rise. If the wire radius is
decreased even further, the donor binding energy increases and
reaches a maximum value at about a quarter of Bohr radius of GaAs.
After this point, if the wire radius is decreased slightly further, most
of the wave function of electron penetrates into the barrier region,
so that the donor binding energy begins to decrease rapidly with
decreasing wire radius. To see the effect of the temperature, we
compare Fig. 1(a)(c) with each other. It is clear that as the
temperature is increased, the binding energy drops slightly. This
effect leads to the weakening of electron localization near the
impurity with the increasing temperature due to decreasing effec-
tive mass and increasing dielectric constant as illustrated in Fig. 2.
Therefore the binding energy decreases for all values of increasing
temperature. Note that for Tr200 K the binding energy decreases
more slowly than that for TZ200 K. This is caused by the difference
between the temperature coefcients in the dielectric constant for
the two regions of temperature given in Eq. (5).
To investigate the hydrostatic pressure effects, we plot all
important pressure dependent parameters that play important role
in binding energy, such as wire radius RP, effective mass m

w,b
P,T,
the dielectric constant e
w,b
P,T and potential height V
o
P, T
in Figs. 2(a),(b),(c) and (d), respectively. It is important to notice
that all the parameters except effective mass drop with increasing
pressure whereas the effective mass rst increases, then passes from
a maximum at about P17 GPa for all temperature values and
nally drops almost symmetrically, as seen in Fig. 2(b). In addition
to this change, the percentage of change is the greatest in effective
mass compared to the others. This is also explains why in Fig. 1
some binding energy lines cross out the others for high pressure
values. It is the combined effect of the change in wire radius,
effective mass, dielectric constant and potential height with the
hydrostatic pressure that causes this peculiar behavior in binding
energy as seen from Figs. 1(a)(c) for higher pressure values
(P420 GPa). In order to understand this phenomenon clearly, we
have also plotted the donor binding energies as functions of the wire
radius for four given pressure values, P20, 25, 30 and 35 GPa at
300 K in Fig. 3 and binding energy vs. P in Fig. 4. From both gures it
is clearly seen that for hydrostatic pressure values 20 and 25 GPa the
binding energy increases with increasing pressure, but as we
increase the hydrostatic pressure values further to 30 and 35 GPa,
the binding energy decreases. Obviously this change originates from
the behavior of the effective mass, since it is the only one that does
not behave monotonically as pressure increases. While the other
physical parameters decrease, it is the only variable that displays
parallel characteristics to those of
~
E
b
as pressure increases. That is, it
increases and then decreases with increasing hydrostatic pressure.
To clarify further, for P425 GPa in Fig. 3, further increasing
0 1 2 3 4
2
4
6
8
10
12
P= 5 GPa
P=10 GPa
P=15 GPa
P=20 GPa
P=25 GPa
P=30 GPa
P=35 GPa
R(P)/a
B
E
b
/
R
y

(
P
,

T
)
T = 4 K
0 1 2 3 4
2
4
6
8
10
12
P= 5 GPa
P=10 GPa
P=15 GPa
P=20 GPa
P=25 GPa
P=30 GPa
P=35 GPa
T = 200 K
E
b
/
R
y

(
P
,

T
)
0 1 2 3 4
2
4
6
8
10
12
P= 5 GPa
P=10 GPa
P=15 GPa
P=20 GPa
P=25 GPa
P=30 GPa
P=35 GPa
T = 300 K
E
b
/
R
y

(
P
,

T
)
R(P)/a
B
R(P)/a
B
Fig. 1. Binding energy vs. wire radius for different P values and at (a) T4 K,
(b) T200 K and (c) T300 K.
H.D. Karki et al. / Physica B 406 (2011) 21162120 2118
hydrostatic pressure results in the effective mass decreasing at
about 17 GPa, which causes an increase in the kinetic energy of
the particle, therefore weakening the binding, in turn
~
E
b
starts to
decrease as well. However we start seeing the decrease in
~
E
b
at
about 25 GPa. The reason for this is that the effective mass is not the
only parameter that affects the binding energy, so from 17 to 25 GPa
this decrease is compensated by the increase that is caused by other
terms such as decreasing dielectric constant. Therefore, we start
seeing the effect of decrease in the effective mass at somewhat
higher pressure values. Another point worth mentioning in Fig. 3 is
that as the hydrostatic pressure increases, the maximum point of
donor binding energy shifts towards the right. This can be explained
by the decrease in potential height and increase in kinetic energy, so
that particle starts penetrating at higher wire radii.
Finally, Fig. 4 sums up clearly that as hydrostatic pressure
value is increased, binding energy increases at rst and then
decreases for higher pressure values.
In this study, we have calculated the effects of hydrostatic
pressure and temperature on the binding energy of a donor impurity
in QWWs using a variational procedure and the effective mass
approximation. The results show that the binding energy increases
when the hydrostatic pressure increases up to 25 GPa and then it
decreases for hydrostatic pressure values higher than 25 GPa.
Detailed discussion leads us to conclude that this is due to the
change of physical parameters with P. Especially, the pressure
dependency of effective mass is responsible for this behavior, even
without including the GX crossover effects. It must be pointed out
that the pressure dependent dielectric constants also play an
important role so
~
E
b
starts decreasing at higher P values compared
to the pressure values at which effective mass also starts to
P (GPa)
0 5 10 15 20 25 30 35
R
(
P
)
/
a
B
0.84
0.88
0.92
0.96
1.00
P (GPa)
0 5 10 15 20 25 30 35
M
a
s
s

(
x

1
0
-
2
9


k
g
)
5
6
7
8
9
10
11
12
Wire
Barrier
T= 4 K
T= 300 K
T= 200 K
P (GPa)
0 5 10 15 20 25 30 35
D
i
e
l
e
c
t
r
i
c

C
o
n
s
t
a
n
t
0
6
8
10
12
14
300 K
4 K
200 K
Wire
Barrier
P (GPa)
0 5 10 15 20 25 30 35
P
o
t
e
n
t
i
a
l

H
e
i
g
h
t

(
m
e
V
)
150
175
200
225
250
T = 4 K
T = 100 K
T = 200 K
T = 300 K
Fig. 2. Change of physical parameters vs. hydrostatic pressure P. (a) wire radius R,
(b) effective mass m* for wire and barrier, (c) dielectric constant for wire and
barrier and (d) potential barrier height V
o
.
0,1 0,2 0,3 0,4
0
5
6
7
8
9
10
P=20 GPa
P=25 GPa
P=30 GPa
P=35 GPa
R(P)/a
B
T= 300 K
E
b
/
R
y

(
P
,

T
)
Fig. 3. Binding energy vs. wire radius for high hydrostatic pressure values P20,
25, 30 and 35 GPa at T300 K.
P (GPa)
0 5 10 15 20 25 30 35
E
b

(
P
,

T
)
/
R
y
0
4
5
6
7
8
9
10
11
T=300 K
Fig. 4. Binding energy vs. hydrostatic pressure for wire radius value0.2 a
B
at
T300 K.
H.D. Karki et al. / Physica B 406 (2011) 21162120 2119
decrease. We also observed that when temperature increases,
binding energy decreases, as similar to other studies available in
the literature.
Acknowledgements
This work is partly supported by TUBITAK 108T015 and CUBAP
F-238.
References
[1] G. Bastard, Phys. Rev. 45 (1981) 4714.
[2] G. Bastard, Surf. Sci. 113 (1982) 165.
[3] R.L. Greene, K. Bajaj, Phys. Rev. B 37 (1988) 4604.
[4] L.E. Oliveira, R. Pe rez-Alvarez, Phys. Rev. B 40 (1989) 10460.
[5] S.V. Branis, et al., Phys. Rev. B 47 (1993) 1316.
[6] J.W. Brown, H.N. Spector, J. Appl. Phys. 59 (4) (1986) 1179.
[7] G. Weber, P.A. Schulz, L.E. Oliveira, Phys. Rev. B 38 (1988) 2179.
[8] H.D. Karki, S. Elagoz, P. Baser, R. Amca, I. Sokmen, Superlattices Microstruct.
41 (2007) 227.
[9] P. Villamil, N. Porras-Montenegro, J. Phys: Condens. Matter 10 (1998) 10599;
P. Villamil, N. Porras-Montenegro, Phys. Rev. B 59 (1999) 1605.
[10] A.I. Mese, S.E. Okan, Phys. Status. Solidi B 241 (15) (2004) 3525.
[11] J.D. Correa, O. Cepeda-Giraldo, N. Porras-Montenegro, C.A. Duque, Phys.
Status. Solidi B 241 (14) (2004) 3311.
[12] A.J. Peter, Physica E 39 (2007) 115.
[13] E. Tangarife, S.Y. Lope z, M. De Dios-Leyva, L.E. Oliveira, C.A. Duque, Micro-
electron. J. 39 (2008) 431 N.
[14] Porras-Montenegro, S.T. Perez-Merchancano, Phys. Rev. B 46 (1992) 9780.
[15] S.T. Porras-Montenegro, Perez-Merchancano, A. Latge , J. Appl. Phys. 74 (1993)
7624.
[16] M. Elabsy, Phys. Scr. 48 (1993) 376.
[17] M. Elabsy, J. Phys: Condens. Matter 6 (1994) 10025.
[18] G.J. Zhao, X.X. Liang, S.L. Ban, Phys. Lett. A 319 (2003) 191.
[19] A.J. Peter, K. Navaneethakrishnan, Superlattices Microstruct. 43 (2008) 63.
[20] H. Panahi, M. Maleki, Phys. Status. Solidi B 245 (5) (2008) 967.
[21] P. Baser, S. Elagoz, D. Kartal, Physica B 405 (2010) 3239.
[22] H. Ehrenrich, J. Appl. Phys. 32 (1961) 2155.
[23] D.E. Aspnes, Phys. Rev. B 14 (1976) 5331;
B. Welber, M. Cardona, C.K. Kim, S. Rodriquez, Phys. Rev. B 12 (1975) 5729.
[24] S. Adachi, J. Appl. Phys. 58 (1985) R1.
[25] G.A. Samara, Phys. Rev. B 27 (1983) 3494;
R.F. Kopf., M.H. Herman, M. Lamont Schnoes, A.P. Perley, G. Livescu,
M. Ohring, J. Appl. Phys. 71 (1992) 5004.
[26] J.M. Mercy, C. Bousquet, J.L. Robert, A. Raymond, G. Gregoris, J. Berens,
J.C. Portal, P.M. Frijlink, P. Delescluse, J. Chevrier, N.T. Linh, Surf. Sci. 142
(1984) 298.
[27] P.Y. Yu, M. Cardona, Fundamentals of Semiconductors, Springer-Verlag,
Berlin, 1988.
[28] S.V. Branis, G. Li, K.K. Bajaj, Phys. Rev. B 47 (1993) 1316.
H.D. Karki et al. / Physica B 406 (2011) 21162120 2120

You might also like