Sei sulla pagina 1di 11

Superlattices and Microstructures 85 (2015) 581–591

Contents lists available at ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

Theoretical investigation of single dopant in core/shell


nanocrystal in magnetic field
A. Talbi a, E. Feddi b, A. Oukerroum c, E. Assaid d, F. Dujardin e,⇑, M. Addou a
a
Laboratoire Optoélectronique et Physico-chimie des Matériaux, Université Ibn Tofail, Faculté des Sciences, BP 133, Kenitra 14000, Morocco
b
Université Mohamed V, Ecole Normale Supérieure de l’ Enseignement Technique, Département De Physique, Rabat (ENSET), Morocco
c
University of Hassan II-Mohammedia, Faculty of Sciences and Techniques, Laboratory of Condensed Matter, B.P. 146, 20800 Mohammedia, Morocco
d
Laboratoire d’électronique et Optique des Nanostructures de Semiconducteurs, Faculté des Sciences, B.P. 20, El Jadida, Morocco
e
LCP-A2MC, Institut de Chimie, Physique et Matériaux, Université de Lorraine, 1 boulevard Arago, 57070 Metz, France

a r t i c l e i n f o a b s t r a c t

Article history: The control of single dopant or ‘‘solitary dopant’’ in semiconductors constitute a challenge
Received 16 June 2015 to achieve new range of tunable optoelectronic devices. Knowing that the properties of
Accepted 20 June 2015 doped monocrystals are very sensitive to different external perturbations, the aim of this
Available online 22 June 2015
study is to understand the effect of a magnetic field on the ground state energy of an
off-center ionized donor in a core/shell quantum dot (CSQD). The binding energies with
Keywords: and without an applied magnetic field are determined by the Ritz variational method tak-
Single dopant
ing into account the electron-impurity correlation in the trial wave function deduced from
Core–shell
Magnetic field
the second-order perturbation. It has been found that the external magnetic field affects
Donor strongly the binding energy, and its effect varies as a function of the core radius and the
shell thickness. We have shown the existence of a threshold ratio ða=bÞcrit which represents
the limit between the tridimensional and the spherical surface confinement. In addition
our analysis demonstrates the important influence of the position of ionized donor in
the shell material.
Ó 2015 Published by Elsevier Ltd.

1. Introduction

Due to their surprising optoelectronic properties, the nanomaterials made in semiconductors (quantum wells, wires, and
dots) continue to attract extensively experimental and theoretical interest [1–3]. In such structures the confinement effect
restricts the motions of charge carriers (electrons and holes) in one, two or three dimensions leading to a partial or total
quantization of the energy levels where some new photoluminescent transitions appears and new applications become pos-
sible [4,5].
In recent years, the progress in chemical growth process have permit to fabricate a new class of coated spherical quantum
dots called core/shell quantum dots (CSQD). They are composed by two semiconductor materials with different band align-
ment: the core with a large band gap which play the role of substrate is coated by a spherical shell with a small band gap or
vice versa (Fig. 1). These manufacturing processes allowed to manipulate several combinations of II–IV, III–V or IV–VI semi-
conductors [6–10]. The originality of these double layer structures is that their physical properties can be controlled by
changing the nature or the size of the core or shell and consequently manipulated in advance by tuning the electron and hole

⇑ Corresponding author.
E-mail address: francis.dujardin@univ-lorraine.fr (F. Dujardin).

http://dx.doi.org/10.1016/j.spmi.2015.06.029
0749-6036/Ó 2015 Published by Elsevier Ltd.
582 A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591

Fig. 1. Simplified potential energies representation of a core/shell quantum dot.

energy levels. This technical process so-called ‘‘wave function engineering’’ offers more emergent opportunities for large
optoelectronic applications in different domains: solar cells [11,12], optical encoding [13], high resolution cellular imaging
[14], quantum computing [15], tumor targeting [16] and biosensors [17].
During the growth process, it is possible to willingly add impurities into nanodots [18]. The control of nature and number
of dopant atoms provides extensive tunable optoelectronic properties. It is well known that the binding of both ionized
donors or acceptors with excitons plays a key role in the interpretation of photoluminescence spectrum [19–22]. Some
remarkable properties are attributed to these complex systems such as the origin of UV in ZnO [20,21] or low-threshold las-
ing in ZnSe [23]. The inclusion of a single dopant or ‘‘solitary dopant’’ in a nanocrystal constitute a challenge for several sci-
entists. The systematic explorations using a combination of optical measurements and scanning tunnelling spectroscopy in
the control of the band gap and Fermi energy of doped semiconductor nanocrystals reveal that the properties due to solitary
dopant lead to a new generation of optoelectronic devices such as solar cells and thin-film transistors [24]. We note that
other breakthrough applications of the single dopant in confined medium are now possible or about to emerge; for more
details we refer the reader to Refs. [25,26].
Generally speaking, the engineering of optoelectronic devices based on single dopant depends on several external pertur-
bations (electric field, magnetic field, strain, temperature) and on the impurity position. Many theoretical and experimental
studies have been devoted to the impurities in different configurations. In a series of papers, Cristea and Niculescu [27–29]
have exposed some interesting theoretical results concerning the electronic properties of CdSe/ZnS and ZnS/CdSe CSQD tak-
ing into account the dielectric mismatch and the photoionization charges. Their investigation of the photoionization have
shown a pronounced red shift for low dielectric constant. In addition, they have found that the binding energy of donor states
in ZnS/CdSe under an electric field is blue shifted due to the dielectric confinement. On the other hand the theoretical studies
of linear and nonlinear optical properties of ZnO/ZnS and ZnS/ZnO core–shell dots [30] have shown that in both structures
the increasing of the shell thickness reinforces the nonlinear absorption coefficients and provokes a change in refractive
index independently on the dielectric environment. The theoretical investigation of the oscillator strength of intra and inter-
band quantum transitions in GaAs=Alx Ga1x As spherical QD with an ionized donor placed at the center has shown that the
existence of the impurity in the center of the core prevents the possibility to fabricate light source with multicolor emission
[31]. The linear and third-order nonlinear optical absorption and refractive index have been studied using the compact den-
sity approach and show that the impurity transition is strongly affected by the shell thickness and the transitions between
orbital with bigger l value shift to a higher photon energy region [32]. The position and electric field effects on single dopant
has been subject of some of our last works in which we have attempted to explain the behavior of an off-center single dopant
in thin quantum disk submitted to electric field [33]. The binding energy depends on disk size, donor position, strength and
orientation of the field. We have also analyzed the interactions of a single ionized donor with an exciton [34], our theoretical
prediction shows that the polarizability of the exciton bound to an ionized donor is proportional to R3;5 and the Haynes rule
remains valid even in the presence of electric field. In core/shell structures we have investigated the effect of dielectric
A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591 583

mismatch on the ground state binding energy for both exciton and off-center donor in spherical core/shell nanostructures
[35,36]. We have shown that in the case of intermediate and large size core/shell, off-center donor binding energy is larger
than exciton binding energy for impurity position near the core surface. However for small sizes the off-center donor is more
stable than the exciton for impurity position near the core center. Consequently, it is provided that the core surface effects
will influence optical qualities of the core/shell nanostructures which will foster non-radiative extrinsic transitions to the
detriment of radiative intrinsic transitions. There does not exist to our knowledge any theoretical complete analysis of
the behavior of the off-center donor impurity in core/shell structures immersed in a magnetic field.
In this paper, we report a calculation of the ground-state binding energy of a single donor impurity placed anywhere in a
spherical core/shell nanodot in the presence of a magnetic field directed along the z-axis. We have used the variational
approach within the effective-mass approximation. We have chosen a trial wave function taking into consideration the effect
of the magnetic field and the Colombian interaction between the electron and ionized donor. First, we determine the binding
energy at zero field to confirm our results thus we investigate the influence of a uniform magnetic field and the impurity
position in the shell for strong and moderate confinement regimes. The paper is organized as follows: in the next section,
we present our theoretical approach based on Ritz variational calculation in order to evaluate the ground-state of binding
energy and understand the effect of the position of the donor dopant and shell thickness. The all of our numerical results
and discussion are reported in Section 3.

2. Background theory

Let us consider a spherical core with radius a, coated by another spherical shell with radius b (Fig. 1). A donor impurity is
localized in the shell along the z-axis. We suppose that the two materials used for core and shell have large band offsets, as
CdS/HgS (Eg;CdS ¼ 2:5 eV and Eg;HgS ¼ 0:5 eV) [37]. In this situation the electron can be supposed completely confined in the
shell material. CSQD are typically surrounded by a wide-band dielectric such as a curable resin [38] or a polymer [39] or
another substance with a very large gap which is considered as an infinite wall. Our system is submitted to a magnetic field
!
B ð0; 0; BÞ directed along the z-axis.
In the effective-mass approximation, the Hamiltonian of the system is given by:
H ¼ H0 þ M þ V w : ð1Þ
H0 denotes the Hamiltonian of the unperturbed system, it can be written in spherical coordinates ðr; h; uÞ as:

h2 e2
H0 ¼  Dðr; h; uÞ   
 ð2Þ
2me e~r  ~
d

!
where me is the electron effective mass, e is the dielectric constant taking into account all polarization of the system, ~
r and d
are respectively the positions of the electron and the impurity from the center of the core and e is the electron charge. M
! ! !
describes the magnetic contribution in the Coulomb gauge A ¼ 12 B  r , it can be expressed as:

eB
h e2 B 2 2 2
M¼ L z þ r sin ðhÞ: ð3Þ
2me c 8me c2

h @@u is the z-component of the angular momentum operator. V w is the well potential chosen as an infinite confinement
Lz ¼ i
due to the large difference of the band offsets:

0 a6r6b
V w ðrÞ ¼ ð4Þ
1 elsewhere
In atomic units, the Hamiltonian of the system can be simplified as:
1 2 2
H ¼ D þ cLz þ c2 r 2 sin h    þ V w:
 ð5Þ
4 ~r ~d

We have used the donor units: aD ¼ h  2 e=me e2 , as unit of length and the effective Rydberg RD ¼ e2 =2eaD as unit of energy.
We have also introduced the dimensionless parameter c ¼  hxc =2RD which characterize the magnetic field strength, where
xc ¼ eB=me c is the effective cyclotron frequency. The magnetic field is obtained in teslas by using the following transforma-
  

tion: BðTÞ ¼ me =me l1 B R c, where lB ¼ e h=2me c is the Bohr magneton and me is the free electron mass. The ground state
energy of donor ED0 and the corresponding envelope wave function are solution of the Schrödinger equation:
HWðr; h; uÞ ¼ ED0 ðcÞWðr; h; uÞ: ð6Þ
This equation can not be solved analytically, then the solution is sought using a variational method. The trial wave func-
tion is chosen by following the second perturbation procedure which provides a good approach to treat problems involving
more than one perturbation [41,42,40,43], then we can write:
584 A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591


1
Wðr; h; uÞ ¼ NWe ðrÞ expðar ed Þ exp  cr 2 sin2 ðhÞ ð7Þ
4
  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
r ~
2
where N is the normalization constant, red ¼ ~ d ¼ r 2 þ d  2rd cos h is the electron position relative to the donor impu-
rity and a is the variational parameter to be determined. The first exponential term exp ðared Þ describes the

2
electron-impurity correlation while the second term exp  14 cr2 sin h [41,42,40,46] corresponds to the influence of the
magnetic field on the free electron. We ðrÞ is the Bessel function solution of the Schrödinger equation corresponding to an
electron in a spherical quantum well [44]. By respecting the boundary condition at r ¼ a and r ¼ b [45], we obtain:

sinðkðr  aÞÞ
We ðrÞ ¼ ð8Þ
r
with k ¼ p=ðb  aÞ. The ground state energy of D0 can be obtained by minimizing the expectation value of H with respect to
the variational parameter a:

hWjHjWi
ED0 ðcÞ ¼ min : ð9Þ
a hWjWi
The binding energy Eb ¼ Ee ðcÞ  ED0 ðcÞ is defined as the difference between the first Landau level of the electron without
 p 2
taking into account the coulomb interaction Ee ðcÞ ¼ ba þ c and the impurity ground state energy ED0 ðcÞ obtained by (9),
then we have:
p 2 hWjHjWi
Eb ¼ þ c  min : ð10Þ
ba a hWjWi

3. Results and discussion

Firstly, we start by studying the donor impurity in the absence of the magnetic field (c ¼ 0). In Fig. 2a we display the
donor binding energy Eb with different representative values of the outer radius b as a function of the core/shell ratio a=b
ð0 6 a=b 6 1Þ. We suppose that the donor is placed in the center of the shell d ¼ ða þ bÞ=2. In order to verify the validity
of our calculations we have analyzed the two limits cases a=b ¼ 0 and 1. The first situation (a=b ¼ 0) corresponds to a spher-
ical homogeneous quantum dot (HQD) of radius b, it has been found that the binding energy decreases as the outer radius ðbÞ
increases and finally approaches to 1RD which corresponds to the binding energy of bulk materials (3D results). The second
situation (a=b ¼ 1), corresponds to a very narrow width well. In this case the problem is reduced to a donor-impurity con-
fined on a spherical layer considered as a 2D system. We can see that for all values of the outer radius b the donor binding

4.8
d=(a+b)/2 γ=0
4.4

4.0
b=0.8

3.6
Binding Energy (RD)

3.2
b=1
2.8

2.4 b=2

2.0

1.6 b=4

1.2
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 2a. Variation of the binding energy of a single dopant localized at d ¼ ða þ bÞ=2 as function of the ratio a=b for different confinement regimes
ðb ¼ 0:8; 1; 2; 4aD Þ and for c ¼ 0.
A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591 585

1.6
d=(a+b)/2 γ=0
1.5

1.4
b=4

Ion-electron mean distance <red>


1.3

1.2

1.1 b=2

1.0

0.9
b=1
0.8

0.7
b=0.8
0.6

0.5
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 2b. The ionized donor electron mean distance as function of the ratio a=b without magnetic field for different confinement regimes ðb ¼ 0:8; 1, 2 and
4aD Þ.

energy tends to 4RD which corresponds to the well known 2D limit [47]. We remark that for the strong confinement the
binding energies have a minimum for a critical value ða=bÞcrit , this value moves toward the small values of ratio a=b and van-
ish for larger values of outer radius b. To confirm this description, we have drawn in Fig. 2b the average value of the distance
between the electron and donor impurity hred i as a function of the ratio a=b at zero field. It can be seen that the binding
energy is inversely related to the behavior of the mean radius hr ed i. In the first case (a=b ¼ 0) the mean radius increases
as the outer radius b increases. In the second case (a=b ¼ 1) the electron is localized very close to the ionized impurity center
and the distance between the two charges tends to 0:6aD whatever the shell radius b. We note that our curves are in good
agreement with those obtained in the exciton case [48,49].
Fig. 3 show the simultaneous effect of the magnetic field and the shell size on the donor binding energy Eb . In the strong
spatial confinement case (b < aD , Fig. 3a), the binding energy strongly depends on the magnetic field strength and the shell
size. For a given magnetic field strength the binding energy decreases as the ratio a=b increases and reaches the minimum at

5.4 8,0

7,5
b=0.5 d=(a+b)/2
5.2
7,0
Binding Energy (R D)

5.0 6,5

6,0

4.8
Binding energy (RD)

5,5 γ= 0.5
5,0
4.6
γ= 0
4,5

4.4 4,0
0,0 0,2 0,4 0,6 0,8 1,0

a/b
4.2

4.0
γ=0.5
3.8 b=0.8 γ=0.3

3.6
d=(a+b)/2 γ=0
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 3a. Variation of the binding energy as a function of the ratio a=b for the strong confinement for 2 values of magnetic field strength (c ¼ 0; 0:3 and 0:5Þ.
586 A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591

ða=bÞcrit ’ 0:7aD , beyond this value it increases and tends to 4RD (2D limit). The dotted lines in the inset of Fig. 3a present the
results cited in Ref. [50] using an inadequate wave function, we remark that their approach leads to an anomalous behavior
and does not recover the 2D limit when a=b ! 1. Back to our curves, we can see that for a fixed a=b the binding energies
increases with increasing magnetic field strength. The shift DEb ¼ Eb ðcÞ  Eb ðc ¼ 0Þ as function of the ratio a=b plotted in
Fig. 3b clarifies the influence of the magnetic field. Indeed for a given c; DEb decreases quickly until a=b ’ 0:8aD as the ratio
a=b increases. After this threshold we assist to a slow variation of DEb . Furthermore it is clear that the magnetic field effect is
more pronounced when the core radius is small ða ! 0Þ in comparison to what happened in the spherical surface ða ! bÞ,
this is due to the fact that in the strong confinement case the geometric confinement effect is more important than the mag-
netic field effect. We can conclude that in the surface limit (a ’ b) the binding energy is less sensitive to the magnetic field
effect.
The behavior in the weak confinement is completely different. Fig. 3c shows that when a=b ! 0 the binding energy tends
to the 3D limit for c ¼ 0. With increasing ratio a=b, the binding energy becomes important and tends to 4RD . We remark that

0.50
b=0.8 d=(a+b)/2

0.45

γ=0.5
0.40
ΔEb (RD)

0.35

0.30

γ=0.3
0.25

0.20
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 3b. Variation of the energy shift DEb ¼ Eb ðcÞ  Eb ðc ¼ 0Þ as a function of the ratio a=b for the strong confinement ðb ¼ 0:8Þ for two values of magnetic
field (c ¼ 0:3 and 0:5).

4.5
d=(a+b)/2 b=4

4.0

3.5
Binding Energy (RD)

3.0

2.5

γ=0.5
2.0

γ=0.3
1.5
γ=0

1.0
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 3c. Variation of the binding energy as a function of the ratio a=b for the weak confinement for 3 values of magnetic field strength (c ¼ 0; 0:3 and 0.5).
A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591 587

0.50
b=4 d=(a+b)/2

0.45

0.40 γ=0.5

ΔEb (RD)
0.35

0.30

0.25 γ=0.3

0.20
0.0 0.2 0.4 0.6 0.8 1.0
a/b

Fig. 3d. Variation of the energy shift DEb ¼ Eb ðcÞ  Eb ðc ¼ 0Þ as a function of the ratio a=b for the weak confinement ðb ¼ 0:8aD Þ and for two values of
magnetic field (c ¼ 0:3 and 0.5).

for a=b ranging between 0 and 0.8, the binding energy is slowly increasing, but when the system begins to convert to a spher-
ical layer ð0:8 < a=b < 1Þ this variation becomes more acute. In this situation the electron wave function is more localized
around the impurity center so that the distance between the electron and the ionized donor decreases which increases
the binding energy. Furthermore our results show that the binding energy increases with increasing magnetic field strength
for all values of the ratio a=b. This is due to a shrinkage of the electron wave function and to a diminution of the cyclotron
radius which confines the electron closer to the impurity. The same results are found in HQD [46,41]. In Fig. 3d we present
the difference DEb as a function of the ratio a=b for the weak confinement. Contrary to what happens in strong confinement,
we can notice that the binding energy is less sensitive to the magnetic field effect for the large well ða=b ! 0Þ, but this effect
becomes more remarkable as the ratio a=b increases. At the surface ða ! bÞ, the effect of the magnetic field decreases weakly,
this is due to the change in the geometrical form of the system: from 3D CSQD to 2D spherical layer.
In order to understand the impact of the impurity position on our system, we present in Fig. 4 the binding energy versus
the ratio a=b for three significative ionized donor positions d ¼ ða þ bÞ=2; d ¼ a and d ¼ b respectively for the strong (Fig. 4a)

6.5
b=0.8 γ=0
6.0

5.5
Binding Energy (RD)

5.0 d=a

4.5

4.0
d=(a+b)/2

3.5

3.0
d=b
2.5

0.0 0.2 0.4 0.6 0.8 1.0


a/b

Fig. 4a. Variation of the binding energy as a function of the ratio a=b for 3 values of impurity position d ¼ ða þ bÞ=2; d ¼ a and d ¼ b for the strong
confinement ðb ¼ 0:8aD Þ.
588 A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591

4.5

b=4 γ=0
4.0

3.5

Binding Energy (RD)


3.0

2.5

2.0
d=(a+b)/2
1.5

1.0 d=a

0.5 d=b

0.0 0.2 0.4 0.6 0.8 1.0


a/b

Fig. 4b. Variation of the binding energy as a function of the ratio a=b for 3 values of impurity position d ¼ ða þ bÞ=2; d ¼ a and d ¼ b for the weak
confinement ðb ¼ 4aD Þ.

and the weak (Fig. 4b) confinement. Firstly, we can notice that in absence of magnetic field, when a=b ! 1 the binding
energy tends to the 2D limit (4RD ) whatever the impurity position. In addition we remark that the binding energy is mainly
affected by the impurity displacement for the strong confinement. In the region a=b ! 0 the binding energy is more impor-
tant when the impurity is placed at d ¼ a which corresponds to a centered donor impurity in HQD and it decreases with
increasing a=b. In this case ðd ¼ aÞ we can remark that the binding energy becomes lower than for the d ¼ ða þ bÞ=2 case
for a=b > 0:4 in the strong confinement and a=b > 0:1 in the weak confinement, this is due to the effect induced by increasing
the core size which leads to a change in the structure of the system: from HQS to CSQD. Furthermore, we can notice that for
the strong confinement ðb 6 1aD Þ all the curves present a minimum except when the impurity is placed on the external side
of the shell ðd ¼ bÞ.
To confirm these results and investigate the effect of magnetic field on the impurity position, we present in Fig. 5 the vari-
ation of binding energy as a function of the impurity position for various values of magnetic field strength, we can see that

4.2
γ=0.5
4.0

3.8
γ=0.3
Binding Energy (RD)

3.6 γ=0

3.4

3.2

a=0.5 b=0.8
3.0

2.8

0.50 0.55 0.60 0.65 0.70 0.75 0.80


Impurity position (aD)

Fig. 5. Variation of the binding energy as a function of the impurity position in the spherical surface defined by (a ¼ 0:5aD and b ¼ 0:8aD ) and for 3 values of
magnetic field strength (c ¼ 0; 0:3 and 0.5).
A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591 589

the binding energy reaches the maximum value when the impurity is almost at the center of the well. Furthermore it is clear
that the binding energy increases as the magnetic field increases whatever the impurity position, but this effect is more pro-
nounced when the impurity is in the middle of the shell, and decreases as the impurity move towards the shell boundaries.
To complete our discussion we present in Fig. 6 the cross section of the radial electron probability density given by
r 2 W2 ðrÞ, with and without magnetic field (c ¼ 0 and c ¼ 0:8), for an outer radius b ¼ 4aD and for two different core sizes
a ¼ 1aD and 3aD . We analyze the three particular situations of the ionized donor position in the shell:
d ¼ ða þ bÞ=2; d ¼ a; d ¼ b. As we can see in Fig. 6a in absence of magnetic field and for large shell thickness: b  a ¼ 3aD (first
row), we assist to different behaviors of the electronic density which is very sensitive to the ionized donor position d. We
begin by the case d ¼ ða þ bÞ=2 , the density presents a spherical distribution around the Dþ ion. However when the ion
moves from b to a the density occupies more space and loose its spherical character. At d ¼ a, the orbital overlap spreads
almost over the entire shell. For the low shell thickness: b  a ¼ 1aD (second row), which corresponds to a spherical layer,
the densities are more localized near the Dþ ion. It should be noted that the displacement of the impurity to the inner wall
leads to a spreading of the electron cloud in the upper cap. The effect of the magnetic field on the probability densities is
shown in Fig. 6b. We underline that the simultaneity of the two positive effects (confinement and magnetic effect) contribute
to a strengthening and shrinkage of the orbitals. In addition we remark that for a spherical surface ða ! bÞ, the densities pre-
sent a ‘‘tail’’. Indeed as already remarked in Fig. 3d, at the surface limit, the effect of magnetic field decreases weakly due to

d=(a+b)/2 d=b d=a

b-a=3aD

b-a=1aD

Fig. 6a. Cross section of the probability density r 2 W2 ðrÞ for the ground state of an off-center donor impurity (for the three Dþ positions: d ¼ ða þ bÞ=2; d ¼ b
and d ¼ a) without magnetic field (c ¼ 0).

d=(a+b)/2 d=b d=a

b-a=3aD

21

b-a=1aD

Fig. 6b. Cross section of the probability density r 2 W2 ðrÞ for the ground state of an off-center donor impurity (for the three Dþ positions: d ¼ ða þ bÞ=2; d ¼ b
and d ¼ a) with magnetic field (c ¼ 0:8).
590 A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591

b=4 d=(a+b)/2
1.6

1.4

Ion-electron mean distance <red >


γ=0

1.2

γ=0.8
1.0

0.8

0.6

0.0 0.2 0.4 0.6 0.8 1.0


a/b

Fig. 7. Variation of the distance between the electron and the ionized donor as a function of the ratio a=b with and without magnetic field (c ¼ 0 and 0:8Þ.

the change geometrical from a 3D CSQD to a 2D spherical layer. This behavior is in harmony with the curves of hr ed i given in
Fig. 7. We observe that the mean radius decreases as the ratio a=b increases, i.e. when the well thickness becomes thinner. In
addition it is clear that hr ed i decreases as the magnetic field strength increases, this effect is more pronounced in a large shell
ða=b ! 0Þ than in a thin spherical layer ða=b ! 1Þ.

4. Conclusion

The properties of single dopant in core shell quantum dots are very sensitive to the external perturbations. Several param-
eters contribute in the change in electronic and optical properties: core size and shell thickness, perturbation strength,
dielectric environment and inclusion of dopant atoms. To our knowledge this is the first complete study concerning the sin-
gle dopant in core shell structure submitted to a magnetic field. Our approach have been achieved in the framework of the
effective mass approximation and by considering an infinite deep potential barrier. We have numerically solved the
Schrödinger equation by using the Ritz variational method. The analysis of the dependence of the binding energy and the
cross section of the densities on the simultaneous effects: shell thickness, magnetic field strength and dopant position
has shown some interesting results which we hope that they will contribute in the comprehension of the behavior of single
dopant and allow to explain the optical phenomena related to the magnetic field for eventual applications.

References

[1] P. Harrison, Quantum Wells, Wires and Dots, Wiley, NY, 2006.
[2] D. Bimberg, M. Grundmann, N.N. Ledentsov, Quantum Dot Heterostructures, Wiley, Chichester, New York, 1999.
[3] G. Schmid, Nanoparticles From Theory to Applications, second ed., Wiley-VCH Verlag GmbH & Co. KGaA, 2010.
[4] M.P. Bruchez, C.Z. Hotz, Quantum Dots Applications in Biology, Quantum dot Corporation Hayward, CA, Humana Press Inc., Totowa, New Jersy, 2007.
[5] T. Steiner, Semiconductor Nanostructure for Optoelectronic Applications, Artech House Semiconductor Materials and Devices Library, 2004.
[6] P. Mélinon, S. Begin-Colin, J.L. Duvail, F. Gauffre, N.H. Boime, G. Ledoux, J. Plain, P. Reiss, F. Silly, B. Warot-Fonrose, Phys. Rep. 543 (2014) 163–197.
[7] A.R. Kortan, R. Hull, R.L. Opila, M.G. Bawendi, M.L. Steigerwald, P.J. Carroll, L. Brus, J. Am, Chem. Soc. 112 (1990) 1327.
[8] H.S. Zhou, I. Honma, H. Komiyama, J. Phys. Chem. 97 (1993) 895.
[9] A. Mews, A. Eychmuller, M. Giersig, D. Schooss, H. Weller, J. Phys. Chem. 98 (1994) 934.
[10] J.W. Haus, H.S. Zhou, I. Honma, H. Komiyana, Phys. Rev. B 47 (1993) 1359.
[11] Prashant V. Kamat, J. Phys. Chem. C 112 (2008) 18737–18753.
[12] M. Bar, S. Lehmann, M. Rusu, A. Grimm, I. Kotschau, I. Lauermann, P. Pistor, S. Sokoll, Th. Schedel-Niedrig, M.Ch. Lux-Steiner, Ch.-H. Fischerb, L.
Weinhardt, C. Heske, Ch. Jung, Appl. Phys. Lett. 86 (2005) 222107.
[13] Zhenda Lu, Chuanbo Gao, Qiao Zhang, Miaofang Chi, Jane Y. Howe, Yadong Yin, Nano Lett. 11 (2011) 3404–3412.
[14] Liam R. Bradshaw, Andreas Hauser, Emily J. McLaurin, Daniel R. Gamelin, J. Phys. Chem. C 116 (2012) 9300–9310.
[15] L.C.L. Hollenberg, A.S. Dzurak, C. Wellard, A.R. Hamilton, D.J. Reilly, G.J. Milburn, R.G. Clark, Phys. Rev. B 69 (2004) 113301.
[16] Xiaohu Gao, Yuanyuan Cui, Richard M. Levenson, Leland W.K. Chung, Shuming Nie, Nat. Biotechnol. 22 (2004) 969–976.
[17] Igor L. Medintz, Aaron R. Clapp, Hedi Mattoussi, Ellen R. Goldman, Brent Fisher, J. Matthew Mauro, Nat. Mater. 2 (2003) 630.
[18] Takahiro Shinada, Shintaro Okamoto, Takahiro Kobayashi, Iwao Ohdomari, Nature 437 (2005) 1128–1131.
[19] F. Dujardin, E. Feddi, E. Assaid, A. Oukerroum, Eur. Phys. J.B 74 (2010) 507.
[20] V.A. Fonoberov, A.A. Balandin, Appl. Phys. Lett. 85 (2004) 5971.
[21] V.A. Fonoberov, A.A. Balandin, Phys. Rev. B 70 (2004) 195410.
A. Talbi et al. / Superlattices and Microstructures 85 (2015) 581–591 591

[22] L. Dallali, S. Jaziri, J. Martinez-Pastor, Solid State Comm. 209–210 (2015) 33–37.
[23] A. Pawlis, M. Panfilova, D.J. As, K. Lischka, K. Sanaka, T.D. Ladd, Y. Yamamoto, Phys. Rev. B 77 (2008) 153304.
[24] David Mocatta, Guy Cohen, Jonathan Schattner, Oded Millo, Eran Rabani, Uri Banin, Science 77 (2011) 332.
[25] Paul M. Koenraad, Michael E. Flatté, Nat. Mater. 10 (2011) 91–100.
[26] Daniel Moraru, Arief Udhiarto, Miftahul Anwar, Roland Nowak, Ryszard Jablonski, Earfan Hamid, Juli Cha Tarido, Takeshi Mizuno, Michiharu Tabe,
Nanoscale Res. Lett. 6 (2011) 479.
[27] M. Cristea, E.C. Niculescu, Eur. Phys. J.B 85 (2012) 191.
[28] M. Cristea, E.C. Niculescu, Phys. Lett. A 377 (2013) 1221.
[29] E.C. Niculescu, M. Cristea, J. Lumin. 135 (2013) 120–127.
[30] Zaiping Zeng, Christos S. Garoufalis, Andreas F. Terzis, Sotirios Baskoutas, J. Appl. Phys. 114 (2013) 023510.
[31] V. Holovatsky, I. Bernik, O. Voitsekhivska, Acta Phys. Pol., A 125 (2014).
[32] A.R. Jafari, Physica B 456 (2015) 72–77.
[33] F. Dujardin, A. Oukerroum, E. Feddi, J.B. Bailach, J. Mart ínez-Pastor, M. Zazi, J. Appl. Phys. 111 (2012) 034317.
[34] E. Feddi, A. Zouitine, A. Oukerroum, F. Dujardin, E. Assaid, M. Zazoui, J. Appl. Phys. 117 (2015) 064309.
[35] A. Ibral, A. Zouitine, E. Assaid, E. Feddi, F. Dujardin, Physica B 449 (2014) 261–268.
[36] A. Ibral, A. Zouitine, S. Aazou, E. Assaid, E. Feddi, F. Dujardin, J. Optoelectron. Adv. Mater. 15 (2013) 11–12.
[37] D. Schooss, A. Mews, A. Eychmuller, H. Weller, Phys. Rev. B 49 (1994).
[38] A. Joshi, K.Y. Narsingi, M.O. Manasreh, E.A. Davis, B.D. Weaver, Appl. Phys. Lett. 89 (2006) 131907.
[39] T.V. Torchynska, J. Douda, R. Pena Sierra, Phys. Status Solidi C 6 (2009) S143.
[40] J. El Khamkhami, E. Feddi, E. Assaid, F. Dujardin, B. Stébé, M. El Haouaria, Physica E 25 (2005) 366–373.
[41] Z.G. Xiao, J.Q. Zhu, F.L. He, Phys. Stat. Sol. B 191 (1995) 401.
[42] Z. Xiao, J. Appl. Phys. 86 (1999) 4509.
[43] H.X. Jiang, Phys. Rev. B 35 (1987) 9287.
[44] M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions, Dover, New York, 1970.
[45] J.M. Ferreyra, C.R. Proetto, Phys. Rev. B 57 (1998) 9061.
[46] Aram Kh. Manaselyan, Albert A. Kirakosyan, Physica E 22 (2004) 825.
[47] S. Legoff, B. Stébé, J. Phys. B: At. Mol. Opt. Phys. 25 (1992) 5261.
[48] Y. Kayanuma, Solid State Commun. 59 (1986) 405.
[49] J. El Khamkhami, E. Feddi, E. Assaid, F. Dujardin, B. Stébé, J. Diouri, Physica E 15 (2002) 99–106.
[50] Haddou El Ghazi, Anouar Jorio, Izeddine Zorkani, Physica B 427 (2013) 106–109.

Potrebbero piacerti anche