Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science in Semiconductor Processing 124 (2021) 105614

Contents lists available at ScienceDirect

Materials Science in Semiconductor Processing


journal homepage: http://www.elsevier.com/locate/mssp

Full length article

Theoretical study of excitons in GaAs quantum dot molecules obtained by


nanoholes filling
S. Tilouche a, *, A. Sayari b, c, M. Omri d, S. Souilem a, e, L. Sfaxi a, e, R. M’Ghaieth a
a
Laboratoire de Micro-Optoélectronique et Nanostructures LR99ES29, Faculté des Sciences de Monastir, Université de Monastir, Monastir, 5019, Tunisia
b
Department of Physics, Faculty of Science, University of Jeddah, P.O. Box 80327, Jeddah, 21589, Saudi Arabia
c
Université de Tunis El Manar, Faculté des Sciences de Tunis, Unité de Recherche Spectroscopie Raman UR13ES31, Tunis, 2092, Tunisia
d
Deanship of Scientific Research (DSR), King Abdulaziz University, Jeddah, 21589, Saudi Arabia
e
High School of Sciences and Technology of Hammam Sousse, Université de Sousse, Tunisia

A R T I C L E I N F O A B S T R A C T

Keywords: In this work we report a theoretical study of excitons behavior in self-assembled strain-free GaAs quantum dot
GaAs quantum dot molecules (QDMs) obtained by nanoholes filling. More precisely, we study a single molecule formed by a
Nanoholes vertically coupled GaAs/AlGaAs quantum dots pair (QDP). The vertical coupling governed by varying the spacer
Envelope function method
thickness layer reveals how the excitons behave in a QDM. A model based on resolving the three-dimensional
Excitonic complexes
effective-mass Schrödinger equation is adopted to reliably provide a complete overview of the fine structure
of neutral and complexes excitons and to explain how a pair of quantum dots interact in between while varying
the spacer layer thickness mediating the coupling in the barrier. Our calculations can be adopted to improve the
growth conditions and boost the optical properties of strain-free GaAs QDMs for using them in quantum infor­
mation technologies and quantum computing.

1. Introduction To estimate the effect of confinement on the HH – LH mixing, a


commonly used approach is to consider the Kohn-Luttinger Hamiltonian
Analogously to ordinary atoms, quantum dots (QDs) have confined HKL. In order to analytically solve these equations, Fischer and Loss [14]
energy levels. Vertically self-assembled GaAs/AlGaAs QDs can be modeled QDs using a parabolic potential of cylindrical symmetry along
considered as an artificial diatomic molecule when quantum mechanical Oz. From the non-diagonal terms of HLK, they estimated the HH – LH
tunneling of electrons and holes between two QDs creates delocalized mixing to be 0.2%. This result has also been confirmed by Luo et al. [13],
molecular states [1,2]. Optical properties of such quantum dot mole­ in a different approach using the pseudo-potential. Meanwhile, the QDs’
cules (QDMs) have drawn more attention for the last two decades due to shape anisotropy in the plane is often considered to be one of the pa­
the rising interest of using them in tunable devices for being entangled rameters causing mixing [15,16]. Liao et al. [17] estimated the order of
photons sources [3] and to create excitons, qubits to be specific [4–9]. magnitude of this effect to be 0.4%. The little influence of planar shape
Thereby, providing essential components for quantum cryptography anisotropy was also predicted by Luo et al. [13]. Note that in our model
[10] and quantum information processing [6,11,12]. we considered isotropic QDs and therefore the shape anisotropy does not
Most of the studies report that the heavy hole – light hole (HH – LH) intervene in the HH – LH mixing. Furthermore, in self-assembled QDs,
coupling is dependent on several parameters. To discuss the different such as InAs/GaAs, whose formation is based on the difference in lattice
key contributions affecting the mixing of heavy and light holes, we will parameters, the strain brings considerable mixing of HH and LH hole
take the classification recently presented by Luo et al. [13]. Five con­ states that can reach 80% [13]. Moreover, the effect of strain, modeled
tributions can be at the origin of the HH – LH mixing; the 3D confine­ by the Bir-Pikus Hamiltonian HBP, constitutes the most important
ment, and quite particularly along Oz, the planar shape anisotropy, the contribution to the HH – LH mixing [15]. However, for GaAs/AlGaAs
strain effects, the alloy disorder and finally the symmetry breaking; QDs the aforementioned strain effect is almost null on the HH – LH
extrinsic, linked to the interfaces and intrinsic, linked to the crystallo­ coupling given the absence of strain in these nanostructures which is the
graphic structure (Zinc Blende for InAs and GaAs). case in our study being used as a quantum molecule. Note also, that a

* Corresponding author
E-mail address: samirtilouche@gmail.com (S. Tilouche).

https://doi.org/10.1016/j.mssp.2020.105614
Received 18 July 2020; Received in revised form 24 November 2020; Accepted 5 December 2020
Available online 22 December 2020
1369-8001/© 2020 Elsevier Ltd. All rights reserved.
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

Fig. 1. Schematic illustration of the Shape modelled GaAs QDM obtained by AlGaAs nanohole filling.

similar trend was recently observed in Csontosova et al.’s work [18] In this work we provide an approach that allows to model a well-
which deals with a similar system as ours, and it was found that since defined QDM shape with appropriate tunnel barriers for excitonic
GaAs and AlGaAs are nearly lattice matched, the strain is rather small, states accurate calculations. We determine the neutral excitons energies
which results in small value of HH – LH mixing, similarly as in Ref. [19]. and we calculate the energy splitting of the excited P and S states by
In addition, during the growth of QDs, there may be a diffusion of ma­ numerically integrating the three dimensional effective-mass
terials resulting in alloys, whether homogeneous or not. This is partic­ Schrödinger equation [2] for the system consisting of an electron
ularly the case with In1-xGaxAs/GaAs or GaAs/Al1-xGaxAs QDs. In the interacting with a heavy hole through direct Coulomb attraction and by
case of In1-xGaxAs/GaAs QDs (as well as in GaAs/AlGaAs QDs), the taking into account the correlation effects in a QDM consisting of a
alloying effects are important on the exciton’s fine structure [20] or the bottom dot (B-dot) and a top dot (T-dot). Finally, we provide the exci­
optical polarization [21], but are negligible (<0.2%) on the HH – LH tonic complexes binding energies defined with respect to the neutral
mixing, as recently calculated [13]. Nevertheless, the effect of an exciton energy, the wave functions and the position of the recombining
interface was estimated also by Luo et al. [13], by first modeling QDs as electrons and holes as they may be recombined by carriers tunneling in a
cylinders (or disks), thus giving D2d symmetry to the QD shape. The QDM.
pseudo-potential approach predicts an HH – LH mixing, in GaAs/AlGaAs
QDs, with an LH proportion of 2–4%. Similarly, as in Ref. [18], a small 2. Theoretical model
LH proportion is observed in GaAs/AlGaAs QDs as well.
Speaking of the symmetry breaking, QDs have a shape with lower The adopted model consists of modeling two GaAs QDs separated by
symmetry than D2d. Lenses, pyramids or nano-holes have C2v symmetry, an AlGaAs barrier, thus forming a strain free QDM as a system since the
or weaker in case of deformation. When this shape effect is taken into lattice parameters of GaAs and AlGaAs match almost perfectly. From
account, the presence of curved interfaces accentuates the non- recent GaAs QDs AFM images, we have chosen a 64 nm cylinder height
equivalence of the [110] and [110] axes. Luo et al. [13] have thus of radius R and height H as a quantization box in which we embedded
shown that, for QDs with a Gaussian profile, the LH component reached two QDs with shape based on Heyn et al.’s model [22] as illustrated in
10%. In the range of large quantum dots and the considered materials Fig. 1. To avoid side-effect problems, these dimensions are chosen such
(GaAs/AlGaAs), recent theoretical works on the HH – LH mixing as R = 2rnanohole and H = 4hnanohole where r and h are respectively the
(atomistic pseudo-potential, k.p calculations) seems to indicate that the radius and the height of the nanohole. The resulting QDM model is a pair
lowering of intrinsic symmetry is the main cause of the mixing. Finally, of QDs inverted crescents shape like with an elliptic base area where the
in our model, we did not take this effect into account but we propose to base thickness of bottom dot is chosen to be 5 nm, and that of the top dot
provide a shape modelization of GaAs/AlGaAs QDs, obtained by nano­ to be 4.4 nm initially to mimic the fact that experimentally the upper
holes filling, compatible with atomic force microscopy images (AFM) dots are larger than the bottom dots when grown in a nanohole [23]. We
[22]. For completeness, Csontosova and colleagues [18] recently clearly choose a 16 nm nanohole height with a radius of 30 nm and a barrier
stated that the HH – LH coupling is not influenced by the effect of cor­ layer of Al0.36Ga0.44As. We also allow the spacer thickness to vary in
relation and it is weakly affected by the direct and the exchange order to explore possibilities of improving the existing structures.
Coulomb interaction. In order to study the spacer’s effect, we use an effective-mass

2
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

Fig. 2. Electrons and holes confinement energies of the ground state S and the excited state P as function of the spacer thickness.

Hamiltonian based on BenDaniel et al.’s Hamiltonian [24]. √̅̅̅̅ ( ( ))


⎛ ⎞ 2 1 z
( ) Zl (z) = sin lπ −
ℏ2 H 2 H
Н e(h) = ∇⎝ − ( )⎠∇ + Ve(h) →r e(h) (1)
2m*e(h) →r e(h) H H
− ≤z≤ (5)
2 2
Where the sub-indices e and h refer to electron and hole and m*e and m*h 1
are respectively, effective masses of electrons and holes. The values of θm (φ) = √̅̅̅̅̅eimφ

effective masses used in our calculations (in unit of free electron mass)
are m*e = 0.067 and m*h = 0.51, in the quantum well GaAs material and 0 ≤ ϕ ≤ 2π (6)
m*e = 0.093 and m*h = 0.60, in the AlGaAs barrier material. The band
offsets for electrons and holes are respectively, Ve = 313 meV and Vh = Where n, m and l are integers. Jm is the Bessel function of order m and
205 meV [25] and the low temperature GaAs energy gap is Eg = 1519 knmr its nth root.
( ( ) )
meV. At each point inside the cylinder we can associate a position vector ∫
The matrix element Hnml,n′ m′ l′ = Φ*nml ∇ − 2m ℏ2
∇ +V Φn′ m′ l′ dΩ
*
specific to the electrons and to the holes denoted respectively →re and →
rh Ω

and having three variables x, y and z. (7), with Ω the cylinder volume.
We numerically obtain the binding energies by solving Marzin et al.’s Then we expand the model by introducing Coulomb interaction as a
matrix [26] and by considering the cylindrical symmetry of the QDs, the second-order disturbance [27] and then by adding correlation terms.
resulting wave functions are expressed by Fourier-Bessel series devel­ Moreover, we only compute the direct Coulomb integrals because the
oped as: exchange terms are of the order of 10–30 μeV [28], therefore very low
∑ compared to the direct Coulomb terms and the correlation terms. The
ψ= Аnml Φnml (2) aforementioned Coulomb interaction influences essentially the spectral
nml
positions of emission of corresponding electron-hole complexes via its
binding energies [29].
Where Anml are coefficients to be determined and Φnml are periodic and
orthogonal functions written as:
3. Modelling results
Φnml = Rnm (r)Zl (z)θm (φ) (3)
√̅̅̅ We have performed calculations up to the fifth excited exciton state.
Rnm (r) =
2
Jm (knm r) Although, we will only focus on the two lowest energy states denoted En
RJm+1 (knm R) and the confining energy of heavy holes denoted HHn, where n = 1
corresponds to ground state referred as S state and n = 2 corresponds to
0≤r ≤ R (4) the first excited state referred as P state. The results depicted in Fig. 2
show that the electrons confinement energies increase significantly until
the spacer thickness reaches 2.5 nm then, the energies become constant
and independent of the spacer thickness for values up to 6 nm. However,

3
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

Fig. 3. Electrons and holes distribution in the GaAs QDM as function of the spacer thickness.

the heavy holes’ confinement energies seem to slightly increase and regime, the QDM behaves as a single QD of height equal to the sum of the
become moderately constant when the spacer thickness exceeds 1 nm. upper and lower dot heights (9.4 nm). In addition, as the spacer is
We also notice that the electrons confinement energies are approxi­ increased, we notice that electrons wave function tends to be mostly
mately three times higher than those of heavy holes as a result of the localised in the B-dot and it appears to be way sensitive to the barrier
huge effective mass of heavy holes compared to that of electrons. thickness compared to the holes wave function. As stated above, this is
As a complement to the confinement energies, we also provide an mostly due to the higher effective masses of heavy holes. Fig. 4 illus­
overview of the wave functions (WFs) behaviour resulting from trates the side view of the calculated WFs in dependence on the spacer
increasing the spacer thickness. In Fig. 3(a) and (b), the probability of thickness. It will be useful in the following to define the nature of the
finding an electron or a hole whether in the B-dot or the T-dot is illus­ bound states. Speaking of the integral overlap values, as the spacer
trated. As expected, we note that the QDs in our QDM are likely to act as thickness increases, then the overlap integral gets closer to the minimum
a single QD before introducing a spacer layer and while increasing its for both electrons and holes WFs which is quite noticeable in Fig. 4. And
thickness until a certain value which is of 1 nm in our case and act as hence, the WFs are called orthogonal. Yet, as the spacer thickness gets
individual QDs isolated from each other when exceeding 3 nm of closer to zero, then the WFs overlap and the overlap integral increases
thickness in our model. Otherwise, it’s worth noticing that a vertical (Table 1). Furthermore, the aforementioned WF reflects the tunnelling
coupling occurs according to the electrons and holes WFs distribution rate between the upper and lower dot. Thus far, the molecular behaviour
amongst the bottom and top dots. Thus, we distinguish the existence of of our modelled structure is confirmed due to the resulting tunnelling
three coupling regimes and this according to the value of the spacer. A between the QDs sensitive to the spacer thickness variation. Quite
weak coupling regime appears for a spacer thicker than 3 nm. In this similar findings were reported in Křápek et al.’s calculations [30].
regime, the QDM behaves like two completely decoupled QDs. In fact, This part discusses the excitonic energies of the ground state S and
the electrons and holes WFs corresponding to the fundamental state are the P state denoted EsX and EpX calculated by the following expression:
localized in the B-dot (hB-dot = 5 nm) which justifies that the probabil­
ities of the presence of electrons and holes are equal to 1 in the B-dot and EXs(p) = Eg (GaAs) + E1(2) + HH1(2) + Jeh
s(p)
(8)
zero in the T-dot. The intermediate coupling regime is established for a The calculated direct Coulomb attraction term for both S and P states
spacer between 1 and 3 nm. In this regime, the WF describing the ground are summarized in Fig. 5. We find that the excitonic energies increase as
state of the QDM is a linear combination of that of the two QDs. Yet, with the thickness of spacer increases.
1.25 nm of spacer thickness, we report that the electrons wave function Generally, in the case of a strong confinement regime, the direct
is evenly localized in both QDs. Somehow, physically this is not Coulomb attraction term is relatively weak compared to the confine­
consistent and does not relatively agree well with the experiment unless ment potential. Hence, Jeh can be considered as a disturbance [31] given
if we were to choose the same shape, geometry and thickness for the QDs by the following equation:
pair. But theoretically this approach should not worry us too much since
∫ ∫
in this case, we consider that we have two symmetrical QDs, besides, we e2 ⃒ s(p) ⃒2 1
(9)
s(p) ⃒ψ ⃒ s(p) 2
Jeh = − |ψ | dΩe dΩh
are dealing with the envelope function approximation and the WFs’ 4πε0 εr Ωe Ωh e ‖→
re − →rh ‖ h
proportion are being extracted by a Gaussian fit on the WFs. Hereby, we
assume that the tunnelling between the WFs in the upper and lower dot Where ϵr is the GaAs relative dielectric constant [32] and, Ωe and Ωh
is governed by the spacer thickness and we suggest that our hypothetic represent the volume of the outer cylinder associated with electrons and
system can form a QDM. The third regime corresponds to the strong holes respectively. The WFs associated with electrons and holes are
coupling regime. It’s observed for a spacer less than 1 nm. For this described inside the outer cylinder containing the modelled molecule

4
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

the integral is. In our case, we have considered Nx = Ny = Nz = 121.


Fig. 5 reveals the essences of the direct Coulomb attraction values in
dependence on the spacer thickness. As we can see from Fig. 5, the
⃒ s(p) ⃒
thicker the spacer is, the more the value of ⃒J ⃒ increases except when
eh
the spacer reaches 0.5 nm for the ground state, it decreases. Moreover,
for the excited state P, the direct Coulomb attraction effect decreases as
well when the barrier separating the two coupled QDs reaches 1 nm. We
assign this decrease to the wave functions overlap integral evidenced
only for the P state consequence of the WF small spatial extension of the
ground state compared to the excited state. We note also that the impact
of the direct Coulomb attraction is further pronounced for the S state
than for the P state due to the WF spatial extension as mentioned above.
Similar behaviour is reported in Ref. [33].
Control over the spacer thickness was based on Graf et al. experi­
mental data [34]. According to the chosen parameters, we cannot
exceed a thickness of 6 nm taking into account the QDs pair heights. As
assumed before, the spacer thickness appears to be a relevant parameter
to be considered. In the following, we focus on the barrier thickness
effects on the excitonic complexes without and with the correlation ef­
fect. The four main transitions: the neutral exciton (X), the neutral
biexciton (XX), the positively charged exciton (X+), and the negatively
charged exciton (X− ) in our modelled QDM are extracted from our
computational data and plotted in Fig. 6. For each spacer thickness, we
have calculated the exciton binding energies (BEs), taking into account
the correlation effect.
The recombination energies of the neutral exciton, charged excitons
and biexciton are calculated as follows:
⃒ s⃒
ex0 = Eg + E1 + HH1 − ⃒Jeh ⃒ (10)
⃒ s⃒
eX+ = Eg + E1 + HH1 − 2⃒Jeh ⃒ + Js
hh (11)
⃒ s⃒
eX− = Eg + E1 + HH1 − 2⃒Jeh ⃒ + Js
ee (12)
⃒ s⃒
eXX = Eg + E1 + HH1 − 3⃒Jeh ⃒ + Js + Js
ee hh (13)

It’s noteworthy that the biexciton and positive trion exhibit positive
BEs, thus forming an unbound state reaching 3.57 meV and 5.33 meV
respectively. Note that the XX BE values are likely in the same range as
those depicted from Ref. [35,36] where the BE of the XX vary between 1
and 6 meV. This fact can arise from the vertical carriers’ localization
when the barrier separating the two QDs is 1 nm thick. Moreover, un­
bound biexcitons are often noticed in some QDs such as nitride polar
DQs [37–39] and in nonpolar strained QDs [40]. We also note that as the
spacer thickness increases, the BEs decrease. Yet, the BE of X− is always
negative, thus forms a bound state which is in good agreement with the
Fig. 4. Probability densities side view of electrons and holes in the z-direction findings reported in Ref. [41]. However, the correlation effect becomes
in the GaAs QDM for the ground state as function of the spacer thickness. more effective and allows the formation of bound excitonic states where
the BEs considering the correlation effect are calculated as follows:
(Fig. 1). Note that the case where re = rh is excluded. Considering that ⃒ s⃒
s
ΔcX+ = Jhh − ⃒Jeh ⃒ + δcX+ − δcX (14)
the term with |re −1 rh | for re = rh should be the dominant contribution to the
direct Coulomb integral. To compensate for that infinity, we have dis­ s
ΔcX− = Jee
⃒ s⃒
− ⃒Jeh ⃒ + δcX− − δcX (15)
cretized the outer cylinder into Nx intervals along x, Ny intervals along y
and Nz intervals along z. The direct Coulomb term is an integral on the s
ΔcXX = Jee s
+ Jhh
⃒ s⃒
− 2⃒Jeh ⃒ + δcXX − 2δcX (16)
three coordinates of the electron (xe , ye and ze ) and on the three co­
ordinates of the hole (xh , yh and zh ). The numerical calculation of this With δcX , δcX+ , δcX− and δcXX the correlation energies written in the
integral is equivalent to the sum of (Nx .Ny .Nz)2 terms from which we following form:
have excluded the case corresponding to x e = xh and y e = yh and ze =
δcX = Weh (17)
zh . The smaller the discretization step is, the more precise the value of

Table1
Overlap integrals values of both electrons and holes as function of the spacer thickness.
Spacer thickness (nm) 1 1.25 1.5 2 2.5 3 3.5 3.75

Electrons 0.467 0.462 0.452 0.401 0.292 0.175 0.134 0.081


Holes 0.160 0.209 0.228 0.244 0.243 0.211 0.091 0.054

5
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

Fig. 5. Direct Coulomb attraction values of S and P states as function of the spacer thickness.

Fig. 6. The four main transitions: the neutral exciton (X), the neutral biexciton (XX), the positively charged exciton (X+), and the negatively charged exciton (X− ) in
the GaAs modelled QDM as function of the spacer thickness.

δcX+ = 2Weh + Whh (18) δcX− = 2Weh + Wee (19)

6
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

Fig. 7. Variation of correlation energies of X− , X+ excitons and biexciton XX as function of the spacer thickness (a). Bond Energies of excitons X− , X+ and biexciton
XX (Δ without correlation effect (b), Δc with correlation effect (c)) as function of the spacer thickness.

δcXX = δcX+ + δcX− (20) 4. Conclusion

With
In conclusion, by numerically solving the three-dimensional strain-
⃒ ⃒
⃒ ni mj ,00 ⃒2 free Schrödinger equation with inclusion of direct Coulomb attraction
∑ ⃒Uij ⃒
(21) and by taking into account the correlation effect while varying the
Wij = ( i ) ( )
j
=(1,1) E1 + E1 −
(ni ,mj )∕ Eni i + Emj j spacer thickness, we provide a complete description of the dynamics of
excitons in a GaAs quantum dot molecule. Our findings are proved to be
Where in very good agreement with realistic calculations. Moreover, the un­
∫ ∫ derstanding of these properties and the control over the barrier sepa­
1 1 m ( ) ( ) → → rating the quantum dots are of paramount importance for the
Ψ ni i (→
ri )Ψ 0i (→
ri ) → → Ψ j j → rj Ψ 0j → (22)
n mj ,00
Uiji = rj d re d rh
4πεo
Ωe Ωh
‖ re − rh ‖ improvement of growth conditions and optical properties of strain-free
GaAs QDs based devices and for the incoming quantum entanglement-
With 1 ≤ ni(j) ≤ 5; where i and j denote the type of the particle based technologies and quantum information processing.
involved, namely an electron (e) or a hole (h). Ei1 and Ej1 are the ground
state energies of a single particle (i or j). Eini and Ejmj are the energies of Declaration of competing interest
the excited states defined by the indexes ni and mj.
As a result, XX becomes bounded for thick barriers except when the The authors declare that they have no known competing financial
barrier is 1 nm thick, consistently with previous studies established on interests or personal relationships that could have appeared to influence
strained InAs/GaAs QDs systems [42]. We also notice that the X− BE is the work reported in this paper.
further suppressed, consistent with M. Abbarchi et al. findings on GaAs
QDs [33], and the X+ BE shows positive and negative values, exhibiting Acknowledgements
a transition from the unbound state to the bound state when the spacer is
between 1 and 2 nm. We assign this behaviour to the strong coupling This project was funded by the Deanship of Scientific Research
between the B-dot and the T-dot as the holes may tunnel since the (DSR), King Abdulaziz University, Jeddah, under Grant no. (D-505-305-
confinement is stronger for electrons than for holes besides it is in stark 1441). The authors, therefore, gratefully acknowledge DSR technical
contrast to higher dimensional systems. The variation of the correlation and financial support.
energies and the excitonic complexes BE without and with correlation
effect in dependence on the barrier thickness are presented in Fig. 7. References
Moreover, it’s worth noticing that the correlation effect plays a major
[1] M.V. Durnev, M. Vidal, L. Bouet, T. Amand, M.M. Glazov, E.L. Ivchenko, P. Zhou,
role in aligning the excitonic complexes since the changes on the XX and
G. Wang, T. Mano, N. Ha, T. Kuroda, X. Marie, K. Sakoda, B. Urbaszek, Phys. Rev. B
X+ BEs are quite significant and sensitive to the variation of the barrier 93 (2015) 245412.
thickness as we note that |δc (X)| < |δc (X− )| < |δc (X+ )| < |δc (XX)|. [2] M. Bayer, P. Hawrylak, K. Hinzer, S. Fafard, M. Korkusinski, Z.R. Wassilewski,
This was also witnessed in A. Schliwa et al. findings [43]. O. Stern, A. Forchel, Science 291 (5503) (2001) 451.
[3] A.J. Shields, Nat. Photon. 1 (2007) 215.

7
S. Tilouche et al. Materials Science in Semiconductor Processing 124 (2021) 105614

[4] D. Loss, D.P. Divincenzo, Phys. Rev. A 57 (1998) 120. [25] Wei Yi, Venkatesh Narayanamurti, Hong Lu, Michael A. Scarpulla, Arthur
[5] D. Kim, S.G. Carter, A. Greilich, A.S. Bracker, D. Gammon, Nat. Phys. 7 (2011) 223. C. Gossard, et al., Appl. Phys. Lett. 95 (2009) 112102.
[6] G. Cao, H.-O. Li, T. Tu, L. Wang, C. Zhou, M. Xiao, G.-C. Guo, H.-W. Jiang, G.- [26] J.Y. Marzin, G. Bastard, Solid State Commun. 92 (1994) 437.
P. Guo, Nat. Commun. 4 (2013) 1401. [27] Ph Lelong, G. Bastard, Solid State Commun. 98 (1996) 819.
[7] K.M. Weiss, J.M. Elzerman, Y.L. Delley, J. Miguel-Sanchez, A. Imamoglu, Phys. [28] Y.H. Huo, V. Křápek, A. Rastelli, O.G. Schmidt, Phys. Rev. B 90 (2014) 41304(R).
Rev. Lett. 109 (2012) 107401. [29] G. Sallen, S. Kunz, T. Amand, L. Bouet, T. Kuroda, T. Mano, D. Paget, O. Krebs,
[8] A. Greilich, S.G. Carter, D. Kim, A.S. Bracker, D. Gammon, Nat. Photon. 5 (2011) X. Marie, K. Sakoda, B. Urbaszek, Nat. Commun. 5 (2014).
702. [30] V. Křápek, P. Klenovsky, A. Rastelli, O.G. Schmidt, D. Munzar, J. Phys. Conf. 245
[9] G. Bester, A. Zunger, J. Shumway, Phys. Rev. B 71 (2005), 075325. (2010), 012027.
[10] N. Akopian, N.H. Lindner, E. Poem, Y. Berlatzky, J. Avron, D. Gershoni, B. [31] M. Grundmann, O. Stier, D. Bimberg, Phys. Rev. B 52 (1995) 11969.
D. Gerardot, P.M. Petroff, Phys. Rev. Lett. 96 (2006) 130501. [32] W.J. Moore, R.T. Holm, J. Appl. Phys. 80 (1996) 6939.
[11] C.H.H. Schulte, J. Hansom, A.E. Jones, C. Matthiesen, C. Le Gall, M. Atature, [33] M. Abbarchi, T. Kuroda, T. Mano, K. Sadoka, C.A. Mastrandrea, A. Vinattieri,
Nature 525 (2015) 222. M. Gurioli, T. Tsuchiya, Phys. Rev. B 82 (R) (2010) 201301.
[12] F. Henneberger, O. Benson, Semiconductor Quantum Bits, Pan Stanford Publishing [34] A. Graf, D. Sonnenberg, V. Paulava, A. Schliwa, Ch Heyn, W. Hansen, Phys. Rev. B
Pte. Ltd., Singapore, 2008. 89 (2014) 115314.
[13] Jun-Wei Luo, Bester Gabriel, Alex Zunger, Phys. Rev. B 92 (2015) 165301. [35] K. Brunner, G. Abstraiter, C. Bohm, G. Trokle, G. Weimann, Phys. Rev. Lett. 73
[14] J. Fischer, D. Loss, Phys. Rev. Lett. 105 (2010) 266603. (1994) 1138.
[15] A.V. Koudinov, I.A. Akimov, Y.G. Kusrayev, F. Henneberger, Phys. Rev. B 70 (R) [36] M. Bayer, G. Ortner, O. Stern, A. Kuther, A.A. Gorbunov, A. Forchel, P. Hawrylak,
(2004) 241305. S. Fafard, K. Hinzer, T.L. Reinecke, S.N. Walck, J.P. Reithmaier, F. Klopf,
[16] T. Belhadj, T. Amand, A. Kunold, C.-M.M. Simon, T. Kuroda, M. Abbarchi, T. Mano, F. Schäfer, Phys. Rev. B 65 (2002) 195315.
K. Sakoda, S. Kunz, X. Marie, B. Urbaszek, Appl. Phys. Lett. 97 (2010), 051111. [37] S. Kako, K. Hoshino, S. Iwamoto, S. Ishida, Y. Arakawa, Appl. Phys. Lett. 85 (2004)
[17] Y.H. Liao, C.C. Liao, C.H. Ku, Y.C. Chang, S.J. Cheng, M. Jo, T. Kuroda, T. Mano, 64.
M. Abbarchi, K. Sakoda, Phys. Rev. B 86 (2012) 115323. [38] S. Kako, C. Santori, K. Hoshino, S. Götzinger, Y. Yamamoto, Y. Arakawa, Nat.
[18] D. Csontosová, P. Klenovský, Phys. Rev. B 102 (2020) 125412. Mater. 5 (2006) 887.
[19] Y.H. Huo, B.J. Witek, S. Kumar, J.R. Cardenas, J.X. Zhang, N. Akopian, R. Singh, [39] D. Simeonov, A. Dussaigne, R. Butté, N. Grandjean, Phys. Rev. B 77 (2008),
E. Zallo, R. Grifone, D. Kriegner, R. Trotta, F. Ding, J. Stangl, V. Zwiller, G. Bester, 075306.
A. Rastelli, O.G. Schmidt, Nat. Phys. 10 (2014) 46. [40] F. Ding, R. Singh, J.D. Plumhof, T. Zander, V. Křápek, Y.H. Chen, M. Benyoucef,
[20] Jun-Wei Luo, Ranber Singh, Alex Zunger, Bester Gabriel, Phys. Rev. B 86 (R) V. Zwiller, K. Dörr, G. Bester, A. Rastelli, O.G. Schmidt, Phys. Rev. Lett. 104
(2012) 161302. (2010), 067405.
[21] V. Mlinar, M. Bozkurt, J.M. Ulloa, M. Ediger, G. Bester, A. Badolato, P. [41] R.J. Warburton, C. Schaflein, D. Haft, F. Bickel, A. Lorke, K. Karrai, J.M. Garcia,
M. Koenraad, R.J. Warburton, A. Zunger, Phys. Rev. B 80 (2009) 165425. W. Schoenfeld, P.M. Petroff, Nature 405 (2000) 926.
[22] Ch Heyn, A. Stemman, W. Hansen, Appl. Phys. Lett. 95 (2009) 173110. [42] M.F. Doty, J.I. Climente, M. Korkusinski, M. Scheibner, A.S. Bracker, P. Hawrylak,
[23] Lixin He, Alex Zunger, Phys. Rev. B 75 (2006), 075330. D. Gammon, Phys. Rev. Lett. 102 (2009), 047401.
[24] D.J. BenDaniel, C.B. Duke, Phys. Rev. 152 (1966) 683. [43] A. Schliwa, M. Winkelnkemper, D. Bimberg, Phys. Rev. B 79 (2009), 075443.

You might also like