ArticlePDF Available

Role of the wetting layer in the carrier relaxation in quantum dots

Authors:

Abstract and Figures

We present picosecond time resolved photoluminescence measurements of GaAs/AlGaAs quantum dot structures—grown by modified droplet epitaxy—where no wetting layer is connecting the dots. We find a fast carrier relaxation time (30 ps) to the dot ground state, which becomes even faster for increasing the photogenerated carrier injection. This shows that the two–dimensional character of the wetting layer is not relevant in determining the quantum dot capture, in contrast with the conclusions of several models so far presented in literature. We discuss the role of the barrier states as well as the possibility of Auger processes involving the zero-dimensional levels of the quantum dots. © 2002 American Institute of Physics.
Content may be subject to copyright.
Role of the wetting layer in the carrier relaxation in quantum dots
S. Sanguinetti,a) K. Watanabe, T. Tateno, M. Wakaki,b) and N. Koguchi
National Institute for Materials Science, Tsukuba, Ibaraki 305-0047, Japan
T. Kuroda and F. Minami
Department of Physics, Tokyo Institute of Technology, Meguro-ku, Tokyo 152-8551, Japan
M. Gurioli
I.N.F.M. - Department of Materials Science; Via Cozzi 53, I-20125 Milan, Italy and L.E.N.S.,
Via Sansone 1, I50019, Sesto Fiorentino, Italy
Received 8 March 2002; accepted for publication 29 May 2002
We present picosecond time resolved photoluminescence measurements of GaAs/AlGaAs quantum
dot structures—grown by modified droplet epitaxy—where no wetting layer is connecting the dots.
We find a fast carrier relaxation time 30 psto the dot ground state, which becomes even faster for
increasing the photogenerated carrier injection. This shows that the two–dimensional character of
the wetting layer is not relevant in determining the quantum dot capture, in contrast with the
conclusions of several models so far presented in literature. We discuss the role of the barrier states
as well as the possibility of Auger processes involving the zero-dimensional levels of the quantum
dots. © 2002 American Institute of Physics. DOI: 10.1063/1.1495525
The nature of the mechanisms underlying carrier relax-
ation phenomena in quantum dot QDmaterials has been
largely debated in the last years.1–5 These phenomena con-
tinue to attract much attention since they involve fundamen-
tal physical aspects of the zero-dimensional semiconductor
systems, and also due to their relevance for device applica-
tions. In QD lasers, the carriers are injected into the barriers
embedding the QDs. After energy dissipation processes, the
carriers are captured by the QDs and then relax to the fun-
damental lasing state. The efficiency of the relaxation cas-
cade directly affects the device performances, such as thresh-
old current, temperature stability, and so on. The
achievement of a very fast capture rate is therefore a relevant
aspect of the device optimization study.
The self-aggregated dots, grown by Stranski–Krastanov
epitaxy on a two-dimensional 2Dwetting layer WL, sat-
isfy this requirement. In general, the characteristic carrier
relaxation time from the barrier states to the fundamental QD
level is on the order of a few tens of picoseconds and be-
comes even shorter for large injection of carriers. The nature
of the fast carrier relaxation in QDs is not yet completely
understood. The typical energy separation of the QD elec-
tronic levels, which matches neither the longitudinal acoustic
phonons nor the longitudinal optical phonons, suggested the
presence of the well known phonon bottleneck effect.1Sev-
eral mechanisms have been invoked to explain the lack of
phonon bottleneck. The following picture has recently
emerged. In the low injection regime, the fast relaxation time
30–70 pshas been interpreted as a consequence of the
presence of a continuum tail of WL defect states to which the
carriers easily relax the excess energy.4In addition, resonant
multiphonon processes can mediate the relaxation between
the localized states.6By increasing the carrier injection, the
carrier–carrier interaction speeds up the energy relaxation
rate, via Auger-type processes.7,8 In particular, the Auger
processes involving carriers localized in the 2D-WL states
have been claimed to increase the carrier relaxation rate in-
side the QD.9,10 The WL is therefore assumed to have a key
role in the QD carrier relaxation.8
In this letter, we intend to test this picture by means of
picosecond time-resolved photoluminescence PLin a
GaAs/AlGaAs QD grown by modified droplet epitaxy
MDE.11,12 MDE is a nonconventional growth method for
self–assembling semiconductor QDs even in lattice matched
systems, such as GaAs/AlGaAs.12 Among other interesting
features, this alternative method leads to the fabrication of
defect free dots without any WL in the structure.13 Therefore,
we eliminate a priori the role of the WL and investigate the
resulting carrier relaxation processes. We found very similar
results with respect to standard QD with a WL, that is a fast
relaxation time 30 pswhich becomes even faster for in-
creasing the photogenerated carrier injection. These results
show that the 2D character of the WL is not relevant in
determining the QD capture. We discuss the role of the bar-
rier states as well as the possibility of Auger processes in-
volving the zero–dimensional levels of the QDs.
The samples were grown by MDE12 with the following
procedure.14,15 After the growth of 300 nm GaAs buffer layer
and 500 nm Al0.3Ga0.7As barrier layer at 580 °C, the sub-
strate temperature was lowered to 180 °C, the As valve was
closed and the As pressure in the growth chamber was de-
pleted. 1.75 monolayers of Al0.3Ga0.7 alloy was supplied in
order to form a group-III stabilized surface on the c44
reconstruction. The subsequent deposition of 3 monolayers
of Ga at this temperature gives rise to the formation of tiny
Ga droplets on the substrate. Following the deposition of the
droplets, an As4molecular beam was irradiated on the sur-
aPermanent address: Department of Materials Science, Via Cozzi 53,
I-20125 Milan, Italy; electronic mail: stefano.sanguinetti@mater.unimib.it
bAlso at: Department of Electro–Photo Optics, Tokai University, Hiratsuka,
Kanagawa 259-1292, Japan.
APPLIED PHYSICS LETTERS VOLUME 81, NUMBER 4 22 JULY 2002
6130003-6951/2002/81(4)/613/3/$19.00 © 2002 American Institute of Physics
Downloaded 16 Apr 2004 to 144.213.253.14. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp
face. After the complete change of reflection high-energy
electron diffraction pattern from halo to spots, an
Al0.3Ga0.7As barrier layer of 10 nm was grown by migration
enhanced epitaxy at the same temperature of 180 °C. This
temperature was chosen in order to prevent 2D regrowth of
the naked GaAs microcrystals.12 Then, the growth tempera-
ture was raised back to 580 °C and 90 nm of Al0.3Ga0.7As
and 10 nm of GaAs as cap layer were grown by ordinary
molecular-beam epitaxy. The sample was reloaded in the
growth chamber and annealed in 1.5105Torr As4atmo-
sphere at 640 °C. This annealing procedure was shown to
improve the sample quality healing the defects formed in the
low-temperature growth of the Al0.3Ga0.7As barrier layer.13,16
Before the deposition of the capping layer, surface and cross
section high-resolution scanning electron microscope images
of the complete annealed structure demonstrate the formation
of pyramidal shape nanocrystals with typical base sizes of
16 nm20 nm and the absence of the WL.15
Continuous wave cwPL spectra were measured at 17
K and excited with an Arlaser in multiline mode. PL spec-
tra were measured by a grating monochromator operating
with a GaAs photomultiplier. Time-resolved PL measure-
ments, have been performed by using the second-harmonic
(exc400 nm) of a fs mode-locked Ti-Sapphire laser with
76 MHz repetition rate. The PL signal was dispersed by a
polychromator, and time resolved by a streak camera with a
resolution of 1 meV and 9 ps, respectively. All time-resolved
measurements were performed at 5 K.
Figure 1 ashows the cw PL spectra for different exci-
tation power densities. At low-power density, the PL spec-
trum consists in a broad QD band 160 meV, reflecting the
large inhomogeneous distribution of the QD size, and in a
narrower band at 2 eV associated with recombination in
the Al0.3Ga0.7As barriers. This latter consists in a doublet
with intrinsic exciton recombination at 1.97 eV and extrinsic
recombination at 1.94 eV. For Pexc larger that 60 W/cm2, the
QD band shows in the high-energy tail at 1.82 eVthe
appearance of a broad emission associated with the carrier
population of the QD excited levels.
The analysis of the time evolution of the different spec-
tral components in the PL spectra Fig. 1 b兲兴 shows an al-
most constant decay 400 psand rise time 35 psof the QD
of different size within the inhomogeneously broadened PL
band. On the wing of the high-energy tail of the QD-PL band
(1.84 eV), the decay time becomes shorter 200 psand it
is associated with the recombination from the QD excited
states. The comparison between the time evolution of the QD
ground state (Eem1.65 eV) and the time evolution of the
excited-state emission (Eem1.85 eV) is reported in Fig.
1b. Finally, the barrier PL time evolution not shownis
characterized by a fast rise time (10 ps) and a relatively
long decay time 250 ps. Therefore, the barrier PL kinetics
is not determined by the QD capture mechanisms. This can
be explained by assuming that the carriers which recombine
in the barrier are localized in a spatial region far from the QD
layers.
Figure 2 areports the PL time evolution of the funda-
mental optical transition in the QDs for three different exci-
tation power densities. The time evolution of the PL band has
been fitted with a phenomenological model based on the dif-
ference between two exponential decays after a convolution
with the experimental response function. The best fits are
reported in Fig. 2 aas solid lines. The decay time
Dis
almost constant for a large range of Pexc nearly three orders
of magnitude. This shows that
Dstems from the enhanced
optical matrix element of the ground-state transition in
GaAs/Al0.3Ga0.7As QD and not from nonradiative competi-
tive channels. At Pexc higher than 300 W/cm2, a lengthening
is observed due to the QD filling. Aclear reduction of the PL
rise time
Ris also observed when increasing the excitation
power, showing that the carrier capture into the QDs strongly
depends of on the carrier injection. Asummary of the PL rise
time is reported in Fig. 2 b. A sharp decrease of
Ris ob-
served for Pexc larger than 10 W/cm2. At the highest ex-
citation power, the rise time of the PL from the low-lying
transition is almost the same of the rise time of the PL from
the excited state. The solid dashedline in Fig. 2brefers
to a best fit with 1/Pexc (1/Pexc
2) dependence. Despite the
small amount of data, the comparison indicates that the
1/Pexc dependence gives a better agreement with the experi-
FIG. 1. aFour different PL spectra taken at increasing Pexc at steps of one
optical density in the range 10 W/cm2–10 kW/cm2.bPL traces recorded
at 1.65 eV ground-state emission—circlesand 1.85 eV excited-state
emission—squareswith Pexc1 kW/cm2.
FIG. 2. aPL traces of the fundamental QD transition at different Pexc :
3300 W/cm2squares, 330 W/cm2circles, and 3 W/cm2triangles. The
fits are shown by continuous lines. bPL rise time as a function of Pexc
circles. Continuous and dashed lines show the best in ps1/Pexc and
1/Pexc
2fits, respectively.
614 Appl. Phys. Lett., Vol. 81, No. 4, 22 July 2002 Sanguinetti
et al.
Downloaded 16 Apr 2004 to 144.213.253.14. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp
mental data. Finally, we note that in the range of Pexc corre-
sponding to the speeding up of
RRef. 17, the QD PL
spectrum shows the insurgence of the recombination from
excited levels.
Let us now compare our results with the findings re-
ported in literature on standard InAs/GaAs QDs which nucle-
ate on a 2D InAs WL. As already discussed, several recent
papers assume that the presence of the 2D layer connecting
the QDs has relevant consequences on the QD carrier relax-
ation. In particular, Morris et al.8reports a PL rise time
R
35 ps, for Pexc10 W/cm2, and a decrease of
Rtoward a
value of 10 ps when Pexc100 W/cm2with a 1/Pexc depen-
dence. The 1/Pexc rise time behavior has been attributed to
carrier relaxation processes involving two-particle Coulomb
scattering pseudo–Auger processbetween carriers in the
QDs and in the WL.9,10,8 The very same results are observed
here in QDs that are not connected by the 2D WL. We,
therefore, conclude that the picture in which the WL states
play a relevant role in speeding up the QD relaxation time is
not completely correct. Our data show that both the fast car-
rier relaxation time, and the dependence of
Ron Pexc , are
general properties of self-aggregated QDs.
In our sample, the continuum of states associated with
the barriers may play the role attributed to the WL in Refs. 4
and 8. That is, the presence of defect states in the
Al0.3Ga0.7As barrier can facilitate the carrier relaxation inside
the QDs.3Similarly, the pseudo–Auger processes could in-
volve the electronic states of the barriers. The most relevant
difference between the barrier and the WL states is, a priori,
the dimensionality. The barrier states have a three
dimensional character while the electronic states of the WL
are 2D states. Our results show that the 2D character of the
continuum of states over the QD does not play any role in
the carrier capture and relaxation. This can be justified, in
principle, by considering that the electronic wave functions
associated with the WL states largely penetrate into the bar-
riers thus washing out a predominant 2D character of the WL
states.
On the other hand, the observed behavior may stem from
intrinsic relaxation mechanisms in a QD. Fast relaxation pro-
cesses, promoted by polaronic electron–phonon, anhar-
monic phonon–phonon,18 and resonant multiphonon
processes6have been proposed. As far as the dependence of
the relaxation rate on the carrier injection is concerned, we
find that the speeding up of the relaxation rate with Pexc
corresponds to a large QD filling, which leads to PL emission
from the QD excited states. We remark that the same effect
was also present in the previous studies,7,8,19 even if the au-
thors did not stress it. The onset of the Auger effect therefore
occurs when the QD is filled by a large number of carriers.
This strongly suggests that the nonlinear mechanisms that
are effective in the QD carrier relaxation processes are pro-
moted by the carrier population inside the QD rather than
involving the carrier population in the continuum of states
above the QD. Note also that the 1/Pexc dependence of the
carrier relaxation time is a common feature to any two-
particle scattering mechanisms. Fast intraband relaxation
processes due to two–particle Coulomb scattering intraband
Auger, active in a QD for dot carrier population larger than
one, have been recently suggested.5,20 We, therefore, believe
that the Auger-type relaxation in QDs has to be associated
with Coulomb scattering processes involving carriers con-
fined in the QD. This is also supported by the recent obser-
vation of fast relaxation processes in isolated CdSe dots.21
In conclusion, we have studied the carrier relaxation dy-
namics in GaAs QDs embedded in three–dimensional
Al0.3Ga0.7As barriers. The removal of the 2D WL does not
modify the carrier relaxation processes. We believe that the
possibility of growing, and therefore investigate, a defect
free QD without a WL gives a further relevant degree of
freedom for obtaining deeper understanding of the carrier
dynamics and optical properties of self–assembled dots.
One of the authors S.S.acknowledges partial financial
support from INFM-LENS.
1H. Benisty, C. M. Sotomayor-Torres, and C. Weisbuch, Phys. Rev. B 44,
10 945 1991.
2A. L. Efros, V. A. Kharchenko, and M. Rosen, Solid State Commun. 93,
281 1995.
3D. F. Schroeter, D. J. Griffiths, and P. C. Sercel, Phys. Rev. B 54, 1486
1996.
4Y. Yoda, O. Moriwaki, M. Nishioka, and Y. Arakawa, Phys. Rev. Lett. 82,
4114 1999.
5R. Ferreira and G. Bastard, Appl. Phys. Lett. 74, 2818 1999.
6R. Heitz, M. Veit, N. N. Ledentsov, A. Hoffmann, D. Bimberg, V. M.
Ustinov, P. S. Kop’ev, and Z. I. Alferov, Phys. Rev. B 56, 10435 1997.
7B. Ohnesorge, M. Albrecht, J. Oshinowo, andA. Forchel, Phys. Rev. B 54,
11532 1996.
8D. Morris, N. Perret, and S. Fafard, Appl. Phys. Lett. 75, 3593 1999.
9U. Bockelmann and T. Egeler, Phys. Rev. B 46, 15 574 1992.
10 A. V. Uskov, F. Adler, H. Schweizer, and M. H. Pilkuhn, J.Appl. Phys. 81,
7895 1997.
11 N. Koguchi, S. Takahashi, and T. Chikyow, J. Cryst. Growth 111,688
1991.
12 K. Watanabe, N. Koguchi, and Y. Gotoh, Jpn. J. Appl. Phys., Part 2 39,
L79 2000.
13 K. Watanabe, S. Tsukamoto, Y. Gotoh, and N. Koguchi, J. Cryst. Growth
227, 1073 2001.
14 Y. Horikoshi, M. Kawashima, and H. Yamaguchi, Jpn. J.Appl. Phys., Part
127, 169 1988.
15 T. Tateno, K. Watanabe, S. Sanguinetti, S. Tsukamoto, M. Wakaki, Y.
Gotoh, and N. Koguchi, in Growth of Well Defined Nanostructures, edited
by N. Koguchi National Research Institute for Metals, Tsukuba, Japan,
2001,p.298.
16 S. Sanguinetti, K. Watanabe, T. Kuroda, F. Minami, Y. Gotoh, and N.
Koguchi, J. Cryst. Growth to be published.
17 The ratio between the carrier injection in cw and time-resolved TRcon-
ditions is given by Ncw /NTR
rep
PLPcw /PTR , where Pcw (PTR)isthe
excitation power density in the cw the TRcondition,
PL is the PL decay
time and
rep is the laser repetition rate.
18 O. Verzelen, R. Ferreira, and G. Bastard, Phys. Rev. B 62, 4809 2000.
19 S. Raymond, K. Hinzer, S. Fafard, and J. L. Merz, Phys. Rev. B 61, 16331
2000.
20 J. L. Pan, Phys. Rev. B 49, 11272 1994.
21 V. I. Klimov, D. W. McBranch, C. A. Leatherdale, and M. G. Bawendi,
Phys. Rev. B 60, 13740 1999.
615Appl. Phys. Lett., Vol. 81, No. 4, 22 July 2002 Sanguinetti
et al.
Downloaded 16 Apr 2004 to 144.213.253.14. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp
... 5 The high carrier density strengthens the Auger capture process where the electron transfers from the WL into a QD, accompanied by WL electrons that transition to high WL states. 6 Subsequently, WL states work as a reservoir: the pumping processes take carriers from the WL to the QD, and then, this type of transition cannot be avoided and must be considered in the QD simulations. 7 Coulomb effects in the WL-QD structure are studied in 6 for shallow dots assuming their confining potential as a simple harmonic oscillator type. ...
... 6 Subsequently, WL states work as a reservoir: the pumping processes take carriers from the WL to the QD, and then, this type of transition cannot be avoided and must be considered in the QD simulations. 7 Coulomb effects in the WL-QD structure are studied in 6 for shallow dots assuming their confining potential as a simple harmonic oscillator type. Auger-and phononassisted tunneling for a WL-QD structure are studied using the Fermi golden rule without considering the OPW. ...
Article
Full-text available
This work studies the carrier capture rates due to the Coulomb scattering process in the double quantum dot (QD) structure using the Fermi golden rule and considering the orthogonalized plane wave for transitions between the wetting layer (WL) and QD states. It took a realistic wave function that is not covered with these considerations together elsewhere. The results show that by applying very high tunneling, the in- and out-scattering processes of the double QD are slowed by state filling, while moderate tunneling was accelerated. The incoherent pumping makes the processes faster, except at high WL occupation. Bandgap engineering, which is of central importance for material researchers, experiences two possibilities: the wide-bandgap QDs have fast QD processes, while deep QDs are associated with slow processes. The high momenta correspond with fast QD processes.
... 7 They also discussed the role of the WL on the PL properties of GaAs QDs grown by a droplet epitaxy. 11,12 If the WL was not present, the redshift of the PL peak energy and the linewidth narrowing could not be observed. It was concluded that no carrier connecting channels among the size-distributed QDs, such as the WL, inhibited the appearance of the peculiar temperature behavior of the PL signals. ...
... The decrease in the FWHM at low temperature has also been reported when the coupling of the carriers in the size-distributed QD occurred through channels such as the WL. 11 However, the further increase at higher temperatures, above 200 K, has been first observed in the present study and will be discussed as follows. ...
Article
Photoluminescence (PL) measurements at a wide temperature range, up to room temperature, for high-quality and high-density GaAs/AlGaAs quantum dots (QDs) fabricated by droplet epitaxy were carried out to investigate the anomalous temperature dependence of the PL peak energy from the QD ensemble. In addition to a reported redshift that deviated from the so-called Varshni's curve of the PL peak energy in the low temperature region, a new blueshift was observed above 200 K. We analyzed the experimental results using a steady-state rate equation model and observed a good agreement. The distribution of the QD sizes and the presence of the AlGaAs barrier layer as a carrier coupling channel were considered in this model. This means that the wetting layer proposed thus far is not a necessary condition for explaining the anomalous temperature behavior of the PL properties. In addition, it was found that the anomalous temperature behavior was smeared out by the insertion of a GaAs height adjustment layer in order to homogenize the apparent QD size. We found that sufficient control of the QD size is a necessary factor for high temperature stability of QD devices.
... 20 In addition, the capability of DE to realize QDs without the presence of a wetting layer at the base of the QD, a typical characteristic of the InAs/GaAs QDs, also allows the study of a pure QD system, with the IB made up of three-dimensionally confined states only. 21,22 Leveraging on the extremely flexible shape control permitted by DE for the QD self-assembling, here, we extend the concepts of IB-CB decoupling by quantum design of the dot states 16 ...
Article
Full-text available
We report the effect of the quantum dot aspect ratio on the sub‐gap absorption properties of GaAs/AlGaAs quantum dot intermediate band solar cells. We have grown AlGaAs solar cells containing GaAs quantum dots made by droplet epitaxy. This technique allows the realization of strain‐free nanostructures with lattice matched materials, enabling the possibility to tune the size, shape, and aspect ratio to engineer the optical and electrical properties of devices. Intermediate band solar cells have been grown with different dot aspect ratio, thus tuning the energy levels of the intermediate band. Here, we show how it is possible to tune the sub‐gap absorption spectrum and the extraction of charge carriers from the intermediate band states by simply changing the aspect ratio of the dots. The tradeoff between thermal and optical extraction is in fact fundamental for the correct functioning of the intermediate band solar cells. The combination of the two effects makes the photonic extraction mechanism from the quantum dots increasingly dominant at room temperature, allowing for a reduction of the open circuit voltage of only 14 mV, compared to the reference cell.
... It is worth mentioning that from the experimental point of view, during the growth of the SiGe material, QDs arise at the wetting layer. Thus, the well-known growth techniques (Starsky Kratsanov or Metalo-Organic Chemical Vapor Deposition) lead to the formation of nonhomogenous islands with a large dispersion of the sizes and shapes of SiGe QDs [25][26][27][28][29]. Indeed, due to the lattices mismatch of different materials and increasing the wetting layer, the strain energy increases rapidly, and dislocated islands formation can occur to stabilize the structure. ...
Article
Semi oblate and semi prolate are among the most probable self-organized nanostructures shapes. The optoelectronic properties of such nanostructures are not just manipulated with the height and lateral size but also with the wetting layer element. The practical interest of derivatives of germanium and silicon has a great important role in optoelectronic devices. This study is a contribution to the analysis of linear and nonlinear optical properties of Si0.7Ge0.3/Si. In the framework of the effective mass approximation, we solve numerically the Schrödinger equation relative to one particle confined in Si0.7Ge0.3/Si semi prolate and semi oblate quantum dots by using the finite element method and by taking into consideration the effect of the wetting layer. The energy spectrum of the lowest states and the dipolar matrix for the fourth allowed transitions are determined and discussed. We also calculate the detailed optical properties, including absorption coefficients, refractive index changes, second and third harmonic generation as a function of the quantum dot sizes. We found that with the change in the size of prolate and oblate quantum dots, there is a shift in the resonance peaks for the absorption coefficient and refractive index. It is due to the modification in the energy levels with changing size. The study proves a redshift in the second harmonic generation and third harmonic generation coefficients with an increase in the height/radius of the oblate/prolate quantum dot, respectively. We also demonstrated the variation of wavefunction inside the quantum dot with the change in wetting layer thickness.
... Individual QDs; QDs diads [2,36,37]; single and multiple quantum rings (QRs) [16,21,[38][39][40][41][42][43]; hybrid QD-QR structures, such as ring-on-adisk [44], dot in-a-ring [45], or dot-on-a-disk [46]; and quantum wires [47] are remarkable examples of the possibilities offered by this self-assembly technique. In addition to this, DE allows forming QDs without a wetting layer [47][48][49][50][51][52][53][54][55], and their size and density can be independently tuned [56]. These features are not easily obtained with other methods such as conventional Stranski-Krastanov growth based on accumulation and relaxation of strain [57]. ...
Article
Full-text available
We provide an extensive and systematic investigation of exciton dynamics in droplet epitaxial quantum dots comparing the cases of (311)A, (001), and (111)A surfaces. Despite a similar s-shell exciton structure common to the three cases, the absence of a wetting layer for (311)A and (111)A samples leads to a larger carrier confinement compared to (001), where a wetting layer is present. This leads to a more pronounced dependence of the binding energies of s-shell excitons on the quantum dot size and to the strong anti-binding character of the positive-charged exciton for smaller quantum dots. In-plane geometrical anisotropies of (311)A and (001) quantum dots lead to a large electron-hole fine interaction (fine structure splitting (FSS) ∼100 μeV), whereas for the three-fold symmetric (111)A counterpart, this figure of merit is reduced by about one order of magnitude. In all these cases, we do not observe any size dependence of the fine structure splitting. Heavy-hole/light-hole mixing is present in all the studied cases, leading to a broad spread of linear polarization anisotropy (from 0 up to about 50%) irrespective of surface orientation (symmetry of the confinement), fine structure splitting, and nanostructure size. These results are important for the further development of ideal single and entangled photon sources based on semiconductor quantum dots.
... The versatility of this method allowed to grow many different semiconductor alloys (GaInSb [32], AlGaAs [33][34][35][36][37], InGaAs [38][39][40][41][42][43][44], and InGaP [26,45,46]), forming a plethora of nanostructures [47] such as quantum dots (QDs); QDs diads [48][49][50]; multiple-concentric quantum rings [18,23,[51][52][53][54][55][56]; coupled structures such as ring-on-a-disk [57], dot in-a-ring [58], or dot-on-a-disk [59]; as well as elongated structures such as nanowires [60]. This technique allows to independently tune the size and density of the nanostructures [61] and to grow them with or without a wetting layer [35,39,40,60,[62][63][64][65][66], aspects that are not matched by the conventional Stranski-Krastanov approach based on strain [67]. ...
Article
Full-text available
Droplet epitaxy allows the efficient fabrication of a plethora of 3D, III-V-based nanostructures on different crystalline orientations. Quantum dots grown on a (311)A-oriented surface are obtained with record surface density, with or without a wetting layer. These are appealing features for quantum dot lasing, thanks to the large density of quantum emitters and a truly 3D lateral confinement. However, the intimate photophysics of this class of nanostructures has not yet been investigated. Here, we address the main optical and electronic properties of s-shell excitons in individual quantum dots grown on (311)A substrates with photoluminescence spectroscopy experiments. We show the presence of neutral exciton and biexciton as well as positive and negative charged excitons. We investigate the origins of spectral broadening, identifying them in spectral diffusion at low temperature and phonon interaction at higher temperature, the presence of fine interactions between electron and hole spin, and a relevant heavy-hole/light-hole mixing. We interpret the level filling with a simple Poissonian model reproducing the power excitation dependence of the s-shell excitons. These results are relevant for the further improvement of this class of quantum emitters and their exploitation as single-photon sources for low-density samples as well as for efficient lasers for high-density samples.
... The versatility of this method allowed to grow many different semiconductor alloys (GaInSb [32], AlGaAs [33][34][35][36][37], InGaAs [38][39][40][41][42][43][44], and InGaP [26,45,46]), forming a plethora of nanostructures [47] such as quantum dots (QDs); QDs diads [48][49][50]; multiple-concentric quantum rings [18,23,[51][52][53][54][55][56]; coupled structures such as ring-on-a-disk [57], dot in-a-ring [58], or dot-on-a-disk [59]; as well as elongated structures such as nanowires [60]. This technique allows to independently tune the size and density of the nanostructures [61] and to grow them with or without a wetting layer [35,39,40,60,[62][63][64][65][66], aspects that are not matched by the conventional Stranski-Krastanov approach based on strain [67]. ...
Preprint
Full-text available
Droplet epitaxy allows the efficient fabrication of a plethora of 3D, III-V-based nanostructures on different crystalline orientations. Quantum dots grown on (311)A-oriented surface are obtained with record surface density, with or without a wetting layer. These are appealing features for quantum-dot lasing, thanks to the large density of quantum emitters and a truly 3D lateral confinement. However, the intimate photophysics of this class of nanostructures has not yet been investigated. Here we address the main optical and electronic properties of s-shell excitons in individual quantum dots grown on (311)A substrates with photoluminescence spectroscopy experiments. We show the presence of neutral exciton and biexciton as well as positive and negative charged excitons. We investigate the origins of spectral broadening, identifying them in spectral diffusion at low temperature and phonon-interaction at higher temperature, the presence of fine interactions between electron and hole spin, and a relevant heavy-hole/light-hole mixing. We interpret the level filling with a simple Poissonian model reproducing the power excitation dependence of the s-shell excitons. These results are relevant for the further improvement of this class of quantum emitters and their exploitation as single photon sources for low density samples as well as for efficient lasers for high density samples.
... optical) phonon scattering, Auger process (high excitation density), multiphonon process (low excitation density) and role of WL [13][14][15][16][17]. There are also many other reasons behind the reduced V oc in case of QDs based solar devices. ...
Article
Quantum dot solar cell (QDSC) is still far below the celebrated Shockley- Queisser limit due to an enhancement in photocurrent only. However, there is very less number of literatures that reports regarding voltage preservation in QDSC, which is one of the most important paradigms in achieving high efficiency. Here a single junction QD embedded intermediate band solar cell (SJ QDIBSC) is designed, which not only preserves the open circuit voltage, but also shows an improvement in short circuit current at the same time. In this work a number of 1-D quantum confined highly mismatched GaSb superlattice is used in the bulk GaAs solar cell, which results an increment in carrier lifetime of minority photo carriers as well as two-photon sub-bandgap photocurrent generation. This model is simulated and validated using Silvaco ATLAS TCAD tool.
Chapter
The review of carrier dynamics in quantum dots (QDs) in this chapter often refers to ensembles of QDs, but we will highlight the properties single quantum dots wherever possible. The discussion starts with the capture of carriers into and their relaxation within the QD potential. We examine the phonon-bottleneck problem and continue with the role of Coulomb scattering. This review of ultra-fast relaxation at low densities closes with pointing out the importance of considering polaron formation. With increasing excitation density charged excitons, multi-excitons and Pauli blocking affect the relaxation dynamics. The treatment of the recombination dynamics of electron-hole pairs has to consider first the mostly unwanted effects of blinking, photodarkening and spectral diffusion. Then the focus is on the interpretation of the decay of the photoluminescence signal. A quantitative discussion of intrinsic parameters like lifetime, oscillator strength and dipole moment of various excitonic species is given. We compare these quantities for the various types of quantum dots. Important for the carrier dynamics in QDs is the role of dark states which will be considered in detail. Finally we will highlight the generation of single photons using semiconductor QDs for possible applications in quantum information. This chapter also includes the description of experimental techniques like charge loading of quantum dots or the Hanbury-Brown and Twiss (HBT) correlation experiment.
Article
We present a microscopic model for a singly charged quantum dot (QD) ensemble to reveal the origin of the long-range effective interaction between the electron spins in the QDs. Wilson’s numerical renormalization group (NRG) is used to calculate the magnitude and the spatial dependency of the effective spin-spin interaction mediated by the growth-induced wetting layer. Surprisingly, we found an antiferromagnetic Heisenberg coupling for very short inter-QD distances that is caused by the significant particle-hole asymmetry of the wetting layer band at very low filling. Using the NRG results obtained from realistic parameters as input for a semiclassical simulation for a large QD ensemble, we demonstrate that the experimentally reported phase shifts in the coherent spin dynamics between single- and two-color laser pumping can be reproduced by our model, solving a long-standing open problem of the microscopic origin of the inter-QD electron spin-spin interaction.
Article
Full-text available
We show that the anharmonicity driven instability of optical phonons leads in semiconductor quantum dots to a decay of polaron states which otherwise would be everlasting. By such a mechanism a single electron in an excited dot state can relax down to the ground state even if the electron energy difference markedly differs from the optical phonon energy.
Article
We have investigated post-annealing effects of photoluminescence (PL) properties in GaAs/AlGaAs QDs fabricated by modified droplet epitaxy. The annealing temperatures were changed between 5208C and 7608C. The PL intensity of QDs increased drastically with the increase of annealing temperature. The PL intensity of QDs after the annealing at 7608C was enhanced by two orders of magnitude as compared to that of before post-annealing. This sample showed a distinct PL peak even at the room temperature. With the increase of annealing temperatures, the peak energy shifted from 1.646 to 1.749 eV, continuously. These effects may be caused by improving the crystallinity of QDs systems and the size reduction and/or changing the composition of QDs by the post-annealing. # 2001 Elsevier Science B.V. All rights reserved.
Article
Femtosecond (fs) broad-band transient absorption (TA) is used to study the intraband relaxation and depopulation dynamics of electron and hole quantized states in CdSe nanocrystals (NC’s) with a range of surface properties. Instead of the drastic reduction in the energy relaxation rate expected due to a “phonon bottleneck,” we observe a fast subpicosecond 1P-to-1S electron relaxation, with the rate exceeding that due to phonon emission in bulk semiconductors. The energy relaxation is enhanced with reducing the NC’s radius, and does not show any dependence on the NC surface properties (quality of the surface passivation). These data indicate that electron energy relaxation occurs by neither multiphonon emission nor by coupling to surface defects, but is likely meditated by Auger-type electron-hole energy transfer. We use fs infrared TA to probe electron and hole intraband transitions, which allows us to distinguish between electron and hole relaxation pathways leading to the depopulation of NC quantized states. In contrast to the electron relaxation, which is controlled by NC surface passivation, the depopulation of hole quantized states is extremely fast (sub-ps-to-ps time scales) in all types of samples, independent of NC surface treatment (including NC’s overcoated with a ZnS layer). Our results indicate that ultrafast hole dynamics are not due to trapping at localized surface defects such as a vacancy, but rather arise from relaxation into intrinsic NC states or intrinsically unpassivated interface states.
Article
We propose a new MBE growth method for InSb microcrystals on CdTe which has a nearly equal lattice constant to InSb. The average size of the InSb microcrystals was about 150 nm × 200 nm × 70 nm. This method is based on the Sb incorporation into In droplets and thought to be useful for fabricating quantum well boxes.
Article
We propose a modified droplet epitaxy method for fabricating self-organized GaAs/AlGaAs quantum dots (QDs) with a high As flux irradiation and a low substrate temperature. By our novel method, GaAs QDs were successfully formed, retaining their pyramidal shape, original base size and density of droplets, and preventing layer-by-layer growth. Quantum size effects of the QDs were distinctly observed by photoluminescence measurements. It was confirmed that this new modified droplet epitaxy method is promising for fabricating a high-quality GaAs/AlGaAs QD system.
Article
Comparison of near-field and far-field photoluminescence excitation (PLE) spectra gives new insight into the carrier relaxation process in InGaAs/GaAs self-assembled quantum dots. The near-field PLE spectra of single quantum dots clearly show 2D-like continuum states and a number of sharp lines, between a large zero-absorption region due to the quasi-0D density of states and the 2D wetting layer absorption edge. The results reveal an efficient intradot relaxation mechanism, proceeding as follows: The carriers can relax easily within continuum states, and make transitions to the excitonic ground state by resonant emission of localized phonons.
Article
Carrier capture and relaxation processes in self-assembled 15-nm In0.5Ga0.5As/GaAs quantum dots are investigated by means of time-resolved photoluminescence spectroscopy. In a systematic study of photoluminescence rise times and barrier decay times (variation of temperature, excitation energy, and excitation density) we aim to identify the physical mechanisms responsible for fast carrier capture and relaxation in quantum dots. Both processes are separated by using appropriate excitation energies. Carrier capture and relaxation are shown to proceed with rates as high as ∼2×1010 s-1 at low temperature even if less than one electron-hole pair per dot and excitation pulse is created. We interpret our results in terms of multiphonon processes at low excitation densities and in terms of Auger processes at high excitation densities. © 1996 The American Physical Society.
Article
Carrier relaxation and recombination in self-organized InAs/GaAs quantum dots (QD’s) is investigated by photoluminescence (PL), PL excitation (PLE), and time-resolved PL spectroscopy. We demonstrate inelastic phonon scattering to be the dominant intradot carrier-relaxation mechanism. Multiphonon processes involving up to four LO phonons from either the InAs QD’s, the InAs wetting layer, or the GaAs barrier are resolved. The observation of multiphonon resonances in the PLE spectra of the QD’s is discussed in analogy to hot exciton relaxation in higher-dimensional semiconductor systems and proposed to be intricately bound to the inhomogeneity of the QD ensemble in conjunction with a competing nonradiative recombination channel observed for the excited hole states. Carrier capture is found to be a cascade process with the initial capture into excited states taking less than a few picoseconds and the multiphonon (involving three LO phonons) relaxation time of the first excited hole state being 40 ps. The |001〉 hole state presents a relaxation bottleneck that determines the ground-state population time after nonresonant excitation. For the small self-organized InAs/GaAs QD’s the intradot carrier relaxation is shown to be faster than radiative (>1 ns) and nonradiative (≈100 ps) recombination explaining the absence of a “phonon bottleneck” effect in the PL spectra.
Article
Carrier relaxation in self-assembled quantum dots due to Coulomb interaction with two dimensional (2D) carriers is studied theoretically. Auger coefficients for carrier relaxation rates are calculated in the dipole approximation for Coulomb interaction. The dipole approximation allows one to derive selection rules for Auger relaxation in a cylindrical quantum dot, and to describe a general picture of Auger relaxation via energy levels in self-assembled quantum dots. A numerical example for InAs/GaAs self-assembled quantum dots demonstrates that the Auger effect may lead to relaxation times in the order of 1–10 ps at 2D carrier densities of 1011–1012 cm−2. This result demonstrates the possibility of fast carrier relaxation in quantum dots if the carrier density in the surrounding barrier is sufficiently high. Analytical formulas for Auger coefficients are derived for moderate temperatures of the 2D carriers. © 1997 American Institute of Physics.
Article
Carrier relaxation processes are investigated in self-assembled InAs/GaAs quantum dots using time-resolved photoluminescence spectroscopy. The quantum-dot photoluminescence rise time has been measured as functions of carrier excitation density and excitation wavelengths. The measured relaxation time is about 32 ps at low excitation density and decreases by 1 over the excitation density from about 3  W/cm <sup> 2 </sup>, under nonresonant laser excitation. The threshold of this density-dependent regime occurs at a slightly higher density as the excitation wavelength increases and it disappears when the photon pumping energy is below the wetting layer barrier energy. These results clearly establish the regime where Auger processes become the dominant carrier relaxation mechanism in these self-assembled quantum dots. © 1999 American Institute of Physics.