Sie sind auf Seite 1von 14

08/11/17 1

Quantum Dots: Theory, Fabrication and Properties.


by Thomas Turner

Abstract

This report discusses the physics of confining electrons to zero dimensions, a quantum dot. It describes
the energy of states within the dot and its density of states. Fabrication techniques old and new are described,
with an assessment of their respective advantages and disadvantages. The second half of the report focusses
on the optical and transport properties of quantum dots - with an explanation of their underlying physics
and an assessment of their importance for current and future technologies.

I. I NTRODUCTION

This report aims to give the reader a strong understanding of quantum dots, focusing on the underlying
physics which gives rise to many interesting optical and transport properties. Some of these properties are
already being exploited, such as in the production of thinner, low power displays, whilst other properties
hold the potential to revolutionise sectors in the future - such as in quantum-information technology.
Before these unique properties can be understood, one must first understand the geometry of a quantum
dot, or as they are often referred to as ‘artificial atoms’ [1]. Quantum dots are artificially fabricated pools
of charge, which can contain between one and many thousand electrons [1]. ‘Quantum’ refers to the size
regime that dots exist in, which ranges between a few nanometers to microns. Their size and shape can
be varied during fabrication, which in turn changes their properties; making the fabrication process vital
to obtain the desired properties of the dot.

II. T HEORY OF Z ERO - DIMENSIONAL ELECTRON SYSTEMS

When the charge carriers within a semiconductor are confined in all three spacial dimensions and
thus have zero-degrees of freedom, the system is zero-dimensional. A quantum dot is a zero-dimensional
electron system [2]. A quantum wire is one such system that is confined in two dimensions, a quantum
well is confined in one dimension and a bulk has zero dimensions confined [2].
08/11/17 2

A. Energy

The energy levels in a quantum dot are discrete. This occurs as the de Broglie wavelength of the
electron is similar to the dimensions of the quantum dot in which it is confined - approximately 100Å.
The discrete energy levels of a one dimensional quantum well are defined as [3],

n2 h2 h̄2 k 2
E= = (1)
8mef f a2 2mef f
where n is an integer energy state and is an integer, h = Planck’s constant, mef f is the effective mass
of the electron, k = nπ/a and a is the length of the box.
The quantum dot however is confined in 3 dimensions. Using the particle in a box model for a 0-d
system, the confinement energy of an exciton1 is,

h̄2 π 2 1 1 n2x n2y n2


Enx ,ny ,nz = ( + )( 2 + 2 + z2 ) (2)
2 me mh Lx Ly Lz
where Lx ,Ly and Lz are lengths of the dot in the x, y and z direction, n is the discrete energy level,
me is the mass of the electron and mh is the mass of the hole.
However, in reality the quantum dot does not have an infinite potential, so the particle in a box model
is not accurate. A more useful way to describe quantum dots is their density of states.

B. Density of States

The density of states (DOS) is a function, that when multiplied by an interval of energy provides the
total concentration of available states within the energy range [3].
The equations for the density of states for the four cases are written below, with units per unit energy
per unit volume.
Bulk density of states:
1 2mef f 3 √
ρ3D (E) = ( )2 E (3)
2π 2 h2
Quantum well density of states:
mef f
ρ2D (E) = (4)
πh2
Nanowire density of states: r
1 2mef f 1
ρ1D (E) = √ (5)
π h2 E
Quantum dot density of states
ρ0D (E) = 2δ(E − Enx ,ny ,nz ) (6)
1
Exciton: An electron-hole pair
08/11/17 3

Fig. 1. A plot of Density of states vs Energy in: (a) a bulk, (b) a quantum well, (c) a quantum wire and (d) and quantum dot [4].

where, E is the energy of the carrier2 , h is Planck’s constant, mef f is the effective mass of the carrier
and Enx ,ny ,nz is the confined energies of the carrier.
The DOS dependence in Eq. 3-6 are shown in Fig. 1. The plot of the bulk’s density of states is

proportional to E as in Eq. 3. The quantum well DOS is a stepwise function, as the confinement
leads to discrete energy levels. A sharp increase or ‘step’, occurs when the energy level increases by an
integer. The plateaus arise as the DOS in Eq. 4 is independent of energy. The nanowire’s inverse square
dependence in Eq. 5 is seen as the DOS falls away after a peak, which arise when there is an integer
increase in energy level, as with the quantum wire. Lastly, the quantum dot exhibits infinitely high spikes
at each discrete energy level transition and is zero elsewhere. These spikes follow the Dirac delta function
as described by Eq. 6, where the factor of 2 accounts for spin degeneracy [3].

III. FABRICATION TECHNIQUES

A. Etching

Etching was one of the first methods of producing quantum dots, developed in 1987 [5]. A diagram of
the process is shown in Fig. 2. An electron beam scans the surface of a semi conductor, such as GaAs,
coated with a layer of polymer such as polymethylmethacrylate (PMMA) [6], the polymer is called a resist.
The resist is removed where the electron beam scans. Next, a metal layer is deposited onto the surface.
A solvent dissolves the surface leaving the metal just where the electron beam scanned on the resist.
The metal layer serves the purpose of a reactive-ion etch mask (explained below) and an Ohmic contact
2
Carrier: An electron or hole
08/11/17 4

Fig. 2. The process of etching. (1) electron beam scanning the resist layer. (2) Resist [5]

between the metal and semiconductor. To obtain both properties, layers of: 500Å Au/Ge, 150Å Ni and
600Å Au are evaporated, in that order, onto the sample [6]. High-energy ions from plasma i.e evaporated
BCl3 , ‘attack’ the sample perpendicular to the surface, etching downwards into the semiconductor area
that is unprotected by the metal. A typical etching rate of GaAs is approximately 350Åmin−1 [6]. The
result is a column of material containing a quantum dot. An alternative method is ‘patterning’ instead of
etching. Electrodes are placed on the surface of the dot such that when a voltage is applied, the resulting
fields expel electrons [5]. The pattern of the electrodes means the resulting field confines electrons within
a layer of semiconductor such as GaAs. The electron is now confined in 3 dimensions - the requirement
for a quantum dot.

B. Molecular Beam Epitaxy (MBE)

1) The Physics of MBE: A more popular method is epitaxial growth. It is the formation of a crystal film
layer by layer on a crystal substrate surface [7]. MBE occurs under ultra-high vacuum conditions, which
can achieve a base pressure of approximately 1011 Torr [8]. This is required to reduce the background
pressure of contaminants and increase the mean free path of the molecules so they do not collide when
travelling to the substrate; which increases both the accuracy and the precision of the beam. Molecular
beams can be interrupted rapidly, in the order of a fraction of a second, which allows for almost atomically
abrupt transitions from one material to another [9]. Reflection High-Energy Electron Diffraction equipment
08/11/17 5

Fig. 3. Schematic of the MBE set up, showing the layout of the RHEED gun and the molecular beam from the effusion cells [10].

(RHEED) is used to study the substrate. A collimated electron beam produced from a ∼ 50 keV RHEED
electron gun is aimed at a shallow angle onto the substrate’s surface. The RHEED electron gun is aligned
such that the electron beam does not interfere with the molecular beam [8]. Accelerated electrons incident
on the substrate diffract from its surface atoms and produce a diffraction pattern on a RHEED fluorescent
screen.
2) RHEED Patterns: RHEED patterns are of great use during growth, allowing one to monitor the
growth on the substrate. Constructive interference patterns appear as spots/stripes on the screen, their
shape is characteristic of the roughness of the surface. Stripes occur for a flat substrate surface, whilst
spots are characteristic of a rough surface [8], this is the pattern observed when producing quantum dots
and can be seen in Fig. 4. The distance between neighbouring constructive interference spots is inversely
proportional to surface lattice constant, allowing to monitor the distance between dots. Furthermore, the
change in intensity of the pattern allows one to count the number of monolayers that have been deposited.
An incomplete monolayer will have a low intensity due to increased scattering from the surface, whereas
a fully finished monolayer will have less scatter and produce higher intensity interference.
3) Growth Mechanism: Quantum dots most commonly form due to Stranski-Krastanow growth, at
temperatures around 500◦ C [11]. The initial atoms arrange themselves in a planar layer, called the wetting
08/11/17 6

Fig. 4. left: RHEED patterns for (a) a flat surface and (b) a rough surface. Right: Schematic cross-section of the substrate. Increased surface
roughness produces dot shaped constructive interference, a flat surface produces striped constructive interference. [10]

layer, which will link the dots. Quantum dots form when a semiconductor layer is grown on a substrate
material of smaller lattice constant [8]. After this, atoms bunch up into islands as this a more energetically
favourable formation due to a reduction in strain energy between the islands [9]. This mechanism is self
assembling, since quantum dots appear spontaneously, growing at a rate of ∼ 5Ås−1 [9].
By changing the amount of deposited dot material, the size of the islands are adjusted and thus the
optical properties of the dot change - which will be discussed later. For now though, smaller dots emit
higher frequency photons [9].
This technique is preferred to etching as it can produce smaller QDs. Etched quantum dots are found
as small as 30nm [12] whereas whereas self-organised growth can produce dots as small as 5nm [13],
hence MBE can create dots which produce higher frequency photons. Furthermore, etching is a much
slower process. The biggest disadvantage of the S-K method, is the size fluctuations between each QD;
which is about ±5% for both diameter and height [11].

C. Droplet Epitaxy

The Stranski-Krastanov method works using lattice-mismatched material, leading to the quantum dots
that experience strain and consequently suffer from size and shape irregularities [14]. An exciting new
growth technique which could rival the Stranski-Krastanov method is a type of MBE called droplet epitaxy
[15]. Self-assembled QDs grown this way have no strain energy stored in them reducing the probability of
structural defects and have a comparable size distribution to the Stranski-Krastanov method (around 10%)
08/11/17 7

[14]. This method allows for larger QDs to be grown with little defects. The property of symmetrical
quantum dots is highly desirable for optical properties which will be discussed in Sec. V-B.

D. Metal organic chemical vapor deposition

Another technique for fabricating quantum dots is metal-organic chemical vapor deposition (MOCVD).
This report will not delve into the physics of MOCVD in too much detail, but it is important to note the
difference in the process compared to MBE. MOCVD uses heated gas flow to invoke a chemical reaction
at the substrate surface, this means temperatures in MOCVD operate at 500-1500◦ C, up to 1000◦ C hotter
than MBE. This means the process can also occur near atmospheric pressure [16]. The results of both
MBE and MOCVD are similar and are both more widely used than etching.

IV. T RANSPORT PROPERTIES

The Coulomb blockade is the increased resistance at certain small bias voltages in a QD. It is easiest
to understand this effect once the single-electron transistor is understood.

A. The single electron-transistor

In a single electron transistor (SET) there are two metal contacts: the source, where electrons enter the
transistor and the drain, where they leave. Between these contacts a thin layer of GaAs is sandwiched
between two layers of insulating AlGaAs. The insulating layers act as a barrier which electrons must
tunnel through in order to enter from the source contact, or leave the quantum dot via the drain contact

Fig. 5. A schematic displaying the layout of a single transistor. The source and drain electrodes connect to the island via insulating
semiconducting, acting as a tunnel junction. The gate electrode changes the voltage across the dot via a capacitor [17].
08/11/17 8

Fig. 6. Schematic energy diagram for the Coulomb Blockade. As voltage is increased, the discrete energy levels in the QD are lowered.
When an energy level is equal to the chemical potential of the source, an electron enters the dot [18].

[18]. A schematic diagram of this set up is shown in Fig. 5. The gate is an electrode which allows the
voltage across the dot to be changed and is connected to the dot via a gate capacitor. By changing the
voltage across the dot, the energy levels within the dot can be shifted up (increasing gate voltage) or
down (decreasing gate voltage).

B. Coulomb blockade

The number of electrons in a dot must be an integer N. In order for an electron to tunnel from the
source contact into the QD, it requires a Coulomb energy of

e2
E= (7)
2C
where e is the elementary charge and C is the capacitance between the carrier and the rest of the system.
This energy allows the electron to overcome the Coulomb repulsion between the electrons within the dot.
This energy barrier is the Coulomb blockade [18].
This means there are quantised energy levels, of separation ∆E, within the dot which can be seen in
Fig. 6 [18]. When applying a positive voltage to the gate, the energy levels of the island are lowered by
eVg [18], this can be seen in Fig. 6. The energy of a charge carrier in a QD is thus,
e2
E = −eVg + (8)
2C
where Vg is the gate voltage. The first term is the attractive interaction between the positive gate electrode
and the charge on the island, and the second term is the repulsive interaction between the charges on the
island [19]. When a quantised energy level is just lower than the chemical potential of the source, an
electron can tunnel from the source contact to the island and occupy the previously vacant energy level
[20]. The electron can now tunnel to the drain. This results in pulses of current with increasing voltage,
which is seen in Fig. 7.
08/11/17 9

Fig. 7. Conductance of a quantum dot in a SET with increasing voltage, at temperature = 60mK. The gap between each peak, ∆V ∼ e/Cg
. Thermal broadening occurs at high Vg , giving a Lorentzian fit after ∼ 290mV. [18]

V. O PTICAL PROPERTIES

A. Photoluminescence

As mentioned in Sec. II-A, in quantum dots, electrons and holes are confined to a size similar to their
de Broglie wavelength and such the energy levels within a dot are discrete.
Photoluminescence (PL) occurs when an electron is excited into the conduction band. When excited, a
hole is left behind in the valence band. This electron-hole pair is a quasiparticle called an exciton. When
the exciton recombines, the electron relaxes back into the valence band and a photon of a wavelength
characteristic of the bandgap is emitted [21].
PL lines are the convolution of every individual emission line from the total number of quantum dots
measured [10]. The lines broaden homogeneously and (within every dot) and inhomogeneously (from dot
to dot). Homeogenous broadening occurs due to phonon scattering and impurities, whilst inhomogeneous
broadening occurs due to the subtle differences in shape and size of each dot [21].
The PL line for individual quantum dots has also been studied. They exhibit much sharper lines, as
inhomogeneous PL broadening does not occur. Broader background emission is also present in these lines
which originate from delocalised excitons [22]. The majority of PL lines are also linearly polarized in the
direction of the plane, oriented orthogonally [22].

B. Fine-Structure Splitting And Entangled Photons

An interesting effect which can be observed in the PL line of a single quantum dot is fine-structure
splitting (FSS) [23]. It is found that the potential within a ‘real’ quantum dot is slightly asymmetrical
08/11/17 10

Fig. 8. Schematic of the decay from a biexciton state (XX) to an exciton state (X) to the ground state. (a) shows a standard quantum dot
with fine-structure splitting at the exciton state. Photons generated in recombination are vertically or horizontally polarized. (b) shows a
quantum dot with a degenerate exciton state, the decay from a biexciton state generates superposed circularly polarised photon pairs, which
are entangled. [24]

due to small irregularities in the QD’s shape, chemical composition, strain and piezoelectricity [25]. The
exciton’s wavefunction within most quantum dots is therefore asymmetrical and causes the exciton state
to split into two polarization dependant states - vertically polarised and horizontally polarized [23]. The
energy difference between these two polarization states is the fine-structure splitting (FSS). This can be
seen in Fig. 8 (a).
If the FSS can be made negligible, by making the potential across the dot symmetric, then the exciton
energy level becomes 2-fold degenerate i.e there are two states for an exciton energy level. It is then
possible for a cascade from a biexciton3 to the degenerate exciton state, to generate entangled photon
pairs. Entanglement occurs since the path of each cascade is spectrally indistinguishable for a degenerate
exciton state. This process can be seen in Fig. 8 (b).
Quantum dots grown using droplet epitaxy, discussed in Sec. III-C, have no strain and are therefore more
symmetrical than QDs produced via S-K growth. As a result, the average fine-structure splitting of QDs
grown by droplet epitaxy is much smaller than any other method to date and the existence of degenerate
exciton states has been observed in InAs/InAlAs dots [26]. This could make droplet epitaxy grown QDs
integral in quantum information networks, due to their ability to produce on-demand, entangled photons.
3
A biexciton is a quasiparticle formed from two excitons and is therefore two electrons and two holes.
08/11/17 11

C. Fermi’s Golden Rule

Fermi’s golden rule is a formula for the transition rate from a quantum state to a collection of final
states [27]. It shows that total transition probability increases linearly with time and hence the transition
probability per unit time is constant. The transition rate wif , is defined as,


wif = |Mif |2 ρf (Ef ) (9)

where h̄ is the reduced Planck’s constant, ρf (Ef ) is the density of final states and the ‘matrix element’,
Mif is

Z
Mif = Ψ∗f V Ψi dV (10)

here, ΨF is the wavefunction of the final state, Ψi is the wavefunction of the initial state and V is the
potential which operates on the initial state of the wavefunction.
Mif is called the ‘matrix element’ of the transition, it is a measure of the coupling between the initial
and final state in the transition [27]. It is shown in Eq. 9 that the transition probability is proportional to
the square of this integral and the density of final states. This rule holds true for all transitions within a
quantum dot, such as the biexciton cascades shown in Fig. 8.
Fermi’s golden rule explains an interesting effect in QDs for an electron that relaxes from the wetting
layer in the presence of a sufficiently high number of thermal phonons. An electron that has relaxed from
the wetting layer enters its excited state - from the excited state the electron can relax again, into the
ground state, or become re-excited back into the wetting layer. As in Eq. 9, the transition rate of the
electron is proportional to the transitions final density of states. The DOS of the wetting layer is much
higher than the effective DOS of the ground state (degeneracy) meaning an electron in its excited state
has a higher chance of thermal re-excitation back into the wetting layer than relaxing into the ground
state [28]. Once in the ground state, re-excitation is not as likely.
As a result, as the temperature of the QD increases, the relaxation time of an electron increases. The
relaxation time at low temperatures is less than a few picoseconds, but at around 300K this is of the
order ∼ 30 picoseconds [28]. This is can be a limiting factor in technologies that use quantum dots, for
example the modulation response in QD lasers.

D. Oscillator Strength

1) Definition: Oscillator strength (OS) in a quantum dot is a measure of the interaction strength between
the quantum dot and light [29]. Therefore, the OS is proportional to the optical transition rate from Fermi’s
golden rule [30]. It is defined as [31],
08/11/17 12

2m
fif = (Ef − Ei )|xif |2 (11)
h̄2
where m is the mass of the particle in the transition, Ef is the energy of the final state, Ei is the energy
of the final state and xif is the transition dipole moment, defined as

Z
xif = Ψ∗f x̂Ψi dx (12)

Oscillator strength varies with the size and shape of each quantum dot. In general, OS increases with
the size of the QD [32]. For smaller dots, such as the Stranski-Krastanow grown self-assembled dots
discussed in Sec. III-B3, OS is around 6, whereas for larger dots ∼ 100nm, OS increases to ∼ 75 [32].
In the strong confinement regime4 (around 15nm) the OS is less dependant on size, showing little change
with dot diameter [32].
2) Thomas-Reiche-Kuhn Rule: The sum of all OS’s over all the transitions from an initial quantum
state equals one; this is the Thomas-Reiche-Kuhn sum rule and is expressed as [31]

2m X
(Ef − Ei )|xif |2 = 1 (13)
h̄2 f

A common misconception is that an individual transitions OS cannot be greater than 1. This is untrue,
since the OS can be positive (an absorbing transition) or negative (an emitting transition)5 . The sum of
the positive and negative transitions will then sum to 1.

VI. C ONCLUSION

This report has shown that quantum dots are zero-dimensional systems, that posses discrete energy
levels due to their confinement of electrons and holes. They can be fabricated multiple ways such etching,
molecular beam epitaxy or MOCVD, although the latter two are more widely used. The unique effect
of the Coulomb blockade has been discussed in single electron transistors, where period conductance
resonances occur with increasing gate voltage. Photoluminescence as a result of exciton recombination is
already being used in displays to decrease thickness and improve colour quality. Whilst, future uses are
likely to exploit the narrow fine-structure splitting of droplet epitaxy grown dots, which could prove vital
in producing entangled photon pairs for the quantum technology industry.
4
Strong confinement regime: The quantum dots radius is less than the exciton Bohr radius
5
Provided Ef > Ei , the opposite is true for Ef < Ei
08/11/17 13

R EFERENCES

[1] L. Kouwenhoven, C. Marcus, ‘Quantum Dots’, Physics World, 11, 35 (1998)


[2] M. A. Reed, J.N Randall, R. J. Aggarwal et al. Observation of Discrete Electronic States in Zero-Dimensional Semiconductor
Nanostructure, Phys Rev Lett 60 (6), 535 (1998)
[3] M. Kuno, Introductory Nanoscience [Garland Science, 2011] ch. 9
[4] E. U. Rafailov, The Physics and Engineering of Compact Quantum Dot-based Lasers for Biophotonics [Wiley-VCH, 2014]
[5] M. Reed, Quantum Dots, Scientific American 268, 1, 118 (1993)
[6] J. N. Randall, M. A. Reed, M. Moore, R. J Matyi and J. W. Lee, Microstructure fabrication and transport through quantum dots, p. 1
(August 1987)
[7] R. Eason et. al, ‘Pulsed Laser Deposition of Thin Films’, [Wiley-Interscience, 2007]
[8] J. R. Arthur, Molecular beam epitaxy, Surface Science, 500, p. 189-217 (2002)
[9] V. M. Ustinov, A. E. Zhukov, A. Y. Egorov and N. A. Maleev, ‘Quantum Dot Lasers’, [Oxford Scholarship Online, January 2010]
[10] S. Franchi, G. Trevisi, L. Seravalli and P. Frigeri, Quantum dot nanostructures and molecular beam epitaxy, Progress in Crystal Growth
and Characterization of Materials, 47 23 (2003)
[11] P.M. Petroff, S.P. DenBaars, MBE and MOCVD growth and properties of self-assembling quantum dot arrays in III-V semiconductor
structures, In Superlattices and Microstructures, 15, 1, p. 15 (1994)
[12] A. P. Alisatos, ‘Semiconductor clusters, nanocrystals, and quantum dots.’, science 271, 933 (Feb 1996)
[13] A. J. Griffiths, Exploring MBE G rowth Of Quantum Dots: Low Density Growth For Quantum Information Devices (Dec 2014)
[14] J. G. Keizer, J. Bocquel, P. M. Koenraad, T. Mano, T. Noda and K. Sakoda, Atomic scale analysis of self assembled GaAs/AlGaAs
quantum dots grown by droplet epitaxy, Appl. Phys. Lett, 96 (2010)
[15] P. Tighineanu, R. Daveau, E. H. Lee, J. D. Song, S. Stobbe and P. Lodahl, Decay dynamics and exciton localization in large GaAs
quantum dots grown by droplet epitaxy, Phys Rev B, 88 (2013)
[16] https://www.k-space.com/applications/mocvd/ (accessed 01/11/2017)
[17] Anurag Srivastava et al., Influence of Boron Substitution on Conductance of Pyridine- and Pentane-Based Molecular Single Electron
Transistors, Journal of Electronic Materials, 45, 4 (2016)
[18] M. Kastner, ‘Artificial Atoms’, Physics Today, 46, p. 24 (Jan 1993)
[19] M. A. Kastner, The single-electron transistor, Rev. Mod. Phys., 64, 849 (1992)
[20] E. B. Foxman, M. Kastner et al., Effects of quantum levels on transport throu gh a Coulomb island, Phys Rev B, 47, 15 (1993)
[21] M. Wolf, J. Berezovsky, Homogeneous and inhomogeneous sources of optical transition broadening in room temperature CdSe/ZnS
nanocrystal quantum dots (July 2014)
[22] D. Gammon, E. S. Snow, B. V. Shanabrook, D. S. Katzer, and D. Park, Fine Structure Splitting in the Optical Spectra of Single GaAs
Quantum Dots, Phys. Rev, 76, 16 (April 1996)
[23] A. Hogele, B. Alen, F.Bickel et al., Exciton structure splitting of single InGaAs self-assembled quantum dots, In Physica E: Low-
dimensional Systems and Nanostructures, 21, 24, p. 175-179 (March 2004)
[24] Robert J. Young, R. Mark Stevenson, Paola Atkinson, Ken Cooper, David A. Ritchie, Andrew J. Shields, Entangled photons on-demand
from single quantum dots
[25] M. Klingbeil, C. Strelow, A. Stemmann, S. Mendach and W. Hansen, Single-dot Spectroscopy of GaAs Quantum Dots Fabricated by
Filling of Self-assembled Nanoholes, Nanoscale Res Lett, 5 (July 2010)
[26] X. Liu, N. Ha, H. Nakajima, T. Mano, T. Kuroda, B. Urbaszek, et al., Vanishing fine-structure splittings in telecommunication-wavelength
quantum dots grown on (111)A surfaces by droplet epitaxy, Phys Rev B, 90 (2014)
[27] E. Merzbacher, Quantum Mechanics, [John Wiley & Sons Inc., 1998], p. 503-515
[28] Fermis Golden Rule, Nonequilibrium Electron Capture From the Wetting Layer, and the Modulation Response in P-Doped Quantum-Dot
Lasers, IEEE Journal of Quantum Electronics,42, 3, (March 2006)
08/11/17 14

[29] M. D. Leistikow, J. Johansen, A. J. Kettelarij, P. Lodahl, and W. L. Vos, Size-dependent oscillator strength and quantum efficiency of
CdSe quantum dots controlled via the local density of states, Phys Rev B, 79 (2009)
[30] O. Gywat, G. Burkard, and D. Loss, Biexcitons in coupled quantum dots as a source of entangled photons, Phys Rev. B, 65
[31] Z. Hens, Can the oscillator strength of the quantum dot bandgap transition exceed unity?, Chem Phys Lett, 463, 391395 (2008)
[32] F. Henneberger, Semiconductor Quantum Bits

Das könnte Ihnen auch gefallen