Sunteți pe pagina 1din 127

Quantum dot

Colloidal quantum dots irradiated with a UV light.


Different sized quantum dots emit different color light
due to quantum confinement.

Quantum dots (QD) are very small


semiconductor particles, only several
nanometres in size, so small that their
optical and electronic properties differ
from those of larger particles. They are a
central theme in nanotechnology. Many
types of quantum dot will emit light of
specific frequencies if electricity or light is
applied to them, and these frequencies
can be precisely tuned by changing the
dots' size,[1][2] shape and material, giving
rise to many applications.

In the language of materials science,


nanoscale semiconductor materials tightly
confine either electrons or electron holes.
Quantum dots are also sometimes
referred to as artificial atoms, a term that
emphasizes that a quantum dot is a single
object with bound, discrete electronic
states, as is the case with naturally
occurring atoms or molecules.[3][4]

Quantum dots exhibit properties that are


intermediate between those of bulk
semiconductors and those of discrete
molecules. Their optoelectronic properties
change as a function of both size and
shape.[5][6] Larger QDs (radius of 5–6 nm,
for example) emit longer wavelengths
resulting in emission colors such as
orange or red. Smaller QDs (radius of 2–
3 nm, for example) emit shorter
wavelengths resulting in colors like blue
and green, although the specific colors
and sizes vary depending on the exact
composition of the QD.[7]

Because of their highly tunable properties,


QDs are of wide interest. Potential
applications include transistors, solar
cells, LEDs, diode lasers and second-
harmonic generation, quantum computing,
and medical imaging.[8] Additionally, their
small size allows for QDs to be suspended
in solution which leads to possible uses in
inkjet printing and spin-coating. [9]Another
technique where QDs have been used is
Langmuir-Blodgett. [10][11][12] These
processing techniques result in less
expensive and less time-consuming
methods of semiconductor fabrication.

Production

Quantum Dots with gradually stepping emission from


violet to deep red are being produced in a kg scale at
PlasmaChem GmbH

There are several ways to prepare


quantum dots, the principal ones involving
colloids.

Colloidal synthesis
Colloidal semiconductor nanocrystals are
synthesized from solutions, much like
traditional chemical processes. The main
difference is the product neither
precipitates as a bulk solid nor remains
dissolved.[5] Heating the solution at high
temperature, the precursors decompose
forming monomers which then nucleate
and generate nanocrystals. Temperature is
a critical factor in determining optimal
conditions for the nanocrystal growth. It
must be high enough to allow for
rearrangement and annealing of atoms
during the synthesis process while being
low enough to promote crystal growth.
The concentration of monomers is another
critical factor that has to be stringently
controlled during nanocrystal growth. The
growth process of nanocrystals can occur
in two different regimes, "focusing" and
"defocusing". At high monomer
concentrations, the critical size (the size
where nanocrystals neither grow nor
shrink) is relatively small, resulting in
growth of nearly all particles. In this
regime, smaller particles grow faster than
large ones (since larger crystals need
more atoms to grow than small crystals)
resulting in "focusing" of the size
distribution to yield nearly monodisperse
particles. The size focusing is optimal
when the monomer concentration is kept
such that the average nanocrystal size
present is always slightly larger than the
critical size. Over time, the monomer
concentration diminishes, the critical size
becomes larger than the average size
present, and the distribution "defocuses".

Cadmium sulfide quantum dots on cells


There are colloidal methods to produce
many different semiconductors. Typical
dots are made of binary compounds such
as lead sulfide, lead selenide, cadmium
selenide, cadmium sulfide, cadmium
telluride, indium arsenide, and indium
phosphide. Dots may also be made from
ternary compounds such as cadmium
selenide sulfide. These quantum dots can
contain as few as 100 to 100,000 atoms
within the quantum dot volume, with a
diameter of ≈10 to 50 atoms. This
corresponds to about 2 to 10 nanometers,
and at 10 nm in diameter, nearly 3 million
quantum dots could be lined up end to end
and fit within the width of a human thumb.
Ideallized image of colloidal nanoparticle of lead
sulfide (selenide) with complete passivation by oleic
acid, oleyl amine and hydroxyl ligands (size ≈5nm)

Large batches of quantum dots may be


synthesized via colloidal synthesis. Due to
this scalability and the convenience of
benchtop conditions, colloidal synthetic
methods are promising for commercial
applications. It is acknowledged to be the
least toxic of all the different forms of
synthesis.

Plasma synthesis

Plasma synthesis has evolved to be one of


the most popular gas-phase approaches
for the production of quantum dots,
especially those with covalent
bonds.[13][14][15] For example, silicon (Si)
and germanium (Ge) quantum dots have
been synthesized by using nonthermal
plasma. The size, shape, surface and
composition of quantum dots can all be
controlled in nonthermal plasma.[16][17]
Doping that seems quite challenging for
quantum dots has also been realized in
plasma synthesis.[18][19][20] Quantum dots
synthesized by plasma are usually in the
form of powder, for which surface
modification may be carried out. This can
lead to excellent dispersion of quantum
dots in either organic solvents[21] or
water[22] (i. e., colloidal quantum dots).

Fabrication

Self-assembled quantum dots are


typically between 5 and 50 nm in size.
Quantum dots defined by
lithographically patterned gate
electrodes, or by etching on two-
dimensional electron gasses in
semiconductor heterostructures can
have lateral dimensions between 20 and
100 nm.
Some quantum dots are small regions
of one material buried in another with a
larger band gap. These can be so-called
core–shell structures, e.g., with CdSe in
the core and ZnS in the shell, or from
special forms of silica called ormosil.
Sub-monolayer shells can also be
effective ways of passivating the
quantum dots, such as PbS cores with
sub-monolayer CdS shells.[23]
Quantum dots sometimes occur
spontaneously in quantum well
structures due to monolayer fluctuations
in the well's thickness.
Self-assembled quantum dots nucleate
spontaneously under certain conditions
during molecular beam epitaxy (MBE)
and metallorganic vapor phase epitaxy
(MOVPE), when a material is grown on a
substrate to which it is not lattice
matched. The resulting strain produces
coherently strained islands on top of a
two-dimensional wetting layer. This
growth mode is known as Stranski–
Krastanov growth. The islands can be
subsequently buried to form the
quantum dot. This fabrication method
has potential for applications in
quantum cryptography (i.e. single
photon sources) and quantum
computation. The main limitations of
this method are the cost of fabrication
and the lack of control over positioning
of individual dots.
Individual quantum dots can be created
from two-dimensional electron or hole
gases present in remotely doped
quantum wells or semiconductor
heterostructures called lateral quantum
dots. The sample surface is coated with
a thin layer of resist. A lateral pattern is
then defined in the resist by electron
beam lithography. This pattern can then
be transferred to the electron or hole
gas by etching, or by depositing metal
electrodes (lift-off process) that allow
the application of external voltages
between the electron gas and the
electrodes. Such quantum dots are
mainly of interest for experiments and
applications involving electron or hole
transport, i.e., an electrical current.
The energy spectrum of a quantum dot
can be engineered by controlling the
geometrical size, shape, and the
strength of the confinement potential.
Also, in contrast to atoms, it is relatively
easy to connect quantum dots by tunnel
barriers to conducting leads, which
allows the application of the techniques
of tunneling spectroscopy for their
investigation.

The quantum dot absorption features


correspond to transitions between
discrete, three-dimensional particle in a
box states of the electron and the hole,
both confined to the same nanometer-size
box.These discrete transitions are
reminiscent of atomic spectra and have
resulted in quantum dots also being called
artificial atoms.[24]

Confinement in quantum dots can also


arise from electrostatic potentials
(generated by external electrodes,
doping, strain, or impurities).
Complementary metal-oxide-
semiconductor (CMOS) technology can
be employed to fabricate silicon
quantum dots. Ultra small (L=20 nm,
W=20 nm) CMOS transistors behave as
single electron quantum dots when
operated at cryogenic temperature over
a range of −269 °C (4 K) to about
−258 °C (15 K). The transistor displays
Coulomb blockade due to progressive
charging of electrons one by one. The
number of electrons confined in the
channel is driven by the gate voltage,
starting from an occupation of zero
electrons, and it can be set to 1 or
many.[25]

Viral assembly

Genetically engineered M13 bacteriophage


viruses allow preparation of quantum dot
biocomposite structures.[26] It had
previously been shown that genetically
engineered viruses can recognize specific
semiconductor surfaces through the
method of selection by combinatorial
phage display.[27] Additionally, it is known
that liquid crystalline structures of wild-
type viruses (Fd, M13, and TMV) are
adjustable by controlling the solution
concentrations, solution ionic strength,
and the external magnetic field applied to
the solutions. Consequently, the specific
recognition properties of the virus can be
used to organize inorganic nanocrystals,
forming ordered arrays over the length
scale defined by liquid crystal formation.
Using this information, Lee et al. (2000)
were able to create self-assembled, highly
oriented, self-supporting films from a
phage and ZnS precursor solution. This
system allowed them to vary both the
length of bacteriophage and the type of
inorganic material through genetic
modification and selection.

Electrochemical assembly
Highly ordered arrays of quantum dots
may also be self-assembled by
electrochemical techniques. A template is
created by causing an ionic reaction at an
electrolyte-metal interface which results in
the spontaneous assembly of
nanostructures, including quantum dots,
onto the metal which is then used as a
mask for mesa-etching these
nanostructures on a chosen substrate.

Bulk-manufacture

Quantum dot manufacturing relies on a


process called "high temperature dual
injection" which has been scaled by
multiple companies for commercial
applications that require large quantities
(hundreds of kilograms to tonnes) of
quantum dots. This reproducible
production method can be applied to a
wide range of quantum dot sizes and
compositions.

The bonding in certain cadmium-free


quantum dots, such as III-V-based
quantum dots, is more covalent than that
in II-VI materials, therefore it is more
difficult to separate nanoparticle
nucleation and growth via a high
temperature dual injection synthesis. An
alternative method of quantum dot
synthesis, the “molecular seeding”
process, provides a reproducible route to
the production of high quality quantum
dots in large volumes. The process utilises
identical molecules of a molecular cluster
compound as the nucleation sites for
nanoparticle growth, thus avoiding the
need for a high temperature injection step.
Particle growth is maintained by the
periodic addition of precursors at
moderate temperatures until the desired
particle size is reached.[28] The molecular
seeding process is not limited to the
production of cadmium-free quantum
dots; for example, the process can be used
to synthesise kilogram batches of high
quality II-VI quantum dots in just a few
hours.

Another approach for the mass production


of colloidal quantum dots can be seen in
the transfer of the well-known hot-injection
methodology for the synthesis to a
technical continuous flow system. The
batch-to-batch variations arising from the
needs during the mentioned methodology
can be overcome by utilizing technical
components for mixing and growth as well
as transport and temperature
adjustments. For the production of CdSe
based semiconductor nanoparticles this
method has been investigated and tuned
to production amounts of kg per month.
Since the use of technical components
allows for easy interchange in regards of
maximum through-put and size, it can be
further enhanced to tens or even hundreds
of kilograms.[29]

In 2011 a consortium of U.S. and Dutch


companies reported a "milestone" in high
volume quantum dot manufacturing by
applying the traditional high temperature
dual injection method to a flow system.[30]

On January 23, 2013 Dow entered into an


exclusive licensing agreement with UK-
based Nanoco for the use of their low-
temperature molecular seeding method
for bulk manufacture of cadmium-free
quantum dots for electronic displays, and
on September 24, 2014 Dow commenced
work on the production facility in South
Korea capable of producing sufficient
quantum dots for "millions of cadmium-
free televisions and other devices, such as
tablets". Mass production is due to
commence in mid-2015.[31] On 24 March
2015 Dow announced a partnership deal
with LG Electronics to develop the use of
cadmium free quantum dots in
displays.[32]

Heavy metal-free quantum


dots

In many regions of the world there is now


a restriction or ban on the use of heavy
metals in many household goods, which
means that most cadmium based
quantum dots are unusable for consumer-
goods applications.

For commercial viability, a range of


restricted, heavy metal-free quantum dots
has been developed showing bright
emissions in the visible and near infra-red
region of the spectrum and have similar
optical properties to those of CdSe
quantum dots. Among these systems are
InP/ZnS and CuInS/ZnS, for example.
Peptides are being researched as potential
quantum dot material.[33] Since peptides
occur naturally in all organisms, such dots
would likely be nontoxic and easily
biodegraded.

Health and safety


Some quantum dots pose risks to human
health and the environment under certain
conditions.[34][35][36] Notably, the studies on
quantum dot toxicity are focused on
cadmium containing particles and has yet
to be demonstrated in animal models after
physiologically relevant dosing.[36] In vitro
studies, based on cell cultures, on
quantum dots (QD) toxicity suggests that
their toxicity may derive from multiple
factors including its physicochemical
characteristics (size, shape, composition,
surface functional groups, and surface
charges) and environment. Assessing their
potential toxicity is complex as these
factors include properties such as QD size,
charge, concentration, chemical
composition, capping ligands, and also on
their oxidative, mechanical and photolytic
stability.[34]

Many studies have focused on the


mechanism of QD cytotoxicity using model
cell cultures. It has been demonstrated
that after exposure to ultraviolet radiation
or oxidation by air, CdSe QDs release free
cadmium ions causing cell death.[37]
Group II-VI QDs also have been reported to
induce the formation of reactive oxygen
species after exposure to light, which in
turn can damage cellular components
such as proteins, lipids and DNA.[38] Some
studies have also demonstrated that
addition of a ZnS shell inhibit the process
of reactive oxygen species in CdSe QDs.
Another aspect of QD toxicity is the
process of their size dependent
intracellular pathways that concentrate
these particles in cellular organelles that
are inaccessible by metal ions, which may
result in unique patterns of cytotoxicity
compared to their constituent metal
ions.[39] The reports of QD localization in
the cell nucleus[40] present additional
modes of toxicity because they may
induce DNA mutation, which in turn will
propagate through future generation of
cells causing diseases.

Although concentration of QDs in certain


organelles have been reported in in vivo
studies using animal models, interestingly,
no alterations in animal behavior, weight,
hematological markers or organ damage
has been found through either histological
or biochemical analysis.[41] These finding
have led scientists to believe that
intracellular dose is the most important
deterring factor for QD toxicity. Therefore,
factors determining the QD endocytosis
that determine the effective intracellular
concentration, such as QD size, shape and
surface chemistry determine their toxicity.
Excretion of QDs through urine in animal
models also have demonstrated via
injecting radio-labeled ZnS capped CdSe
QDs where the ligand shell was labelled
with 99mTc.[42] Though multiple other
studies have concluded retention of QDs in
cellular levels,[36][43] exocytosis of QDs is
still poorly studied in the literature.
While significant research efforts have
broadened the understanding of toxicity of
QDs, there are large discrepancies in the
literature and questions still remains to be
answered. Diversity of this class material
as compared to normal chemical
substances makes the assessment of
their toxicity very challenging. As their
toxicity may also be dynamic depending
on the environmental factors such as pH
level, light exposure and cell type,
traditional methods of assessing toxicity
of chemicals such as LD50 are not
applicable for QDs. Therefore, researchers
are focusing on introducing novel
approaches and adapting existing
methods to include this unique class of
materials.[36] Furthermore, novel strategies
to engineer safer QDs are still under
exploration by the scientific community. A
recent novelty in the field is the discovery
of carbon quantum dots, a new generation
of optically-active nanoparticles potentially
capable of replacing semiconductor QDs,
but with the advantage of much lower
toxicity.

Optical properties
Fluorescence spectra of CdTe quantum dots of various
sizes. Different sized quantum dots emit different color
light due to quantum confinement.

In semiconductors, light absorption


generally leads to an electron being
excited from the valence to the conduction
band, leaving behind a hole. The electron
and the hole can bind to each other to
form an exciton. When this exciton
recombines (i.e. the electron resumes its
ground state), the exciton's energy can be
emitted as light. This is called
fluorescence. In a simplified model, the
energy of the emitted photon can be
understood as the sum of the band gap
energy between the highest occupied level
and the lowest unoccupied energy level,
the confinement energies of the hole and
the excited electron, and the bound energy
of the exciton (the electron-hole pair):

As the confinement energy depends on the


quantum dot's size, both absorption onset
and fluorescence emission can be tuned
by changing the size of the quantum dot
during its synthesis. The larger the dot, the
redder (lower energy) its absorption onset
and fluorescence spectrum. Conversely,
smaller dots absorb and emit bluer (higher
energy) light. Recent articles in
Nanotechnology and in other journals have
begun to suggest that the shape of the
quantum dot may be a factor in the
coloration as well, but as yet not enough
information is available. Furthermore, it
was shown [44] that the lifetime of
fluorescence is determined by the size of
the quantum dot. Larger dots have more
closely spaced energy levels in which the
electron-hole pair can be trapped.
Therefore, electron-hole pairs in larger
dots live longer causing larger dots to
show a longer lifetime.

To improve fluorescence quantum yield,


quantum dots can be made with "shells" of
a larger bandgap semiconductor material
around them. The improvement is
suggested to be due to the reduced
access of electron and hole to non-
radiative surface recombination pathways
in some cases, but also due to reduced
auger recombination in others.

Potential applications
Quantum dots are particularly promising
for optical applications due to their high
extinction coefficient.[45] They operate like
a single electron transistor and show the
Coulomb blockade effect. Quantum dots
have also been suggested as
implementations of qubits for quantum
information processing.

Tuning the size of quantum dots is


attractive for many potential applications.
For instance, larger quantum dots have a
greater spectrum-shift towards red
compared to smaller dots, and exhibit less
pronounced quantum properties.
Conversely, the smaller particles allow one
to take advantage of more subtle quantum
effects.
A device that produces visible light, through energy
transfer from thin layers of quantum wells to crystals
above the layers.[46]

Being zero-dimensional, quantum dots


have a sharper density of states than
higher-dimensional structures. As a result,
they have superior transport and optical
properties. They have potential uses in
diode lasers, amplifiers, and biological
sensors. Quantum dots may be excited
within a locally enhanced electromagnetic
field produced by gold nanoparticles,
which can then be observed from the
surface plasmon resonance in the
photoluminescent excitation spectrum of
(CdSe)ZnS nanocrystals. High-quality
quantum dots are well suited for optical
encoding and multiplexing applications
due to their broad excitation profiles and
narrow/symmetric emission spectra. The
new generations of quantum dots have
far-reaching potential for the study of
intracellular processes at the single-
molecule level, high-resolution cellular
imaging, long-term in vivo observation of
cell trafficking, tumor targeting, and
diagnostics.

CdSe nanocrystals are efficient triplet


photosensitizers.[47] Laser excitation of
small CdSe nanoparticles enables the
extraction of the excited state energy from
the Quantum Dots into bulk solution, thus
opening the door to a wide range of
potential applications such as
photodynamic therapy, photovoltaic
devices, molecular electronics, and
catalysis.

Computing
Quantum dot technology is potentially
relevant to solid-state quantum
computation. By applying small voltages
to the leads, current through the quantum
dot can be controlled and thereby precise
measurements of the spin and other
properties therein can be made. With
several entangled quantum dots, plus a
way of performing operations, quantum
calculations and the computers that would
perform them might be possible.

Biology

In modern biological analysis, various


kinds of organic dyes are used. However,
as technology advances, greater flexibility
in these dyes is sought.[48] To this end,
quantum dots have quickly filled in the
role, being found to be superior to
traditional organic dyes on several counts,
one of the most immediately obvious
being brightness (owing to the high
extinction coefficient combined with a
comparable quantum yield to fluorescent
dyes[49]) as well as their stability (allowing
much less photobleaching).[50] It has been
estimated that quantum dots are 20 times
brighter and 100 times more stable than
traditional fluorescent reporters.[48] For
single-particle tracking, the irregular
blinking of quantum dots is a minor
drawback. However, there have been
groups which have developed quantum
dots which are essentially nonblinking and
demonstrated their utility in single
molecule tracking experiments.[51][52]

The use of quantum dots for highly


sensitive cellular imaging has seen major
advances.[53] The improved photostability
of quantum dots, for example, allows the
acquisition of many consecutive focal-
plane images that can be reconstructed
into a high-resolution three-dimensional
image.[54] Another application that takes
advantage of the extraordinary
photostability of quantum dot probes is
the real-time tracking of molecules and
cells over extended periods of time.[55]
Antibodies, streptavidin,[56] peptides,[57]
DNA,[58] nucleic acid aptamers,[59] or
small-molecule ligands [60] can be used to
target quantum dots to specific proteins
on cells. Researchers were able to observe
quantum dots in lymph nodes of mice for
more than 4 months.[61]

Semiconductor quantum dots have also


been employed for in vitro imaging of pre-
labeled cells. The ability to image single-
cell migration in real time is expected to be
important to several research areas such
as embryogenesis, cancer metastasis,
stem cell therapeutics, and lymphocyte
immunology.

One application of quantum dots in


biology is as donor fluorophores in Förster
resonance energy transfer, where the large
extinction coefficient and spectral purity of
these fluorophores make them superior to
molecular fluorophores[62] It is also worth
noting that the broad absorbance of QDs
allows selective excitation of the QD donor
and a minimum excitation of a dye
acceptor in FRET-based studies.[63] The
applicability of the FRET model, which
assumes that the Quantum Dot can be
approximated as a point dipole, has
recently been demonstrated[64]

The use of quantum dots for tumor


targeting under in vivo conditions employ
two targeting schemes: active targeting
and passive targeting. In the case of active
targeting, quantum dots are functionalized
with tumor-specific binding sites to
selectively bind to tumor cells. Passive
targeting uses the enhanced permeation
and retention of tumor cells for the
delivery of quantum dot probes. Fast-
growing tumor cells typically have more
permeable membranes than healthy cells,
allowing the leakage of small
nanoparticles into the cell body. Moreover,
tumor cells lack an effective lymphatic
drainage system, which leads to
subsequent nanoparticle-accumulation.

Quantum dot probes exhibit in vivo toxicity.


For example, CdSe nanocrystals are highly
toxic to cultured cells under UV
illumination, because the particles
dissolve, in a process known as photolysis,
to release toxic cadmium ions into the
culture medium. In the absence of UV
irradiation, however, quantum dots with a
stable polymer coating have been found to
be essentially nontoxic.[61][65] Hydrogel
encapsulation of quantum dots allows for
quantum dots to be introduced into a
stable aqueous solution, reducing the
possibility of cadmium leakage.Then
again, only little is known about the
excretion process of quantum dots from
living organisms.[66]

In another potential application, quantum


dots are being investigated as the
inorganic fluorophore for intra-operative
detection of tumors using fluorescence
spectroscopy.

Delivery of undamaged quantum dots to


the cell cytoplasm has been a challenge
with existing techniques. Vector-based
methods have resulted in aggregation and
endosomal sequestration of quantum dots
while electroporation can damage the
semi-conducting particles and aggregate
delivered dots in the cytosol. Via cell
squeezing, quantum dots can be efficiently
delivered without inducing aggregation,
trapping material in endosomes, or
significant loss of cell viability. Moreover, it
has shown that individual quantum dots
delivered by this approach are detectable
in the cell cytosol, thus illustrating the
potential of this technique for single
molecule tracking studies.[67]

Photovoltaic devices
The tunable absorption spectrum and high
extinction coefficients of quantum dots
make them attractive for light harvesting
technologies such as photovoltaics.
Quantum dots may be able to increase the
efficiency and reduce the cost of today's
typical silicon photovoltaic cells.
According to an experimental proof from
2004,[68] quantum dots of lead selenide
can produce more than one exciton from
one high energy photon via the process of
carrier multiplication or multiple exciton
generation (MEG). This compares
favorably to today's photovoltaic cells
which can only manage one exciton per
high-energy photon, with high kinetic
energy carriers losing their energy as heat.
Quantum dot photovoltaics would
theoretically be cheaper to manufacture,
as they can be made "using simple
chemical reactions."

Quantum dot only solar cells

Aromatic self-assembled monolayers


(SAMs) (e.g. 4-nitrobenzoic acid) can be
used to improve the band alignment at
electrodes for better efficiencies. This
technique has provided a record power
conversion efficiency (PCE) of 10.7%.[69]
The SAM is positioned between ZnO-PbS
colloidal quantum dot (CQD) film junction
to modify band alignment via the dipole
moment of the constituent SAM molecule,
and the band tuning may be modified via
the density, dipole and the orientation of
the SAM molecule.[69]

Quantum dot in hybrid solar cells

Colloidal quantum dots are also used in


inorganic/organic hybrid solar cells. These
solar cells are attractive because of
potentially their low-cost fabrication and
relatively high efficiency.[70] Incorporation
of metal oxides, such as ZnO, TiO2, and
Nb2O5 nanomaterials into organic
photovoltaics have been commercialized
using full roll-to-roll processing.[70] A 13.2%
power conversion efficiency is claimed in
Si nanowire/PEDOT:PSS hybrid solar
cells.[71]

Quantum dot with nanowire in


solar cells

Another potential use involves capped


single-crystal ZnO nanowires with CdSe
quantum dots, immersed in
mercaptopropionic acid as hole transport
medium in order to obtain a QD-sensitized
solar cell. The morphology of the
nanowires allowed the electrons to have a
direct pathway to the photoanode. This
form of solar cell exhibits 50-60% internal
quantum efficiencies.[72]

Nanowires with quantum dot coatings on


silicon nanowires (SiNW) and carbon
quantum dots. The use of SiNWs instead
of planar silicon enhances the antiflection
properties of Si.[73] The SiNW exhibits a
light-trapping effect due to light trapping in
the SiNW. This use of SiNWs in
conjunction with carbon quantum dots
resulted in a solar cell that reached 9.10%
PCE.[73]

Graphene quantum dots have also been


blended with organic electronic materials
to improve efficiency and lower cost in
photovoltaic devices and organic light
emitting diodes (OLEDs) in compared to
graphene sheets. These graphene
quantum dots were functionalized with
organic ligands that experience
photoluminescence from UV-Vis
absorption.[74]

Light emitting diodes

Several methods are proposed for using


quantum dots to improve existing light-
emitting diode (LED) design, including
"Quantum Dot Light Emitting Diode" (QD-
LED) displays and "Quantum Dot White
Light Emitting Diode" (QD-WLED) displays.
Because Quantum dots naturally produce
monochromatic light, they can be more
efficient than light sources which must be
color filtered. QD-LEDs can be fabricated
on a silicon substrate, which allows them
to be integrated onto standard silicon-
based integrated circuits or
microelectromechanical systems.[75]

Quantum dot displays

Quantum dots are valued for displays,


because they emit light in very specific
gaussian distributions. This can result in a
display with visibly more accurate colors.
A conventional color liquid crystal display
(LCD) is usually backlit by fluorescent
lamps (CCFLs) or conventional white LEDs
that are color filtered to produce red,
green, and blue pixels. An improvement is
using conventional blue-emitting LEDs as
the light sources and converting part of
the emitted light into pure green and red
light by the appropriate quantum dots
placed in front of the blue LED or using a
quantum dot infused diffuser sheet in the
backlight optical stack. This type of white
light as the backlight of an LCD panel
allows for the best color gamut at lower
cost than a RGB LED combination using
three LEDs.
The ability of QDs to precisely convert and
tune a spectrum makes them attractive for
LCD displays. Previous LCD displays can
waste energy converting red-green poor,
blue-yellow rich white light into a more
balanced lighting. By using QDs, only the
necessary colors for ideal images are
contained in the screen. The result is a
screen that is brighter, clearer, and more
energy-efficient. The first commercial
application of quantum dots was the Sony
XBR X900A series of flat panel televisions
released in 2013.[76]

In June 2006, QD Vision announced


technical success in making a proof-of-
concept quantum dot display and show a
bright emission in the visible and near
infra-red region of the spectrum. A QD-LED
integrated at a scanning microscopy tip
was used to demonstrate fluorescence
near-field scanning optical microscopy
(NSOM) imaging.[77]

Photodetector devices

Quantum dot photodetectors (QDPs) can


be fabricated either via solution-
processing,[78] or from conventional single-
crystalline semiconductors.[79]
Conventional single-crystalline
semiconductor QDPs are precluded from
integration with flexible organic
electronics due to the incompatibility of
their growth conditions with the process
windows required by organic
semiconductors. On the other hand,
solution-processed QDPs can be readily
integrated with an almost infinite variety of
substrates, and also postprocessed atop
other integrated circuits. Such colloidal
QDPs have potential applications in
surveillance, machine vision, industrial
inspection, spectroscopy, and fluorescent
biomedical imaging.

Photocatalysts
Quantum dots also function as
photocatalysts for the light driven
chemical conversion of water into
hydrogen as a pathway to solar fuel. In
photocatalysis, electron hole pairs formed
in the dot under band gap excitation drive
redox reactions in the surrounding liquid.
Generally, the photocatalytic activity of the
dots is related to the particle size and its
degree of quantum confinement.[80] This is
because the band gap determines the
chemical energy that is stored in the dot in
the excited state. An obstacle for the use
of quantum dots in photocatalysis is the
presence of surfactants on the surface of
the dots. These surfactants (or ligands)
interfere with the chemical reactivity of the
dots by slowing down mass transfer and
electron transfer processes. Also,
quantum dots made of metal
chalcogenides are chemically unstable
under oxidizing conditions and undergo
photo corrosion reactions.

Theory
Quantum dots are theoretically described
as a point like, or a zero dimensional (0D)
entity. Most of their properties depend on
the dimensions, shape and materials of
which QDs are made. Generally QDs
present different thermodynamic
properties from the bulk materials of
which they are made. One of these effects
is the Melting-point depression. Optical
properties of spherical metallic QDs are
well described by the Mie scattering
theory.

Quantum confinement in
semiconductors

3D confined electron wave functions in a quantum dot.


Here, rectangular and triangular-shaped quantum dots
are shown. Energy states in rectangular dots are more
s-type and p-type. However, in a triangular dot the wave
functions are mixed due to confinement symmetry.
(Click for animation)

In a semiconductor crystallite whose size


is smaller than twice the size of its exciton
Bohr radius, the excitons are squeezed,
leading to quantum confinement. The
energy levels can then be predicted using
the particle in a box model in which the
energies of states depend on the length of
the box. Comparing the quantum dots size
to the Bohr radius of the electron and hole
wave functions, 3 regimes can be defined.
A 'strong confinement regime' is defined
as the quantum dots radius being smaller
than both electron and hole Bohr radius,
'weak confinement' is given when the
quantum dot is larger than both. For
semiconductors in which electron and hole
radii are markedly different, an
'intermediate confinement regime' exists,
where the quantum dot's radius is larger
than the Bohr radius of one charge carrier
(typically the hole), but not the other
charge carrier.[81]

Splitting of energy levels for small quantum dots due


to the quantum confinement effect. The horizontal axis
is the radius, or the size, of the quantum dots and ab*
is the Exciton Bohr radius.
Band gap energy
The band gap can become smaller in
the strong confinement regime as the
energy levels split up. The Exciton Bohr
radius can be expressed as:

where ab is the Bohr radius=0.053 nm, m


is the mass, μ is the reduced mass, and
εr is the size-dependent dielectric
constant (Relative permittivity). This
results in the increase in the total
emission energy (the sum of the energy
levels in the smaller band gaps in the
strong confinement regime is larger
than the energy levels in the band gaps
of the original levels in the weak
confinement regime) and the emission
at various wavelengths. If the size
distribution of QDs is not enough
peaked, the convolution of multiple
emission wavelengths is observed as a
continuous spectra.
Confinement energy
The exciton entity can be modeled using
the particle in the box. The electron and
the hole can be seen as hydrogen in the
Bohr model with the hydrogen nucleus
replaced by the hole of positive charge
and negative electron mass. Then the
energy levels of the exciton can be
represented as the solution to the
particle in a box at the ground level
(n = 1) with the mass replaced by the
reduced mass. Thus by varying the size
of the quantum dot, the confinement
energy of the exciton can be controlled.
Bound exciton energy
There is Coulomb attraction between
the negatively charged electron and the
positively charged hole. The negative
energy involved in the attraction is
proportional to Rydberg's energy and
inversely proportional to square of the
size-dependent dielectric constant[82] of
the semiconductor. When the size of the
semiconductor crystal is smaller than
the Exciton Bohr radius, the Coulomb
interaction must be modified to fit the
situation.

Therefore, the sum of these energies can


be represented as:

where μ is the reduced mass, a is the


radius, me is the free electron mass, mh is
the hole mass, and εr is the size-
dependent dielectric constant.

Although the above equations were


derived using simplifying assumptions,
they imply that the electronic transitions of
the quantum dots will depend on their
size. These quantum confinement effects
are apparent only below the critical size.
Larger particles do not exhibit this effect.
This effect of quantum confinement on the
quantum dots has been repeatedly verified
experimentally[83] and is a key feature of
many emerging electronic structures.[84]
The Coulomb interaction between
confined carriers can also be studied by
numerical means when results
unconstrained by asymptotic
approximations are pursued.[85]

Besides confinement in all three


dimensions (i.e., a quantum dot), other
quantum confined semiconductors
include:

Quantum wires, which confine electrons


or holes in two spatial dimensions and
allow free propagation in the third.
Quantum wells, which confine electrons
or holes in one dimension and allow free
propagation in two dimensions.

Models

A variety of theoretical frameworks exist


to model optical, electronic, and structural
properties of quantum dots. These may be
broadly divided into quantum mechanical,
semiclassical, and classical.

Quantum mechanics

Quantum mechanical models and


simulations of quantum dots often involve
the interaction of electrons with a
pseudopotential or random matrix.[86]
Semiclassical

Semiclassical models of quantum dots


frequently incorporate a chemical
potential. For example, the thermodynamic
chemical potential of an N-particle system
is given by

whose energy terms may be obtained as


solutions of the Schrödinger equation. The
definition of capacitance,

with the potential difference


may be applied to a quantum dot with the
addition or removal of individual electrons,

and .

Then

is the "quantum capacitance" of a


quantum dot, where we denoted by I(N)
the ionization potential and by A(N) the
electron affinity of the N-particle
system.[87]
Classical mechanics

Classical models of electrostatic


properties of electrons in quantum dots
are similar in nature to the Thomson
problem of optimally distributing electrons
on a unit sphere.

The classical electrostatic treatment of


electrons confined to spherical quantum
dots is similar to their treatment in the
Thomson,[88] or plum pudding model, of
the atom.[89]

The classical treatment of both two-


dimensional and three-dimensional
quantum dots exhibit electron shell-filling
behavior. A "periodic table of classical
artificial atoms" has been described for
two-dimensional quantum dots.[90] As well,
several connections have been reported
between the three-dimensional Thomson
problem and electron shell-filling patterns
found in naturally-occurring atoms found
throughout the periodic table.[91] This
latter work originated in classical
electrostatic modeling of electrons in a
spherical quantum dot represented by an
ideal dielectric sphere.[92]

History
The term “quantum dot” was coined in
1986.[93] They were first discovered in a
glass matrix and in colloidal solutions[94]
by Alexey Ekimov[95][96][97][98] and Louis
Brus.[99][100]

See also
Core–shell semiconductor nanocrystal
Cadmium-free quantum dots
Nanocrystal solar cell
Quantum dot laser
Quantum point contact
Superatom
Trojan wave packet
Langmuir-Blodgett film

Additional reading
Photoluminescence of a QD vs particle
diameter.[101]
Methods to produce quantum-confined
semiconductor structures (quantum
wires, wells and dots via grown by
advanced epitaxial techniques),
nanocrystals by gas-phase, liquid-phase
and solid-phase approaches.[102]

References
1. Sabaeian, Mohammad; Khaledi-Nasab, Ali
(2012-06-20). "Size-dependent intersubband
optical properties of dome-shaped
InAs/GaAs quantum dots with wetting
layer" . Applied Optics. 51 (18): 4176–4185.
Bibcode:2012ApOpt..51.4176S .
doi:10.1364/AO.51.004176 . ISSN 2155-
3165 .
2. Khaledi-Nasab, Ali; Sabaeian,
Mohammad; Sahrai, Mostafa; Fallahi, Vahid
(2014). "Kerr nonlinearity due to
intersubband transitions in a three-level
InAs/GaAs quantum dot: the impact of a
wetting layer on dispersion curves" .
Journal of Optics. 16 (5): 055004.
Bibcode:2014JOpt...16e5004K .
doi:10.1088/2040-8978/16/5/055004 .
ISSN 2040-8986 .
3. Ashoori, R. C. (1996). "Electrons in
artificial atoms". Nature. 379 (6564): 413–
419. Bibcode:1996Natur.379..413A .
doi:10.1038/379413a0 .
4. Kastner, M. A. (1993). "Artificial Atoms".
Physics Today. 46 (1): 24–31.
Bibcode:1993PhT....46a..24K .
doi:10.1063/1.881393 .
5. Murray, C. B.; Kagan, C. R.; Bawendi, M. G.
(2000). "Synthesis and Characterization of
Monodisperse Nanocrystals and Close-
Packed Nanocrystal Assemblies". Annual
Review of Materials Research. 30 (1): 545–
610. Bibcode:2000AnRMS..30..545M .
doi:10.1146/annurev.matsci.30.1.545 .
6. Brus, L.E. (2007). "Chemistry and Physics
of Semiconductor Nanocrystals" (PDF).
Retrieved 7 July 2009.
7. "Quantum Dots" . Nanosys – Quantum
Dot Pioneers. Retrieved 2015-12-04.
8. Ramírez, H. Y., Flórez J., and Camacho A.
S., (2015). "Efficient control of coulomb
enhanced second harmonic generation
from excitonic transitions in quantum dot
ensembles". Phys. Chem. Chem. Phys. 17
(37): 23938.
Bibcode:2015PCCP...1723938R .
doi:10.1039/C5CP03349G .
9. Coe-Sullivan, S.; Steckel, J. S.; Woo, W.-K.;
Bawendi, M. G.; Bulović, V. (2005-07-01).
"Large-Area Ordered Quantum-Dot
Monolayers via Phase Separation During
Spin-Casting". Advanced Functional
Materials. 15 (7): 1117–1124.
doi:10.1002/adfm.200400468 .
10. Xu, Shicheng; Dadlani, Anup L.; Acharya,
Shinjita; Schindler, Peter; Prinz, Fritz B.
"Oscillatory barrier-assisted Langmuir–
Blodgett deposition of large-scale quantum
dot monolayers" . Applied Surface Science.
367: 500–506.
doi:10.1016/j.apsusc.2016.01.243 .
11. Gorbachev, I. A.; Goryacheva, I. Yu;
Glukhovskoy, E. G. (2016-06-01).
"Investigation of Multilayers Structures
Based on the Langmuir-Blodgett Films of
CdSe/ZnS Quantum Dots" .
BioNanoScience. 6 (2): 153–156.
doi:10.1007/s12668-016-0194-0 .
ISSN 2191-1630 .
12. Achermann, Marc; Petruska, Melissa A.;
Crooker, Scott A.; Klimov, Victor I. (2003-12-
01). "Picosecond Energy Transfer in
Quantum Dot Langmuir−Blodgett
Nanoassemblies" . The Journal of Physical
Chemistry B. 107 (50): 13782–13787.
doi:10.1021/jp036497r . ISSN 1520-6106 .
13. Mangolini, L.; Thimsen, E.; Kortshagen,
U. (2005). "High-yield plasma synthesis of
luminescent silicon nanocrystals". Nano
Letters. 5 (4): 655–659.
Bibcode:2005NanoL...5..655M .
doi:10.1021/nl050066y . PMID 15826104 .
14. Knipping, J.; Wiggers, H.; Rellinghaus, B.;
Roth, P.; Konjhodzic, D.; Meier, C. (2004).
"Synthesis of high purity silicon
nanoparticles in a low Pressure microwave
reactor". Journal of Nanoscience and
Technology. 4 (8): 1039–1044.
doi:10.1166/jnn.2004.149 .
15. Sankaran, R. M.; Holunga, D.; Flagan, R.
C.; Giapis, K. P. (2005). "Synthesis of blue
luminescent Si nanoparticles using
atmospheric-pressure microdischarges".
Nano Letters. 5 (3): 537–541.
Bibcode:2005NanoL...5..537S .
doi:10.1021/nl0480060 . PMID 15755110 .
16. Kortshagen, U (2009). "Nonthermal
plasma synthesis of semiconductor
nanocrystals". J. Phys. D: Appl. Phys. 42
(11): 113001.
Bibcode:2009JPhD...42k3001K .
doi:10.1088/0022-3727/42/11/113001 .
17. Pi, X. D.; Kortshagen, U. (2009).
"Nonthermal plasma synthesized
freestanding silicon–germanium alloy
nanocrystals". Nanotechnology. 20 (29):
295602. Bibcode:2009Nanot..20C5602P .
doi:10.1088/0957-4484/20/29/295602 .
18. Pi, X. D.; Gresback, R.; Liptak, R. W.;
Campbell, S. A.; Kortshagen, U. (2008).
"Doping efficiency, dopant location, and
oxidation of Si nanocrystals". Applied
Physics Letters. 92 (2): 123102.
Bibcode:2008ApPhL..92b3102S .
doi:10.1063/1.2830828 .
19. Ni, Z. Y.; Pi, X. D.; Ali, M.; Zhou, S.;
Nozaki, T.; Yang, D. (2015). "Freestanding
doped silicon nanocrystals synthesized by
plasma". J. Phys. D: Appl. Phys. 48 (31):
314006. Bibcode:2015JPhD...48E4006N .
doi:10.1088/0022-3727/48/31/314006 .
20. Pereira, R. N.; Almeida, A. J. (2015).
"Doped semiconductor nanoparticles
synthesized in gas-phase plasmas". J.
Phys. D: Appl. Phys. 48 (31): 314005.
Bibcode:2015JPhD...48E4005P .
doi:10.1088/0022-3727/48/31/314005 .
21. Mangolini, L.; Kortshagen, U. (2007).
"Plasma-assisted synthesis of silicon
nanocrystal inks". Advanced Materials. 19
(18): 2513–2519.
doi:10.1002/adma.200700595 .
22. Pi, X. D.; Yu, T.; Yang, D. (2014). "Water-
dispersible silicon-quantum-dot-containing
micelles self-assembled from an
amphiphilic polymer". Part. Part. Syst.
Charact. 31 (7): 751–756.
doi:10.1002/ppsc.201300346 .
23. Clark, Pip; Radtke, Hanna; Pengpad,
Atip; Williamson, Andrew; Spencer, Ben;
Hardman, Samantha; Neo, Darren;
Fairclough, Simon; et al. (2017). "The
Passivating Effect of Cadmium in PbS / CdS
Colloidal Quantum Dot Solar Cells Probed
by nm-Scale Depth Profiling". Nanoscale. 9
(18): 6056–6067.
doi:10.1039/c7nr00672a .
24. Silbey, Robert J.; Alberty, Robert A.;
Bawendi, Moungi G. (2005). Physical
Chemistry, 4th ed. John Wiley &Sons.
p. 835.
25. Prati, Enrico; De Michielis, Marco; Belli,
Matteo; Cocco, Simone; Fanciulli, Marco;
Kotekar-Patil, Dharmraj; Ruoff, Matthias;
Kern, Dieter P; et al. (2012). "Few electron
limit of n-type metal oxide semiconductor
single electron transistors".
Nanotechnology. 23 (21): 215204.
arXiv:1203.4811  .
Bibcode:2012Nanot..23u5204P .
doi:10.1088/0957-4484/23/21/215204 .
PMID 22552118 .
26. Lee SW, Mao C, Flynn CE, Belcher AM
(2002). "Ordering of quantum dots using
genetically engineered viruses". Science.
296 (5569): 892–5.
Bibcode:2002Sci...296..892L .
doi:10.1126/science.1068054 .
PMID 11988570 .
27. Whaley SR, English DS, Hu EL, Barbara
PF, Belcher AM (2000). "Selection of
peptides with semiconductor binding
specificity for directed nanocrystal
assembly". Nature. 405 (6787): 665–8.
Bibcode:2000Natur.405..665W .
doi:10.1038/35015043 . PMID 10864319 .
28. Jawaid A.M.; Chattopadhyay S.; Wink
D.J.; Page L.E.; Snee P.T. (2013). "A". ACS
Nano. 7 (4): 3190–3197.
doi:10.1021/nn305697q . PMID 23441602 .
29. Continuous Flow Synthesis Method for
Fluorescent Quantum Dots . Azonano.com
(2013-06-01). Retrieved on 2015-07-19.
30. Quantum Materials Corporation and the
Access2Flow Consortium (2011). "Quantum
materials corp achieves milestone in High
Volume Production of Quantum Dots" .
Retrieved 7 July 2011.
31. The Times (25 September 2014).
"Nanoco and Dow tune in for sharpest
picture yet" . Retrieved 9 May 2015.
32. MFTTech (24 March 2015). "LG
Electronics Partners with Dow to
Commercialize LGs New Ultra HD TV with
Quantum Dot Technology" . Retrieved
9 May 2015.
33. Hauser, Charlotte A. E.; Zhang,
Shuguang (2010). "Peptides as biological
semiconductors". Nature. 468 (7323): 516–
517. Bibcode:2010Natur.468..516H .
doi:10.1038/468516a . PMID 21107418 .
34. Hardman, R. (2006). "A Toxicologic
Review of Quantum Dots: Toxicity Depends
on Physicochemical and Environmental
Factors" . Environmental Health
Perspectives. 114 (2): 165–72.
doi:10.1289/ehp.8284 . PMC 1367826  .
PMID 16451849 .
35. Pelley, J. L.; Daar, A. S.; Saner, M. A.
(2009). "State of Academic Knowledge on
Toxicity and Biological Fate of Quantum
Dots" . Toxicological Sciences. 112 (2):
276–296. doi:10.1093/toxsci/kfp188 .
PMC 2777075  . PMID 19684286 .
36. Tsoi, Kim M.; Dai, Qin; Alman, Benjamin
A.; Chan, Warren C. W. (2013-03-19). "Are
Quantum Dots Toxic? Exploring the
Discrepancy Between Cell Culture and
Animal Studies". Accounts of Chemical
Research. 46 (3): 662–671.
doi:10.1021/ar300040z . PMID 22853558 .
37. Derfus, Austin M.; Chan, Warren C. W.;
Bhatia, Sangeeta N. (2004-01-01). "Probing
the Cytotoxicity of Semiconductor Quantum
Dots". Nano Letters. 4 (1): 11–18.
Bibcode:2004NanoL...4...11D .
doi:10.1021/nl0347334 .
38. Liu, Wei; Zhang, Shuping; Wang, Lixin;
Qu, Chen; Zhang, Changwen; Hong, Lei;
Yuan, Lin; Huang, Zehao; Wang, Zhe (2011-
09-29). "CdSe Quantum Dot (QD)-Induced
Morphological and Functional Impairments
to Liver in Mice" . PLoS ONE. 6 (9): e24406.
Bibcode:2011PLoSO...624406L .
doi:10.1371/journal.pone.0024406 .
PMC 3182941  . PMID 21980346 .
39. Parak, W.j.; Boudreau, R.; Le Gros, M.;
Gerion, D.; Zanchet, D.; Micheel, C.m.;
Williams, S.c.; Alivisatos, A.p.; Larabell, C.
(2002-06-18). "Cell Motility and Metastatic
Potential Studies Based on Quantum Dot
Imaging of Phagokinetic Tracks". Advanced
Materials. 14 (12): 882–885.
doi:10.1002/1521-
4095(20020618)14:12<882::AID-
ADMA882>3.0.CO;2-Y .
40. Green, Mark; Howman, Emily (2005).
"Semiconductor quantum dots and free
radical induced DNA nicking". Chemical
Communications (1): 121.
doi:10.1039/b413175d .
41. Hauck, T. S.; Anderson, R. E.; Fischer, H.
C.; Newbigging, S.; Chan, W. C. W. (2010).
"In vivo Quantum-Dot Toxicity Assessment".
Small. 6 (1): 138–44.
doi:10.1002/smll.200900626 .
PMID 19743433 .
42. Soo Choi, Hak; Liu, Wenhao; Misra,
Preeti; Tanaka, Eiichi; Zimmer, John P.; Itty
Ipe, Binil; Bawendi, Moungi G.; Frangioni,
John V. (2007-10-01). "Renal clearance of
quantum dots" . Nature Biotechnology. 25
(10): 1165–1170. doi:10.1038/nbt1340 .
PMC 2702539  . PMID 17891134 .
43. Fischer, Hans C.; Hauck, Tanya S.;
Gómez-Aristizábal, Alejandro; Chan, Warren
C. W. (2010-06-18). "Exploring Primary Liver
Macrophages for Studying Quantum Dot
Interactions with Biological Systems".
Advanced Materials. 22 (23): 2520–2524.
doi:10.1002/adma.200904231 .
PMID 20491094 .
44. Van Driel; A. F. (2005). "Frequency-
Dependent Spontaneous Emission Rate
from CdSe and CdTe Nanocrystals:
Influence of Dark States" (PDF). Physical
Review Letters. 95 (23): 236804.
arXiv:cond-mat/0509565  .
Bibcode:2005PhRvL..95w6804V .
doi:10.1103/PhysRevLett.95.236804 .
PMID 16384329 .
45. Leatherdale, C. A.; Woo, W. -K.; Mikulec,
F. V.; Bawendi, M. G. (2002). "On the
Absorption Cross Section of CdSe
Nanocrystal Quantum Dots". The Journal of
Physical Chemistry B. 106 (31): 7619–
7622. doi:10.1021/jp025698c .
46. Achermann, M.; Petruska, M. A.; Smith,
D. L.; Koleske, D. D.; Klimov, V. I. (2004).
"Energy-transfer pumping of semiconductor
nanocrystals using an epitaxial quantum
well". Nature. 429 (6992): 642–646.
Bibcode:2004Natur.429..642A .
doi:10.1038/nature02571 .
PMID 15190347 .
47. Mongin C.; Garakyaraghi S.;
Razgoniaeva N.; Zamkov M.; Castellano F.N.
(2016). "Direct observation of triplet energy
transfer from semiconductor nanocrystals".
Science. 351 (6271): 369–372.
Bibcode:2016Sci...351..369M .
doi:10.1126/science.aad6378 .
PMID 26798011 .
48. Walling, M. A.; Novak, Shepard (February
2009). "Quantum Dots for Live Cell and In
Vivo Imaging" . Int. J. Mol. Sci. 10 (2): 441–
491. doi:10.3390/ijms10020441 .
PMC 2660663  . PMID 19333416 .
49. Michalet, X.; Pinaud, F. F.; Bentolila, L. A.;
Tsay, J. M.; Doose, S.; Li, J. J.; Sundaresan,
G.; Wu, A. M.; Gambhir, S. S.; Weiss, S.
(2005). "Quantum Dots for Live Cells, in
Vivo Imaging, and Diagnostics" . Science.
307 (5709): 538–44.
Bibcode:2005Sci...307..538M .
doi:10.1126/science.1104274 .
PMC 1201471  . PMID 15681376 .
50. Juan Carlos Stockert, Alfonso Blázquez-
Castro (2017). "Chapter 18 Luminescent
Solid-State Markers". Fluorescence
Microscopy in Life Sciences . Bentham
Science Publishers. pp. 606–641. ISBN 978-
1-68108-519-7. Retrieved 24 December
2017.
51. Marchuk, K.; Guo, Y.; Sun, W.; Vela, J.;
Fang, N. (2012). "High-Precision Tracking
with Non-blinking Quantum Dots Resolves
Nanoscale Vertical Displacement". Journal
of the American Chemical Society. 134 (14):
6108–11. doi:10.1021/ja301332t .
PMID 22458433 .
52. Lane, L. A.; Smith, A. M.; Lian, T.; Nie, S.
(2014). "Compact and Blinking-Suppressed
Quantum Dots for Single-Particle Tracking
in Live Cells" . The Journal of Physical
Chemistry B. 118 (49): 14140–7.
doi:10.1021/jp5064325 . PMC 4266335  .
PMID 25157589 .
53. Spie (2014). "Paul Selvin Hot Topics
presentation: New Small Quantum Dots for
Neuroscience". SPIE Newsroom.
doi:10.1117/2.3201403.17 .
54. Tokumasu, F; Fairhurst, Rm; Ostera, Gr;
Brittain, Nj; Hwang, J; Wellems, Te; Dvorak,
Ja (2005). "Band 3 modifications in
Plasmodium falciparum-infected AA and CC
erythrocytes assayed by autocorrelation
analysis using quantum dots". Journal of
Cell Science (Free full text). 118 (Pt 5):
1091–8. doi:10.1242/jcs.01662 .
PMID 15731014 .
55. Dahan, M. (2003). "Diffusion Dynamics
of Glycine Receptors Revealed by Single-
Quantum Dot Tracking". Science. 302
(5644): 442–5.
Bibcode:2003Sci...302..442D .
doi:10.1126/science.1088525 .
PMID 14564008 .
56. Howarth, M.; Liu, W.; Puthenveetil, S.;
Zheng, Y.; Marshall, L. F.; Schmidt, M. M.;
Wittrup, K. D.; Bawendi, M. G.; Ting, A. Y.
(2008). "Monovalent, reduced-size quantum
dots for imaging receptors on living cells" .
Nature Methods. 5 (5): 397–9.
doi:10.1038/nmeth.1206 . PMC 2637151  .
PMID 18425138 .
57. Akerman, M. E.; Chan, W. C. W.;
Laakkonen, P.; Bhatia, S. N.; Ruoslahti, E.
(2002). "Nanocrystal targeting in vivo" .
Proceedings of the National Academy of
Sciences. 99 (20): 12617–21.
Bibcode:2002PNAS...9912617A .
doi:10.1073/pnas.152463399 .
PMC 130509  . PMID 12235356 .
58. Farlow, J.; Seo, D.; Broaders, K. E.; Taylor,
M. J.; Gartner, Z. J.; Jun, Y. W. (2013).
"Formation of targeted monovalent
quantum dots by steric exclusion" . Nature
Methods. 10 (12): 1203–5.
doi:10.1038/nmeth.2682 . PMC 3968776  .
PMID 24122039 .
59. Dwarakanath, S.; Bruno, J. G.; Shastry,
A.; Phillips, T.; John, A.; Kumar, A.;
Stephenson, L. D. (2004). "Quantum dot-
antibody and aptamer conjugates shift
fluorescence upon binding bacteria".
Biochemical and Biophysical Research
Communications. 325 (3): 739–43.
doi:10.1016/j.bbrc.2004.10.099 .
PMID 15541352 .
60. Zherebetskyy D.; Scheele M.; Zhang Y.;
Bronstein N.; Thompson C.; Britt D.;
Salmeron M.; Alivisatos P.; Wang L.W.
(2014). "Hydroxylation of the surface of PbS
nanocrystals passivated with oleic acid".
Science. 344 (6190): 1380–1384.
Bibcode:2014Sci...344.1380Z .
doi:10.1126/science.1252727 .
PMID 24876347 .
61. Ballou, B.; Lagerholm, B. C.; Ernst, L. A.;
Bruchez, M. P.; Waggoner, A. S. (2004).
"Noninvasive Imaging of Quantum Dots in
Mice". Bioconjugate Chemistry. 15 (1): 79–
86. doi:10.1021/bc034153y .
PMID 14733586 .
62. Resch-Genger, Ute; Grabolle, Markus;
Cavaliere-Jaricot, Sara; Nitschke, Roland;
Nann, Thomas (28 August 2008). "Quantum
dots versus organic dyes as fluorescent
labels". Nature Methods. 5 (9): 763–775.
doi:10.1038/nmeth.1248 .
PMID 18756197 .
63. Algar, W. Russ; Krull, Ulrich J. (7
November 2007). "Quantum dots as donors
in fluorescence resonance energy transfer
for the bioanalysis of nucleic acids,
proteins, and other biological molecules".
Analytical and Bioanalytical Chemistry. 391
(5): 1609–1618. doi:10.1007/s00216-007-
1703-3 . PMID 17987281 .
64. Beane, Gary; Boldt, Klaus; Kirkwood,
Nicholas; Mulvaney, Paul (7 August 2014).
"Energy Transfer between Quantum Dots
and Conjugated Dye Molecules". The
Journal of Physical Chemistry C. 118 (31):
18079–18086. doi:10.1021/jp502033d .
65. Pelley, J. L.; Daar, A. S.; Saner, M. A.
(2009). "State of Academic Knowledge on
Toxicity and Biological Fate of Quantum
Dots" . Toxicological Sciences. 112 (2):
276–96. doi:10.1093/toxsci/kfp188 .
PMC 2777075  . PMID 19684286 .
66. Soo Choi, H.; Liu, W.; Misra, P.; Tanaka,
E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.;
Frangioni, J. V. (2007). "Renal clearance of
quantum dots" . Nature Biotechnology. 25
(10): 1165–70. doi:10.1038/nbt1340 .
PMC 2702539  . PMID 17891134 .
67. Sharei, A.; Zoldan, J.; Adamo, A.; Sim, W.
Y.; Cho, N.; Jackson, E.; Mao, S.; Schneider,
S.; Han, M. -J.; Lytton-Jean, A.; Basto, P. A.;
Jhunjhunwala, S.; Lee, J.; Heller, D. A.; Kang,
J. W.; Hartoularos, G. C.; Kim, K. -S.;
Anderson, D. G.; Langer, R.; Jensen, K. F.
(2013). "A vector-free microfluidic platform
for intracellular delivery" . Proceedings of
the National Academy of Sciences. 110 (6):
2082–7. Bibcode:2013PNAS..110.2082S .
doi:10.1073/pnas.1218705110 .
PMC 3568376  . PMID 23341631 .
68. Schaller, R.; Klimov, V. (2004). "High
Efficiency Carrier Multiplication in PbSe
Nanocrystals: Implications for Solar Energy
Conversion". Physical Review Letters. 92
(18): 186601. arXiv:cond-mat/0404368  .
Bibcode:2004PhRvL..92r6601S .
doi:10.1103/PhysRevLett.92.186601 .
PMID 15169518 .
69. Kim, Gi-Hwan; Arquer, F. Pelayo García
de; Yoon, Yung Jin; Lan, Xinzheng; Liu,
Mengxia; Voznyy, Oleksandr; Yang, Zhenyu;
Fan, Fengjia; Ip, Alexander H. (2015-11-02).
"High-Efficiency Colloidal Quantum Dot
Photovoltaics via Robust Self-Assembled
Monolayers". Nano Letters. 15 (11): 7691–
7696. Bibcode:2015NanoL..15.7691K .
doi:10.1021/acs.nanolett.5b03677 .
PMID 26509283 .
70. Krebs, Frederik C.; Tromholt, Thomas;
Jørgensen, Mikkel (2010). "Upscaling of
polymer solar cell fabrication using full roll-
to-roll processing". Nanoscale. 2 (6): 873–
86. Bibcode:2010Nanos...2..873K .
doi:10.1039/b9nr00430k .
PMID 20648282 .
71. Park, Kwang-Tae; Kim, Han-Jung; Park,
Min-Joon; Jeong, Jun-Ho; Lee, Jihye; Choi,
Dae-Geun; Lee, Jung-Ho; Choi, Jun-Hyuk
(2015-07-15). "13.2% efficiency Si
nanowire/PEDOT:PSS hybrid solar cell using
a transfer-imprinted Au mesh electrode" .
Scientific Reports. 5: 12093.
Bibcode:2015NatSR...512093P .
doi:10.1038/srep12093 . PMC 4502511  .
PMID 26174964 .
72. Leschkies, Kurtis S.; Divakar,
Ramachandran; Basu, Joysurya; Enache-
Pommer, Emil; Boercker, Janice E.; Carter, C.
Barry; Kortshagen, Uwe R.; Norris, David J.;
Aydil, Eray S. (2007-06-01).
"Photosensitization of ZnO Nanowires with
CdSe Quantum Dots for Photovoltaic
Devices". Nano Letters. 7 (6): 1793–1798.
Bibcode:2007NanoL...7.1793L .
doi:10.1021/nl070430o . PMID 17503867 .
73. Xie, Chao; Nie, Biao; Zeng, Longhui;
Liang, Feng-Xia; Wang, Ming-Zheng; Luo,
Linbao; Feng, Mei; Yu, Yongqiang; Wu, Chun-
Yan (2014-04-22). "Core–Shell
Heterojunction of Silicon Nanowire Arrays
and Carbon Quantum Dots for Photovoltaic
Devices and Self-Driven Photodetectors".
ACS Nano. 8 (4): 4015–4022.
doi:10.1021/nn501001j . PMID 24665986 .
74. Gupta, Vinay; Chaudhary, Neeraj;
Srivastava, Ritu; Sharma, Gauri Datt;
Bhardwaj, Ramil; Chand, Suresh (2011-07-
06). "Luminscent Graphene Quantum Dots
for Organic Photovoltaic Devices". Journal
of the American Chemical Society. 133 (26):
9960–9963. doi:10.1021/ja2036749 .
PMID 21650464 .
75. "Nano LEDs printed on silicon" .
nanotechweb.org. 3 July 2009.
76. Bullis, Kevin. (2013-01-11) Quantum
Dots Produce More Colorful Sony TVs | MIT
Technology Review .
Technologyreview.com. Retrieved on 2015-
07-19.
77. Hoshino, Kazunori; Gopal, Ashwini; Glaz,
Micah S.; Vanden Bout, David A.; Zhang,
Xiaojing (2012). "Nanoscale fluorescence
imaging with quantum dot near-field
electroluminescence". Applied Physics
Letters. 101 (4): 043118.
Bibcode:2012ApPhL.101d3118H .
doi:10.1063/1.4739235 .
78. Konstantatos, G.; Sargent, E. H. (2009).
"Solution-Processed Quantum Dot
Photodetectors". Proceedings of the IEEE.
97 (10): 1666–1683.
doi:10.1109/JPROC.2009.2025612 .
79. Vaillancourt, J.; Lu, X.-J.; Lu, Xuejun
(2011). "A High Operating Temperature
(HOT) Middle Wave Infrared (MWIR)
Quantum-Dot Photodetector". Optics and
Photonics Letters. 4 (2): 1–5.
doi:10.1142/S1793528811000196 .
80. Zhao, Jing; Holmes, Michael A.;
Osterloh, Frank E. (2013). "Quantum
Confinement Controls Photocatalysis: A
Free Energy Analysis for Photocatalytic
Proton Reduction at Cd Se Nanocrystals".
ACS Nano. 7 (5): 4316–25.
doi:10.1021/nn400826h . PMID 23590186 .
81. Ramírez, H. Y., Lin C. H., Chao, C. C.,
Hsu, Y., You, W. T., Huang, S. Y., Chen, Y. T.,
Tseng, H. C., Chang, W. H., Lin, S. D., and
Cheng, S. J., (2010). "Optical fine structures
of highly quantized InGaAs/GaAs self-
assembled quantum dots" . Phys. Rev. B. 81
(3): 245324.
Bibcode:2010PhRvB..81x5324R .
doi:10.1103/PhysRevB.81.245324 .
82. Brandrup, J.; Immergut, E.H. (1966).
Polymer Handbook (2 ed.). New York: Wiley.
pp. 240–246.
83. Khare, Ankur, Wills, Andrew W.,
Ammerman, Lauren M., Noris, David J., and
Aydil, Eray S. (2011). "Size control and
quantum confinement in Cu2ZnSnX4
nanocrystals". Chem. Commun. 47 (42):
11721. doi:10.1039/C1CC14687D .
84. Greenemeier, L. (5 February 2008). "New
Electronics Promise Wireless at Warp
Speed" . Scientific American.
85. Ramírez, H. Y. & Santana, A., (2012).
"Two interacting electrons confined in a 3D
parabolic cylindrically symmetric potential,
in presence of axial magnetic field: A finite
element approach". Comp. Phys. Comm.
183 (8): 1654.
Bibcode:2012CoPhC.183.1654R .
doi:10.1016/j.cpc.2012.03.002 .
86. Zumbühl DM, Miller JB, Marcus CM,
Campman K, Gossard AC (December 2002).
"Spin-orbit coupling, antilocalization, and
parallel magnetic fields in quantum dots".
Phys. Rev. Lett. 89 (27): 276803. arXiv:cond-
mat/0208436  .
Bibcode:2002PhRvL..89A6803Z .
doi:10.1103/PhysRevLett.89.276803 .
PMID 12513231 .
87. Iafrate, G. J.; Hess, K.; Krieger, J. B.;
Macucci, M. (1995). "Capacitive nature of
atomic-sized structures". Phys. Rev. B. 52
(15): 10737–10739.
Bibcode:1995PhRvB..5210737I .
doi:10.1103/physrevb.52.10737 .
88. Thomson, J.J. (1904). "On the Structure
of the Atom: an Investigation of the Stability
and Periods of Oscillation of a number of
Corpuscles arranged at equal intervals
around the Circumference of a Circle; with
Application of the Results to the Theory of
Atomic Structure" (extract of paper).
Philosophical Magazine. Series 6. 7 (39):
237–265.
doi:10.1080/14786440409463107 .
89. Bednarek, S.; Szafran, B. & Adamowski,
J. (1999). "Many-electron artificial atoms".
Phys. Rev. B. 59 (20): 13036–13042.
Bibcode:1999PhRvB..5913036B .
doi:10.1103/PhysRevB.59.13036 .
90. Bedanov, V. M. & Peeters, F. M. (1994).
"Ordering and phase transitions of charged
particles in a classical finite two-
dimensional system". Physical Review B. 49
(4): 2667–2676.
Bibcode:1994PhRvB..49.2667B .
doi:10.1103/PhysRevB.49.2667 .
91. LaFave, T. Jr. (2013). "Correspondences
between the classical electrostatic
Thomson Problem and atomic electronic
structure". Journal of Electrostatics. 71 (6):
1029–1035.
doi:10.1016/j.elstat.2013.10.001 .
92. LaFave, T. Jr. (2011). "The discrete
charge dielectric model of electrostatic
energy". Journal of Electrostatics. 69 (5):
414–418.
doi:10.1016/j.elstat.2013.10.001 .
93. Reed, M. A.; Bate, R. T.; Bradshaw, K.;
Duncan, W. M.; Frensley, W. R.; Lee, J. W.;
Shih, H. D. (1986-01-01). "Spatial
quantization in GaAs–AlGaAs multiple
quantum dots" . Journal of Vacuum
Science & Technology B: Microelectronics
Processing and Phenomena. 4 (1): 358–
360. Bibcode:1986JVSTB...4..358R .
doi:10.1116/1.583331 . ISSN 0734-211X .
94. Kolobkova, E. V.; Nikonorov, N. V.; Aseev,
V. A. (2012). "Optical Technologies Silver
Nanoclusters Influence on Formation of
Quantum Dots in Fluorine Phosphate
Glasses" . Scientific and Technical Journal
of Information Technologies, Mechanics
and Optics. 5 (12).
95. Екимов АИ; Онущенко АА (1981).
"Квантовый размерный эффект в
трехмерных микрокристаллах
полупроводников" (PDF). Письма в
ЖЭТФ. 34: 363–366.
96. Ekimov AI, Onushchenko AA (1982).
"Quantum size effect in the optical-spectra
of semiconductor micro-crystals". Soviet
Physics Semiconductors-USSR. 16 (7):
775–778.
97. Ekimov AI, Efros AL, Onushchenko AA
(1985). "Quantum size effect in
semiconductor microcrystals". Solid State
Communications. 56 (11): 921–924.
Bibcode:1985SSCom..56..921E .
doi:10.1016/S0038-1098(85)80025-9 .
98. "Nanotechnology Timeline" . National
Nanotechnology Initiative.
99. Rossetti, R.; Nakahara, S.; Brus, L. E.
(1983-07-15). "Quantum size effects in the
redox potentials, resonance Raman spectra,
and electronic spectra of CdS crystallites in
aqueous solution" . The Journal of
Chemical Physics. 79 (2): 1086–1088.
Bibcode:1983JChPh..79.1086R .
doi:10.1063/1.445834 . ISSN 0021-9606 .
100. Brus, L. E. (1984-05-01). "Electron–
electron and electron‐hole interactions in
small semiconductor crystallites: The size
dependence of the lowest excited electronic
state" . The Journal of Chemical Physics.
80 (9): 4403–4409.
Bibcode:1984JChPh..80.4403B .
doi:10.1063/1.447218 . ISSN 0021-9606 .
101. Norris, D.J. (1995). "Measurement and
Assignment of the Size-Dependent Optical
Spectrum in Cadmium Selenide (CdSe)
Quantum Dots, PhD thesis, MIT".
hdl:1721.1/11129 .
102. Delerue, C. & Lannoo, M. (2004).
Nanostructures: Theory and Modelling.
Springer. p. 47. ISBN 3-540-20694-9.

External links
Quantum Dots: Technical Status and
Market Prospects
Quantum dots that produce white light
could be the light bulb's successor
Single quantum dots optical properties
Quantum dot on arxiv.org
Quantum Dots Research and Technical
Data
Simulation and interactive visualization
of Quantum Dots wave function

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Quantum_dot&oldid=825764071"

Last edited 10 hours ago by Jyrkorp…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

S-ar putea să vă placă și