ArticlePDF Available

Homogenous, far-reaching tuning and highly emissive QD-silica core-shell nanocomposite synthesized via a delay photoactive procedure; Their applications in two-photon imaging of human mesenchymal stem cells

Authors:

Abstract and Figures

In this article, we present the exploration of a facile synthetic tactic incorporating delay-photo-oxidation to recover the loss in emission frequently encountered after encapsulating quantum dots (QDs) inside a silica shell. This facile synthesis procedure reproducibly increases emissive intensity of QDs (core)-SiO(2) (shell) (60 nm) nanomaterials by >5 fold (QY from 3% to >15%). The resulting QDs (core)-SiO(2) proved to be a single quantum dot in single SiO(2), homogeneous and highly monodispered; their emissions have been successfully fine-tuned from visible to the near infrared region. We then demonstrate their power in biological imaging by labeling human mesenchymal stem cells under two-photon confocal microscopy. The results of low cytotoxicity, efficient labeling, and specific location nearby the nucleus characters of these nanoparticles should spark an intensive relevant research within a living system.
Content may be subject to copyright.
Homogenous, far-reaching tuning and highly emissive QD–silica core–shell
nanocomposite synthesized via a delay photoactive procedure; their
applications in two-photon imaging of human mesenchymal stem cells
Chih-Wei Lai,
a
Yu-Hsiu Wang,
a
Yu-Chun Chen,
a
Cheng-Chih Hsieh,
a
Borade Prajakta Uttam,
b
Jong-Kai Hsiao,*
bc
Cheng-Chih Hsu
a
and Pi-Tai Chou*
a
Received 2nd April 2009, Accepted 28th August 2009
First published as an Advance Article on the web 28th September 2009
DOI: 10.1039/b906575j
In this article, we present the exploration of a facile synthetic tactic incorporating delay-photo-
oxidation to recover the loss in emission frequently encountered after encapsulating quantum
dots (QDs) inside a silica shell. This facile synthesis procedure reproducibly increases emissive
intensity of QDs (core)–SiO
2
(shell) (60 nm) nanomaterials by >5 fold (QY from 3% to >15%). The
resulting QDs (core)–SiO
2
proved to be a single quantum dot in single SiO
2
, homogeneous and highly
monodispered; their emissions have been successfully fine-tuned from visible to the near infrared
region. We then demonstrate their power in biological imaging by labeling human mesenchymal stem
cells under two-photon confocal microscopy. The results of low cytotoxicity, efficient labeling, and
specific location nearby the nucleus characters of these nanoparticles should spark an intensive relevant
research within a living system.
Introduction
Quantum dots (QDs), synonym of semiconducting nano-
particles, possess a number of advantageous optical properties
such as high extinction absorption coefficient, size-tunable
wavelength, sharp emission bandwidth, and superior photo-
stability, etc.
1,2
Structurally, composition of QDs has evolved
from basic homo-QDs to versatile hetero-QDs with core–shell
configuration. The addition of inorganic shell growth on the
central core expands the possible interplay of energy levels in
QDs.
3
Depending on the energetic relationship between the core
and shell, hetero-QDs frequently encountered can be classified as
type-I and type-II.
4
In type-I QDs, the conduction band of the
shell is higher in energy than that of the core, while the relative
energy level is opposite in the valence band. In contrast, the
energy levels of both valence and conduction bands in type-II
QDs’ cores are either lower or higher than that of the shells. Such
a configuration permits spatially separated interband transitions.
Having a direct band-gap transition, type-I QDs, usually have
high quantum yield (QY) in visible domain, while type II QDs
can be largely tuned to bear emissions in far visible to near-
infrared (NIR) region by utilizing the indirect interband transi-
tion character. Regardless of the difference in their emissive
mechanisms, however, both types of QDs have demonstrated
great potential in bio-imaging.
5
Unless the excitons are trapped
in the surface defect sites, in theory, QDs should exhibit intense
emission due to the lack of high-frequency vibrational quench-
ing. This is particularly important regarding the imaging appli-
cation toward NIR since in this region classical organic dyes
greatly suffer from high frequency vibrational (overtone) deac-
tivation; a universal quenching mechanism for excited molecules
dubbed as the energy gap law.
6
Despite their superiority, a major obstacle for QDs in bio-
imaging is known to be their cytotoxicity. Among numerous
attempts to circumvent this disadvantage, QDs coated with silica
shell, QDs–SiO
2
, have proven to possess less cytotoxicity and
versatile functionalities towards practical applications, allowing
biological molecules to form covalent bonding with the silica
material surface.
7,8
However, QDs–SiO
2
core–shell materials
prepared via this prevailing microemulsion method suffer
considerable loss in emission QY.
9
Recently, several synthetic
approaches have been proposed to achieve high QY QDs–SiO
2
,
including layer-by-layer assembly and outer inorganic shell
thickening, etc.
10,11
One convenient approach should be credited
to photo-oxidation of QDs, which involves light exposure to
QDs after surface modification (prior to silica shell coating).
Although this protocol seems to alleviate the QY deprivation, the
performance of the resulting QDs remains sceptical due to the
lack of bio-compatible evidence, inferior homogeneity, and poor
dispersibility.
12
In an aim to improve the QDs–SiO
2
quality, using cysteine
(–Cy) as a capping ligand in phase transfer, we then proposed an
alternative, facile protocol, in which the as prepared QDs-Cy
surface was subject to a photo-oxidation process only when QDs-
Cy had been located inside the SiO
2
shell. As a result, highly
emissive QDs-Cy–SiO
2
for both type I and type II QDs were
prepared, and the emission wavelength could be fine-tuned from
the visible to NIR region of 1000 nm. It is noteworthy that to
a
Department of Chemistry, National Taiwan University, Taipei 106,
Taiwan. E-mail: chop@ntu.edu.tw
b
Department of Medical Imaging, National Taiwan University Hospital
and College of Medicine, Taipei, Taiwan. E-mail: jongkai@gmail.com
c
Institute of Biomedical Engineering, National Taiwan University, Taipei,
106, Taiwan
† Electronic supplementary information (ESI) available: Experimental
section, including further experimental detail procedure, TEM picture,
emission spectra, instrument information, EDX and MTT data. See
DOI: 10.1039/b906575j/
8314 | J. Mater. Chem., 2009, 19, 8314–8319 This journal is ªThe Royal Society of Chemistry 2009
PAPER www.rsc.org/materials | Journal of Materials Chemistry
our knowledge, QDs–SiO
2
with NIR emission has not yet been
reported.
13
The resulting QDs-Cy–SiO
2
are homogeneous with
great dispersibility. We then investigated their two photon
absorption (TPA) effect as well as in-vitro bio-viability. Using
stem cells as a cellular model, their capability for two-photon cell
imaging is successfully demonstrated.
Experimental
Materials
Tri-n-octylphosphine oxide (TOPO, 99%, Aldrich), tri-n-butyl-
phosphine (TBP, technical grade 98%, SHOWA), di-n-octyl-
amine (DOA, 98%, ACROS), hexadecylamine (HDA, 90%,
TCI), CdO (99.99%, Strem), selenium (Se) powder (99.5%,
200 mesh, Aldrich), CdCl
2
(99.99%, Aldrich), tellurium (Te)
powder (99.8%, 200 mesh, Aldrich), and zinc stearate
(RiedeldeHa
en), hexanol (98%, ACROS), Triton X-100
(ACROS), tetraethyl orthosilicate (98%, ACROS), and ammo-
nium hydroxide (28–30 wt%, Fluka) were used without further
purification.
Phase transfer
According to our previous synthetic protocol, CdSe/ZnS and
CdTe/CdSe/ZnS were prepared for further surface modification.
4
(see ESI†). The water-soluble CdSe/ZnS and CdTe/CdSe/ZnS
QDs were made using a stepwise procedure reported by
Chen et al. and Mattoussi et al.
1d,14
Briefly, TBP/TOPO-capped
CdSe-ZnS core–shell particles were prepared from the growth
and annealing of CdO. 20 mg of the resulting TBP/TOPO-cap-
ped CdSe-ZnS and 80 mg of cysteine were then placed in
a reaction vessel containing 15 mL of methanol, the pH of which
was adjusted to >10 with tetramethylammonium hydroxide
pentahydrate to aid the dissolution process. Under dark condi-
tions and regular airflow, the mixture was heated under reflux at
65 C overnight. After cooling to room temperature, the cysteine-
capped CdSe/ZnS (QDs-Cy) nanocrystals were then precipitated
with ether. For further purification, methanol was added to
dissolve the precipitate, followed by the addition of ether to
re-precipitate the nanocrystals. The vacuum dried QDs-Cy were
diluted in 10 ml DI water.
Synthesis of QDs-Cy–SiO
2
QDs-Cy–SiO
2
with different CdSe/ZnS core–shell sized nano-
particles were prepared from reverse micelles.
15
The as-prepared
QDs-Cy nanoparticles (see above) were used as seeds for growth
of the SiO
2
shell. Briefly, TEOS (100 mL), and various amounts of
QDs-Cy (3 5.5 10
5
M) within water (520 mL) were added to
a solution containing cyclohexane (7.4 mL), hexanol (1.8 mL),
and Triton X-100 (1.8 mL), which the QDs concentration was
measured with an UV-Vis absorption instrument.
16
In the
microemulsion system, a mixture of silica precursor, TEOS, and
the CdSe/ZnS core–shell nanoparticles of capped water-soluble
ligand was stirred over 6 h. NH
4
OH was then added in the dark.
After 24 h polymerization process, CdSe/ZnS-Cy were coated
with silica, extracted by centrifugation, and re-precipitated twice
with ethanol. The QDs-Cy–SiO
2
was then suspended in deion-
ized water. Under oxygen gas, the reaction solution was
illuminated by a 366 nm UV lamp (0.72 mW/cm) light source for
several hours (1–6 h).
Cytotoxicity examination
Human mesenchymal stem cells (hMSC) were cultured in
Dulbecco’s modified Eagle’s medium (DMEM) (Cellgro, Hern-
don, VA, USA), supplemented with 10% heat-inactivated fetal
bovine serum (FBS), penicillin (50 U/ml), and streptomycin
(0.05 mg/ml). All cultures were kept in moist atmosphere of 5%
CO
2
, 95% air at 37 C. Cells were passaged through trypsiniza-
tion, and nucleated cells were centrifuged at 100 g for harvesting.
Phosphate buffer saline (PBS) used for washing and rinsing were
composed of 137 mM NaCl, 2.68 mM KCl, 10 mM Na
2
HPO
4
,
1.76 mM KH
2
PO
4
, and tuned to pH 7.4. For cytotoxicity eval-
uation, cells were seeded in a 24 well plate at 5 10
3
cells/well
density in 500 mL culturing medium 24 h prior to the feeding of
the particles. Including the control, six different dosages of
particles were given to each well and the incubation concentra-
tion was confirmed with inductively coupled plasma-mass spec-
trometry (ICP-MS): 0, 0.19, 0.37, 0.75, 1.49, and 2.98 mM Cd.
After 24 h incubation, each well was washed with phosphate
buffer saline (PBS) twice, and replenished with 500 mL culturing
medium with 10% of MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-
diphenyltetrazolium bromide, Sigma) agent. After 1 h incubation
and medium removal, the newly formed purple MTT-formazan
was dissolved in 200 mL dimethyl sulfoxide (DMSO, Sigma) and
the absorbance was measured at 570 nm by a multi-plate spec-
trophotometer.
In vitro Imaging
hMSCs were seeded in a 6 well plate at 5 10
3
cell/well density in
2 mL of culturing medium 24 h prior to particle feeding. After
24 h incubation with 2.98 mM of QD-Cy–SiO
2
, cells were washed
three times with PBS and then fixed in a 3.7% paraformaldehyde
solution in PBS at room temperature for 5 min. The cells were
then washed twice with PBS and incubated with 0.5% Triton
X-100 (Sigma-Aldrich) plus 1% bovine serum albumin (BSA;
Sigma-Aldrich) in PBS for 5 min. Then, fixed cells were incu-
bated with 5 mg/mL 40,6-diamidino-2-phenylindole (DAPI,
molecular probe) in PBS for nucleus staining for 10 min at room
temperature. Afterwards, 7.5 mL methanolic stock solution of
rhodamine phalloidin (Invitrogen) was diluted into 1 mL PBS
and used for actin staining for 20 to 30 min. The sample slide was
examined by a TCS SP2 confocal spectral microscope and MP
two-photon imaging system with a X40 water immersion objec-
tive, and using 800 nm two-photon diode laser and 543 nm
He–Ne laser as the excitation source. Since DAPI and CdSe/Zn-
Cy–SiO
2
with emission wavelengths of 530 nm could both be
excited through an 800 nm two-photon laser as DAPI features
a broad emission spectrum, an appropriate dye separation
procedure was performed.
Results and discussion
Freshly prepared type-I CdSe/ZnS-Cy and type-II CdTe/CdSe/
ZnS-Cy nanocrystals, made by phase-transfer procedures with
a cysteine (Cy) water-soluble ligand, commonly suffer from
rather weak emission intensity (QY #3%). To circumvent this, it
This journal is ªThe Royal Society of Chemistry 2009 J. Mater. Chem., 2009, 19, 8314–8319 | 8315
has been known that photo-oxidation of QDs to reduce the
surface defects plays a key role in enhancing the luminescence of
QDs, particularly in the aqueous solution.
12
Although this
workup procedure is rather succinct, the emphasis of this tactic
lies in the selection of the occasion. For example, applying photo-
oxidation at this stage could successfully regain the QY;
however, the effort would be devastated as significant quenching
re-occurred during silica encapsulation under a microemulsion
system. Therefore, in our viewpoint, a more appropriate
proposal to initiate photo induced oxidation reaction would be
after silane polymerization, referred here as delay photo-oxida-
tion. The advantage mainly lies in the fact that the cysteine
ligands, having been locked inside SiO
2
, remain unperturbed
during photo-oxidation. We then expect the as prepared
QDs-Cy–SiO
2
to be homogeneous for QDs (inside SiO
2
), well
dispersive and highly emissive.
As a result, the delay photo-oxidation strategy did effectively
preserve the emission intensity for both type I and type II
QDs-Cy–SiO
2
, and the degree of emission preservation depended
on the time of illumination. As shown in Fig. 1A, the immediate
photoluminescence spectra of freshly prepared silica (60 nm)
coated CdSe/ZnS (4.9 nm/0.7 nm) core–shell nanoparticles under
oxygen atmosphere, water as the solvent, and increasing period
of UV-366 nm (0.72 mW/cm) exposure demonstrates positive
enhancement in QY recoveries at 610 nm. The optimum expo-
sure period appeared to be 6 h as 5-fold recovery (QY 15%)
was achieved in comparison to the control. In addition to
emission gain, a blue shift of about 3 nm/h was also observed
within the photo-oxidation period of 6 h and then remained
unchanged after prolonged illumination. The shifting in emission
wavelength was the consequence of size reduction, as the rough
surface of the nanoparticles was etched after photo-oxidation.
12
To verify the origin of the emission as well as to demonstrate the
wavelength tunability of our synthetic approach, photo-
luminescence (PL) spectra of various sizes of silica coated QDs
(type I CdSe/ZnS-Cy, type II CdTe/CdSe/ZnS-Cy) subject to
photo-oxidation were measured. From the PL spectra shown in
Fig. 1B, several important remarks can be pointed out: first of all,
in response to the increase in size of QDs (core, vide infra), the
band-edge emission peaks systematically shift to the lower band
gap energy. As shown in Fig. 1B, emissions from different sizes
(measured by transmission electron microscopy (TEM), vide
infra) of CdSe/ZnS-Cy–SiO
2
at 530, 545, 580, 610 nm and
different sizes of CdTe/CdSe/ZnS-Cy–SiO
2
at 780 nm and
1027 nm demonstrate that the synthetic protocol employed is
capable of generating QDs with emissions ranging from visible to
the NIR region. This size dependent PL emission for both
CdSe/ZnS-Cy–SiO
2
and CdTe/CdSe/ZnS-Cy–SiO
2
, together
with similar emission spectral features of the bare QDs-Cy
(without SiO
2
, not shown here), warrants the origin of the
emission from the band gap. Note that each peak exists as a clean
single shape indicating that CdSe/ZnS-Cy–SiO
2
and CdTe/CdSe/
ZnS-Cy–SiO
2
nanocrystals maintained their unique optical
properties and dispersed homogeneously. More importantly, the
quantum yields of all QDs–SiO
2
prepared via delay photo-
oxidation are higher than 15% versus those of <0.05 prepared by
either non-photo-oxidation or photo-oxidation on non-SiO
2
coated QDs-Cy. Finally, this protocol is facile and reproducible,
producing monodispersity and high emission intensity of single
QDs at the center of smooth silica nanoparticles.
In addition to the exploitation of the cysteine ligand, 3-mer-
captopropionic acid (3-MPA) was also selected as a candidate for
QD surface modification to study the influence of the surface
ligand and the effectiveness of delay photo-oxidation. As both
cysteine and 3-MPA contain reducing functional groups, they
are expected to enhance the QY of the conjugated QDs.
17
Nonetheless, under identical treatment of 6 h illumination in an
oxygen atmosphere, the QDs-3-MPA–SiO
2
nanomaterials did
not receive comparable emission recovery as obtained by
QDs-Cy–SiO
2
(Fig. S1, ESI†). The difference in QY enhance-
ment may be rationalized by the stronger reducing power of the
amine functionality compared to that of the carboxylic group.
This finding reinforces the importance of amine functional
groups on the surface, and is in agreement with pervious reports,
stating that the amine functional ligand modification on e.g. ZnO
nanoparticles and alkyl-amine modification on Zn
x
Cd
1x
Se alloy
nanoparticles augments emission intensity much more efficiently
compared to 3-MPA or thiolglycolic acid species. It is also noted
that SiO
2
–QDs modified with 3-amino- propyltrimethoxysilane
carry a stronger emission intensity than uncoated SiO
2
–QDs.
9a,18,19
In brief, amino functional ligands seem to serve as an indis-
pensable capping agent in respect to delay photo-oxidation
enhancing emission.
As for the morphology, the TEM images of cysteine ligand
capped QDs, including CdSe/ZnS and CdTe/CdSe/ZnS nano-
crystals are shown in Fig. S2. (ESI†). Compared to the TEM
pictures of CdSe/ZnS with TOPO and HDA, QDs-Cy was less
clear in imaging because of the water-soluble capping ligand on
the QDs surface. The use of water-soluble QDs-Cy assured that
the nanoparticles were located in the water pool before
polymerization, which is an important criterion for excellent
silica encapsulation. As one can perceive in Fig. 2A, the size of
the core (3.5 nm) in the silica shell provides positive identification
of QDs-Cy. Insert of Fig. 2 clearly indicates that CdSe/ZnS
(3.5–5.8 nm) and CdTe/CdSe/ZnS (7.1 nm) nanocrystals
(Fig. 2A–2D) are successfully encapsulated within the silica
shells. Although QDs embedded in silica appeared as an amor-
phous structure, a more detailed examination by high resolution
TEM (HRTEM) revealed the crystal lattice of QDs nanocrystals
Fig. 1 (A) Photoluminescence spectra of freshly prepared CdSe/ZnS-
Cy–SiO
2
(4.9 nm/0.7 nm) core–shell nanoparticles with 610 nm emission
after illumination. Increases in luminescence were obtained with
increasing period of UV-366 nm (0.72 mW/cm) exposure. (B) The QDs-
Cy–SiO
2
PL peaks resulted from 530, 545, 580, 610, 780, (3.5–5.8 nm in
diameter) and 1027 nm (7.1 nm in diameter) correlating to the CdSe/ZnS,
CdTe/CdSe/ZnS cores.
8316 | J. Mater. Chem., 2009, 19, 8314–8319 This journal is ªThe Royal Society of Chemistry 2009
(see insert of Fig. 2C). The monodisperisty and the overall size of
the nanomaterials of 60 nm measured by TEM show that the
shape of nanocrystals were retained after the polymerization and
photo-oxidation process. The advanced verification of the
composition of QDs-Cy–SiO
2
nanoparticles was then analyzed by
energy dispersive X-ray spectroscopy (EDX). The results shown
in Fig. S3A† (CdSe/ZnS-Cy) and S3B† (CdTe/CdSe/ZnS-Cy)
unveil the elemental characteristic peaks and the corresponding
atomic percentage ratios, which render sufficient evidence for
successful encapsulation of the CdSe/ZnS-Cy and CdTe/CdSe/
ZnS-Cy in the SiO
2
shell.
Previously, our group has demonstrated that QDs such as
CdSe, and CdTe possessed outstanding two-photon absorption
cross-section values (>10 000 GM), and it would be eligible to
assume that QDs-Cy–SiO
2
is applicable in two-photon imaging
as well.
20
As for the naked eye view, the QDs-Cy–SiO
2
two-
photon visible imaging pictures (e.g. 530 nm) excited by NIR
laser (800 nm) in aqueous solution were taken and depicted in the
insert (lower right corner) of Fig. 2A. Also, the insert (lower right
corner) of Fig. 2D shows the two-photon emission of
CdTe/CdSe/ZnS-Cy–SiO
2
(l
ex
1200 nm) captured by
a conventional hand-held infrared viewer coupled with a CCD
detector. The results affirm the potential of QDs-Cy–SiO
2
in two-
photon bio-imaging.
We then applied the above silica nanoparticle embedded QDs
with in vitro optical imaging; particular focus is on two-photon
imaging. Prior to the experiment, a cytotoxicity test of the QDs-Cy
SiO
2
nanocrystals is necessary, in which human mesenchymal stem
cells (hMSCs) were adopted as a target. Stem cells are self-renewabl e
and capable of differentiating into different kinds of cells that are
essential for tissue transplantation. Such unidirectional
differentiation forbids the reversible process once the cells are in the
designated locations. Consequently, labeling of stem cells with
QDs-Cy–SiO
2
is biomedically important and will facilitate the
understanding of stem cell biology. The hMSCs were immortalized
by the gene transfer of a combination of human telomerase reverse
transcriptase (hTERT) with human papillomavirus. The differen-
tiation ability of these hMSCs has been previously demonstrated.
21
To examine the cytotoxicity of the as prepared particles, cells
were incubated with six different dosages of particles for 24 h, the
concentration of which was confirmed with inductively coupled
plasma-mass spectrometry (ICP-MS): 0, 0.19, 0.37, 0.75, 1.49,
and 2.98 mM Cd. The feedback suggested that bare CdSe/ZnS-Cy
are detrimental to cells as the cellular viability dropped down to
70% at higher incubation concentrations (2.98 mM, see ESI†
for QD/cell calculation). In contrast, both CdSe/ZnS-Cy–SiO
2
and CdTe/CdSe/ZnS-Cy-SiO
2
were harmless toward cells as the
viabilities of the cells within the prescribed dosages remained well
above 90% even after 24 h incubation (Fig. S4, ESI†).The uptake
efficiency of the nanoparticles, using CdSe/ZnS-Cy–SiO
2
as the
representing QDs, was found to be 13%, in equivalence of
1.40 10
7
QDs-Cy–SiO
2
per cell. Note that a harsh condition of
1:1 HNO
3
/H
2
O acid digestion was performed on a hot plate for
6 h to digest the silica protected core before ICP-MS measurement.
To ensure the event of internalization and the location of the
particles inside the cell, confocal microscopy with two-photon
laser and fluorescence staining has been performed. In this
experiment, hMSCs incubated with 2.98 mM of QD-Cy–SiO
2
for
24 h were fixed with paraformaldehyde solution and stained with
molecular dyes DAPI and rhodamine phalloidin for nucleus and
cytoskeleton labeling, respectively. The confocal image shown in
Fig. 3 clearly indicates that QDs-Cy–SiO
2
of green fluorescence
was internalized into the cells and mainly resided in the
Fig. 2 TEM pictures of core–shell–shell nanomaterials (CdSe/ZnS-Cy–
SiO
2
) (2A–C) and core–shell–shell–shell (CdTe/CdSe/ZnS-Cy–SiO
2
) (2D)
nanoparticles after illumination. Upper right insets: emission of the
corresponding nanoparticles under UV lamp. Lower right insets in (A)
and (D): emission of the corresponding nanoparticles with two-photon
laser excitation. Lower right inset in (C): HRTEM picture of CdSe/ZnS-
Cy–SiO
2
.
Fig. 3 Microscopic observation of the QDs-Cy–SiO
2
internalization.
hMSCs were treated with silica coated QDs for 24 h and then processed
for two photon and confocal microscopic examination. Cell nuclei were
stained with 40,6-diamidino-2-phenylindole (blue color). Actin fibers were
stained with rhodamine phalloidin to confirm the cell boundary
(red color). QDs-Cy–SiO
2
exhibit a two photon emission (green fluo-
rescence, 530 nm) for in vitro bio-applications. Fluorescence overlay
image demonstrates the internalization of QDs-Cy–SiO
2
which resided
near the nucleus.
This journal is ªThe Royal Society of Chemistry 2009 J. Mater. Chem., 2009, 19, 8314–8319 | 8317
cytoplasm nearby the nucleus. Although the exact location of the
particle confinement was not fully determined, collective images
from Z-stack scanning (see Fig. 4) advised that no particles were
located on the cell membrane or inside the nucleus. The confocal
imaging illustrated efficient labeling and specific location char-
acters of the CdSe/ZnS-Cy–SiO
2
. The biocompatibility of
CdSe/ZnS-Cy–SiO
2
with 530 nm was ensured in two ways:
(1) negligible cytotoxicity from cell proliferation test; and
(2) total preservation of two-photon signal after internalization
into the cell as observed under the confocal microscopy. It should
be noted here that although Prasad and co-workers have
reported that QD nanocomposites after silica shell modification
could be uptaken by cancer cells for two-photon imaging, the
structural robustness of the core–shell to avoid heavy element
release remains sceptical.
22
Our findings provide solid and strong
evidence that QD nanocomposites after silica shell modification
are very practical and versatile in biological two-photon imaging.
Conclusions
By using a bottom-up synthesis approach, this article illustrates
a synthetic strategy to prepare QDs–SiO
2
nanocomposites and
their potential applications for bio-imaging. This facile and
simple method incorporating a delay photoactive procedure not
only embeds core–shell nanoparticles, such as CdSe/ZnS and
CdTe/CdSe/ZnS, into SiO
2
for reducing its cytotoxicity, but also
makes semiconductor nanocrystals retaining their unique and
excellent optical properties. We then successfully demonstrate
their power in two-photon imaging of human mesenchymal stem
cells. For future work, QDs-Cy–SiO
2
(shell) featuring both NIR
emission and two photon absorption effects simultaneously will
be explored. We also believe that this facile synthetic protocol
can be exploited in other QD species with unique optical
properties to preserve their own emission wavelength and
intensity after being embedded in silica nanoparticles. This
protocol should be vital for the synthesized nanomaterials to be
non-toxic with minimal dosage required to generate effective
fluorescent signal within a living system.
Acknowledgements
This work was funded by the National Science Council of
Taiwan, R.O.C. We thank the excellent technical assistance of
Technology Commons, College of Life Science, NTU (Taiwan)
with the confocal laser scanning microscopy (CLSM).
Notes and references
1(a) A. P. Alivisatos, Science, 1996, 271, 933; (b) M. Danek,
K. F. Jensen, C. B. Murray and M. G. Bawendi, Chem. Mater.,
1996, 8, 173; (c) M. Nirmal and L. Brus, Acc. Chem. Res., 1999, 32,
407; (d) H. Mattoussi, J. M. Mauro, E. R. Goodman,
G. P. Anderson, V. C. Sundar, F. V. Mikulec and M. G. Bawendi,
J. Am. Chem. Soc., 2000, 122, 12142.
2(a) B. Murray, D. J. Norris and M. G. Bawendi, J. Am. Chem. Soc.,
1993, 115, 8706; (b) J. K. Jaiswal, H. Mattoussi, J. M. Mauro and
S. M. Simon, Nat. Biotechnol., 2003, 21, 47; (c) A. P. Alivisatos,
J. Phys. Chem., 1996, 100, 13226; (d) C. B. Murray, C. R. Kagan
and M. G. Bawendi, Annu. Rev. Mater. Sci., 2000, 30, 545;
(e) H. Weller, Angew. Chem., Int. Ed. Engl., 1993, 32, 41;
(f) J. M. Klostranec and W. C. W. Chan, Adv. Mater., 2006, 18, 1953.
3(a) S. Kim, B. Fisher, H. J. Eisler and M. G. Bawendi, J. Am. Chem.
Soc., 2003, 125, 11466; (b) C.-Y. Chen, C.-T. Cheng, J.-K. Yu,
S.-C. Pu, Y.-M. Cheng, P.-T. Chou, Y.-H. Chou and H.-T. Chiu,
J. Phys. Chem. B, 2004, 108, 10687; (c) P.-T. Chou, C.-Y. Chen,
C. T. Cheng, S.-C. Pu, K.-C. Wu, Y.-M. Cheng, C.-W. Lai,
Y.-H. Chou and H.-T. Chiu, ChemPhysChem, 2006, 7, 222;
(d) B. Blackman, D. M. Battaglia, T. D. Mishima, M. B. Johnson
and X. G. Peng, Chem. Mater., 2007, 19, 3815; (e) C.-Y. Chen,
C.-T. Cheng, C.-W. Lai, Y.-H. Hu, P.-T. Chou, Y.-H. Chou and
H.-T. Chiu, Small, 2005, 1, 1215; (f) C.-T. Cheng, C.-Y. Chen,
C.-W. Lai, W.-H. Liu, S.-C. Pu, P.-T. Chou, Y.-H. Chou and
H.-T. Chiu, J. Mater. Chem., 2005, 15, 3409.
4(a) B. O. Dabbousi, J. Rodriguez-Viejo, F. V. Mikulec, J. R. Heine,
H. Mattoussi, R. Ober, K. F. Jensen and M. G. Bawendi, J. Phys.
Chem. B, 1997, 101, 9463; (b) D. V. Talapin, A. L. Rogach,
A. Kornowski, M. Haase and H. Weller, Nano Lett., 2001, 1, 207;
(c) P. Reiss, J. Bleuse and A. Pron, Nano Lett., 2002, 2, 781.
5(a) S. Kim, Y. T. Lim, E. G. Soltesz, A. M. De Grand, J. Lee,
A. Nakayama, J. A. Parker, T. Mihalijevic, R. G. Laurence,
D. M. Dor, L. H. Cohn, M. G. Bawendi and J. V. Frangioni, Nat.
Biotechnol., 2004, 22, 93; (b) L. Bakueva, I. Gorelikov, S. Musikhin,
X. S. Zhao, E. H. Sargent and E. Kumacheva, Adv. Mater., 2004, 16,
926; (c) W. Jiang, S. Mardyani, H. Fischer and W. C. W. Chan,
Chem. Mater., 2006, 18, 872; (d) Y. Lei, H. Y. Tang, L. D. Yao,
R. C. Yu, M. F. Feng and B. S. Zou, Bioconjugate Chem., 2008, 19, 421.
6(a) W. Siebrand, J. Chem. Phys., 1967, 47, 2411; (b) K.-Y. Chen,
C.-C. Hsieh, Y.-M. Cheng, C.-H. Lai and P.-T. Chou, Chem.
Commun., 2006, 4395.
7(a) Jr. M. Bruchez, M. Moronne, P. Gin, S. Weiss and
A. P. Alivisatos, Science, 1998, 281, 2013; (b) N. Gaponik,
D. V. Talapin, A. L. Rogach, K. Hoppe, E. V. Shevchenko,
A. Kornowski, A. Eychmu1ller and H. Weller, J. Phys. Chem. B,
2002, 106, 7177; (c) T. Zhang, J. L. Stilwell, D. Gerion, L. Ding,
O. Elboudwarej, P. A. Cooke, J. W. Gray, A. P. Alivisatos and
F. F. Chen, Nano Lett., 2006, 6, 800; (d) A. Derfus, W. Chan and
S. Bhatia, Nano Lett., 2004, 4, 11; (e) C. Kirchner, T. Liedl,
S. Kudera, T. Pellegrino, A. Munoz Javier, H. E. Gaub, S. Stolzle,
N. Fertig and W. J. Parak, Nano Lett., 2005, 5, 331.
8 S. Santra, P. Zhang, K. Wang, R. Tapec and W. H. Tan, Anal. Chem.,
2001, 73, 4988.
9(a) S. Selvan and T. Tan, Adv. Mater., 2005, 17, 1620; (b) M. Darbandi,
R. Thomann and T. Nann, Chem. Mater., 2005, 17, 5720.
10 Z. Zhelev, H. Ohba and R. Bakalova, J. Am. Chem. Soc., 2006, 128,
6324.
Fig. 4 Confocal microscopy examination of QDs-Cy–SiO
2
labeled
hMSCs. For confirmation of the exact intracellular location of QDs-Cy–
SiO
2
, hMSCs treated with QDs-Cy–SiO
2
for 48 h were stained with
40,6-diamidino-2-phenylindole (DAPI) for visualization of the cell
nucleus and the actin fibers were stained with rhodamine phalloidin. The
QDs-Cy–SiO
2
was located at the cytoplasm near the nucleus. No
QDs-Cy/SiO
2
was detected at the cell membrane or nucleus. The finding
indicates efficient labeling and specific location characters of the
QDs-Cy/SiO
2
.
8318 | J. Mater. Chem., 2009, 19, 8314–8319 This journal is ªThe Royal Society of Chemistry 2009
11 R. Koole, M. M. van Schooneveld, J. Hilhorst, C. D. Doneg
a, C. D. ’t
Hart, A. van Blaaderen and D. Vanmaekelbergh, Chem. Mater., 2008,
20, 2503.
12 (a) A. A. Bol and A. Meijerink, J. Phys. Chem. B, 2001, 105, 10203;
(b) L. Manna, E. C. Scher, L.-S. Li and A. P. Alivisatos, J. Am.
Chem. Soc., 2002, 124, 7136; (c) Y. Wang, Z. Tang, M. A. Correa-
Duarte, I. Pastoriza-Santos, M. Giersig, N. A. Kotov and
L. M. Liz-Marz
an, J. Phys. Chem. B, 2004, 108, 15461–15469.
13 (a) M. Darbandi, W. Lu, J. Fang and T. Nann, Langmuir, 2006, 22,
4371; (b) T. T. Tan, S. T. Selvan, L. Zhao, S. Gao and J. Y. Ying,
Chem. Mater., 2007, 19, 3112.
14 C.-Y. Chen, C.-T. Cheng, C.-W. Lai, P.-W. Wu, K.-C. Wu, P.-
T. Chou, Y.-H. Chou and H.-T. Chiu, Chem. Commun., 2006, 263.
15 (a) S. Santra, R. P. Bagwe, D. Dutta, J. T. Stanley, G. A. Walter,
W. Tan, B. M. Moudgil and R. A. Mericle, Adv. Mater., 2005, 17,
2165; (b) C.-W. Lai, Y.-H. Wang, C.-H. Lai, M.-J. Yang,
C.-Y. Chen, P.-T. Chou, C.-S. Chan, Y. Chi, Y.-C. Chen and
J.-K. Hsiao, Small, 2008, 4, 218.
16 (a) W. W. Yu, L. Qu, W. Guo and X. Peng, Chem. Mater., 2003, 15,
2854; (b) B. O. Dabbousi, J. Rodrigez-Viejo, F. V. Mikulec,
J. R. Heine, H. Mattoussi, R. Ober, K. J. Jensen and
M. G. Bawendi, J. Phys. Chem., 1997, 101, 9463.
17 Y.-J. Chen and X.-P. Tan, Small, 2009, 5, 2012.
18 F.-C. Liu, T.-L. Cheng, C.-C. Shen, W.-L. Tseng and M. Y. Chiang,
Langmuir, 2008, 24, 2162.
19 M. L. Kahn, T. Cardinal, B. Bousquet, M. Monge, V. Jubera and
B. Chaudret, ChemPhysChem, 2006, 7, 2392.
20 S.-C. Pu, M.-J. Yang, C.-C. Hsu, C.-W. Lai, C.-C. Hsieh, S. H. Lin,
Y.-M. Cheng and P.-T. Chou, Small, 2006, 2, 1308.
21 (a) J.-K. Hsiao, M.-F. Tai, H.-H. Chu, S.-T. Chen, H. Li, D.-M. Lai,
S.-T. Hsieh, J.-L. Wang and H.-M. Liu, Magn. Reson. Med., 2007, 58,
717; (b) T. Okamoto, T. Aoyama, T. Nakayama, T. Nakamata,
T. Hosaka, K. Nishijo, T. Nakamura, T. Kiyono and J. Toguchida,
Biochem. Biophys. Res. Commun., 2002, 295, 354.
22 W.-C. Law, K.-T. Yong, I. Roy, G. X. Xu, H. Ding, E. J. Bergey,
H. Zeng and P. N. Prasad, J. Phys. Chem. C, 2008, 112, 7972.
This journal is ªThe Royal Society of Chemistry 2009 J. Mater. Chem., 2009, 19, 8314–8319 | 8319
... QDs possess excellent optical properties, such as high extinction absorption coefficient, size-tunable wavelength, sharp emission bandwidth, and superior photostability, which make them good candidates to be used as probes (Bruchez et al. 1998, Chan and Nie 1998, Dubertret et al. 2002. Their cytotoxicity, hydrophobicity, and the property that makes them prone to aggregate were also improved significantly after being carried with the SiNPs, which is vital for them to be used in biological analysis (Lai et al. 2009 ). In summary, probes that are loaded on or encapsulated into the SiNPs possess excellent property and have attracted great attention in bioanalytical chemistry (Santra et al. 2001b, Mahtab et al. 2011. ...
... For example, the near-infrared (NIR)-emitting rare earth metals are of particular interest because biological tissues are transparent in this spectral range. Thus, more near-infrared-to-visible up-conversion SiNPs is expected to contribute in specific targeting, cell sorting, and bioimaging (Lai et al. 2009, Fedorenko et al. 2010, Gorelikov et al. 2011. ...
Article
Full-text available
Owing to the small size, high surface-to-volume ratio, and good biocompatibility, silica nanoparticles (SiNPs) have become the normally used carrier for biomolecule immobilization. Label-based detection techniques have been used extensively for a wide range of applications in biological analysis. Organic dye, enzyme, electroactive species, and quantum dots (QDs) with unique optical and electrochemical performances have been used as probes for chemical and biological analysis. The common problem is that only a limited amount of probes can be linked to a biomolecule, which limited the extensive use of these probes in trace amounts of analyte detection. New technologies can provide efficient means to carry them with SiNPs, which possess a three-dimensional network suitable for load or encapsulating a large amount of probes. As a result, composite SiNPs can improve the detection sensitivity significantly because they carry a large amount of probes, which is beneficial for trace analysis. Also, SiNPs are nontoxic, highly water soluble, suitable for many biomolecules to conjugate, and increase the chemical and physical stability of the probes, which is very important for biological analysis. The goal of this review is to describe recent progress in the methods that SiNPs are used as a carrier of dyes, QDs, and enzyme and their application in biological analysis. Among these, SiNPs as a biolabel reagent for signal amplification in immunoassay, DNA, protein, and small molecule detection are the main topic in this review. We focus on two types of methods of carrying signal probes with SiNPs. One is to modify probes on the outer surface of the SiNPs and another is to encapsulate signal probes into the SiNPs. Particular attention is paid to update reported advances since 2009.
... In particular, silica capped nanoparticles are water soluble, optically transparent and being biocompatible, silica capped QDs can be designed as probes for bioanalysis [34][35][36][37]. The usual methods for SiO 2 modified QDs (Stober method) [38][39][40] and reverse microemulsion method [35,41] involves multiple complicated steps. Recently, one-pot large scale synthesis of silica hybridized CdTe QDs was achieved by Zhou et al. [42], where both the growth of QDs and silica hybridization were performed in the same reactor but at higher temperature (100 C) and longer duration (10 h). ...
... Mesoporous silicas have been widely applied to protect Electronic supplementary material The online version of this article (doi:10.1007/s00604-015-1481-0) contains supplementary material, which is available to authorized users. nanoparticles from the external environment and thereby improve the photostability and biocompatibility of the nanoparticles [19][20][21][22]. They are also often used to incorporate dyes or photosensitizers into the silica framework due to their large surface area and being able to be chemically functionalized [23,24]. ...
Article
A new multifunctional nanocomposite is described that displays both upconversion and downconversion luminescence. It is based on upconversion nanoparticles (UCNPs) with a mesoporous coating that is boned with a terbium(III) complex. The modified mesoporous nanocomposite was characterized by transmission electron microscopy, X-ray powder diffraction, nitrogen adsorption-desorption isotherms, and photoluminescence spectra. They display dual (green and red) fluorescence under 980-nm excitation, and green fluorescence under 365-nm excitation. Both emissions can be easily detected by bare eyes. The material has low cytotoxicity and good biocompatibility as proven by the methyl thiazolyltetrazolium (MTT) assay. The nanocomposite was successfully applied to upconversion luminescence based in-vitro confocal imaging of the cytosol of murine macrophage cells (RAW264.7), and this resulted in images of excellent contrast. In addition, the particles display strongly temperature-dependent luminescence in the range from 280 K to 330 K. Graphical Abstract A new multifunctional nanocomposite was designed and prepared with dual green luminescence based on the upconversion and downconversion emission for in vitro bioimaging and temperature sensing
Article
Nanotechnology is an emerging area of research for the exploitation of various scientific advancements for controlling material configuration at a reduced dimensional scale. With the advent of nanotechnology, a new era for the development of biological products has opened. Multifunctional mesoporous silica-based particles fulfil critical requisites for the preparation and characterization of nanotechnology products. Mesoporous silica nanoparticles (MSNPs) have distinctive structural features, such as large surface areas and tunable pore sizes, with particular suitability to deliver hydrophobic cargo intracellularly and tunable surface properties. Combinations of various nanostructured materials will enable the development of a multifunctional MSNP hybrid. Mesoporous silica nanoparticles have recently attracted much attention due to their significant potential for tumor treatment and imaging. Functionalized mesoporous silica materials have recently shown rapid growth in imagistic and curative applications. Multifunctional mesoporous silica nanocomposites have a higher capacity for drug loading and chemical release compared to amorphous colloidal silica. Therefore, MSNPS is an ideal platform for constructing multipurpose tools that incorporate various functional nanostructured materials. This review summarized the most recent advances in mesoporous materials research and discussed the use of various nanocarriers conjugated with MSNPs. High potential characteristics and exciting practical applications highlight different manufacturing strategies for these hybrids (noble metals, carbohydrates, up-conversion, and magnetic nanoparticles (NPs), quantum dots (QDs), and photodynamic therapy (PDT)).
Article
Full-text available
The prime goal of this work is to synthesize free-standing polyvinyl alcohol (PVA) films doped with Ag2S, CuS, Ag2S/CuS alloy, and Ag2S/CuS nanocomposites through the sol–gel route. The dependence of Ag2S content in the PVA nanocomposite films on both the real and imaginary parts of the complex permittivity and loss tangent values was examined. An enhanced dielectric constant was achieved with minimum dielectric loss due to the insulating silica layer. By changing the Ag2S content in Ag2S/CuS PVA films, the AC conductivity is improved with pure Ag2S nanoparticles exhibiting highest values of the order of 10−6−10−9 S/cm. The Cole–Cole parameters were calculated and the semicircles observed in the plots indicate a single relaxation process. The results suggest that these composite films are potential materials for embedded capacitor applications. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016, 133, 43568.
Article
Understanding the relationships between the physicochemical properties of engineered nanomaterials and their toxicity is critical for environmental and health risk analysis. However, this task is confounded by material diversity, heterogeneity of published data and limited sampling within individual studies. Here, we present an approach for analysing and extracting pertinent knowledge from published studies focusing on the cellular toxicity of cadmium-containing semiconductor quantum dots. From 307 publications, we obtain 1,741 cell viability-related data samples, each with 24 qualitative and quantitative attributes describing the material properties and experimental conditions. Using random forest regression models to analyse the data, we show that toxicity is closely correlated with quantum dot surface properties (including shell, ligand and surface modifications), diameter, assay type and exposure time. Our approach of integrating quantitative and categorical data provides a roadmap for interrogating the wide-ranging toxicity data in the literature and suggests that meta-analysis can help develop methods for predicting the toxicity of engineered nanomaterials.
Article
A novel synthetic protocol is developed to prepare highly emissive zinc copper indium sulphide (ZCIS) core/shell nanoparticles of configuration of (ZnS) x -Cu0.1 InS1.55 /ZnS via one-pot reaction of Zn(S2 CNEt2 2, CuI, 1-dodecanthiol and In(CH3 COO)3. Although each of these reagents has been exploited for the ZCIS synthesis in previous studies, the superiority of current method lies in that the ZCIS nucleation won't proceed at temperature as high as 180°C until addition of oleylamine ligand to trigger the overall reaction. The as prepared ZCIS core is able to maintain the compositional homogeneity (emission fwhm 40%) and emission is tunable from green to red depending on elemental composition. In particular, the green one (∼530 nm), with f ∼;42% is the highest recorded to date. This surfactant induced one-pot reaction also offers scale-up advantage, with 1.6 (green), 2.27 (yellow) and 2.52 (red) grams per reaction being successfully made.
Article
CuInS2 quantum dots encapsulated in ZnS shells have emerged as less-toxic alternatives to cadmium-based nanoparticles. A well-established synthesis route for hydrophobic CIS/ZnS core/shell QDs has been developed using dodecanethiol as a ligand, solvent and source of sulfur, although the transfer of these nanoparticles into the aqueous phase still proves to be non-trivial. In this study we demonstrate that coating CIS/ZnS with an amphipol, poly(maleic anhydride-alt-1-tetradecene), 3-(dimethylamino)-1-propylamine derivative is an effective method to disperse CIS/ZnS QDs in the aqueous environment. The polymer-coated QDs display good colloidal stability over a wide pH range and low sensitivity to the presence of various metal salts, except Cu2+ for which some sensitivity is observed. Cytotoxic effects were tested in HaCat cells and compared to thioglycolic acid modified CdTe/ZnS QDs. Neither CdTe nor CIS QDs decreased viability until above 10 μg mL−1 in a WST-1 assay. By also assessing the toxicity of the PMAL-d polymer it is clear that remaining toxicity of the CIS/ZnS QDs is predominately due to the polymer coating rather than the QD core. Interestingly, MTT assays seem to artificially exaggerate cytotoxicity of the QDs, which we ascribe to the QDs interfering with the MTT or formazan crystals.
Article
Full-text available
We report the development of multi-functional uniform nanoparticles comprising a hollow manganese oxide core and a functionalized mesoporous silica shell (H–MnO@mSiO2 (Ir)@–FA, Ir: an emissive iridium complex, FA: folic acid) for use as theranostic nanocomposites that combine four-in-one functionality, i.e., dual imaging (both magnetic resonance and optical), specific targeting and therapy in a single unit. The H–MnO@mSiO2 (Ir)–FA nanoparticles (NPs) demonstrate its versatility in various applications. (1) The magnetic core provides the capability for T 1magnetic resonance imaging. (2) Phosphorescent iridium complex serves as both photosensitizer and the luminescent chromophore. (3) The nano-channels in the mesoporous silica shell increase the free diffusion of oxygen and water permeability, enhancing photodynamic therapy (PDT) and T 1 signal, respectively. (4) Folic acid anchored on the outermost surface is able to target at specific cells for bimodal imaging and PDT. Proof of concept is provided in the in-vitro cell testing of H–MnO@mSiO2 (Ir)–FA NPs, which demonstrates excellent targeting specificity, minimal collateral damage and highly potent therapeutic effect.
Article
Full-text available
CdSe/CdS-Quantum-dots-quantum-rods (QDQRs) with an aspect ratio of ∼6 are prepared via the seeded growth method, encapsulated within a shell of crosslinked poly(isoprene)-block-poly(ethylene glycol) (PI-b-PEG) diblock copolymer, and transferred from the organic phase into aqueous media. Their photoluminescence quantum yield (PLQY) of 78% is not compromised by the phase transfer. Within a period of two months the PLQY of QDQRs in aqueous solution at neutral pH decreases only slightly (to ∼65%). The two-photon (TP) action cross sections of QDQRs (∼10(5) GM) are two orders of magnitude higher than those of CdSe/CdS/ZnS-core/shell/shell quantum dots (QDs, ∼10(3) GM) with comparable diameter (∼5 nm). After applying PI-b-PEG encapsulated QDQRs onto the small intestinal mucosa of mice in vivo, their strong red fluorescence can easily be observed by two-photon laser scanning microscopy (TPLSM) and clearly distinguished from autofluorescent background. Our results demonstrate that PI-b-PEG encapsulated CdSe/CdS-QDQRs are excellent probes for studying the uptake and fate of nanoparticles by two-photon imaging techniques in vivo.
Article
Full-text available
Solution phase syntheses and size-selective separation methods to prepare semiconductor and metal nanocrystals, tunable in size from similar to 1 to 20 nn acid monodisperse to less than or equal to 5% are presented. Preparation of monodisperse samples enables systematic characterization of the structural, electronic, and optical properties of materials as they evolve from molecular to bulk in the nanometer size range. Sample uniformity makes it possible to manipulate nanocrystals into close-packed, glassy, and ordered nanocrystal assemblies (superlattices, colloidal crystals, supercrystals). Rigorous structural characterization is critical to understanding the electronic and optical properties of both nanocrystals and their assemblies. At inter-particle separations 5-100 Angstrom, dipole-dipole interactions lead to energy transfer between neighboring nanocrystals, and electronic tunneling between proximal nanocrystals gives rise to dark and photoconductivity. At separations <5 Angstrom, exchange interactions cause otherwise insulating assemblies to become semiconducting, metallic, or superconducting depending on nanocrystal composition. Tailoring the size and composition of the nanocrystals and the length and electronic structure of the matrix may tune the properties of nanocrystal solid-state materials.
Article
Full-text available
A simple route to the production of high-quality CdE (E = S, Se, Te) semiconductor nanocrystallites is presented. Crystallites from approximately 12 angstrom to approximately 115 angstrom in diameter with consistent crystal structure, surface derivatization, and a high degree of monodispersity are prepared in a single reaction. The synthesis is based on the pyrolysis of organometallic reagents by injection into a hot coordinating solvent. This provides temporally discrete nucleation and permits controlled growth of macroscopic quantities of nanocrystallites. Size selective precipitation of crystallites from portions of the growth solution isolates samples with narrow size distributions (<5% rms in diameter). High sample quality results in sharp absorption features and strong ''band-edge' emission which is tunable with particle size and choice of material. Transmission electron microscopy and X-ray powder diffraction in combination with computer simulations indicate the presence of bulk structural properties in crystallites as small as 20 angstrom in diameter.
Article
In semiconductor particles of nanometer size, a gradual transition from solid-state to molecular structure occurs as the particle size decreases. Consequently, a splitting of the energy bands into discrete, quantized levels occurs. Particles that exhibit these quantization effects are often called “Q-particles” or, generally, quantized material. The optical, electronic and catalytic properties of Q-particles drastically differ from those of the corresponding macrocrystalline substance. The band gap, a substance-specific quantity in macrocrystalline materials, increases by several electron volts in Q-particles with decreasing particle size. In Q-particles there are approximately as many molecules on the surface as in the interior of the particle. Therefore, the nature of the surface as well as the particle size is also largely responsible for the physico-chemical properties of the particle. Q-particles of many materials can be prepared in the form of colloidal solutions or embedded in porous matrices and are stable over a long period of time. In sandwich colloids, in which Q-particles of different materials are coupled, as well as in porous semiconductor electrodes containing Q-particles in the pores, very efficient primary charge separation is observed. As a result, sandwich colloids have greatly enhanced photocatalytic activity relative to the individual particles, while electrodes modified with Q-particles show high photocurrents. This article deals with the size quantization effect, the synthesis and characterization of Q-particles, as well as with the spectroscopic, electrochemical, and electron-microscopic investigation of these particles.
Article
The theory of Part I is applied to nonradiative transitions from the lowest triplet state to the ground state of aromatic hydrocarbons. A previously communicated empirical relation between triplet energy, triplet lifetime, and the relative number of hydrogen atoms per molecule is substantiated and its physical implications are discussed. It is transformed into a relation between the Franck‐Condon factor of the transitions and the triplet energy. In this form it can be compared with the theoretical expressions derived in Part I. No satisfactory theoretical representation of the empirical formula could be obtained on the basis of a harmonic‐oscillator description of the normal modes of the molecules. However, introduction of anharmonicity leads to excellent agreement between theory and experiment. A one‐parameter formula is derived which accounts with good accuracy for the dependence of the triplet lifetime on the triplet energy and the number of carbon and hydrogen atoms in the molecule. This formula shows that the Franck‐Condon factors relevant to the radiationless triplet‐ground‐state transition are governed by CH or CD stretching modes, which behave in this respect as completely degenerate for a given molecule. The single adjustable parameter is related to the anharmonicities of these modes in the ground and triplet state. The analysis confirms that the purely radiative triplet lifetime of most if not all aromatic hydrocarbons is close to 30 sec and that the Franck‐Condon factor is the only parameter in the expression for the nonradiative triplet decay constant which varies considerably between different aromatic hydrocarbons. Finally the temperature dependence of the Franck‐Condon factors is considered. A theoretical treatment indicates a very small temperature dependence below 400°K. This result seems to be borne out by an analysis of recent experiments, leading to the conclusion that the observed temperature effects are associated with bimolecular processes. A notable exception is benzene, where the triplet lifetime is temperature‐dependent down to very low temperatures, possibly due to a Jahn‐Teller distortion.
Article
With their bright, photostable fluorescence, semiconductor quantum dots (QDs) show promise as alternatives to organic dyes for biological labeling. Questions about their potential cytotoxicity, however, remain unanswered. While cytotoxicity of bulk cadmium selenide (CdSe) is well documented, a number of groups have suggested that CdSe QDs are cytocompatible, at least with some immortalized cell lines. Using primary hepatocytes as a liver model, we found that CdSe-core QDs were indeed acutely toxic under certain conditions. Specifically, we found that the cytotoxicity of QDs was modulated by processing parameters during synthesis, exposure to ultraviolet light, and surface coatings. Our data further suggest that cytotoxicity correlates with the liberation of free Cd2+ ions due to deterioration of the CdSe lattice. When appropriately coated, CdSe-core QDs can be rendered nontoxic and used to track cell migration and reorganization in vitro. Our results provide information for design criteria for the use of QDs in vitro and especially in vivo, where deterioration over time may occur.
Article
A simple synthetic route for the preparation of high-quality CdSe/ZnSe core/shell nanocrystals without the use of any pyrophoric organometallic precursors is presented. Effective surface passivation of monodisperse CdSe nanocrystals is achieved by overcoating them with a ZnSe shell, applying zinc stearate as a zinc source. The resulting core/shell nanocrystals exhibit high room temperature photoluminescence efficiencies (60−85%) in organic solvents as well as in water after functionalization with mercaptoundecanoic acid (MUA).
Article
Highly monodisperse CdSe nanocrystals were prepared in a three-component hexadecylamine−trioctylphosphine oxide−trioctylphosphine (HDA−TOPO−TOP) mixture. This modification of the conventional organometallic synthesis of CdSe nanocrystals in TOPO−TOP provides much better control over growth dynamics, resulting in the absence of defocusing of the particle size distribution during growth. The room-temperature quantum efficiency of the band edge luminescence of CdSe nanocrystals can be improved to 40−60% by surface passivation with inorganic (ZnS) or organic (alkylamines) shells.
Article
We report a synthesis of highly luminescent (CdSe)ZnS composite quantum dots with CdSe cores ranging in diameter from 23 to 55 Å. The narrow photoluminescence (fwhm ≤ 40 nm) from these composite dots spans most of the visible spectrum from blue through red with quantum yields of 30−50% at room temperature. We characterize these materials using a range of optical and structural techniques. Optical absorption and photoluminescence spectroscopies probe the effect of ZnS passivation on the electronic structure of the dots. We use a combination of wavelength dispersive X-ray spectroscopy, X-ray photoelectron spectroscopy, small and wide angle X-ray scattering, and transmission electron microscopy to analyze the composite dots and determine their chemical composition, average size, size distribution, shape, and internal structure. Using a simple effective mass theory, we model the energy shift for the first excited state for (CdSe)ZnS and (CdSe)CdS dots with varying shell thickness. Finally, we characterize the growth of ZnS on CdSe cores as locally epitaxial and determine how the structure of the ZnS shell influences the photoluminescence properties.
Article
We report here solution-phase synthesis of multifunctional nanoprobes for imaging live cancer cells. More specifically, quantum dots (QDs) and magnetite (Fe3O4) nanoparticles were coencapsulated within organically modified silica (ORMOSIL) nanoparticles. These optically and magnetically doped ORMOSIL nanoparticles were systematically characterized. These ORMOSIL nanoparticles were used for magnetically guided in vitro delivery to a human cancer cell line. Two-photon imaging was used to confirm the uptake of nanoparticle-doped ORMOSIL nanoparticles into the cancer cells. We also demonstrate that synergistic interaction of both magnetic targeting and transferrin-receptor (TfR)-mediated targeting results in robust in vitro cellular uptake, which is significantly higher than that obtained using these two targeting modes individually. Our cell viability study indicates them to be nontoxic. The results illustrate a promising biocompatible nanoparticle platform where multiple imaging and targeting agents can be coincorporated, with the purpose of efficient two-photon bioimaging and magnetically assisted targeting.
Article
CdSe and CdSe@CdS semiconductor nanocrystals have been synthesized in aqueous solutions, using sodium citrate as a stabilizer. Although initially these quantum dots display photoluminescence with very low quantum yields, upon prolonged illumination with visible light, enhancements up to 5000% have been measured. This leads to aqueous quantum dots with high luminescence, which can have important implications in biological and other applications. A distinct correlation between the photocorrosion process and the photoactivation process is observed. The primary reason for luminescence enhancement is considered to be the smoothing of the CdSe core surface. Importantly, even stronger activation was observed in silica- and CdS-coated nanocolloids where the CdSe core was expected to be shielded from photocorrosion. Preferential adsorption of oxygen molecules in the porous silicate shell accelerates the photocorrosion process. In CdS-coated particles, incomplete coating of the original particles is postulated, which is accompanied by the reforming of the CdS coat because of ionic diffusion at the interface on the newly opening areas with smoother surfaces.