ArticlePDF Available

Abstract and Figures

Luminescence spectroscopy experiments were realized for single colloidal quantum dots CdSe/ZnS in a broad temperature range above room temperature in a nitrogen atmosphere. Broadening and shifts of spectra due to the temperature change as well as due to spectral diffusion processes were detected and analyzed. A linear correlation between the positions of maxima and the squared linewidths of the spectra was found. This dependence was explained by a model that takes into account the slow variation of the electron-phonon coupling strength.
Content may be subject to copyright.
J. Chem. Phys. 151, 174710 (2019); https://doi.org/10.1063/1.5124913 151, 174710
© 2019 Author(s).
Contribution of electron-phonon coupling
to the luminescence spectra of single
colloidal quantum dots
Cite as: J. Chem. Phys. 151, 174710 (2019); https://doi.org/10.1063/1.5124913
Submitted: 20 August 2019 . Accepted: 21 October 2019 . Published Online: 07 November 2019
Eduard A. Podshivaylov, Maria A. Kniazeva, Aleksei A. Gorshelev, Ivan Yu. Eremchev, Andrei V.
Naumov, and Pavel A. Frantsuzov
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
Contribution of electron-phonon coupling
to the luminescence spectra of single colloidal
quantum dots
Cite as: J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913
Submitted: 20 August 2019 Accepted: 21 October 2019
Published Online: 7 November 2019
Eduard A. Podshivaylov,1Maria A. Kniazeva,1Aleksei A. Gorshelev,2Ivan Yu. Eremchev,2,a) Andrei V. Naumov,2,3
and Pavel A. Frantsuzov1,4,b)
AFFILIATIONS
1Lomonosov Moscow State University, 119991 Moscow, Russia
2Institute of Spectroscopy RAS, 108840 Moscow, Russia
3Moscow State Pedagogical University, 119991 Moscow, Russia
4Voevodsky Institute of Chemical Kinetics and Combustion SB RAS, 630090 Novosibirsk, Russia
Note: This paper is part of the JCP Special Topic on Colloidal Quantum Dots.
a)eremchev@isan.troitsk.ru
b)pavel.frantsuzov@gmail.com
ABSTRACT
Luminescence spectroscopy experiments were realized for single colloidal quantum dots CdSe/ZnS in a broad temperature range above
room temperature in a nitrogen atmosphere. Broadening and shifts of spectra due to the temperature change as well as due to spectral
diffusion processes were detected and analyzed. A linear correlation between the positions of maxima and the squared linewidths of the
spectra was found. This dependence was explained by a model that takes into account the slow variation of the electron-phonon coupling
strength.
Published under license by AIP Publishing. https://doi.org/10.1063/1.5124913
., s
I. INTRODUCTION
Colloidal semiconductor quantum dots (QDs) are very interest-
ing objects because of their unique optical properties such as a wide
absorption spectrum, a narrow emission line, a size-tunable emis-
sion wavelength, high photostability, and high fluorescence quan-
tum yield. The very first spectroscopic measurements of single CdSe
quantum dot photoluminescence revealed interesting phenomena
such as long-term fluctuations of the emission intensity (blinking)1
and very slow spectral diffusion (SD)2–4 with characteristic time
scales of up to hundreds of seconds. It was shown that at cryo-
genic temperatures, the observed emission spectrum linewidth of
single QDs depends on the signal accumulation time due to spectral
shifts,2,5 while the linewidths and the peak positions of the spectra
are correlated.6
Spectral diffusion at higher temperatures was observed by
Muller et al.7,8 in single CdSe QDs capped by a CdS rodlike shell.
The linewidth and the peak position of the emission spectrum are
found to be correlated at 5 K, 50 K, and room temperature. These
correlations at all temperatures were explained7,8 by the motion of
the net surface charge, which induces a Stark shift of the emis-
sion energy depending on the distance to the CdSe core, while
the spatial jitter of the charge density causes spectral line broad-
ening. Gomez et al.9noted that this hypothesis does not apply to
the spherically symmetric QDs. Besides, it should lead to varia-
tions of the linewidth with a change in the dielectric properties
of the medium. A series of spectroscopic experiments were per-
formed on single spherical QDs spin-coated on top of thin films
of various polymer matrices at room temperature. It was shown9
that there is a correlation between the linewidth and the peak posi-
tion of the emission spectrum in these particles without a signifi-
cant dependence on the dielectric permittivity of the matrix. Based
on this, it was concluded in Ref. 9that the mechanism respon-
sible for the correlated broadening and the peak position shift
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-1
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
of the emission spectra in the PL has to be intrinsic to the QD
core.
Note that the broadenings of single QD emission spectra at
5 K and at room temperature are different in nature. At 5 K,
the zero-phonon line is observed and its width is much smaller
than the longitudinal optical (LO) phonon energy.10 The linewidth
at room temperature becomes greater than the energy of the LO
phonons, which means that the multiphonon nature of the broad-
ening should be taken into account.11,12 While the electron-phonon
coupling and spectral diffusion contributions to the spectra of chro-
mophore molecules in solid matrices have been studied in detail,13–15
the same contributions in QDs are still of much interest. Here, we
present an in-depth experimental and theoretical study of the dis-
cussed spectral characteristics of single colloidal semiconductor QDs
CdSe/ZnS in the context of their feasible relation to electron-phonon
coupling.
II. EXPERIMENT
We performed a set of spectroscopic experiments with single
QDs, including measurements with slow heating and cooling of a
sample.
Fluorescence images and spectra of single quantum dots were
recorded using a home-built fluorescence microscope equipped
with a prism spectrometer.13,16 Two optical schemes—a wide-field
scheme and a scanning confocal one (see Fig. 1)—were combined in
the microscope in order to simplify the procedure of single quan-
tum dot preliminary searching (by using fluorescence image pro-
cessing and antibunching identification) and to perform sequential
measurements of the fluorescence spectra of the selected QD. Quan-
tum dots (CdSe/ZnS from Sigma-Aldrich with the fluorescence peak
at 620 nm) were dispersed in a toluene solution of polyisobuty-
lene of low concentration and then spin coated onto a cover glass.
The thickness of the polymer films with single quantum dots var-
ied within the range of several tens of nanometers. The sample was
placed onto the piezo-driven stage (NanoScan Technology), which
allowed one to move the selected QD to the laser spot position
with high (nanometer) precision. Between the sample and the piezo-
driven stage, a thermoinsulating (fluoroplastic) substrate a few mil-
limeters thick was placed, with a hole in the center allowing the
microscope objective to approach the plane of the sample at the
required distance. The thermoinsulating substrate contained a tem-
perature sensor that had good thermal contact with the sample.
On top of the sample, a three-stage thermoelectric module was
pressed, which was used to heat or cool the sample. This opti-
cal scheme (including the piezo-driven stage with the sample, the
microscope objective, and the thermoelectric module) was mounted
inside a special home-built chamber, allowing measurements both
in a vacuum and in a gas nitrogen/helium atmosphere. In this
particular case, the measurements were performed in a nitrogen
atmosphere. The sample temperature was controlled by using a
LakeShore temperature controller. A tunable dye laser (Coherent
CR599) or solid state laser Coherent Verdi was used to excite quan-
tum dots at the wavelength of 580 nm (near the quantum dot
absorption band edge) or at 532 nm, respectively. The excitation
laser intensity (100 W/cm2in a focused spot) was attenuated by
neutral spectral density filters (Standa) and controlled by using a
Newport power meter. A set of interference filters (Semrock and
Thorlabs) was used for the separation of the QD fluorescence sig-
nal from the scattered laser radiation. Two highly sensitive cooled
electron multiplying charge-coupled device (EMCCD) cameras were
utilized to record single quantum dot images (Andor Luca) and
spectra (Andor Ixon Ultra). The Hanbury Brown and Twiss scheme
with broadband 50% splitter (Thorlabs) and two identical single-
photon avalanche diode (SPAD) detectors (EG&G SPCM-200PQ,
time resolution 1.3 ns, dead time 200 ns, QE 65%) was used to
measure the autocorrelation function for QD fluorescence intensity.
Each fluorescence spectrum from a single quantum dot was mea-
sured with an exposure time of 200 ms and a spectral resolution
of 0.7 nm, which was sufficient to achieve a good signal-to-noise
ratio.
FIG. 1. A schematic picture of the experimental setup.
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-2
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
FIG. 2. Spectral traces (left panel) where the peak position is shown by the red line, time dependencies of the normalized PL intensity (central panel), and the linewidth (right
panel) for two different single CdSe/ZnS quantum dots (a) and (b) measured at room temperature.
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-3
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
III. RESULTS
In the experiments at room temperature for each studied sin-
gle QD, we registered 2500–3000 emission spectra with 200 ms
accumulation time. The presence of both blinking and spectral dif-
fusion processes can be clearly seen. Spectral traces for two QDs are
shown in Fig. 2.
Each spectrum was fitted with a Gaussian function
G(ϵ)=G0
2πσ exp(ϵϵ0)2
2σ2+b,
whose four parameters were peak emission photon energy ϵ0,
linewidth σ, amplitude G0, and background level b. An example of a
typical spectrum fitting is shown in Fig. 3.
The correlation between the peak energy and the linewidth
was found for all studied QDs. As seen in Fig. 4, the peak energy
dependence of the linewidth squared can be fitted by the linear
function
σ2=αkT(E0ϵ0), (1)
where Tis the absolute temperature, kis the Boltzmann constant,
E0is the energy gap for particular QD, and the parameter αis
the linear dependence coefficient between the squared line width
and peak energy in the units of kT. The values of αare found
to be in the range from 0.48 to 0.63 for varied QDs at room
temperature.
In order to characterize the shift of the spectra, we found a
squared peak energy displacement of a typical single QD emission
spectrum
FIG. 3. The emission spectrum of a single CdSe/ZnS QD (blue line) and its fit
with a Gaussian (dashed line). The parameters of the fit are ϵ0= 2.029 eV and
σ= 18.3 meV.
D2(τ)=(ϵ0(t)ϵ0(t+τ))2
as a function of time τfollowing Ref. 17. As can be seen in Fig. 5, the
averaged spectral shift squared D2is less than the σ2of a typical sin-
gle QD spectrum for all delay times, but more importantly, it is much
smaller than linewidth squared when τis equal to the signal accu-
mulation time τ= 200 ms. Thus, we can conclude that the observed
linewidth is not related to the spectral shifts during this time
period.
FIG. 4. The peak energy vs the linewidth
squared for different single QDs: (a)–(d)
at room temperature (black points) and
the linear fit Eq. (1) with T= 300 K (red
lines). The statistical error in the value of
alpha in each fit is less than 0.014.
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-4
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
FIG. 5. Time dependence of D2function for a single QD at room temperature (black
diamonds). The red line is τβ. The value of βis 0.603.
IV. THEORY AND DISCUSSION
Such a large value of the linewidth can be explained by multi-
phonon excitation.11,12 Let us consider the following Hamiltonian of
the QD electronic system interacting with Nphonon modes:
ˆ
H=ˆ
H0+A
N
i=1
ˆ
qi(aiee+bigg), (2)
where
ˆ
H0=N
i=1
ˆ
p2
i
2+ω2
i
2ˆ
q2
i+E0ee, (3)
E0is the energy gap, and |gand |eare the ground state and
excited electronic state of the QD, respectively. The parameter A
characterizes the electron-phonon interaction strength. ˆ
qiand ˆ
pi
are the coordinate and momentum operators of ith phonon mode,
characterized by the frequency ωi. Both the excited state and the
ground state are connected with the photon modes in the model.
The interaction of the excited state and the ground state with the
ith phonon is described by the dimensionless coefficients aiand bi,
correspondingly.
The emission spectrum at a given value of Ahas a Gaussian
form with the following parameters (see details of the derivation in
the Appendix):
ϵ0=E0A2S
i=1
ai(aibi)
ω2
i
, (4)
σ2=kTA2S
i=1
(aibi)2
ω2
i
. (5)
Fluctuations of the linewidth within the model are explained by
a slow variation of the parameter A. Such variations of the electron-
phonon interaction were observed experimentally in single colloidal
QDs6,8 as well as in single chromoprotein molecules.18,19
Variations of the parameter Awith time lead to shifts in the
position of the maximum (spectral diffusion) correlated with the
linewidth. Equations (4) and (5) give the linear dependence of Eq. (1)
where
α=N
i=1
ai(aibi)
ω2
i1N
i=1
(aibi)2
ω2
i
.
As seen in Fig. 4, the parameter αvaries from one QD to another,
as well as E0. The model predictions are consistent with the exper-
imental results of Gomez et al.9since the electron-phonon inter-
action in QDs is not related to the dielectric properties of the
environment.
In order to check the theory at various temperatures, the spec-
troscopic experiment on one quantum dot was performed with
heating and cooling of the sample. Twenty spectra were measured
sequentially at each selected temperature in the range from 305.5 K
to 353.6 K. It was found that all the data can be well fitted by Eq. (1),
provided that the energy gap E0depends on temperature and αkeeps
constant, as seen in Fig. 6. The E0value proves to decrease with
temperature rise. The changes in the effective band gap presum-
ably occur due to thermal expansion. Importantly, this result does
not depend on the process which led the system to that temperature
(heating or cooling). Note that this dependence of the “pure” energy
gap E0on temperature is not due to the electron-phonon interaction
as is usually considered.20–22
Therefore, the suggested model explains the fluctuations of the
linewidth of a single QD emission at temperatures 300 K and above,
but at the same time, it predicts the spectral shifts correlated with the
linewidth.
What is the mechanism of the variations in the magnitude of
the electron-phonon interaction? The estimate shows that at a given
excitation intensity of 100 W/cm2, the average time between the
absorption of photons of one QD is about 1 μs. Thus, the effect of
multiexciton states can be excluded from consideration. An increase
in the temperature of a single QD after absorption of photons
does not exceed 2 K, as estimated by Kuno et al.,23 while thermal
relaxation is of the order of 10 ps. This means that local heating
can also be excluded from the possible causes of the phenomenon.
FIG. 6. The peak energy vs the linewidth squared for a single QD at various tem-
peratures (points). The solid lines indicate the theoretical prediction at four different
temperatures. Parameter α= 0.63.
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-5
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
Plakhotnik et al.17 showed that the squared energy displacement
of a single QD emission at cryogenic temperatures has an anoma-
lous (sublinear) behavior at short times D2τβ, where β<1. It
was explained by introducing a number of stochastic two-level sys-
tems (TLSs) having a wide distribution of flipping rates. The squared
energy displacement calculated with the use of our experimental
data at room temperature also shows a similar sublinear time depen-
dence (Fig. 5). This means that the TLS based model can be used to
describe the fluctuations of the electron-phonon interaction value at
high temperatures as well. A possible microscopic origin of the con-
formation change in the TLS could be due to the jumps of the surface
or interface atom between two quasistable positions.24 Note the gen-
eral interest regarding the microscopic nature of the SD processes
that were observed for most single quantum emitters: single organic
dye molecules,25–27 single light harvesting complexes and proteins,28
color centers in diamonds,29 and single rare-earth ions in crystals.30
In many cases, SD has been attributed to the tunneling processes in
an emitter and/or its local surroundings. At the same time, the rela-
tion between SD and phonon-assisted optical dephasing was always
under discussion.
Empedocles and Bawendi6observed changes in the electron-
phonon interaction parameter upon application of an external elec-
tric field. It can be assumed that the shift of the peak of the spectrum
in an external electric field is partially determined by a change in the
electron phonon interaction. To verify this assumption, additional
experiments could be performed.
Single QD blinking can also be explained within a TLS based
model of Ref. 31. The Multiple Recombination Center (MRC) model
suggested by Frantsuzov et al.31 reproduces the key properties of
single QD blinking, such as the ON and OFF time distribution func-
tions,31 the power spectral density,32 and the long-term correlations
between subsequent blinking times.33 The similarity in temporal
fluctuations of the spectrum and the emission intensity of a single
QD allows one to make the assumption that both phenomena can be
explained by a unified mechanism.34
In conclusion, our experiments show a linear correlation
between the position of the maximum and the linewidth squared of
a single QD emission spectrum at room temperature and above. In
order to explain the experimental results, we consider a model of QD
emission spectrum linewidth fluctuations based on a slow variation
of the electron-phonon interaction. The model was tested using the
data of a unique single QD spectroscopy experiment under heating
and cooling conditions.
ACKNOWLEDGMENTS
This study was supported by the Russian Foundation for Basic
Research, Project No. 16-02-00713. The measurements were car-
ried out under the State Contract of the Institute of Spectroscopy
RAS. The luminescence microscopy technique with detection of
single quantum dots with nanometer spatial resolution is devel-
oped under support of the Russian Science Foundation (Project No.
17-72-20266, head I. Yu. Eremchev).
APPENDIX: DERIVATION OF EQUATIONS (4) AND (5)
Potential energy of the excited electronic state is given by the
following formula:
Ue(q)=N
i=1ω2
i
2q2
i+Aqiai.
In the classical limit
hωikT, the probability distribution function
of the coordinates is given by the Boltzmann distribution,
P(q)=1
Zexp1
kT
N
i=1ω2
i
2q2
i+Aqiai, (A1)
where Zis the partition function,
Z=dq1dq2dqNexp1
kT
N
i=1ω2
i
2q2
i+Aqiai.
It is assumed that the thermal relaxation is much faster than the vari-
ations of the parameter A. The energy of the emitted photon at given
values of the phonon coordinates is equal to the difference between
the energies of the excited and ground states,
ϵ=E0+A
N
i=1(aibi)qi. (A2)
Equation (A1) is a multidimension Gaussian distribution, and
Eq. (A2) is a linear function of the coordinates qi. Consequently, the
distribution of ϵis also Gaussian,
p(ϵ)=1
2πσ exp(ϵϵ0)2
2σ2,
where the parameters ϵ0and σ2can be found by averaging over the
distribution (A1),
ϵ0=¯
ϵ=E0+A
N
i=1(aibi)qi, (A3)
σ2=(ϵ¯
ϵ)2=A2N
i=1(aibi)2(qi¯
qi)2. (A4)
The mean values for the coordinates can be easily found by integra-
tion over distribution (A1)
¯
qi=Aai
ω2
i
(qi¯
qi)2=kT
ω2
i
.
Substituting these expressions into Eqs. (A3) and (A4) gives
Eqs. (4) and (5).
REFERENCES
1M. Nirmal, B. O. Dabbousi, M. G. Bawendi, J. J. Maklin, J. K. Trautman, T. D.
Harris, and L. E. Brus, Nature 383, 802 (1996).
2S. A. Empedocles, D. J. Norris, and M. G. Bawendi, Phys. Rev. Lett. 77, 3873
(1996).
3S. A. Blanton, M. A. Hines, and P. Guyot-Sionnest, Appl. Phys. Lett. 69, 3905
(1996).
4A. P. Beyler, L. F. Marshall, J. Cui, X. Brokmann, and M. G. Bawendi, Phys. Rev.
Lett. 111, 177401 (2013).
5S. A. Empedocles and M. G. Bawendi, J. Phys. Chem. B 103, 1826 (1999).
6S. A. Empedocles and M. G. Bawendi, Science 278, 3873 (1997).
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-6
Published under license by AIP Publishing
The Journal
of Chemical Physics ARTICLE scitation.org/journal/jcp
7J. Muller, J. M. Lupton, A. L. Rogach, J. Feldmann, D. V. Talapin, and H. Weller,
Phys. Rev. Lett. 93, 167402 (2004).
8J. Muller, J. M. Lupton, A. L. Rogach, J. Feldmann, D. V. Talapin, and H. Weller,
Phys. Rev. B 72, 205339 (2005).
9D. E. Gomez, J. van Embden, and P. Mulvaney, Appl. Phys. Lett. 88, 154106
(2006).
10V. M. Dzhagan, Yu. M. Azhniuk, A. G. Milekhin, and D. R. T. Zahn, J. Phys. D
51, 503001 (2018).
11K. Huang and A. Rhys, Proc. R. Soc. London, Ser. A 204, 406 (1950).
12R. Kubo and Y. Toyozawa, Prog. Theor. Phys. 13, 160 (1955).
13I. Y. Eremchev, M. Y. Eremchev, and A. V. Naumov, Phys. Usp. 62, 294 (2019).
14I. Y. Eremchev, A. V. Naumov, Y. G. Vainer, and L. Kador, J. Chem. Phys. 130,
184507 (2009).
15K. R. Karimullin and A. V. Naumov, J. Lumin. 152, 15 (2014).
16I. Y. Eremchev, N. A. Lozing, A. A. Baev, A. O. Tarasevich, M. G. Gladush,
A. A. Rozhentsov, and A. Naumov, JETP Lett. 108, 30 (2018).
17T. Plakhotnik, M. J. Fernee, B. Littleton, H. Rubinsztein-Dunlop, C. Potzner,
and P. Mulvaney, Phys. Rev. Lett. 105, 167402 (2010).
18R. Kunz, K. Timpmann, J. Southall, R. J. Cogdell, A. Freiberg, and J. Kohler,
J. Phys. Chem. B 116, 11017 (2012).
19R. Kunz, K. Timpmann, J. Southall, R. J. Cogdell, A. Freiberg, and J. Kohler,
Angew. Chem., Int. Ed. 52, 8726 (2013).
20K. P. O’Donnell and X. Chen, Appl. Phys. Lett. 58, 2924 (1991).
21X. Wen, A. Sitt, P. Yu, Y.-R. Toha, and J. Tang, Phys. Chem. Chem. Phys. 14,
3505 (2012).
22K. A. Magaryan, K. R. Karimullin, I. A. Vasil’eva, and A. V. Naumov, Opt.
Spectrosc. 126, 41 (2019).
23M. Kuno, D. P. Fromm, H. F. Hammann, A. Gallagher, and D. J. Nesbitt,
J. Chem. Phys. 115, 1028 (2001).
24O. Voznyy and E. H. Sargent, Phys. Rev. Lett. 112, 157401 (2014).
25W. E. Moerner and M. Orrit, Science 283, 1670 (1999).
26E. Barkai, A. V. Naumov, Y. G. Vainer, M. Bauer, and L. Kador, Phys. Rev. Lett.
91, 075502 (2003).
27P. D. Reilly and J. L. Skinner, Phys. Rev. Lett. 71, 4257 (1993).
28C. Hofmann, H. Michel, M. van Heel, and J. Kohler, Phys. Rev. Lett. 94, 195501
(2005).
29J. Wolters, N. Sadzak, A. W. Schell, T. Schroder, and O. Benson, Phys. Rev. Lett.
110, 027401 (2013).
30E. Eichhammer, T. Utikal, S. Gotzinger, and V. Sandoghdar, New J. Phys. 17,
083018 (2015).
31P. A. Frantsuzov, S. Volkán-Kacsó, and B. Jankó, Phys. Rev. Lett. 103, 207402
(2009).
32P. A. Frantsuzov, S. Volkán-Kacsó, and B. Jankó, Nano Lett. 13, 402
(2013).
33S. Volkán-Kacsó, P. A. Frantsuzov, and B. Jankó, Nano Lett. 10, 2761 (2010).
34V. K. Busov and P. A. Frantsuzov, Opt. Spectrosc. 126, 70 (2019).
J. Chem. Phys. 151, 174710 (2019); doi: 10.1063/1.5124913 151, 174710-7
Published under license by AIP Publishing
... The temperature dependence of the spectral characteristics of QDs can become the basis for temperature sensors (15,16) and nanosensors (17,18). ...
... Температурная зависимость спектральных характеристик КТ может стать основой температурных сенсоров [15,6] и наносенсоров [17,18]. ...
Article
Full-text available
Полупроводниковые нанокристаллы (квантовые точки, КТ) обладают уникальными фотофизическими свой­ствами, что открывает широкие возможности их прикладного использования в методах и инструментах современной фотоники. В данной статье рассматриваются возможные приложения КТ. Обсуждаются как существующие устройства, так и перспективы разработки новых методов и приборов фотоники. Рассмотрены инновационные подходы применения КТ в различных областях современных фотонных технологий: оптоэлектронике, биофизике, квантовой оптике, сенсорике, фотовольтаике.
... [1][2][3][4][5][6][7] The possibility of applying nanocrystals to design light-emitting devices, laser sources, photocells, phosphors, biotags, and sensors is being actively investigated. [8][9][10] As an alternative to nanosized cadmium-containing structures, colloidal core/shell III-V compound-based QDs with low toxicity and higher photostability are particularly interesting. 11,12 At the same time, QDs with an InP core and a ZnS shell are among the most efficient and have gained widespread acceptance in practice. ...
Article
The utilization of InP-based biocompatible quantum dots (QDs) necessitates a comprehensive understanding of the structure-dependent characteristics influencing their optical behavior. The optimization of core/shell QDs for practical applications is of particular interest due to their reduced toxicity, enhanced photostability, and improved luminescence efficiency. This optimization involves analyzing thermally activated processes involving exciton and defect-related energy levels. This study investigates water-soluble colloidal InP/ZnS QDs with varying shell thicknesses and stabilizing coatings using temperature-dependent optical absorption (OA) and photoluminescence (PL). Our results indicate that all samples experience temperature-induced shifts in exciton absorption and luminescence peaks due to interactions with acoustic phonons. Despite the wide size distribution of nanocrystals, the halfwidth of the bands remains constant. We observe a temperature-dependent Stokes shift in InP/ZnS QDs, revealing the fine structure of exciton states across different configurations. Furthermore, our findings demonstrate common mechanisms underlying PL thermal quenching in these QDs, regardless of the shell thickness or coating type. Specifically, defect-related emissions arise from localized energy levels at the core/shell interface. At the same time, exciton PL quenching primarily occurs through thermally activated electron migration from the InP core to the ZnS shell. Overall, our study highlights the potential for tailoring the temperature response of InP/ZnS QDs by adjusting shell thickness, offering opportunities to optimize their performance for specific applications.
... These trap states can participate in the formation of both radiative [4,5] and non-radiative recombination channels [4,6,7]. Trap states play a decisive role in the blinking effect of single QD luminescence [8][9][10][11][12]. ...
Article
The paper presents the results of a study of shallow non-radiative trap states in colloidal Ag2S QDs and Ag2S/SiO2 core/shell QDs, passivated with 2-mercaptopropionic acid in ethylene glycol carried out the thermally stimulated luminescence technique. For Ag2S QDs, the presence of thermoluminescence peaks in the region of 150–330 K was established. The formation of the SiO2 shells leads to the disappearance of the thermoluminescence peak at 250 K. It is interpreted as a “healing” of defects, localized on Ag2S QD surface. The depths of the detected trap states were estimated using a kinetic model. They showed that the observed thermoluminescence bands are due to the presence of shallow holes trap states in Ag2S QDs with depths of 0.117 eV and 0.135 eV for the thermoluminescence peaks at 190 K and 250 K, respectively.
Article
For decades, the thermal broadening of zero-phonon lines of single organic molecules embedded in amorphous solids was associated with their weak electron-phonon coupling to quasilocal vibrational modes. In this paper we show that the generally accepted perturbative approach for the pure dephasing problem is not applicable to the studied chromophore-doped glass, just as there is no evidence of weak coupling to individual quasilocal modes. Experimental and theoretical analysis of the temperature-dependent fluorescence excitation spectra of single tetra-tert-butylterrylene molecules in a polymer matrix of polyisobutylene is used as an example. Analysis of wide-range temperature measurements (9–60 K) with the nonperturbative theory of electron-phonon coupling demonstrates that the observed zero-phonon line broadening is due to the appearance of a wide spectrum of resonant modes induced by the impurity molecules themselves. This spectrum and the associated thermal broadening are obtained in the long-wave approximation for propagating waves using the density of the vibrational states for the host matrix and two quadratic coupling constants describing the interaction between the host and guest molecules. We also test the model considering the resonant modes arising from the host normal modes of a quasilocal character and find it completely incompatible with the experimental results.
Article
Full-text available
Quantum dots are the most exciting representatives of nanomaterials. They are synthesized by modern methods of nanotechnology pertaining to both inorganic and organic chemistry. Quantum dots possess unique physical and chemical properties; therefore, they are used in very different fields of physics, chemistry, biology, engineering and medicine. It is not surprising that the Nobel Prize in chemistry in 2023 was given for discovery and synthesis of quantum dots. In this review, modern methods of synthesis of quantum dots, their optical properties and practical applications are analyzed. In the beginning, a short historical background of quantum dots is given. Many gifted scientists, including chemists and physicists, were engaged in these studies. The synthesis of quantum dots in solid and liquid matrices is described in detail. Quantum dots are well-known owing to their unique optical properties, that is why the attention in the review is focused on the quantum-size effect. The causes for fascinating blinking of quantum dots and techniques for observation of a single quantum dot are explained. The last part of the review describes important applications of quantum dots in biology, medicine and quantum technologies. Bibliography — 772 references.
Article
Colloidal quantum dots (CQDs) are promising candidates for single photon sources (SPSs), pivotal for quantum technology. This review explores their applications, advancements, and potential in quantum photonics.
Article
In spite of decades of comprehensive studies, the phenomenon of photoluminescence (PL) blinking in single semiconductor colloidal quantum dots (QDs) still requires theoretical retreatment. Here we present an enhanced model which proposes that the blinking phenomenon is caused by fluctuations in the rate of nonradiative relaxation due to temporal variations in the electron–phonon interaction coupling. This model quantitatively reproduces the results of single CdSeS/ZnS core/shell QD spectroscopy experiments. In order to analyze the temporal properties of blinking, a new method of power spectral density estimation is proposed, based on the second-order cross-correlation function of the PL intensity, obtained experimentally. The proposed method extends the frequency range up to 5–6 orders of magnitude.
Article
The results of studies on the optical properties of colloidal InP/ZnS nanocrystals stabilized by a heterobifunctional polymer based on polyvinylpyrrolidone are presented. The optical absorption and photoluminescence spectra of (i) solutions containing different concentrations of nanocrystals and (ii) film samples, as well as the temperature dependences of these spectra in the range of 6.5–296 K are analyzed. An inhomogeneous broadening of the exciton optical bands was observed, which is associated with a broad size distribution of nanocrystals. It was established that the temperature-induced shift of the exciton absorption and emission maxima is mainly due to the interaction with acoustic phonons. It was shown that quenching of defect-related luminescence involves the local energy levels of the dangling bonds of phosphorus atoms at the core—shell interface, while the temperature stability of exciton emission is governed by the thickness of the ZnS shell.
Article
The cylindrical quasi-one-dimensional shape of colloidal semiconductor nanorods (NRs) gives them unique electronic structure and optical properties. In addition to the band gap tunability common to nanocrystals, NRs have polarized light absorption and emission and high molar absorptivities. NR-shaped heterostructures feature control of electron and hole locations as well as light emission energy and efficiency. We comprehensively review the electronic structure and optical properties of Cd-chalcogenide NRs and NR heterostructures (e.g., CdSe/CdS dot-in-rods, CdSe/ZnS rod-in-rods), which have been widely investigated over the last two decades due in part to promising optoelectronic applications. We start by describing methods for synthesizing these colloidal NRs. We then detail the electronic structure of single-component and heterostructure NRs and follow with a discussion of light absorption and emission in these materials. Next, we describe the excited state dynamics of these NRs, including carrier cooling, carrier and exciton migration, radiative and nonradiative recombination, multiexciton generation and dynamics, and processes that involve trapped carriers. Finally, we describe charge transfer from photoexcited NRs and connect the dynamics of these processes with light-driven chemistry. We end with an outlook that highlights some of the outstanding questions about the excited state properties of Cd-chalcogenide NRs.
Preprint
The present paper deals with the results of a research work on the optical properties of colloidal InP/ZnS nanocrystals stabilized with a heterobifunctional polyvinylpyrrolidone polymer. We have analyzed the absorption and photoluminescence spectra of the samples as solutions with various concentrations and deposited films, as well as the patterns of their temperature changes in the range of 6.5 - 296 K. An inhomogeneous broadening of exciton optical bands has been observed to be related to a wide distribution of nanocrystals in size. A temperature shift of the exciton absorption and emission maxima has been found to be due to the interaction with acoustic phonons. It has been shown that the quenching of defect-related luminescence is carried out through local energy levels of dangling bonds of phosphorus atoms involved at the core-shell interface, and the temperature stability of exciton emission is determined by the thickness of the ZnS shell.
Article
Full-text available
Three models of single colloidal quantum dot emission fluctuations (blinking) based on spectral diffusion were considered analytically and numerically. It was shown that the only one of them, namely the Frantsuzov and Marcus model reproduces the key properties of the phenomenon. The other two models, the Diffusion-Controlled Electron Transfer (DCET) model and the Extended DCET model predict that after an initial blinking period, most of the QDs should become permanently bright or permanently dark which is significantly different from the experimentally observed behavior.
Article
Full-text available
The temperature dependences of the positions of maxima of exciton bands in the luminescence spectra of liquid crystal nanocomposites with CdSe quantum dots with sizes of 1.8 and 2.3 nm at T = 77–300 K have been analyzed. The analysis under the theoretical model taking into account the electron–phonon interaction inside quantum dots has made it possible to calculate the values of the Huang–Rhys factor and average phonon energy in nanocrystals under study.
Article
Full-text available
A review of recent applications of Raman spectroscopy as a fast, sensitive, and non-destructive technique for exploring II-VI semiconductor nanocrystals fabricated by various methods (colloidal chemistry, Langmuir-Blodgett method, diffusion-limited growth) is presented. Specific size-related features revealed in the nanocrystal Raman spectra (phonon confinement, surface phonons) are analysed, as well as more complicated size effects for ultrasmall nanocrystals (NCs) related to the activation of the phonon density of states modified by surface reconstruction. Similarities and differences of the Raman scattering in II-VI and III-V or elemental (Si) semiconductor NCs are briefly analysed. Implementation of resonant conditions and application of infrared absorption analysis, complementary to the Raman spectroscopy - resulting in the observation of phonon modes 'silent' in conventional Raman scattering processes - are discussed. Furthermore, Raman spectroscopy is employed for fast and efficient assessment of the composition of matrix-embedded ternary II-VI nanocrystals, as well as more complicated multimode quaternary II-VI systems. Selective probing of electronic and vibrational spectra of different parts of heterogeneous NCs (such as core-shell systems) by tuning the excitation wavelength in resonant Raman scattering is considered. The analysis of phonon spectra is applied to the quantitative estimation of strain in the core and shell, and degree of interface intermixing, as well as to checking the surface oxidation. The above approaches and phenomena are further explored in more complex compound NCs beyond II-VI, such as CuInS2/ZnS. Recent results in the field of surface- and tip-enhanced Raman spectroscopy and surface-enhanced infrared absorption are analysed showing the perspectives of Raman spectroscopy as a tool for investigation of single-nanocrystal phonon spectra.
Article
Full-text available
Rare earth ions in crystals exhibit narrow spectral features and yperfine-split ground states with exceptionally long coherence times. These features make them ideal platforms for quantum information processing in the solid state. Recently, we reported on the first high-resolution spectroscopy of single Pr3+ ions in yttriumorthosilicate nanocrystals via the 3H4 − 3P0 transition at a wavelength of 488 nm. Here we show that individual praseodymium ions can also be detected on the more commonly studied 3H4 - 1D2 transition at 606 nm. In addition,we present the first measurements of the second-order autocorrelation function, fluorescence lifetime, and emission spectra of single ions in this system as well as their polarization dependencies on both transitions. Furthermore, we demonstrate that by a proper choice of the crystallite, one can obtain narrower spectral lines and, thus, resolve the hyperfine levels of the excited state.Weexpect our results to make single-ion spectroscopy accessible to a larger scientific community.
Article
Full-text available
Optoelectronic applications of colloidal quantum dots demand a high emission efficiency, stability in time, and narrow spectral bandwidth. Electronic trap states interfere with the above properties but understanding of their origin remains lacking, inhibiting the development of robust passivation techniques. Here we show that surface vacancies improve the fluorescence yield compared to vacancy-free surfaces, while dynamic vacancy aggregation can temporarily turn fluorescence off. We find that infilling with foreign cations can stabilize the vacancies, inhibiting intermittency and improving quantum yield, providing an explanation of recent experimental observations.
Article
Far-field fluorescence spectromicroscopy of single quantum emitters (SQEs) (single molecules, quantum dots, color centers in crystals) is an actively developing field of modern photonics, which is in widespread demand in various applications in physics, chemistry, material sciences, life sciences, and quantum technologies. In this paper, we present a description of a multifunctional experimental setup which was developed in recent years at the Institute for Spectroscopy of the Russian Academy of Sciences. It allows measuring optical spectra and fluorescence images of SQEs, as well as their temporal behavior and luminescence kinetics, in a broad range of temperatures (from cryogenic to ambient). It is shown that the spatial coordinates of SQEs can be reconstructed with subdiffractional accuracy (up to a few angstroms). Some examples of the developed methods for multiparameter superresolution microscopy (nanoscopy) of materials and nanostructures are presented.
Article
The possibility of determining the distance between two closely spaced blinking single colloidal quantum dots using super-resolution far-field luminescence microscopy is considered. Numerical simulation of microscope images of a single pair of point light emitters is performed for different modes of luminescence blinking observed in experiments with CdSe/ZnS core/shell quantum dots. The distance between the emitters is varied within the radiation wavelength λ, on the scale where effects associated with dipole–dipole interaction are typically manifested (from λ/5 to λ/30). An algorithm to determine the spatial arrangement of emitters in a pair with subdiffraction accuracy is developed and applied in model and laboratory experiments for different types of blinking dynamics of single quantum dots.
Article
The phenomenon of the zero-phonon spectral lines (ZPL), which corresponds to pure electronic transitions of impurity dye chromophore centers in solids, has already been the basis of high resolution spectroscopy for a few decades. The unique properties of the phonon-less emission in dye-doped solids (very narrow ZPLs and their extreme sensitivity to local environment) opens qualitatively new instrumental possibilities for sequentially-parallel separate spectromicroscopy of giant ensembles of single dopants. On this base the technique can be realized for multi-color far-field optical nanodiagnostics by phonon-less optical reconstruction single-molecule spectromicroscopy (PLORSM). Here we overview and systematize the already published data on photo-physical properties of a number of dyes (organic fluorescent molecules and quantum dots) in the context of their application in PLORSM in a broad range of temperatures. We show that some of dyes are suitable for realization of PLORSM technique at high (even room) temperatures.
Article
A quantitative theory for the shapes of the absorption bands of F-centres is given on the basis of the Franck-Condon principle. Underlying the treatment are two simplifying assumptions: namely, (a) that the lattice can be approximately treated as a dielectric continuum; (b) that in obtaining the vibrational wave functions for the lattice, the effect of the F-centre can be considered as that of a static charge distribution. Under these assumptions, it is shown that the absorption constant as a function of frequency and temperature can be expressed in terms of the Bessel functions with imaginary arguments. The theoretical curves for the absorption constant compare very favourably with the experimental curves for all temperatures. Also considered in the paper are the probabilities of non-radiative transitions, which are important in connexion with the photo-conductivity observed following light absorption by F-centres. The treatment given differs from the qualitative considerations hitherto in one important aspect, namely, the strength of the coupling between the electron and the lattice is taken into account. The adiabatic wave functions for the F-centre electron required for the discussion are obtained by perturbation methods. The probability for an excited F-centre to return to its ground state by non-radiative transitions is shown to be negligible; similar transitions to the conduction band are, however, important if the excited state is separated from the conduction band by not much more than 0cdot 1 eV. The temperature dependence of such transitions is complicated, but, for a wide range of temperatures, resembles eW/kT^-W/kT. Tentative estimates show that the result is consistent with the observed steep drop of the photo-conductive current with temperature.
Article
Spherical CdSe nanocrystals capped by a CdS rod-like shell exhibit interesting spectral dynamics on the single particle level. Spectral boxcar averaging reveals a high degree of correlation between the emission energy, spectral linewidth, phonon coupling strength, and emission intensity of the single nanocrystal. The results can be described in terms of a spatially varying surface charge density in the vicinity of the exciton localized in the CdSe core, leading to a quantum confined Stark effect which modifies the transition energy and the radiative rate. Whereas internal charging of the particle results in a change in the nonradiative rate, surface charges primarily influence the radiative rate. Additionally, we observe characteristic spectral dynamics in frequency space, the magnitude of which depends slightly on temperature and strongly on excitation density. By distinguishing between continuous spectral jitter and discrete spectral jumping associated with a reversible particle ionization event, we can attribute the spectral dynamics to either a slowly varying surface charge density or a rapidly occurring polarization change due to a reversible expulsion of a charge carrier from the semiconductor nanostructure. Whereas the former exhibits universal Gaussian statistics, the latter is best characterized by a Lorentzian noise spectrum. The Gaussian spectral noise increases with spectral redshift of the emission and with increasing proximity of the surface charge to the localized exciton. The observation of a high degree of correlation between peak position and linewidth right up to room temperature suggests applications of the nanocrystals as extremely sensitive single charge detectors in both solid state devices and in biomolecular labeling, where highly local measurements of the dielectric environment are required. Nanoscale control of the physical shape of nanocrystals provides a versatile test bed for studying electronic noise, making the approach relevant to a wide range of conducting and emissive solid state systems.