You are on page 1of 331

Lecture Notes in Nanoscale Science and Technology 28

Xin Tong
Zhiming M. Wang  Editors

Core/Shell
Quantum
Dots
Synthesis, Properties and Devices
Lecture Notes in Nanoscale Science and
Technology

Volume 28

Series editors
Zhiming M. Wang, Chengdu, China
Greg Salamo, Fayetteville, USA
Stefano Bellucci, Frascati RM, Italy
Lecture Notes in Nanoscale Science and Technology (LNNST) aims to report latest
developments in nanoscale science and technology research and teaching–quickly,
informally and at a high level. Through publication, LNNST commits to serve
the open communication of scientific and technological advances in the creation
and use of objects at the nanometer scale, crossing the boundaries of physics,
materials science, biology, chemistry, and engineering. Certainly, while historically
the mysteries in each of the sciences have been very different, they have all required
a relentless step-by-step pursuit to uncover the answer to a challenging scientific
question, but recently many of the answers have brought questions that lie at the
boundaries between the life sciences and the physical sciences and between what
is fundamental and what is application. This is no accident since recent research
in the physical and life sciences have each independently cut a path to the edge of
their disciplines. As both paths intersect one may ask if transport of material in a
cell is biology or is it physics? This intersection of curiosity makes us realize that
nanoscience and technology crosses many if not all disciplines. It is this market that
the proposed series of lecture notes targets.

More information about this series at http://www.springer.com/series/7544


Xin Tong • Zhiming M. Wang
Editors

Core/Shell Quantum Dots


Synthesis, Properties and Devices
Editors
Xin Tong Zhiming M. Wang
Institute of Fundamental and Frontier Institute of Fundamental and Frontier
Sciences Sciences
University of Electronic Science and University of Electronic Science and
Technology of China Technology of China
Chengdu, China Chengdu, China

ISSN 2195-2159 ISSN 2195-2167 (electronic)


Lecture Notes in Nanoscale Science and Technology
ISBN 978-3-030-46595-7 ISBN 978-3-030-46596-4 (eBook)
https://doi.org/10.1007/978-3-030-46596-4

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2020
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Colloidal semiconductor quantum dots (QDs) are small-sized fluorescent nanoparti-


cles covered with surfactant ligands and dispersed in solution. The typical diameter
of QDs in nanoscale dimension is comparable to that of the Bohr exciton radius
of the material, which results in the quantum confinement effect that enables the
size-dependent and tunable optical and electrical properties of QDs. This attractive
feature allows the realization of optoelectronic materials with properties that are
not available in conventional bulk semiconductors, making QDs as emerging next-
generation semiconducting materials in various photovoltaic and light-emission
devices. However, the extremely small size of QDs can result in a very high surface-
to-volume ratio and lead to a highly sensitive surface region, wherein the lower
coordination of surface atoms is prone to induce the formation of the surface
traps. These surface traps/defects can act as nonradiative recombination centers
of photogenerated charge carriers, thereby reducing the photoluminescence (PL)
efficiency and photo-/chemical/thermal stability of QDs, which are detrimental for
the relevant optoelectronic applications.
The growth of another semiconductor shell on core QDs to form core/shell
QDs is one of the most promising approaches to efficiently passivate the surface
defects/traps, leading to enhanced PL quantum yield, suppressed photobleaching,
and photoblinking as well as largely improved photo-/chemical-/thermal-stability
with respect to the organic ligands-capped bare QDs. More importantly, the
band structure of core/shell QDs can be tailored via tuning the core and shell
compositions and shell thickness etc., which is typically classified as type I, type
II, and quasi-type II according to the spatial distribution of photogenerated charge
carriers in the core/shell architecture. In type I core/shell QDs, both electron
and hole are confined within the core, giving rise to the significantly enhanced
PLQY and photostability that are favorable for light-emitting applications. The
photogenerated electrons and holes are spatially separated in type II and quasi-type
II core/shell QDs, leading to long-lived PL lifetime that is favorable for photovoltaic
(PV) applications requiring efficient charge separation/transfer. These optimized
and tunable optoelectronic properties render core/shell QDs as promising building
blocks for various photonic technologies.

v
vi Preface

This book consists of the latest research progress in the rational synthesis,
optoelectronic properties, and device applications of core/shell QDs. Specifically,
Chaps. 1, 2 and 3 mainly focus on the investigations of synthesis of binary II-VI and
ternary I-III-VI2 semiconductor core/shell QDs. Chapter 1 presents the synthesis
and microstructural and optical properties of binary II-VI semiconductor core/shell
QDs, such as CdSe/ZnS, CdSe/CdS, CdTe/ZnS etc., and the relevant applications of
these II-VI semiconductor core/shell QDs are briefly introduced as well. Chapter 2
aims to deal with the synthesis of core/shell, core/multi-shell, and core/doped shell
QDs using heavy metal-free ternary I-III-VI2 chalcogenides of CuInS2 and AgInS2
as core QDs. These core/shell structures can dramatically improve the optical and
luminance properties of QDs, which act as an alternative to the existing toxic II-VI
systems. In Chap. 3, a synthetic overview of stabilizing AgInS2 and AgInS2 /ZnS
QDs in aqueous colloids by small multi-functional molecular ligands is illustrated,
and the dependences of optical characteristics of the non-stoichiometric AgInS2
and AgInS2 /ZnS QDs on their composition and size as well as temperature are
discussed.
The optical and electrical properties of core/shell QDs are the main subjects of
Chaps. 4, 5 and 6. A theoretical study of single-electron, two-electron, and impurity
states as well as interband optical absorption and single-electron current in spherical
and cylindrical core/shell QDs is presented in Chap. 4, showing the possibility
of flexible manipulation of their spectral optical and spin characteristics. Chap-
ter 5 studies the temperature-dependent optical properties in a wide temperature
range for colloidal InP/ZnS core/shell QDs with different size distributions, and
corresponding exciton-phonon interactions are determined by using derivative spec-
trophotometry approach. The Auger recombination in core/shell QDs can induce
emission intermittency, which is detrimental for their optoelectronic applications
such as lasers. Chapter 6 investigates the optical properties of thick-shell core/shell
QDs, showing that the thick shells can suppress the Auger recombination and photo-
blinking and decrease the surface nonradiative channel and increase PLQYs and
absorption cross-sections, thus resulting in superior optical gain performance and
photostability.
Chapters 7, 8 and 9 are dedicated to several representative solar energy con-
version devices based on core/shell QDs. QDs-sensitized solar cells (QDSCs) are
considered to be one of the most appealing approaches that directly convert solar
radiation into electricity. Chapter 7 offers an overview of the recent progress
in the development of type-I, type-II, quasi type-II core/shell QDs as promising
light-harvesting materials to boost the performance of QDs-sensitized solar cells
(QDSCs). Particularly, the band alignment between core and shell materials is
demonstrated to efficiently tune the optoelectronic properties of QDs and reduce the
carrier recombination within QDSCs. Colloidal QDs-based photoelectrochemical
(PEC) cells are cost-effective devices showing remarkable solar-to-fuel conver-
sion efficiency. The advances in core/shell QDs-based solar-driven PEC cells are
presented in Chap. 8, highlighting the design of core/alloyed shell and heavy
metal–free, near-infrared (NIR) core/shell QDs for enhanced optical absorption and
optimized charge dynamics to achieve high performance QDs-based PEC devices.
Preface vii

Luminescent solar concentrators (LSCs) with large-area sunlight collection can be


employed for building-integrated PV devices. Chapter 9 reports the LSCs based on
core/shell QDs with tailored optical properties including large Stokes shift and high
PLQY, the relationship between QD’s structure/optical properties and the device
performance, and a perspective on the remaining key issues and open opportunities
in the field are provided as well.
The editors are grateful to all the authors for their significant contributions and
efforts to make this book a valuable guide to future optimization and developments
of colloidal semiconductor QDs-based optoelectronic applications. We would like
to thank Mr. Xin Li for his indispensable editorial assistance. Lastly, the editors
acknowledge the financial support of the National Key Research and Development
Program of China (2019YFB2203400), the “111 Project” (B20030), and the
UESTC Shared Research Facilities of Electromagnetic Wave and Matter Interaction
(Y0301901290100201).

Chengdu, China Xin Tong


Zhiming M. Wang
Contents

Synthesis, Properties, and Applications of II –VI Semiconductor


Core/Shell Quantum Dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Amar Nath Yadav, Ashwani Kumar Singh, and Kedar Singh
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based
Core-Multishell Nanocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
V. Renuga and C. Neela Mohan
Unique Luminescent Properties of Composition-/Size-Selected
Aqueous Ag-In-S and Core/Shell Ag-In-S/ZnS Quantum Dots . . . . . . . . . . . . . 67
Oleksandr Stroyuk, Oleksandra Raievska, and Dietrich R. T. Zahn
Electronic and Optical Characteristics of Core/Shell Quantum Dots . . . . . . 123
D. A. Baghdasaryan, H. T. Ghaltaghchyan, D. B. Hayrapetyan,
E. M. Kazaryan, and H. A. Sarkisyan
Exciton –Phonon Interactions and Temperature Behavior of Optical
Spectra in Core/Shell InP/ZnS Quantum Dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Sergey Savchenko, Alexander Vokhmintsev, and Ilya Weinstein
Thick-Shell Core/Shell Quantum Dots. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
Lei Zhang, Wenbin Xiang, and Jiayu Zhang
Core/Shell Quantum-Dot-Sensitized Solar Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Gurpreet Singh Selopal
Core/Shell Quantum-Dot-Based Solar-Driven
Photoelectrochemical Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Ali Imran Channa, Xin Li, Xin Tong, and Zhiming M. Wang
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators . . . . . . . 287
Guiju Liu, Xiaohan Wang, Guangting Han, and Haiguang Zhao

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315

ix
Synthesis, Properties, and Applications
of II–VI Semiconductor Core/Shell
Quantum Dots

Amar Nath Yadav, Ashwani Kumar Singh, and Kedar Singh

Abstract Semiconductor core/shell quantum dots (QDs) are composed of at least


two semiconducting materials having a structure like an onion. In the recent past, the
synthesis of these systems impelled significant progress, as by growing an epitaxial
shell, we can easily tune basic optical properties such as fluorescence quantum
yield, emission wavelength, and carrier lifetime. The significance of developing
an epitaxial shell over the surface of core QDs is making the nanocrystal less
sensitive to environmental changes and photo-oxidation. Another advantage is the
enhancement of fluorescence quantum yield by passivating surface trap states of
core QDs. These properties are essential for the application of semiconductor
core/shell QDs in light-emitting diodes, solar cell, and biological labelling. This
chapter discusses the synthesis and microstructural and optical properties of mostly
II–VI semiconductor, core/shell QDs. Moreover, various applications of core/shell
QDs in solar cells, light-emitting diode, and biomedical have been also discussed in
detail.

Keywords Core/shell quantum dots · Photoluminescence · Semiconductors ·


Synthesis

1 Introduction

In recent decades, semiconductor nanostructured materials are significantly cher-


ished because they can link the gap between small molecules and bulk materials
[1–4]. The nanostructured materials show distinct optical and electronic properties
when their size varies in the range of 1–100 nm. With the variation of dimension,
they can be classified as (1) two-dimensional, e.g., nanosheets or thin films or
quantum wells; (2) one-dimensional, e.g., quantum wires; and (3) zero-dimensional,

A. N. Yadav · A. K. Singh · K. Singh ()


School of Physical Sciences, Jawaharlal Nehru University, New Delhi, India
e-mail: kedar@mail.jnu.ac.in

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 1
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_1
2 A. N. Yadav et al.

Fig. 1 Schematic representation of the idealized density of states for semiconductor nanostructure
with reduced dimensionality (3d, 2d, 1d, and 0d represent three-dimensional, two-dimensional,
one-dimensional, and zero-dimensional). For 3d bulk structure, the energy levels are continuous,
whereas for 0d QDs energy levels are discrete. (Reprinted with permission from Ref. [5])

e.g., quantum dots. Thus, compared to the bulk semiconductor material, a quantum
dot (QD) is zero-dimensional and has a limited number of atoms, displaying discrete
energy states (Fig. 1) [5, 6]. Bulk semiconductor materials have continuous valance
and conduction energy states with composition-dependent bandgap (Eg ), which
is defined as minimal energy required to create an electron in conduction band
from the valance band. In this process, the electron leaves a hole in valance band
upon excitation energy higher than Eg . In the presence of the electric field, these
opposite charge carriers may be mobilized and hence carry current. A bound state
of electron-hole pair in their minimum energy state is called an exciton, and the
distance between them is exciton Bohr radius (rB ). Further, the excited-state electron
relaxes to annihilate exciton via radiative recombination of electron-hole and emit
their energy in the form of a photon [7]. Two major factors can influence the unique
properties of QDs: quantum confinement and surface effects.
Quantum confinement: When the radius of semiconductor nanocrystal (r)
becomes smaller or equal to exciton Bohr radius, i.e., r ≤ rB , then the motion of
electrons and holes is spatially confined to the dimension of the QD. Also, in this
condition, the energy difference between two levels of QD exceeds the value KB T
(here KB is Boltzmann constant); this results in restriction of the mobility of electron
and hole in the crystal dimensionality. In this regime, depending on the size of the
nanocrystals, the QDs exhibit size-dependent absorption and emission with discrete
electronic transitions [8, 9]. This effect is called quantum confinement effect. It can
affect several properties of semiconductor QDs, including magnetic properties and
conductivity. Nevertheless, the most exciting properties that arise from quantum
confinement is the size- and shape-dependent optoelectronic properties.
Surface effect: As the size of nanocrystal reduces, their surface-to-volume ratio
increases. This leads to an increment in the free energy of the NCs, which makes
NCs more dynamic and reactive as compared to their bulk counterparts. Henceforth,
it modifies the materials basic properties including solubility, reactivity, evaporation
and melting temperatures, plasticity etc. It also makes the NCs to easily disperse
in solvents media and opens up a possibility to functionalize or modify their
surface. This is the most beautiful property of colloidal NCs that can be used to
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 3

(a) (b)

(c)
band
offset
organic
molecule Eg Eg
(core) (shell)
band
offset

Fig. 2 Schematic representation of (a) organically passivated QD, (b) inorganically passivated
QD (core/shell structure), and (c) energy band offsets of core/shell structure. (Reprinted with
permission from Ref. [13])

build optoelectronic devices and targeting bio drugs. However, for very smaller
nanocrystal, the surface-to-volume ratio is very high; results in most of the bonds
associated with the surface are unsatisfied [10]. The energy levels associated with
surface states lie in between valance and conduction band. Further, these surface
states behave as fast non-radiative de-exciton channels and trap the photo-generated
charge carriers. As a result, the optical and optoelectronic properties of QDs are
significantly influenced by surface trap states [11, 12]. Therefore, passivation or
capping of the surface is essential for the evolution of luminescent and stable QDs.
Surface passivation of QDs can be carried out by two types: (1) organically and (2)
inorganically (Fig. 2) [13].
Organic surface passivation involves organic molecules that bond with sur-
face atoms and act as capping agent. The advantages of the organic passivation
include monodispersity, colloidal suspension, and bioconjugation of QDs [14–16].
However, full coverage of surface atoms and simultaneous passivation of both
cation and anion surface sites are still complicated because of shape distortion
and larger size of the organic capping molecules [17, 18]. The second strategy is
inorganically passivation, which involves full passivation of surface trap states. In
this approach, surface passivation has been carried out by overgrowth of inorganic
layer, particularly a second semiconductor. The resulting material is known as
core/shell QDs [19]. In this case, fluorescence QY and photostability of the
core/shell QDs drastically improved. It is also possible to tune the absorption and
emission spectra of the material by choosing appropriate core and shell materials.
The proper choice of shell material and its thickness are the essential terms that
contribute to overall properties of QDs. If the core and shell structures have huge
lattice mismatch, then this results in lattice strain that generates defect states within
or at the core/shell interface. In addition, the thicker shell creates misfit dislocations,
which also decreased fluorescence QY by the non-radiative process [20].
4 A. N. Yadav et al.

Fig. 3 Schematic diagram of band alignment for different types of core/shell structure: (a) type I,
(b) type II, (c) quasi type II, and (d) reverse type I. The wave function represented by blue color
is stands for electron wave function, whereas red color represents hole wave function. (Reprinted
with permission from Ref. [21])

1.1 Classification of Core/Shell Quantum Dots

Based on the energy bandgap offset, semiconductor core/shell QDs are generally
categorized into three groups; type I, reverse type I, and type II (Fig. 3). In type
I, the shell material has wider bandgap than the core material, i.e., valance and
conduction band edge of the core lies in between shell material so that the charge
carriers (electrons and holes) are restricted within the core. Further, in reverse type
I, the bandgap of the core material is larger than the shell material. In this case, the
photo-generated charge carriers partially or completely delocalized in the shell, and
by changing shell thickness, the emission wavelength can be tuned. Finally, in type
II, either valance or conduction band edge of shell material lies within the bandgap
of the core material. Consequently, upon photoexcitation, the charge carriers are
spatially isolated in a distinct region of core/shell heterostructure [19, 21].
In addition to above-discussed types of core/shell structure, there is an interme-
diate one identified as quasi type II (or type I1/2 ) core/shell QDs. The most studied
quasi type II system is CdSe/CdS core/shell QDs, although it is a type I core/shell
structure. However, the energy offset of the electron is minimal to confine it in
the CdSe core, and subsequently the electronic wave function delocalized over the
whole nanocrystal, while hole wave function remnant inside the core of the QD
[19, 21].

2 Synthesis of Core and Core/Shell Semiconductor QDs

Core/shell QDs have been mostly synthesized by two-step process: first the synthe-
sis of core QDs and second subsequent shell growth process.
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 5

Fig. 4 Schematic illustration Monodisperse Colloid Growth (La Mer)


of nucleation and growth

Nucleation
Concentration of Precursors
mechanism in nanocrystals
based on La Mer model [23].

(arbitrary units)
(Reprinted with permission
Nucleation Threshold
from Ref. [22])

from Solution
Injection

Growth
Ostwald ripening Staturation

0 200 400 600 800 1000


Time
(Seconds)

2.1 Synthesis of Core QDs

Colloidal semiconductor QDs have been generally synthesized by using three com-
ponents such as solvents, precursor, and organic surfactant. In an inert atmosphere,
heating of reaction mixture up to the required temperature changes the precursor
into monomers (active atomic and molecular species) and is called nucleation.
Further, the monomers are converted into nanocrystals whose continuous growth
depends on the surfactant molecules. Thus, two steps are required for the formation
of nanocrystals: (1) nucleation and (2) growth [22, 23].
In more than two decades, a large number of reviews have been reported related
to the synthesis of II–VI and the other semiconductor QDs [13, 19, 21–24]. In
general, two methods of non-injection and injection have been used for the synthesis
of aqueous and nonaqueous semiconductor QDs by varying solvents, temperatures,
and precursors (Fig. 4).

2.1.1 Injection Method

In 1993, Murray et al. developed a traditional approach for the synthesis of semi-
conductor QDs by fast injection of the precursor into a hot solvent [25]. In a typical
synthesis protocol, they first prepared Cd and Se precursor solutions by mixing
Cd(CH3 )2 and Se powder into tri-n-octyl phosphine (TOP) solvent, distinctively.
These precursors were further injected quickly into hot trioctylphosphine oxide
(TOPO) solution in a three-neck flask and under an inert (N2 ) atmosphere. Here,
TOPO acts as stabilizing agent and allows the reaction mixture to heat at a high
temperature generally up to 320 ◦ C. This synthesis technique was a model for the
preparation of CdSe, CdS, and CdTe QDs with different sizes 1.5–11.5 nm.
6 A. N. Yadav et al.

Later in 2001, Peng et al. have used CdO as a Cd precursor in place of


Cd(CH3 )2 . The reason behind is toxicity, explosive, and expensiveness of Cd(CH3 )2
in comparison to CdO [26, 27]. Moreover, the quality of CdSe QDs remains
the same as in the case of Cd(CH3 )2 . Well ahead, due to environment issue,
many efforts have been made for the synthesis of II–VI semiconductor QDs using
“green” hot injection synthesis method [28, 29]. After that, the expensive and
coordinating solvent TOPO has been replaced by low-cost and non-coordinating
solvent octadecane (ODE) [30–32]. Further, Deng et al. simplified the reaction by
using low-cost solvent oleic acid instead of TOP/TOPO [33].
As described above, the hot injection method is based on a fast injection of
precursors into a hot solution containing another precursor followed by an instant
homogenous reaction. In this method, it was hard to control the precise reaction
temperature upon injection of the precursor solution. Thus, scale-up synthesis and
reproducibility are difficult to achieve using this method. Therefore, for large-scale
synthesis, another method “non-injection method” has been developed.

2.1.2 Non-injection Method

In this method, all reagents are mixed in a three-neck flask, and nucleation and
growth are originated either chemically, thermally, or by physical impact (e.g.,
microwave irradiation) [34–36]. First, Pradhan et al. have developed a versatile and
easy method for the synthesis of extremely good metal sulfide (CdS) nanoparticles
using one-spot and low-temperature process [37]. In this method, heating of
metal xanthate in hexadecylamine (HDA) gives rise to metal sulfide NPs even
at low temperature 70 ◦ C. Moreover, by adjusting reaction temperature, NPs of
various sizes could be obtained using this method. Later, Coe and his collaborators
synthesized high-quality and monodispersed CdS QDs by one-spot non-injection
method [34]. In this method, they have introduced two nucleation initiators, namely,
tetraethylthiuram (I1 ) and 2,2 -dithiobisbenzothiazole (I2 ), in the reaction mixture.
So separate nucleation and growth are accomplished, and the quality of CdS QDs is
quite comparable with the injection method.
Besides the abovementioned non-injection method, another non-injection way
is single-source inorganic cluster approach. In this method, the single-source
precursor (e.g., Li4 [Cd10 Se4 (SPh)16 ] for CdSe) was added in dodecylamine or
hexadecylamine at 80 ◦ C under N2 . Further, after dissolving the cluster, the reaction
mixture was heated up to 220 ◦ C for the growth of QDs. From this method,
nucleation at relatively low temperature can be achieved, and reaction can be scaled
to large quantities (1–50 g/L). Further, other QDs such as CdS, ZnSe, and CdSe/ZnS
also can be prepared via this method [38].
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 7

2.2 Synthesis of Semiconductor Core/Shell QDs

For the growth of the shell, two points are most important: (1) choice of shell
material and (2) thickness of the shell material. In the former one, generally,
semiconductors with small lattice mismatch have been chosen for the core and
shell structure. If both structures have a huge lattice mismatch, then it results in
lattice strain and generates defect states at the core/shell structure or within the
shell. The defect states can act as trap states for the charge carrier and can decrease
PL QY [20]. Table 1 represents a list of material parameters for some selected bulk
semiconductors.
The second point is the thickness of shell material, which also plays a vital role
in the properties of core/shell QDs. If the shell thickness is skinny, then passivation
of surface traps of core QDs may be incomplete. On the other hand, if the shell
thickness is very thick, then it results in the formation of defect states due to lattice
strain. During shell growth, number monolayers of shell material are deposited
on the surface of the core QDs. The required amount of shell thickness can be
calculated as follows.
The volume of the shell material for “m” monolayers of shell thickness can be
calculated as
4  
Vshell =  (rc + m × dML )3 − rc3
3
where Vshell is the volume of the shell material, rc is the radius of core QDs, and
dML is the shell thickness for one monolayer (nm).
If nshell is the amount of shell material (in moles) required to deposit “m” mL
shell, then

Vshell × Dcore × NA × nQD


nshell =
MWcore

Table 1 Material parameter for bulk II–VI semiconductors [19]


Lattice
Semiconductor parameter
materials Structure (300 K) Type Egap (eV) (Å) Density (kg m−3 )
CdSe Wurtzite II–VI 1.74 4.3/7.01 5810
CdTe Zinc blende II–VI 1.43 6.48 5870
CdS Wurtzite II–VI 2.49 4.136/6.714 4820
ZnSe Zinc blende II–VI 2.69 5.668 5266
ZnTe Zinc blende II–VI 2.39 6.104 5636
ZnS Zinc blende II–VI 3.61 5.41 4090
8 A. N. Yadav et al.

where Dcore is the density of core material, NA is the Avogadro number, nQD is the
number of moles of core QDs in solution, and MWcore is the molecular weight of
core material [39].
After calculations of the precise amount of shell thickness, deposition of shell has
been carried out by a technique called successive ionic layer adsorption and reaction
(SILAR) method [40]. By this technique, we can deposit one monolayer at a time
by injecting cationic and anionic precursors into the core solution. After depositing
one layer successively, one can deposit many monolayers of shell thickness on the
surface of core QDs. By using this method, monodisperse and highly luminescent
semiconductor core/shell QDs could be synthesized.

2.2.1 Synthesis and Characterization of Type I Core/Shell QDs

Type I core/shell QDs have been synthesized for the motive to increase fluorescence
QY and to improve photostability of the QDs. The most studied II–VI semicon-
ductor in type I is CdSe/ZnS core/shell QDs. First, Guyot-Sionnest group have
grown ZnS shell of one to two monolayers onto 3 nm CdSe QDs by using growth
temperature 300 ◦ C, and, in this case, they have found QY to be 50% [41]. For the
synthesis of this system, they have used a hot injection method. First, they prepared
Cd and Se stock solutions by dissolving dimethylcadmium and selenium shot in
TOP. Further, using Cd and Se stock solutions, CdSe nanocrystal was synthesized.
The growth of ZnS shell has been achieved by injection of Zn and S stock
solutions which were developed by diethylzinc (Me2 Zn) and hexamethyldisilathiane
[S(TMS)2 ] in TOP solution. Later, Bawendi group have reported a depth study
on CdSe/ZnS core/shell QDs using low-temperature (140–220 ◦ C) synthesis [42].
Moreover, using HDA into TOPO/TOP, Weller group have been able to control
the growth kinetics of CdSe/ZnS core/shell QDs and achieved lower particle size
distribution with high QY up to 60% [43]. Moreover, the blue spectral region is also
accessible using this material with extremely small size as reported by Kundra and
co-authors [44]. They first synthesized CdSe “magic-sized” cluster in the solution of
TOP, DDA, and nonionic acid at 80 ◦ C and further sequential growth of ZnS shell.
In another report, Jun and Jang obtained similar spectral range with overgrowth of
ZnS shell at 300 ◦ C by using zinc acetate and octanthiol as a precursor [45]. In this
case, due to the low reactivity of precursors, the shell material spread into the core
results in a remarkable blue shift of emission peak observed around 470 nm with
QY up to 60%. In another work on the same material by Zhang group, they reached
QY up to 95% by TOP-SILAR method (Fig. 5) [46]. This high QY is obtained by
growing three monolayers of ZnS shell and even maintained after six monolayers of
shell thickness (Fig. 6c).
The second material in the type I system is CdSe/CdS core/shell QDs. In this
case, the lattice mismatch between core and shell material is only 4% and exhibits
different band alignment (holes have larger band offset than electrons). A detailed
study on CdSe/CdS core/shell QDs with varying core diameters 2.3–3.9 nm and QY
up to 50% was first reported by Peng and co-authors [47]. In this system, due to the
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 9

TOP Zn S TOP Zn S
SILAR TOP-SILAR

CdSe core Activated CdSe CdSe/ZnS QDs


Multilayered
CdSe/ZnS QDs

Fig. 5 Growth mechanism of multilayered CdSe/ZnS core/shell QDs by TOP-SILAR synthesis


method. (Reprinted with permission from Ref. [46])

Fig. 6 TEM images of (a) CdSe core and (b) CdSe/ZnS core/shell QDs after three monolayers
of Zn and S precursors. Insets of these figures are their high-resolution transmission electron
microscope (HRTEM) image. (c) Variation of QY with increasing shell thickness, where the core
is synthesized with three emitting colors green, orange, and red, respectively. Inset is photograph
of CdSe/ZnS core/shell QDs under UV light with three monolayers of shell thickness. (Reprinted
with permission from Ref. [46])

delocalization of electronic wave function over the whole core/shell, the absorption
and emission spectra get more shifted in comparison to the CdSe/ZnS system. The
precursors used in this system were dimethylcadmium and bis(trimethylsilyl sulfide)
with solvent TOPO. Later, using SILAR technique the same group synthesized the
same material by air-stable precursors cadmium oleate and elemental sulfur in the
10 A. N. Yadav et al.

low-cost, low toxic, and high boiling solvent octadecane [40]. The SILAR technique
was further extended to synthesize “giant” core/shell QDs as reported by Klimov
group for the synthesis of giant CdSe/CdS core/shell QDs [48]. The shell thickness
was 6 nm by growing 19 monolayers of CdS, and the QY was achieved up to 40%.
Recently, Manna group have reported the synthesis of extremely luminescent giant
CdSe/CdS core/shell QDs by a fast continuous injection method [49]. In this report,
they have synthesized CdSe core of diameter between 2.8 and 5.5 nm and CdS shell
thickness of 7–8 nm (~20 monolayers of CdS). Interestingly, the QY was maintained
up to 90% by using this defect-free QD. The purpose of growth of the giant shell was
to make nanocrystal independent from environment and surface chemistry. Further,
a thicker shell protects the QDs from photobleaching and photoblinking. However,
due to their larger size, they have poor size distribution and broad PL emission.
Therefore, another method was required that could meet all the criteria. Later,
Bawendi group synthesized high-quality CdSe/CdS core/shell QDs by optimizing
growth rate of the shell material, using cadmium oleate and octanthiol as shell
precursors [50]. The obtained core/shell QDs have very narrow emission peak with
high PL QY (Fig. 7) and high uniformity (Fig. 8). Moreover, in contrast to previous
studies, in this case, photoblinking too much suppressed with growing a relatively
thin shell (2.4 nm).
CdSe/ZnSe is another type I core/shell system, where electrons conduction band
offset is larger than the holes valance band offset. The lattice mismatch, in this case,
is significant as 6.3%; however, the anion-type structure is suitable for epitaxial shell
growth. In the early time of synthesis, the QY of this system is very low (<1%) [51].
Afterward, the synthesis method has been modified using zinc stearate rather than
diethylzinc as a zinc precursor and TOP-Se (Se dissolved in TOP) as a selenium

Fig. 7 Optical properties of CdSe/CdS core/shell QDs. a–d: UV-visible absorption (blue) and
photoluminescence (PL) (red) spectra of four different CdSe/CdS core/shell QDs synthesized
with different CdSe core diameters of (a) 2.7 nm, (b) 3.4 nm, (c) 4.4 nm, and (d) 5.4 nm. (e
and f) Temporal evolution of sample as shown in figure (c) during shell growth: (e) emission,
(f) absorption, (g) PL QY, and (h) FWHM of PL peak. Green square shows the original
photoluminescence QY of CdSe QDs. (Reprinted with permission from Ref. [50])
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 11

Fig. 8 TEM images of (a) CdSe core with diameter 4.4 nm and CdSe/ZnS core/shell QDs with
thickness (b) 0.8 nm, (c) 1.6 nm, and (d) 2.4 nm. Scale bar: 50 nm. (Reprinted with permission
from Ref. [50])

precursor [52, 53]. Narrow emission line widths with high QY up to 85% were
obtained using this method. Further, in another study, the concentration of ZnSe
precursor solution was varied to observe changes in optical spectra due to lattice
strain [54].
In CdSe/CdS and CdSe/ZnSe core/shell QDs, the emission wavelength can
be tunable with high fluorescence QY. However, for the high stability of optical
properties against photoblinking and photodegradation, the better choice of shell
material is zinc sulfide (ZnS). It is discussed above that, using a smaller size of
CdSe/ZnS core/shell QDs, it was possible to obtain blue and green emission. In
this context, materials other than CdSe have been developed for the better spam
of UV/blue/green and near-infrared spectrum. QD using alloyed structure is an
alternative for the tuning of emission by changing materials composition rather than
the particles size. For example, in Cd1 − x Znx Se QDs, the emission wavelength can
be tuned by changing the value of x. Subsequently, the alloy Cd1 − x Znx Se core
with ZnS shell has been grown using well-established precursors diethylzinc and
air-stable zinc diethyl xanthate [55, 56].
Shen et al. have described synthesis of high-quality Cd1 − x Znx Se/ZnS core/shell
QDs by using a new method called “nucleation at low temperature/shell growth at
high temperature” [57]. The obtained core/shell QDs are monodispersed with high
QY (up to 100%), high color purity (FWHM<25 nm), and emission tunability from
400 to 470 nm of the optical window (Fig. 9). Moreover, the resulting QDs have
excellent chemical and photochemical stability, and method can be easily applied
to large-scale synthesis up to 37 g in batch synthesis. Since after shell growth the
absorption and emission spectra were shifted toward longer wavelength side. Such
red shift, in this case, can be explained in terms of delocalization of electronic wave
function into the surrounding shell and reduction in quantum confinement effect.
Further, the size of the core/shell QDs varies in between 5.7 and 10.8 nm with
core mean diameter of 4.8 nm (Fig. 10). Recently, the same group successfully
controlled ZnS shell thickness on the ZnCdSe core QDs [58]. Moreover, by growing
the shell thickness up to ten monolayers, the external quantum efficiency of 17% was
achieved.
12 A. N. Yadav et al.

Fig. 9 Evolution of (a) absorption and emission spectra of Cd1 − x Znx Se core and
Cd1 − x Znx Se/ZnS (x = 0.25) core/shell QDs. (b) Normalized PL spectra of Cd1 − x Znx Se core
and Cd1 − x Znx Se/ZnS core/shell QDs (at different x) and inset image is corresponding UV-
illumination photos. (Reprinted with permission from Ref. [57])

In addition to the above discussion, the emission wavelength in blue and near-
UV spectral region was generally obtained by core material with larger bandgap.
As far as we know, the first core/shell structure historically was reported by
Spanhel et al. in 1987 [59]. It was treatment on 4–6 nm CdS core QDs with
sodium hydroxide (NaOH) and cadmium perchlorate [Cd(ClO4 )] in aqueous media,
making cadmium hydroxide shell. Initially, the PL QY of CdS QDs was 1%, but
after surface modification, the PL QY quickly rose to 50% with narrow emission
line widths. Later, in organic media, ZnS shell is grown on CdS core using
the organometallic method with S(TMS)2 and dimethylcadmium as precursors.
The emission wavelength spans over 460–480 nm with PL QY of 20–30% [60].
Moreover, Reiss group developed a new approach for the synthesis of CdS/ZnS
core/shell QDs [61]. In this method, the reagent described above is further replaced
by air-stable precursors like zinc stearate and zinc ethylxanthate. The subsequent
CdS/ZnS core/shell QDs were found to be monodispersed, tunable emission (range
of 440–480 nm) and QY up to 35–45%. In another study, manganese has been used
as a dopant in CdS host and the ZnS shell was either grown by the reverse micelle
or organometallic route [62–64]. In 2009, Peng group developed a new approach
“thermal cycling coupled single precursor” (TC-SP) for the synthesis of high-quality
CdS/ZnS core/shell QDs [65]. In contrast to classical SILAR, this is also a one-spot
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 13

Fig. 10 (a) TEM image for Cd1 − x Znx Se core QDs. TEM images of Cd1 − x Znx Se/ZnS core/shell
QDs, (b) shell growth for 5 min, (c) shell growth for 15 min, (d) shell growth for 45 min, (e)
shell growth for 90 min. Figures (f–j) represent HRTEM images corresponding to a–e images.
(Reprinted with permission from Ref. [57])

method; however, the growth temperature considerably reduced, and the technique
was simple than SILAR.
Meanwhile, ZnSe/ZnS core/shell QDs are also synthesized using organometallic
precursors by Manna group [66]. The observed emission wavelength was in the
range of 400 nm, with FWHM of 20 nm and QY of 17%. Further, a new method
was adopted by using alternative precursors like ZnO and TOPSe for the synthesis
of ZnSe core and zinc laurate and TOPS for ZnS shell [67]. The reaction mixture
was carried out at 180 ◦ C in HDA and the final product ZnSe/ZnS QDs have
QY up to 30%. However, most of the above method entails the use of TOP or
TBP as a precursor solvent. It well accepted that alkylphosphines (TOP or TBP)
are very expensive, hazardous, unstable, and air-unstable solvent. Therefore, a
green approach, i.e., phosphine-free synthesis, will be needed to synthesize these
core/shell QDs. Afterward, Dong et al. described a very-low-temperature (<150 ◦ C),
facile, and two-step process to prepare water-soluble ZnSe/ZnS core/shell QDs [68].
The precursors used for the shell growth were zinc acetate and thiourea in MPA
solvent. The emission wavelength was tuned in the range of 390–460 nm with QY
up to 45% by growing three monolayers of shell thickness (Fig. 11). The diameter of
core/shell QDs is found to be 5.5 nm after three monolayers of ZnS shell thickness
on 3.3 nm ZnSe core.
In this context, a green synthesis was described by Li group, who grow ZnS shell
over ZnSe core by two converse injection methods [69]. In this report, zinc and
14 A. N. Yadav et al.

(a) (b)
50 50
Absorbance, PL Intensity (a.u.)

40 40
4-layer

FWHM (nm)
QY (%)
3-layer
30 30
2-layer

1-layer 20 20
core

10 10
300 400 500 60 0 1 2 3 4
Wavelength (nm) Number of ZnS shell monolayer

Fig. 11 (a) Absorption and emission spectra of ZnSe core and ZnSe/ZnS core/shell QDs with
increasing shell thickness. (b) PL QY and FWHM at different ZnS shell monolayers. (Reprinted
with permission from Ref. [68])

selenium precursors were prepared by using octadecane (ODE: a phosphine-free


solvent) to synthesize ZnSe QDs. After that, stock solutions for shell growth were
prepared by dissolving ZnO in oleic acid and paraffin oil and sulfur powder in ODE.
The optical properties using this method retain the same as in the case of phosphine
solvent used QDs. In another report by the same group were synthesized (via low-
temperature/high-temperature shell growth) highly stable and violet-blue emitting
ZnSe/ZnS core/shell QDs [70]. First zinc stock solution was prepared by mixing
ZnO in oleic acid and paraffin oil at 300 ◦ C. Similarly, selenium stock solution was
adopted by dissolving Se powder in ODE at 220 ◦ C. These two stock solutions were
used to synthesize ZnSe core QDs in paraffin oil at 280 ◦ C. Finally, shell growth
of ZnS was carried out at 320 ◦ C by injecting Zn stock solution and octanethiol
in a solution containing ZnSe QDs. The obtained core/shell QDs retain excellent
optical properties such as high QY (up to 83%), high color saturation (FWHM
between 12 and 20 nm), tunable emission (violet to blue, 400–455 nm), and superior
chemical/photostability.
Cadmium telluride (CdTe) QD is another material in the II–VI group, which has
lower bandgap (1.5 eV) in comparison to CdSe (1.76 eV). This material is a good
candidate for application in red-infrared-emitting LEDs and other optical devices
[19]. Even if the described synthesis method in the case of CdSe also could be
adopted to CdTe QDs, however, some little efforts also were chosen to synthesize
this QD in organic media [71]. Meanwhile, Gao group observed a substantial
enhancement in PL QY up to 85% by the illumination of thioglycolic acid (TGA)-
capped CdTe QDs [72]. As a matter of fact, the improvement in PL QY was due to
photodegradation of TGA rather than CdTe QDs. In this case, the sulfide ions are
released from TGA during illumination and form a shell of CdS on the surface of
CdTe QDs reacting with Cd surface atoms. After that, similar to the previous study,
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 15

the CdS shell was grown on CdTe by microwave-assisted and ultrasonic irradiation,
respectively [73, 74]. In a report by Gu et al., they have proposed a one-spot aqueous
synthesis of highly luminescent CdTe/CdS core/shell QDs [75]. The CdS shell was
grown in a crude solution of MPA-capped CdTe QDs by using thiourea (acts as
sulfur source), which further reacted with excess Cd to form core/shell structure.
The QY was found to be high as 75%, and the stability of QDs quite improved.

2.2.2 Synthesis and Characterization of Type II Core/Shell QDs

Type II core/shell QDs could be recognized by recording optical spectra because


in this case huge spectral shift was observed in absorption as well as in emission
spectra after shell growth. Bawendi group described the first protocol for the
preparation of type II CdTe/CdSe and CdSe/ZnTe core/shell QDs and studied their
optical properties [76]. The emission wavelength was varied in the spectral region
from 700 to 1000 nm by changing CdTe core size and CdSe shell thickness. In
this structure, the average decay lifetime significantly increased in the core/shell
structure (57 ns) in comparison to core QDs (9.6 ns), but the QY was low as
4%. Further, the synthesis of CdTe/CdSe core/shell QDs was carried out without
using organometallic precursors [77]. The shell precursors in this method were
used as CdO and TOPSe in TOP, whereas for core QDs CdO and TOPTe. The
QY approaches up to 40% with growth of shell thickness of 0.5 nm. In a report
on the same material using SILAR technique, Chin et al. observed very high QY
up to 80% and the formation of the anisotropic structure (pyramids and multi-pods)
during shell growth [78]. Xia and co-authors carried out the aqueous synthesis of
CdTe/CdSe core/shell QDs, where CdCl2 and Na2 SeSO3 were used as the shell
precursor [79]. The prepared QDs had high stability, moderate QY up to 20%, and
tunable emission near-infrared region.
Further, different shells of CdTe, CdSe, and CdS were grown onto ZnTe core by
Xie et al., and they studied their properties [80]. The precursors were prepared by
dissolving cadmium oleate, TOPTe, TOPSe, and sulfur in octadecane. The obtained
precursors were further dissolved in crude ZnTe core QDs to form core/shell
QDs. The PL emission was spammed in the range of 500–900 nm with QY up
to 30%. Moreover, after lowering the shell growth temperature from 240 ◦ C to
215 ◦ C, the same group observed a transition from pyramidal to tetrapod using
ZnTe/CdSe core/shell QDs in another report [81]. Similar tradition is also described
by Alivisatos group in another paper [82].
Furthermore, Klimov group have developed a synthesis procedure for highly
luminescent type II CdS/ZnSe core/shell QDs [83]. For the synthesis of this system,
first of all, as-synthesized CdS QDs were added in the mixture of octadecylamine
and ODE. Further, ZnSe shell has been grown by adding precursors zinc oleate
and TOPSe in the as-prepared solution. The emission wavelength was tuned from
500 to 650 nm, with varying core radius and shell thickness. Moreover, Smith
and co-authors have tuned the optical and electronic properties of colloidal QDs
by lattice strain [39]. They deposited compressive shell of ZnS, ZnSe, ZnTe, CdS,
16 A. N. Yadav et al.

(a)
25 1.2
CdTe -350 nm

Fluorescence intensity (a.u.)


(CdTe)ZnSe, 0.5 ML 1.0
20 (CdTe)ZnSe, 1.0 ML
(CdTe)ZnSe, 3.0ML
Absorption (a.u.)

(CdTe)ZnSe, 6.0ML 0.8


15
0.6
10 0.4

5 0.2

0.0
0

300 400 500 600 700 800 900 500 600 700 800 900 1,000
Wavelength (nm) Wavelength (nm)
(b) (c)

Fig. 12 (a) Absorption and emission spectra of CdTe core QDs (size 1.8 nm) and CdTe/ZnSe
core/shell QDs with different shell thickness. (b) HRTEM and fast Fourier transform of CdTe core
QDs (top) and CdSe/ZnSe core/shell QDs with 6 mL of shell thickness (bottom). (c) HRTEM of
CdSe/ZnSe core/shell QDs with 6 mL of shell thickness. (Reprinted with permission from Ref.
[39])

and CdSe onto a CdTe core to form lattice-mismatched QDs. These materials were
synthesized by two-step organometallic method in a high-temperature coordinating
solvent, as described in previous studies. The band structure of obtained core/shell
QDs changes from type I to type II behavior, which was characterized by a large
spectral shift in absorption and emission spectra as well as an increment in excited-
state lifetimes (Fig. 12 a and b). In particular, the emission wavelength was tuned
from visible to near-infrared (500–1050 nm), and PL QY varied in between 25 and
60%.
Bang et al. pioneered a synthesis of nontoxic, water-soluble, and highly stable
ZnTe/ZnSe core/shell QDs [84]. The emission wavelength of the QDs was in the
range of 500–580 nm and PL QY was reached up to 60%. For the synthesis of
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 17

this core/shell structure, zinc and selenium precursors were prepared by dissolving
diethylzinc and selenium pellets in TOP, respectively. After that, ZnTe core QDs
were loaded in QDE, and HDA under N2 . Finally, these precursors were added
dropwise in bare CdTe QDs for 6 h, and final shell growth was achieved at a
temperature in between 200 and 250 ◦ C.

2.2.3 Synthesis and Characterization of Reverse Type I Core/Shell QDs

A reverse type I core/shell QD is also called “inverted” core/shell QDs because in


this case, the core material has a larger bandgap than the shell material. Peng group
developed a synthesis of inverted CdS/CdSe core/shell QD by SILAR technique
[85]. The emission wavelength was tuned from 520 to 650 nm and QY from 20% to
40% by growing another shell of CdS onto CdS/CdSe core/shell QDs. Later, Klimov
group described a continuous transformation of type I to type II and back to the type
I structure in ZnSe/CdSe core/shell QDs [86]. These transformations depend on the
core size and shell thickness, where electron and hole wave functions delocalized
in the distinct region of core/shell structure. The emission wavelength was varied in
the range of 430–600 nm with QY in between 60% and 80%. Moreover, the same
material is synthesized by using cadmium oxide (dissolved in oleic acid and ODE)
and TOPSe as the shell precursors [87]. The emission color was varied from violet
to red (417–678 nm) with QY up to 85% by changing CdSe shell thickness onto
2.8 nm ZnSe core (Fig. 13).

a b c d e f g h
Absorbance (a.u.)

PL intensity (a.u.)

h
g
f
e

d
c
b
a

400 500 600 700 800 400 500 600 700


Wavelength (nm) Wavelength (nm)

Fig. 13 Absorption and corresponding normalized emission spectra of ZnSe/CdSe core/shell QDs
with increasing monolayers of CdSe shell (a) 0, (b) 0.1, (c) 0.2, (d) 0.5, (e) 1, (f) 2, (g) 4, (h) 6.
(Reprinted with permission from Ref. [87])
18 A. N. Yadav et al.

3 Applications of Semiconductor Core/Shell QDs

The semiconductor core/shell QDs can be used in many applications because of


its excellent properties including tunability of bandgap by size, high PL QY with
narrow FWHM, longer excited-state lifetime, and robust stability. Here, we have
discussed some selected applications of QDs in solar cells, light-emitting diodes
(LEDs), and biomedical.

3.1 Solar Cells

In the beginning, II–VI (e.g., CdSe, etc.) and IV–VI (e.g., PbSe, etc.) QDs were
used as light absorber (sensitizer) in the solar cell [88, 89]. Now, the research
extends on core/shell QDs because of its high stability against photobleaching and
easy alternation of its basic properties including bandgap, charge separation, and
exciton recombination rate. All three types of core/shell QDs can be used in solar
cell applications. However, much better efficiency in the solar cell performance has
been found by using type II core/shell QDs as a sensitizer.
In typical QD-sensitized solar cell, there are a photoanode, liquid electrolytes,
and a counter electrode (Fig. 14a) [90]. The photoanode part consists of a light
absorber (QDs) and a large bandgap semiconductor crystalline (e.g., TiO2 or ZnO,
etc.) substituted on FTO (fluorine-doped tin oxide) or ITO (indium-doped tin oxide)
glass. Upon light excitation, electrons and holes are generated in the conduction
and valance band of the QD, respectively. Further, the photo-generated electrons are
quickly injected into the conduction band of the TiO2 (used as electrons acceptor).
The electrons transferred to the FTO substrate through TiO2 and then to counter
electrode through an external wire [90]. In the meantime, the oxidized QD was
neutralized by reduced species of redox couple in the electrolyte, whereas electrons
reduced the oxidized species of redox couple from the counter electrode. The circuit

Fig. 14 (a) Schematic diagram for device structure of QD-sensitized solar cells (QDSCs). (b) J-V
characteristic graph for QDSCs. (Reprinted with permission from Ref. [90])
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 19

is completed and the photocurrent is measured. The photovoltaic performance of


QD-solar cells can be determined from the J-V curve and IPCE test. The J-V graph
has been used to find power conversion efficiency (PCE) and is related through
power density of incident light (Pin ) by the equation as follows:

Imp .Vmp Jsc .Voc .FF


PCE = =
Pin Pin

where Imp is the photocurrent, Vmp is the photovoltage, Jsc is the short-circuit
density, Voc is the open-circuit voltage, and FF is the fill factor defined by the ratio
of specific value Pmax to Voc and Jsc (all factors can be found from Fig. 14b) [90].
In 2009, Lie et al. used first-time core/shell QDs as a sensitizer in solar cell
applications [91]. They observed PCE of 4.22% by using inverted type I CdS/CdSe
core/shell QDs. After that, core/shell QDs have been commonly used in the solar
cell. Further, Parkison et al. have used type I CdSe/ZnS core/shell QDs to sensitize
TiO2 , and, in this case, the stability of photoanode significantly improved [92].
Moreover, in the case of type I CdSeTe/ZnS core/shell QDs, the PCE of solar cells
has been found to be 9.48%, whereas bare CdSeTe QD shows a lower PCE of 8.02%
[93].
Z. Ning et al. used the first-time type II ZnSe/ZnS core/shell QDs as a sensitizer
in the solar cell [94]. In this case, they have found a very high photon-to-current
conversion efficiency due to the improvement of charge separation character. Using
type II CdTe/CdS QDs to sensitize TiO2 electrode, Yu et al. recorded PCE of
3.8% [95]. Q. Meng group used a microwave-assisted aqueous synthesis method
to prepare alloys CdSex Te1 − x /CdS core/shell QDs and recorded PCE of 5.04%
[96]. Furthermore, J. Wang et al. have shown a very high PCE value of 6.76%
using type II CdTe/CdSe core/shell QDs linked with the TiO2 mesoporous electrode.
The reason for the high PCE value was the suppression of the electron-hole
recombination rate and acceleration of electron injection from QD to TiO2 [97].
Moreover, Zhong et al. explored type II ZnTe/CdSe core/shell QDs as a sensitizer
and showed PCE of 7.17% (Fig. 15) [98].

3.2 Light-Emitting Diodes (LEDs)

In recent years, light-emitting diodes using QD (QLEDs) gain much attention of


scientists due to its high brightness (~200,000 cd m−2 ), high color purity (FWHM
<30 nm), low operating voltage (< 2 V), and easy process-ability [99–101]. The
QLEDs generally consist of a cathode, electron transport layers (ETLs), QD layers,
hole transport layers (HTLs), and an anode (Fig. 3a) [102]. When the potential
applied, electrons and holes are injected into charge transport layers (ETLs and
HTLs) from cathode and anode, respectively. Further, the charge carriers are injected
into the QDs where they recombine radiatively [102]. Thus, charge transport
materials play a primary role in the performance and stability of QLEDs.
20 A. N. Yadav et al.

Fig. 15 Photovoltaic performance of ZnTe/CdSe core/shell QDs, compared with CdTe/CdSe and
CdSe QDs: (a) J-V curve and (b) IPCE curve. (Reprinted with permission from Ref. [98])

Fig. 16 (a) Schematic illustration of device structure for QD light-emitting diodes (QLEDs). (b)
Energy-level diagram of a typical QLED displaying charge injection from the cathode and anode.
(Reprinted with permission from Ref. [102])

Usually, type I core/shell QDs have been used in QLEDs because wider bandgap
shell material confines the exciton to the core of the QDs and passivates the surface
defect states [41, 42, 47, 102–105]. This results in an enhancement in external
quantum efficiency (EQE) since it is directly related to PL QY and stability of
the QDs. Meanwhile, increment in PL QY does not mean improvement in the
electroluminescence (EL) performance because Augur recombination in charged
QD and interparticle energy transfer between different QDs can lead to decrease
in EL efficiency. These processes are directly affected by the structure of the
core/shell interface, and thus structure modification becomes an essential issue
[102]. For example, Yang et al. have observed better EL properties for the ZnSe-
rich intermediate shell in comparison to CdS-rich intermediate shell in the case
of CdSe/ZnS core/shell structure [106]. The results can be interpreted in terms
of a low carrier injection barrier of ZnSe-rich QDs in comparison to the CdS-rich
intermediate shell (Fig. 16).
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 21

Another way to make QDs with high stability and less blinking (or non-blinking)
is to grow a thicker shell or multi-shell over the surface of QDs [48, 107]. Thicker
shell in QDs can enhance PL QY of the charged QDs and suppressed the charge
fluctuation. The enhancement in these properties of thick core/shell QDs can lead
to a drastic improvement in device performance. Recently, Hao et al. have observed
excellent luminescence properties using a thick shell of ZnS over the surface of
CdSe [108]. The green-emitting thick CdSe/ZnS core/shell QDs were synthesized
by the TOP-SILAR method.

3.3 Applications in Biology

QDs have the upper hand over the organic fluorophores because of their size,
surface chemistry, spectral properties, large one- and two-photon absorption cross
sections, and stability, making it an excellent material to investigate for in vitro
and in vivo detection/imaging in biology. QDs with stokes shift, tunable emission
wavelengths, and high absorption coefficients can be created simply by altering
the size or composition without compromising the biocompatibility of the QDs.
Starting from the late 1990s, the QDs are playing an essential role in the field of
biology for biomolecular imaging, photodynamic therapy, sensing, gene and drug
delivery, and many more [109–111]. Selected applications of semiconductor QDs
in the biomedical field have been given below.

3.3.1 Biosensing

Semiconductor QDs are excellent material for their applications in sensing due to
their tunable bandgap along with their surface modification ability. It is essential
and beneficial to detect a specific biomarker for early-stage detection and therapy
of disease. Various biological and chemical species have been detected by surface
functionalization of QDs. Luminescence quenching of QDs is the primary technique
used for biosensing.
First of all, using CdSe/ZnS core/shell QDs coupled with biomolecules, Chan
and Nie developed a biosensor for the detection of ultrasensitive nonisotopic [109].
A sensor to investigate heme iron absorption Geng et al. used hemin-conjugated
CdSe/ZnS core/shell QDs as a probe in the biological setup [112]. Yang et al.
developed a sensitive biosensor for probing the interaction of clofazimine with
protein [113]. To make this sensor, they have developed CdZnSeS/ZnS alloyed
core/shell QDs as energy donors in Förster resonance energy transfer (FRET)
applications. They have used a direct ligand method for capping of multifunctional
polymer and further functionalized with cyanine 3-labelled human serum albumin
[113].
22 A. N. Yadav et al.

3.3.2 Gene and Drug Delivery

These days there is a growing trend of gene and drug delivery in the research field of
nanomedicine. Drug-formulated QDs can provide superiorities like a long lifetime,
targeted drug delivery, and enhanced cellular uptake. Various new nano-drugs and
nanoprobes have been designed for in vivo and in vitro study by using the versatility
of bioconjugated QDs [114].
Recently, Olerile et al. have used paclitaxel (PTX: a drug for various human
cancers), in CdTe/CdS/ZnS core/shell QDs along with nanostructured lipid carrier
for the cancer therapy [115]. This drug (PTX) along with hybrid silica nanocapsules
was further loaded with ZnSe:Mn/ZnS core/shell QDs by Gu group for chemother-
apy and fluorescence imaging [116]. In other work, Yang et al. prepared quercetin
(QE)-loaded CdSe/ZnS core/shell QDs as antibacterial and anticancer nano-drug
[117].

3.3.3 Therapy

QDs either separately or in combination with various known photosensitizers such


as porphyrins, phthalocyanines, organic dyes, inorganic complexes, etc. have been
investigated in photodynamic therapy (in vitro and in vivo both). Photosensi-
tized singlet oxygen and other reactive oxygen species (ROS) produced through
photosynthesis involving the transference of excitation energy is the basic step
involved in photodynamic therapy [118]. However, QDs are known to have poor
efficiency in producing singlet oxygen which hinders the active application of
QDs in photodynamic therapy. Later on, it was found that the conjugation of
standard photosensitizers like chlorine e6 with QDs results in an almost 70%
increase in efficiency. In this scenario, photoactivation of QDs provides an efficient
energy transfer to sensitized and followed by intersystem crossing efficiently and
consequently energy transfer to oxygen. Nevertheless, the scientific community
is still facing a major challenge which arises due to cytotoxicity of cadmium-
based QDs in developing QD-based in vivo imaging and photodynamic therapy.
Furthermore, various other issues particularly the surface charge, surface functional
groups, and the size decide the renal clearance of QDs implemented in vivo.

4 Summary and Prospects

This chapter mostly emphasizes how the chemical synthesis and its challenges with
a broad perspective have been developed in recent decades for II–VI semiconductor
core/shell QDs. Further, we intend to address its potential application in the various
research fields. The chemical-based synthesis and its ability to precisely control
shell thickness make highly luminescent semiconductor core/shell QDs for multiple
applications, including in solar cells, LEDs, biomedical, and many more. Due to the
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 23

enormous successes of semiconductor QDs in the recent past, it can be expected


the synthesis of a large number of new materials with excellent properties in the
next few years. In particular, to compete Cd-based QDs, the future research will
be focused on the synthesis of lead-free perovskite and I-III-VI ternary QDs. This
should lead to novel discoveries in biomedical applications such as biosensors,
targeted drug delivery, cell levelling, and bioimaging.

Acknowledgments Authors thank “Science and Engineering Research Board (SERB), Govern-
ment of India,” for financial assistance under the project number EEQ/2016/000652. ANY thanks
UGC, New Delhi, for junior research fellowship.

References

1. Henglein, A.: Small-particle research physicochemical properties of extremely small colloidal


metal and semiconductor particles. Chem. Rev. 89, 1861–1873 (1989)
2. Trindade, T., O’Brien, P., Pickett, N.L.: Nanocrystalline semiconductors: synthesis, proper-
ties, and perspectives. Chem. Mater. 13, 3843–3858 (2001)
3. Kuchibhatla, S., Karakoti, A.S., Bera, D., Seal, S.: One dimensional nanostructured materials.
Prog. Mater. Sci. 52, 699–913 (2007)
4. Bera, D., Kuiry, S.C., Seal, S.: Synthesis of nanostructured materials using template-assisted
electrodeposition. JOM. 56, 49–53 (2004)
5. Alivisatos, A.P.: Perspectives on the physical chemistry of semiconductor nanocrystals. J.
Phys.Chem. 100, 13226–13239 (1996)
6. Bera, D., Qian, L., Holloway, P.H.: Phosphor quantum dots. Wiley, West Sussex (2008)
7. Smith, A.M., Nie, S.: Semiconductor nanocrystals: structure, properties, and band gap
engineering. Acc. Chem. Res. 43, 190–200 (2010)
8. Yoffe, A.D.: Low-dimensional systems–quantum-size effects and electronic-properties of
semiconductor microcrystallites (zero-dimensional systems) and some quasi-2-dimensional
systems. Adv. Phys. 42, 173–266 (1993)
9. Yoffe, A.D.: Semiconductor quantum dots and related systems: electronic, optical, lumines-
cence and related properties of low dimensional systems. Adv. Phys. 50, 1–208 (2001)
10. Wang, Y., Herron, N.: Nanometer-sized semiconductor clusters–materials synthesis, quantum
size effects, and photophysical properties. J. Phys. Chem. 95, 525–532 (1991)
11. Bang, J., Yang, H., Holloway, P.H.: Enhanced and stable green emission of ZnO nanoparticles
by surface segregation of Mg. Nanotechnology. 17, 973–978 (2006)
12. Kucur, E., Bucking, W., Giernoth, R., Nann, T.: Determination of defect states in semicon-
ductor nanocrystals by cyclic voltammetry. J. Phys. Chem. B. 109, 20355–20360 (2005)
13. Bera, D., Qian, L., Tseng, T.K., Holloway, P.H.: Quantum dots and their multimodal
applications: a review. Mater. (Basel). 3, 2260–2345 (2010)
14. Colvin, V.L., Goldstein, A.N., Alivisatos, A.P.: Semiconductor nanocrystals covalently bound
to metal-surfaces with self-assembled monolayers. J. Am. Chem. Soc. 114, 5221–5230
(1992)
15. Dabbousi, B.O., Murray, C.B., Rubner, M.F., Bawendi, M.G.: Langmuir-Blodgett manipula-
tion of size-selected CdSe nanocrystallites. Chem. Mater. 6, 216–219 (1994)
16. Murray, C.B., Kagan, C.R., Bawendi, M.G.: Self-organization of CdSe nanocrystallites into
3-dimensional quantum-dot superlattices. Science. 270, 1335–1338 (1995)
17. Wang, Q., Kuo, Y.C., Wang, Y.W., Shin, G., Ruengruglikit, C., Huang, Q.R.: Luminescent
properties of water-soluble denatured bovine serum albumin-coated CdTe quantum dots. J.
Phys. Chem. B. 110, 16860–16866 (2006)
24 A. N. Yadav et al.

18. Ma, N., Yang, J., Stewart, K.M., Kelley, S.O.: DNA-passivated CdS nanocrystals: lumines-
cence, bioimaging, and toxicity profiles. Langmuir. 23, 12783–12787 (2007)
19. Reiss, P., Protiere, M., Li, L.: Core/Shell semiconductor nanocrystals. Small. 5, 154–168
(2009)
20. Chen, X.B., Lou, Y.B., Samia, A.C., Burda, C.: Coherency strain effects on the optical
response of core/shell heteronanostructures. Nano Lett. 3, 799–803 (2003)
21. Zhou, Y., Zhao, H., Ma, D., Rosei, F.: Harnessing the properties of colloidal quantum dots in
luminescent solar concentrators. Chem. Soc. Rev. 47, 5866–5890 (2018)
22. Farkhani, S.M., Valizadeh, A.: Review: three synthesis methods of CdX (X = Se, S or Te)
quantum dots. IET Nanobiotechnol. 8, 59–76 (2014)
23. Murray, C.B., Kagan, C.R., Bawendi, M.G.: Synthesis and characterization of monodisperse
nanocrystals and close-packed nanocrystal assemblies. Annu. Rev. Mater. Sci. 30, 545–610
(2000)
24. Peng, X.: Band gap and composition engineering on a nanocrystal (BCEN) in solution. Acc.
Chem. Res. 43, 1387–1395 (2010)
25. Murray, C.B., Norris, D.J., Bawendi, M.G.: Synthesis and characterization of nearly Monodis-
perse CdE (E = sulfur, selenium, tellurium) semiconductor Nanocrystallites. J. Am. Chem.
Soc. 115, 8706–8715 (1993)
26. Peng, Z.A., Peng, X.: Formation of high-quality CdTe, CdSe, and CdS nanocrystals using
CdO as precursor. J. Am. Chem. Soc. 123, 183–184 (2001)
27. Qu, L.H., Peng, Z.A., Peng, X.: Alternative routes toward high-quality CdSe nanocrystals.
Nano Lett. 1, 333–337 (2001)
28. Peng, X.: Green chemical approaches toward high-quality semiconductor nanocrystals.
Chem. Eur. J. 8, 334–339 (2002)
29. Ghorpade, U., Suryawanshi, M., Shin, S.W., Gurav, K., Patil, P., Pawar, S., Hong, C.W., Kim,
J.H., Kolekar, S.: Towards environmentally benign approaches for the synthesis of CZTSSe
nanocrystals by a hot injection method: a status review. Chem. Commun. 50, 11258–11273
(2014)
30. Yu, W.W., Peng, X.: Formation of high-quality CdS and other II–VI semiconductor nanocrys-
tals in noncoordinating solvents: tunable reactivity of monomers. Angew. Chem. Int. Ed. 41,
2368–2371 (2002)
31. Bullen, C.R., Mulvaney, P.: Nucleation and growth kinetics of CdSe nanocrystals in
Octadecene. Nano Lett. 4, 2303–2307 (2004)
32. Abe, S., Capek, R.K., De Geyter, B., Hens, Z.: Reaction chemistry/nanocrystal property
relations in the hot injection synthesis, the role of the solute solubility. ACS Nano. 7, 943–949
(2013)
33. Deng, Z., Cao, L., Tang, F., Zou, B.: A new route to zinc-blende CdSe nanocrystals:
mechanism and synthesis. J. Phys. Chem. B. 109, 16671–16675 (2005)
34. Cao, Y.C., Wang, J.: One-pot synthesis of high-quality zinc-blende CdS nanocrystals. J. Am.
Chem. Soc. 126, 14336–14337 (2004)
35. Protière, M., Nerambourg, N., Renard, O., Reiss, P.: Rational design of the gram-scale
synthesis of nearly monodisperse semiconductor nanocrystals. Nanoscale Res. Lett. 6, 472
(2011)
36. Qian, H., Qiu, X., Li, L., Ren, J.: Microwave-assisted aqueous synthesis: a rapid approach to
prepare highly luminescent ZnSe(S) alloyed quantum dots. J. Phys. Chem. B. 110, 9034–9040
(2006)
37. Pradhan, N., Efrima, S.: Single-precursor, one-pot versatile synthesis under near ambient
conditions of Tunable, single and dual band fluorescing metal Sulfide nanoparticles. J. Am.
Chem. Soc. 125, 2050–2051 (2003)
38. Cumberland, S.L., Hanif, K.M., Javier, A., Khitrov, G.A., Strouse, G.F., Woessner, S.M.,
Yun, C.S.: Inorganic clusters as single-source precursors for preparation of CdSe, ZnSe, and
CdSe/ZnS nanomaterials. Chem. Mater. 14, 1576–1584 (2002)
39. Smith, A.M., Mohs, A.M., Nie, S.: Tuning the optical and electronic properties of colloidal
nanocrystals by lattice strain. Nat. Nanotechnol. 4, 56–63 (2009)
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 25

40. Li, J.J., Wang, Y.A., Guo, W.Z., Keay, J.C., Mishima, T.D., Johnson, M.B., Peng, X.G.: Large-
scale synthesis of nearly Monodisperse CdSe/CdS Core/Shell nanocrystals using air-stable
reagents via successive ion layer adsorption and reaction. J. Am. Chem. Soc. 125, 12567–
12575 (2003)
41. Hines, M.A., Guyot-Sionnest, P.: Synthesis and characterization of strongly luminescing ZnS-
capped CdSe nanocrystals. J. Phys. Chem. 100, 468–471 (1996)
42. Dabbousi, B.O., Rodriguez-Viejo, J., Mikulec, F.V., Heine, J.R., Mattoussi, H., Ober, R.,
Jensen, K.F., Bawendi, M.G.: (CdSe)ZnS Core−Shell quantum dots: synthesis and char-
acterization of a size series of highly luminescent nanocrystallites. J. Phys. Chem. B. 101,
9463–9475 (1997)
43. Talapin, D.V., Rogach, A.L., Kornowski, A., Haase, M., Weller, H.: Highly lumines-
cent monodisperse CdSe and CdSe/ZnS nanocrystals synthesized in hexadecylamine -
trioctylphosphine oxide−trioctylphosphine mixture. Nano Lett. 1, 207–211 (2001)
44. Kudera, S., Zanella, M., Giannini, C., Rizzo, A., Li, Y.Q., Gigli, G., Cingolani, R., Ciccarella,
G., Spahl, W., Parak, W.J., Manna, L.: Sequential growth of magic-size CdSe nanocrystals.
Adv. Mater. 19, 548–551 (2007)
45. Jun, S., Jang, E.: Interfused semiconductor nanocrystals: brilliant blue photoluminescence
and electroluminescence. Chem. Commun. 36, 4616–4618 (2005)
46. Hao, J.J., Zhoua, J., Zhang, C.Y.: A tri-n-octylphosphine-assisted successive ionic layer
adsorption and reaction method to synthesize multilayered core–shell CdSe–ZnS quantum
dots with extremely high quantum yield. Chem. Commun. 49, 6346–6348 (2013)
47. Peng, X.G., Schlamp, M.C., Kadavanich, A.V., Alivisatos, A.P.: Epitaxial growth of highly
luminescent CdSe/CdS Core/Shell nanocrystals with photostability and electronic accessibil-
ity. J. Am. Chem. Soc. 119, 7019–7029 (1997)
48. Chen, Y., Vela, J., Htoon, H., Casson, J.L., Werder, D.J., Bussian, D.A., Klimov, V.I.,
Hollingsworth, J.A.: “Giant” multishell CdSe nanocrystal quantum dots with suppressed
blinking. J. Am. Chem. Soc. 130, 5026–5027 (2008)
49. Christodoulou, S., Vaccaro, G., Pinchetti, V., Donato, F.D., Grim, J.Q., Casu, A., Genovese,
A., Vicidomini, G., Diaspro, A., Brovelli, S., Mannaa, L., Moreels, I.: Synthesis of highly
luminescent wurtzite CdSe/CdS giant-shell nanocrystals using a fast continuous injection
route. J. Mater. Chem. C. 2, 3439–3447 (2014)
50. Chen, O., Zhao, J., Chauhan, V.P., Cui, J., Wong, C., Harris, D.K., Wei, H., Han, H.S.,
Fukumura, D., Jain, R.K., Bawendi, M.G.: Compact high-quality CdSe–CdS core–shell
nanocrystals with narrow emission linewidths and suppressed blinking. Nature Mater. 12,
445–451 (2013)
51. Danek, M., Jensen, K.F., Murray, C.B., Bawendi, M.G.: Synthesis of luminescent thin-film
CdSe/ZnSe quantum dot composites using CdSe quantum dots passivated with an overlayer
of ZnSe. Chem. Mater. 8, 173–180 (1996)
52. Reiss, P., Bleuse, J., Pron, A.: Highly luminescent CdSe/ZnSe Core/Shell nanocrystals of low
size dispersion. Nano Lett. 2, 781–784 (2002)
53. Reiss, P., Carayon, S., Bleuse, J.: Large fluorescence quantum yield and low size dispersion
from CdSe/ZnSe core/shell nanocrystals. Phys. E. 17, 95–96 (2003)
54. Lee, Y.J., Kim, T.G., Sung, Y.M.: Lattice distortion and luminescence of CdSe/ZnSe
nanocrystals. Nanotechnology. 17, 3539–3542 (2006)
55. Steckel, J.S., Snee, P., Coe-Sullivan, S., Zimmer, J.R., Halpert, J.E., Anikeeva, P., Kim, L.A.,
Bulovic, V., Bawendi, M.G.: Color-saturated green-emitting QD-LEDs. Angew. Chem. Int.
Ed. 45, 5796–5799 (2006)
56. Protiere, M., Reiss, P.: Highly luminescent Cd1−x Znx Se/ZnS Core/Shell nanocrystals emit-
ting in the blue–green spectral range. Small. 3, 399–403 (2007)
57. Shen, H., Bai, X., Wang, A., Wang, H., Qian, L., Yang, Y., Titov, A., Hyvonen, J.,
Zheng, Y., Li, L.S.: High-efficient deep-blue light-emitting diodes by using high quality Znx
Cd1− x S/ZnS Core/Shell quantum dots. Adv. Funct. Mater. 24, 2367–2373 (2014)
26 A. N. Yadav et al.

58. Li, Z., Chen, F., Wang, L., Shen, H., Guo, L., Kuang, Y., Wang, H., Li, N., Li, L.S.: Synthesis
and evaluation of ideal Core/Shell quantum dots with precisely controlled Shell growth:
nonblinking, single photoluminescence Decay channel, and suppressed FRET. Chem. Mater.
30, 3668–3676 (2018)
59. Spanhel, L., Haase, M., Weller, H., Henglein, A.: Photochemistry of colloidal semiconduc-
tors. 20. Surface modification and stability of strong luminescing CdS particles. J. Am. Chem.
Soc. 109, 5649–5655 (1987)
60. Steckel, J.S., Zimmer, J.P., Coe-Sullivan, S., Stott, N.E., Bulovic, V., Bawendi, M.G.:
Blue luminescence from (CdS)ZnS Core–Shell nanocrystals. Angew. Chem. Int. Ed. 43,
2154–2158 (2004)
61. Protiere, M., Reiss, P.: Facile synthesis of monodisperse ZnS capped CdS nanocrystals
exhibiting efficient blue emission. Nanoscale Res. Lett. 1, 62–67 (2006)
62. Yang, H., Holloway, P.H.: Efficient and photostable ZnS-passivated CdS:Mn luminescent
nanocrystals. Adv. Funct. Mater. 14, 152–156 (2004)
63. Yang, Y.A., Chen, O., Angerhofer, A., Cao, Y.C.: Radial-position-controlled doping in
CdS/ZnS Core/Shell nanocrystals. J. Am. Chem. Soc. 128, 12428–12429 (2006)
64. Hofman, E., Robinson, R.J., Li, Z.-J., Dzikovski, B., Zheng, W.: Controlled dopant migration
in CdS/ZnS core/shell quantum dots. J. Am. Chem. Soc. 139(26), 8878–8885 (2017)
65. Chen, D., Zhao, F., Qi, H., Rutherford, M., Peng, X.: Bright and stable purple/blue emitting
CdS/ZnS Core/Shell nanocrystals grown by thermal cycling using a single-source precursor.
Chem. Mater. 22, 1437–1444 (2010)
66. Lomascolo, M., Creti, A., Leo, G., Vasanelli, L., Manna, L.: Exciton relaxation processes in
colloidal core/shell ZnSe/ZnS nanocrystals. Appl. Phys. Lett. 82, 418–420 (2003)
67. Chen, H.S., Lo, B., Hwang, J.Y., Chang, G.Y., Chen, C.M., Tasi, S.J., Wang, S.J.J.: Colloidal
ZnSe, ZnSe/ZnS, and ZnSe/ZnSeS quantum dots synthesized from ZnO. J. Phys. Chem. B.
108, 17119–17123 (2004)
68. Dong, B., Cao, L., Sua, G., Liu, W.: Facile synthesis of highly luminescent UV-blue emitting
ZnSe/ZnS core/shell quantum dots by a two-step method. Chem. Commun. 46, 7331–7333
(2010)
69. Shen, H., Wang, H., Li, X., Niu, J.Z., Wang, H., Chen, X., Li, L.S.: Phosphine-free synthesis
of high quality ZnSe, ZnSe/ZnS, and Cu-, Mn-doped ZnSe nanocrystals. Dalton Trans. 47,
10534–10540 (2009)
70. Wang, A., Shen, H., Zang, S., Lin, Q., Wang, H., Qian, L., Niu, J., Li, L.S.: Bright,
efficient, and color-stable violet ZnSe-based quantum dot light-emitting diodes. Nanoscale.
7, 2951–2959 (2015)
71. Tsay, J.M., Pflughoefft, M., Bentolila, L.A., Weiss, S.: Hybrid approach to the synthesis of
highly luminescent CdTe/ZnS and CdHgTe/ZnS nanocrystals J. Am. Chem. Soc. 126, 1926–
1927 (2004)
72. Bao, H.B., Gong, Y.J., Li, Z., Gao, M.Y.: Enhancement effect of illumination on the
photoluminescence of water-soluble CdTe nanocrystals: toward highly fluorescent CdTe/CdS
Core−Shell structure. Chem. Mater. 16, 3853–3859 (2004)
73. He, Y., Lu, H.T., Sai, L.M., Lai, W.Y., Fan, Q.L., Wang, L.H., Huang, W.: Microwave-assisted
growth and characterization of water-dispersed CdTe/CdS Core−Shell nanocrystals with high
photoluminescence. J. Phys. Chem. B. 110, 13370–13374 (2006)
74. Wang, C.L., Zhang, H., Zhang, J.H., Li, M.J., Sun, H.Z., Yang, B.: Application of ultrasonic
irradiation in aqueous synthesis of highly fluorescent CdTe/CdS Core−Shell nanocrystals. J.
Phys. Chem. C. 111, 2465–2469 (2007)
75. Gu, Z., Zou, L., Fang, Z., Zhu, W., Zhong, X.: One-pot synthesis of highly luminescent
CdTe/CdS core/shell nanocrystals in aqueous phase. Nanotechnology. 19, 135604 (2008).
(pp 7)
76. Kim, S., Fisher, B., Eisler, H.J., Bawendi, M.: Type-II quantum dots: CdTe/CdSe(Core/Shell)
and CdSe/ZnTe(Core/Shell) heterostructures. J. Am. Chem. Soc. 125, 11466–11467 (2003)
77. Yu, K., Zaman, B., Romanova, S., Wang, D.S., Ripmeester, J.A.: Sequential synthesis of type
II colloidal CdTe/CdSe Core–Shell nanocrystals. Small. 1, 332–338 (2005)
Synthesis, Properties, and Applications of II–VI Semiconductor Core/Shell. . . 27

78. Chin, P.T.K., Donega, C.D.M., Bavel, S.S., Meskers, S.C.J., Sommerdijk, N., Janssen, R.A.J.:
Highly luminescent CdTe/CdSe colloidal Heteronanocrystals with temperature-dependent
emission color. J. Am. Chem. Soc. 129, 14880–14886 (2007)
79. Xia, Y., Zhu, C.: Aqueous synthesis of type-II core/shell CdTe/CdSe quantum dots for near-
infrared fluorescent sensing of copper (II). Analyst. 133, 928–932 (2008)
80. Xie, R.G., Zhong, X.H., Basché, T.: Synthesis, characterization, and spectroscopy of type-II
Core/Shell semiconductor nanocrystals with ZnTe Cores. Adv. Mater. 17, 2741–2744 (2005)
81. Xie, R.G., Kolb, U., Basché, T.: Design and synthesis of colloidal nanocrystal heterostructures
with tetrapod morphology. Small. 2, 1454–1457 (2006)
82. Milliron, D.J., Hughes, S.M., Cui, Y., Manna, L., Li, J.B., Wang, L.W., Alivisatos, A.P.:
Colloidal nanocrystal heterostructures with linear and branched topology. Nature. 430, 190–
195 (2004)
83. Ivanov, S.A., Piryatinski, A., Nanda, J., Tretiak, S., Zavadil, K.R., Wallace, W.O., Werder, D.,
Klimov, V.I.: Type-II Core/Shell CdS/ZnSe nanocrystals: synthesis, electronic structures, and
spectroscopic properties. J. Am. Chem. Soc. 129, 11708–11719 (2007)
84. Bang, J., Park, J., Lee, J.H., Won, N., Nam, J., Lim, J., Chang, B.Y., Lee, H.J., Chon, B.,
Shin, J., Park, J.B., Choi, J.H., Cho, K., Park, S.M., Joo, T., Kim, S.: ZnTe/ZnSe (Core/Shell)
type-II quantum dots: their optical and photovoltaic properties. Chem. Mater. 22, 233–240
(2010)
85. Battaglia, D., Li, J.J., Wang, Y.J., Peng, X.G.: Colloidal two-dimensional systems: CdSe
quantum shells and wells. Angew. Chem. Int. Ed. 42, 5035–5039 (2003)
86. Balet, L.P., Ivanov, S.A., Piryatinski, A., Achermann, M., Klimov, V.I.: Inverted Core/Shell
nanocrystals continuously tunable between type-I and type-II localization regimes. Nano Lett.
4, 1485–1488 (2004)
87. Zhong, X.H., Xie, R.G., Zhang, Y., Basché, T., Knoll, W.: High-quality violet- to red-emitting
ZnSe/CdSe Core/Shell nanocrystals. Chem. Mater. 17, 4038–4042 (2005)
88. Klimov, V.I.: Mechanisms for Photogeneration and recombination of multiexcitons in semi-
conductor nanocrystals: implications for lasing and solar energy conversion. J. Phys. Chem.
B. 110, 16827–16845 (2006)
89. Kamat, P.V.: Quantum dot solar cells. Semiconductor nanocrystals as light harvesters. J. Phys.
Chem. C. 112, 18737–18753 (2008)
90. Pan, Z., Rao, H., Mora-Seró, I., Bisquert, J., Zhong, X.: Quantum dot-sensitized solar cells.
Chem. Soc. Rev. 47, 7659–7702 (2018)
91. Lee, Y.-L., Lo, Y.-S.: Highly efficient quantum-dot-sensitized solar cell based on co-
sensitization of CdS/CdSe. Adv. Funct. Mater. 19, 604–609 (2009)
92. Sambur, J.B., Parkinson, B.A.: CdSe/ZnS Core/Shell quantum dot sensitization of low index
TiO2 single crystal surfaces. J. Am. Chem. Soc. 132, 2130–2131 (2010)
93. Yang, J., Wang, J., Zhao, K., Izuishi, T., Li, Y., Shen, Q., Zhong, X.: CdSeTe/CdS type-I
Core/Shell quantum dot sensitized solar cells with efficiency over 9%. J. Phys. Chem. C. 119,
28800–28808 (2015)
94. Ning, Z., Tian, H., Yuan, C., Fu, Y., Qin, H., Sun, L., Agren, H.: Solar cells sensitized with
type-II ZnSe–CdS core/shell colloidal quantum dots. Chem. Commun. 47, 1536–1538 (2011)
95. Yu, X.-Y., Lei, B.-X., Kuang, D.-B., Su, C.-Y.: Highly efficient CdTe/CdS quantum dot
sensitized solar cells fabricated by a one-step linker assisted chemical bath deposition. Chem.
Sci. 2, 1396–1400 (2011)
96. Luo, J., Wei, H., Li, F., Huang, Q., Li, D., Luo, Y., Meng, Q.: Microwave assisted aqueous
synthesis of core–shell CdSex Te1−x –CdS quantum dots for high performance sensitized solar
cells. Chem. Commun. 50, 3464–3466 (2014)
97. Wang, J., Mora-Sero, I., Pan, Z., Zhao, K., Zhang, H., Feng, Y., Yang, G., Zhong, X., Bisquert,
J.: Core/Shell colloidal quantum dot Exciplex states for the development of highly efficient
quantum-dot-sensitized solar cells. J. Am. Chem. Soc. 135, 15913–15922 (2013)
98. Jiao, S., Shen, Q., Mora-Sero, I., Wang, J., Pan, Z., Zhao, K., Kuga, Y., Zhong, X., Bisquert,
J.: Band engineering in Core/Shell ZnTe/CdSe for Photovoltage and efficiency enhancement
in Exciplex quantum dot sensitized solar cells. ACS Nano. 9, 908–915 (2015)
28 A. N. Yadav et al.

99. Yang, J., Choi, M.K., Kim, D.-H., Hyeon, T.: Designed assembly and integration of colloidal
nanocrystals for device applications. Adv. Mater. 28, 1176–1207 (2016)
100. Dai, X., Deng, Y., Peng, X., Jin, Y.: Quantum-dot light-emitting diodes for large area displays:
towards the dawn of commercialization. Adv. Mater. 29, 1607022 (2017)
101. Pimputkar, S., Speck, J.S., Denbaars, S.P., Nakamura, S.: Prospects for LED lighting. Nat.
Photon. 3, 180–182 (2009)
102. Choi, M.K., Yang, J., Hyeon, T., Kim, D.H.: Flexible quantum dot light-emitting diodes for
next-generation displays. npj Flexible Electronics. 2, 10 (2018)
103. Bae, W.K., Char, K., Hur, H., Lee, S.: Single-step synthesis of quantum dots with chemical
composition gradients. Chem. Mater. 20, 531–539 (2008)
104. Lee, J., Yang, J., Kwon, S.G., Hyeon, T.: Nonclassical nucleation and growth of inorganic
nanoparticles. Nat. Rev. Mater. 1, 16034 (2016)
105. Owen, J., Brus, L.: Chemical synthesis and luminescence applications of colloidal semicon-
ductor quantum dots. J. Am. Chem. Soc. 139, 10939–10943 (2017)
106. Yang, Y., Zheng, Y., Cao, W., Titov, A., Hyvonen, J., Manders, J.R., Xue, J., Holloway,
P.H., Qian, L.: High-efficiency light-emitting devices based on quantum dots with tailored
nanostructures. Nat. Photon. 9, 259–266 (2015)
107. Galland, C., Ghosh, Y., Steinbrück, A., Hollingsworth, J.A., Htoon, H., Klimov, V.I.: Lifetime
blinking in nonblinking nanocrystal quantum dots. Nat. Commun. 3, 908 (2012)
108. Hao, J., Liu, H., Miao, J., Lu, R., Zhou, Z., Zhao, B., Xie, B., Cheng, J., Wang, K.,
Delville, M.H.: A facile route to synthesize CdSe/ZnS thick-shell quantum dots with precisely
controlled green emission properties: towards QDs based LED applications. Sci. Rep. 9,
12048 (2019)
109. Chan, W.C.W., Nie, S.: Quantum dot bioconjugates for ultrasensitive nonisotopic detection.
Science. 281, 2016–2018 (1998)
110. Gao, X., Cui, Y., Levenson, R.M., Chung, L.W.K., Nie, S.: In vivo cancer targeting and
imaging with semiconductor quantum dots. Nat. Biotechnol. 22, 969 (2004)
111. Martynenko, I.V., Litvin, A.P., Purcell-Milton, F., Baranov, A.V., Fedorov, A.V., Gun’ko,
Y.K.: Application of semiconductor quantum dots in bioimaging and biosensing. J. Mater.
Chem. B. 5, 6701–6727 (2017)
112. Geng, L., Duan, X., Wang, Y., Zhao, Y., Gao, G., Liu, D., Chang, Y.Z., Yu, P.: Quantum
dots-hemin: preparation and application in the absorption of heme iron. Nanomedicine. 12,
1747–1755 (2016)
113. Yang, H.Y., Fu, Y., Jang, M.S., Li, Y., Lee, J.H., Chae, H., Lee, D.S.: Multifunctional polymer
ligand interface CdZnSeS/ZnS quantum dot/Cy3-labeled protein pairs as sensitive FRET
sensors. ACS Appl. Mater. Interfaces. 8, 35021–35032 (2016)
114. Matea, C.T., Mocan, T., Tabaran, F., Pop, T., Mosteanu, O., Puia, C., Iancu, C., Mocan, L.:
Quantum dots in imaging, drug delivery and sensor applications. Int. J. Nanomedicine. 12,
5421–5431 (2017)
115. Olerile, L.D., Liu, Y., Zhang, B., Wang, T., Mu, S., Zhang, J., Selotlegeng, L., Zhang, N.:
Near-infrared mediated quantum dots and paclitaxel co-loaded nanostructured lipid carriers
for cancer theragnostic. Colloids Surf. B Biointerfaces. 150, 121–130 (2017)
116. Zhao, T., Liu, X., Li, Y., Zhang, M., He, J., Zhang, X., Liu, H., Wang, X., Gu, H.: Fluorescence
and drug loading proprieties of ZnSe:Mn/ZnS-paclitaxel/SiO2 nanocapsules templated by
F127 micelles. J. Colloid Interface Sci. 490, 436–443 (2017)
117. Yang, X., Zhang, W., Zhao, Z., Li, N., Mou, Z., Sun, D., Cai, Y., Wang, W., Lin, Y.: Quercetin
loading CdSe/ZnS nanoparticles as efficient antibacterial and anticancer materials. J. Inorg.
Biochem. 167, 36–48 (2017)
118. Biju, V., Mundayoor, S., Omkumar, R.V., Anas, A., Ishikawa, M.: Bioconjugated quantum
dots for cancer research: present status, prospects and remaining issues. Biotechnol. Adv. 28,
199–213 (2010)
Design, Synthesis, and Properties
of I-III-VI2 Chalcogenide-Based
Core-Multishell Nanocrystals

V. Renuga and C. Neela Mohan

Abstract This chapter describes the design and properties of ternary CuInS2 - and
AgInS2 -based core/shell and core-multishell nanocrystals (QDs – quantum dots)
considering these ternary materials as core. The focus of this chapter is to deal with
the utilization of the full potential of these ternary chalcogenides as an alternative to
the existing toxic nanocrystals systems and further improve their optical properties
as well as intensity by architecting the core/shell structure. Hot-injection method
plays a vital role to architect such highly crystalline nanocrystals with controllability
of size. To enhance the luminescence property of these core materials, they are either
passivated by wider-bandgap ZnS shell or doped with highly luminescent Mn2+
and Cd2+ ions or combination of these two effects on the surface of core materials.
The influence of dopant as shell materials (MnS and ZnS) are also analyzed by
architecting core-multishell nanocrystals. The crystal structure, optical properties,
and morphologies of the core, core/shell, and core-multishell nanocrystals are
analyzed and described in detail in this chapter.

Keywords Core/shell · Ternary chalcogenides · Nanocrystals ·


Core-multishell · Luminescence

1 Introduction

The most attractive and fascinating properties of nanoclusters are mainly concealed
in their inherent properties such as finite size, large surface-to-volume ratio, and
quantum effects [1]. These highly stable and robust natured materials open a new
avenue to create the possibilities for tailoring their properties. Therefore they rule
the world by providing new building blocks for variety of materials to design many
new and efficient devices.

V. Renuga () · C. Neela Mohan


PG & Research Department of Chemistry, National College (Autonomous), Tiruchirappalli,
Tamil Nadu, India

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 29
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_2
30 V. Renuga and C. Neela Mohan

The two strategies involved in cluster science are to understand the structure-
property relationship of size-dependent nanoclusters and to explore these
nanocluster-assembled materials to practical applications such as novel electronic
and optical devices, chemical sensors, and efficient and selective catalysts [2, 3].
Core/shell nanoclusters have recently received considerable attention owing to
their properties (optical property, fluorescence wavelength, quantum yield, and
lifetime) [4] that are strongly dependent on the structure of the core, shell, and
interface, so that they can control their chemical composition and relative size of
the core and shell. The efficient passivation of shell materials helps to reduce the
surface trap states of core and also enhance the fluorescence quantum yield.
This chapter briefly discusses the fundamental design strategies of core/shell
architecture, the general considerations for the choice of materials and reaction
parameters, and the synthesis methods of core/shell and core-multishell structure
of binary and ternary Cu- and Ag-based chalcogenide nanocrystals.
These colloidal nanocrystals also termed as “quantum dots” are mostly composed
of inorganic core that contains few hundred to few thousand atoms surrounded by
either an organic outer layer of surfactant molecules or inorganic material.

1.1 Classes of Core/Shell Nanocrystals

The relative position of electronic energy levels and their corresponding bandgaps
[5, 6] decide the functions of core/shell systems. Thickness of the shell materials
are distinguished by three cases of core/shell particle as type I, reverse type I, and
type II. Accordingly, it transferred from type I (both the charge carriers electron and
hole wave functions are distributed over the entire core) to type II (electron and hole
are spatially separated between the shell and the core, but the highest probability of
hole resides in core than in shell) and back to reverse type I (both electron and hole
primarily reside in the shell).
The core/shell type consists of wide range of different combinations such
as inorganic/inorganic, inorganic/organic, organic/inorganic, and organic/organic
materials. Among the different groups, inorganic/inorganic core/shells are the most
important of all the different types of core/shell particles. Here both core and shell
are made of metal, metal oxide, other inorganic compounds, or silica.
Apart from the combinations and classifications of semiconductor materials, their
efficiency is decided mainly by quantum yield (Qy) and the response time [7, 8]. A
shell with higher bandgap and semiconductor materials fulfills these possibilities.
The type I and reverse type I systems have low quantum yield, and type II has
low photooxidation stabilities, and these limitations can be improved by forming
a multilayer core/shell/shell (CSS) architecture. The corresponding energy levels
of valence bond (VB) and conduction bond (CB) for core/shell [9] semiconductor
nanocrystals are presented in Fig. 1.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 31

Fig. 1 Types of core/shell QDs based on band alignment: (a) type I, (b) reverse type I, and (c)
type II core/shell semiconductor nanocrystals

In order to form better CSS semiconductor particle as in the case of core/shell


system, bandgap and the lattice structure [10] of the materials involved are
important, and these two factors play a vital role in designing core/shell/shell system
successfully.
The most versatile and alternative semiconductor materials for the existing toxic
Cd and Pb-based materials are binary chalcogenides. For the last two decades, the Pb
and Cd govern the semiconductor society because of their luminescence and higher
light-emitting properties. But these materials’ lower bandgap and lattice structure
are not suitable for most of CS and CSS nanocrystal formation. Moreover its heavy
toxic nature could not withstand for long time.
During recent times considerable attention has been paid over Cu- and Ag-based
ternary I-III-VI2 chalcogenides (e.g., CuInS2 , CuInSe2 , AgInS2 , and AgInSe2 ), and
their alloy system also does greater contribution to the semiconductor field due to
their potential ability by tuning their size and to extend its possibilities as a best
photocatalyst and as an active component in light-emitting and solar cell devices.
Among Cu- and Ag-based chalcogenides, Ag-based systems are less exposed
despite their remarkable optical properties compared to Cu. These Cu- and Ag-
based semiconductor nanocrystals have many fascinating properties such as tunable
absorption and emission, high-efficiency interface charge separation, spatial sepa-
rations, multiexcitons generation, and easy chemical modifications and processing
which make these semiconductor nanocrystals suitable for next-generation photo-
voltaic applications.
The main content of this chapter deals with copper- and silver-based core/shell
and core-multishell nanocrystals. In the case of copper, CuInS2 and for silver
AgInS2 were considered as core and ZnS as shell material. The dopants used in
each of the above cases are Mn2+ ions and Cd2+ ions for CuInS2 and AgInS2
core, respectively. These dopants were doped in each of the above materials
individually and in combined manner as in both core and shell. In all the above cases
(both undoped and doped), type I core/shell architecture is formed. To form core-
multishell nanocrystals, the dopants were also considered as shell materials such
as Mns and CdS and form CuInS2 /MnS/ZnS (type I/type I) and AgInS2 /CdS/ZnS
(type II/type I) multishell nanocrystals. The influence of shell and dopant on optical,
32 V. Renuga and C. Neela Mohan

Scheme 1 Schematic two-dimensional diagram of core-multishell of CuInS2 nanocrystals

crystal structure, phase, and morphology on the core materials CuInS2 and AgInS2
in single core/shell and core-multishell nanocrystals were discussed in detailed
manner.

2 Design and Formation of Core-Multishell and Alloyed


Shell Nanocrystals

At the outset, the design and formation of core/shell and core-multishell


(CuInS2 /MnS/ZnS and AgInS2 /CdS/ZnS or AgInS2 /ZnS/CdS) architecture used
in the present case are discussed briefly herein, with a schematic diagram
(Schemes 1 and 2). Schemes 1 and 2 show the schematic two-dimensional
diagram of core-multishell architecture (CuInS2 /MnS/ZnS and AgInS2 /CdS/ZnS or
AgInS2 /ZnS/CdS) with the bulk bandgap values, successive lattice mismatches, and
valence band maximum (VBM) and conduction band minimum (CBM) values in a
systematic manner.
The lattice mismatch between the core and shell nanocrystals is calculated from
the corresponding d(hkl) using Eq. (1):

% lattice mismatch = 100 × (d (hkl) host − d (hkl) shell)/d(hkl) host (1)

The purpose of epitaxial growth of outer ZnS shell over single core/shell system
has many advantages such as it:
(i) Stabilizes the inner shell
(ii) Protects from oxidization of inner shell against atmospheric conditions
(iii) Increases the quantum yield and optical stability
(iv) Reduces the toxicity
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 33

Scheme 2 Schematic two-dimensional diagram of core-multishell of AgInS2 nanocrystals

2.1 Synthesis of Core/Shell Nanocrystals

In general, core/shell nanoparticles are synthesized using a two-step process, first


synthesis of core and second growth of the shell over the core. In the present
case, both CuInS2 - and AgInS2 -based core/shell, doped core/shell, and multishell
nanocrystals were synthesized by hot-injection method. Here the outline of the
synthesis of core, metal ion-doped core, growth of shell over core and doped core
and also the growth of multishell is discussed. The detailed synthesis was discussed
in our paper [11].
In a typical process, metal precursors with 1-octadecene as a non-coordinating
solvent were loaded into a three-necked flask and degassed at 100 ◦ C for 30 min
by bubbling with nitrogen to remove the oxygen and water. The above mixture was
heated up to 150 ◦ C, and 0.3 mmol of sulfur dissolved in 2 ml oleylamine as a
capping agent was injected rapidly into the flask. Then the temperature was raised
to 230 ◦ C and maintained there for 30 min to allow the growth of core nanocrystals.
It is then cooled and purified with chloroform/acetone co-solvents and centrifuged
(4000 rpm, 10 min).
For the synthesis of metal ion-doped core nanocrystals, 0.02 mmol of dopant
precursor as respective acetates (manganese and cadmium) was added initially in
the first prepared mixture, while keeping the other conditions constant.
For the growth of shell over core and metal ion-doped core, 0.1 mmol of shell
precursor as zinc acetates dissolved in OLA/ODE (0.1/0.9 ml) was injected into the
reaction flask. Subsequently, the temperature was increased to 250 ◦ C and kept at
this temperature for 30 min to allow the growth of shell over core and doped core
nanocrystals. It is then cooled and purified with chloroform/acetone co-solvents by
34 V. Renuga and C. Neela Mohan

Fig. 2 Apparatus used for


the synthesis of nanocrystals
(hot-injection setup)

centrifugation (4000 rpm, 10 min). The schematic representation of the hot-injection


setup is shown in Fig. 2.

2.2 Growth of Multishell Over Core Nanocrystals

For the growth of multishell over the core, the growth procedure for single shell
over core was carried out first as discussed in the previous section followed by the
addition of second shell material precursors over the formed core/shell nanocrystals.
The reaction temperature was maintained at 250 ◦ C for 30 min and then the
contents were allowed to cool to room temperature. The prepared core-multishell
nanocrystals were purified using acetone/chloroform co-solvents and were dispersed
in chloroform for further spectroscopic analyses.

2.3 Optical Properties of Mn2+ Ion-Doped Core


and Core/Shell CuInS2 Nanocrystals

The onset absorption, bandgap, and crystal structures of the above core/shell
material mainly depend on their size, shape, composition, phase, and surface states
[12] and are shown in detail in Fig. 3 and Table 1.
Compared to undoped CuInS2 (837 nm) (ZB), Mn2+ ion-doped core (775 nm)
and ZnS shell formed over Mn2+ -doped CuInS2 core (607 nm) have absorption in
the blue-shift region with wurtzite crystal structure. The high-frequency dielectric
constant Mn2+ ions and the partial cation exchange occurred when ZnS shell formed
over the above doped core causes the Zn2+ to replace In3+ and Cu+ and thus
reduces the size of the core, resulting in a blue shift with change of phase or crystal
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 35

Fig. 3 Absorption spectra of undoped and Mn2+ ion-doped core, core/shell, and both core and
shell of CuInS2 /ZnS nanocrystals

Table 1 Combined absorption (nm), bandgap (eV), and phase structure details of the core CuInS2
and Mn2+ ion-doped core with ZnS shell nanocrystals
S.no Sample UV-Vis (nm) Bandgap (eV) Crystal structure
1 CuInS2 837 1.48 Zinc blende
2 CuInS2 /ZnS 563 2.20 Zinc blende
3 CuInS2 /ZnS:Mn2+ 525 2.36 Zinc blende
4 CuInS2 :Mn2+ 775 1.60 Wurtzite
5 CuInS2 : Mn2+ /ZnS2+ 607 2.04 Wurtzite
6 CuInS2 :Mn2+ / ZnS:Mn2+ 574 2.16 Wurtzite

structure. Similarly the appearance of strong blue shift with wider bandgap with the
same wurtzite phase that occurred when Mn2+ were doped over both core and shell
(573 nm) is again due to the entrance of Zn2+ ions into the CuInS2 the host lattice.
The growth of ZnS over the CuInS2 core (563 nm) and Mn2+ -doped ZnS shell
over CuInS2 core (525 nm) shows slight blue shift with pure core and has the same
zinc blende crystal phase as core. The existence of very high bandgap in the present
case may be attributed to the surface reconstruction of ZnS shell with Mn2+ ions
over the CuInS2 core. This is the cause why the Mn2+ doped both core and shell
exhibits its absorption toward lower end of the blue shift with very high bandgap.
36 V. Renuga and C. Neela Mohan

2.4 Photoluminescence Properties of Mn2+ Ion-Doped Core


and Core/Shell CuInS2 Nanocrystals

The full width at half maximum (FWHM) and emission peak efficiency of the
core CuInS2 , Mn2+ ion-doped core CuInS2 , and Mn2+ -doped ZnS shell over core
CuInS2 exhibit almost similar FWHM as ~35, ~30, and ~32 nm, respectively, with
narrow to reduced emission as shown in Fig. 4. The above consequence is due to
the existence of the inhomogeneity of the CuInS2 with dopants, because the surface
emission from CuInS2 is often overlapping with Mn2+ emission [13]. This may lead
to irregular distribution of vibrational states, so that the emission peaks could not
fall in the wider range [14]. Conversely wider FWHM with higher emissions were
observed for CuInS2 /ZnS, CuInS2 :Mn2+ /ZnS, and CuInS2 :Mn2+ /ZnS:Mn2+ as
~40, ~ 43, and ~45 nm, respectively. The reasons for higher FWHM and appreciable
emissions in all the above cases are due to:
(i) Surface etching and partial interfacial alloying of core CuInS2 with ZnS.
(ii) The existence of strain at the interface of core/shell nanocrystals may induce
anisotropic pressure around Mn2+ present in the core and change their local
environment and consequently their emission properties.
The existence of wider FWHM (~45 nm) with higher emission for
CuInS2 :Mn2+ /ZnS:Mn2+ among all the above cases is mainly due to dual emission

Fig. 4 Photoluminescence spectra of undoped and Mn2+ -doped CuInS2 /ZnS core/shell nanocrys-
tals
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 37

from Mn2+ localized transitions and the existence of excitonic recombination.


According to Yu et al. [15] and Vlaskin et al. [16], both factors such as thermo-
activated energy transfer from Mn2+ excited state 4 T1 and the back transfer from
the Mn2+ to the excitonic state of shell (ZnS) are mainly responsible for this highest
emission.

2.5 XPS Analysis of Mn2+ Ion-Doped Core and Core/Shell


CuInS2 Nanocrystals

X-ray photoelectron spectroscopy (XPS) analysis confirms the existence of Cu, In,
Mn, and Zn as Cu 2p, In 3d, S 2p, Mn 2p, Zn 2p, and C1s peaks in each of Mn2+
ion-doped CuInS2 /ZnS (CuInS2 :Mn2+ /ZnS) core/shell nanocrystals. The oxidation
states for Cu as +1, In as +3, and Mn as +2 states are confirmed by the appearance
of the two peaks split at 444.9 eV (In 3d5/2) and 452.3 eV (In 3d5/2) of In 3d core
with a splitting of 7.8 eV, which is consistent with a valence of +3 of indium ions.
Further, the existence of two strong peaks at 643.1 and 654.9 eV with a splitting of
13.2 eV in the spectrum confirms the presence of Mn2+ in Mn 2p region with the
corresponding binding energy of Mn 2p3/2 and Mn 2p1/2 [17], and the absence of
satellite peak at 944.0 eV responsible for +2 states of Cu2+ confirms the presence
of +1 state of Cu. The corresponding peaks are shown in Fig. 5.

Fig. 5 XPS spectra of CuInS2 :Mn2+ /ZnS:Mn2+ core/shell nanocrystals


38 V. Renuga and C. Neela Mohan

Fig. 6 XRD patterns of Mn2+ ion-doped core and core/shell nanocrystals

2.6 Crystal Structure Analysis of Mn2+ Ion-Doped Core


and Core/Shell CuInS2 Nanocrystals

The crystal structure and peak shifting of the CuInS2 -based nanocrystals from X-ray
diffraction (XRD) analysis reveal that CuInS2 , CuInS2 /ZnS, and CuInS2 /ZnS:Mn2+
exhibit zinc blende structure, whereas CuInS2 :Mn2+ , CuInS2 :Mn2+ /ZnS, and
CuInS2 :Mn2+ /ZnS:Mn2+ exist as wurtzite phase structures. Since, the 10% larger
ionic radius of Mn2+ than that of Cu+ and In3+ ions may force a small change in
the site symmetry, and that might be the reason behind in the phase transformation
of CuInS2 with Mn2+ incorporation. Their diffraction peaks exist at 2θ range with
corresponding planes and are shown in Fig. 6.

2.7 Morphological Properties of Mn2+ Ion-Doped Core


and Core/Shell CuInS2 Nanocrystals

TEM images of the CuInS2 /ZnS core/shell, Mn2+ ion-doped CuInS2 core with ZnS
shell, and Mn2+ ions doped both core and shell of CuInS2 /ZnS nanocrystals confirm
that the distance between the lattice fringes is found to be 3.24 Å with the distance
between the (111) planes of zinc blende phase. The selected area electron diffraction
(SAED) pattern of CuInS2 /ZnS core/shell nanocrystals exhibits bright intense dots,
which clearly reveals higher crystallinity in nature.
Similarly, the fringe space of Mn2+ ions doped on core of CuInS2 core with ZnS
shell has fringe space 3.13 Å, and d-spacing (002) plane is 3.132 Å. The fringe
space and d-spacing of the Mn2+ ions doped both core and shell also matched with
d-spacing of the corresponding (002) plane in XRD. All the above results confirm
the wurtzite structure formation in the formed nanocrystals.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 39

Fig. 7 TEM images and elemental compositions of (a) CuInS2 /ZnS, (b) CuInS2 :Mn2+ /ZnS, and
(c) CuInS2 :Mn2+ /ZnS:Mn2+ nanocrystals with their corresponding SAED patterns

Table 2 Elemental composition analysis of CuInS2 /ZnS core/shell and Mn2+ -doped core/shell
nanocrystals
S.no Sample At%
Cu In Zn Mn S
1 CuInS2 /ZnS 31.90 4.65 10.49 – 52.96
2 CuInS2 :Mn2+ /ZnS 18.10 5.92 19.40 1.06 55.52
3 CuInS2 :Mn2+ /ZnS:Mn2+ 27.88 2.95 15.31 1.35 52.52

From Fig. 7, it is inferred that the spherical shape observed for CuInS2 /ZnS
core/shell nanocrystals and for Mn2+ ions doped on CuInS2 core over the ZnS
shell exhibit hexagonal shape, whereas Mn2+ ions doped both core and shell of
CuInS2 /ZnS nanocrystals exhibited nearly spherical and some random structure
distribution due to the Ostwald ripening. The corresponding elemental composition
analysis from energy-dispersive X-ray spectrometer (EDX) also reveals the forma-
tion of CuInS2 /ZnS with appropriate amount of Cu, In, S, and Mn in CuInS2 /ZnS
core/shell nanocrystals and is shown in Fig. 7 and Table 2.

2.8 Absorption Properties of Core-Multishell


CuInS2 /MnS/ZnS Nanocrystals

The dopant Mn2+ ions in the previous case CuInS2 /ZnS nanocrystal were consid-
ered here as MnS in CuInS2 /MnS/ZnS multishell nanocrystal formation. The effect
40 V. Renuga and C. Neela Mohan

Fig. 8 (a) UV-Vis absorption spectra of core-multishell nanocrystals and (b, c) UV-Vis absorption
spectra of core-multishell nanocrystals with different concentrations of ZnS and MnS shells

of MnS as inner shell and their influence on optical, morphological, and phase
structure are discussed in the following section.
The excitonic absorption peaks (Fig. 8) in all these cases do not show any distinct
absorption features; instead, they exhibit less featured absorption [18]. It may be
due to:
(i) Sole or combined features of a broad and inhomogeneous size distribution
(ii) Unequal composition distribution
(iii) The existence of a special electronic property of CuInS2 nanocrystals
Out of the four onset absorption for CuInS2 , CuInS2 /ZnS, CuInS2 /MnS, and
CuInS2 /MnS/ZnS, both MnS and ZnS shells on CuInS2 exhibit onset absorption in
the blue shifts as shown in (Fig. 8a) with higher bandgap (2.02 and 2.46 eV, Table 3)
due to the existence of quantum confinement effect than the core CuInS2 (1.68 eV).
Compared to Mn2+ ions as dopant on CuInS2 (due to surface reconstruction)
(775 nm), Mn2+ as shell on CuInS2 (613 nm) shifts its absorption on lower side
(due to the formation of interfacial alloy (CuInS2 -ZnS and CuInS2 -MnS)) [18].
Further the shift in absorption is lower for MnS (CuInS2 /MnS) compared to ZnS
(CuInS2 /ZnS). Due to higher ionic radius, Mn2+ diffusion is less effective than
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 41

Table 3 Onset absorption peak and optical bandgap of the synthesized core, core/shell, and
core-multishell nanocrystals
S.no Sample name Onset absorption peak (nm) Optical bandgap (eV)
1. CuInS2 738 1.68
2. CuInS2 /ZnS 504 2.46
3. CuInS2 /MnS 613 2.02
4. CuInS2 /MnS/ZnS 784 1.58

lower ionic radii with higher interdiffusion natured ZnS over CuInS2 . This blue
shift [19] confirms the shell formation, instead of alloy formation, which usually
results in red shift.
But the formation of multishell structure over CuInS2 core shifts its absorp-
tion toward red-shift region compared to single-shell structure (CuInS2 /MnS and
CuInS2 /ZnS). The possibilities contributing to this effect are:
(i) High refractive index of the inner MnS shell [20].
(ii) The inner MnS acts as a medium to delocalize electrons from the core to outer
ZnS and thus reduce the quantum confinement effect, which results in red
shifting.
To find the exact influence of MnS on absorption, the multishell CuInS2 /MnS/ZnS
nanocrystals’ absorption is measured by changing the concentration of either
MnS or ZnS shell on CuInS2 /MnS/ZnS. The absorption spectra and bandgap of
CuInS2 /MnS/ZnS core-multishell nanocrystals with different shell concentrations
(ZnS (0.1–0.5) and MnS (0.1–0.5)) keeping all the other parameters constant are
shown in Fig. 8b, c.
In both the cases, the wider-bandgap ZnS and MnS effectively passivate the
surface; hence it shifts absorption toward blue shift (from 785 to 449 nm with
bandgap 1.58 to 2.76 eV for ZnS and from 784 to 525 nm with bandgap 1.58 to
2.36 eV for MnS).

2.9 Photoluminescence Analysis of Core-Multishell


CuInS2 /MnS/ZnS Nanocrystals

The fluorescence emission spectrometer at excitation wavelength of 360 nm


and 480 nm, for CuInS2 core, CuInS2 /MnS and CuInS2 /ZnS core/shell, and
CuInS2 /MnS/ZnS core-multishell nanocrystals, is shown in Fig. 9a, b.
The core CuInS2 exhibit very low intensity with emission at 742 nm by excitation
at 480 nm. Mn2+ as a dopant on CuInS2 and the doped CuInS2 passivated by ZnS
exhibit nearly the same emission as of core but with very poor intensity. Growth
of MnS shell over CuInS2 core nanocrystals increases intensity considerably with
42 V. Renuga and C. Neela Mohan

Fig. 9 PL spectra of core, core/shell, and core-multishell nanocrystals in chloroform using


excitation wavelength of 360 nm and 480 nm

emission at 726 nm. The existence of higher intensity of MnS shell over CuInS2
compared to Mn2+ dopant on CuInS2 is due to:
(i) Interdiffusion of Mn2+ ions into the core nanocrystals during MnS shell
passivation [21, 22]
(ii) Ion exchange between MnS and CuInS2 , showing reduction of the particle size
and increase of quantum confinement effect [23, 24]
(iii) Slight increase in PL intensity as the consequence of growth of the wider
bandgap of MnS (3.10 eV) over CuInS2 (1.52 eV)
(iv) The radiative recombination pathway of excitons confined within the core
structure promoted by MnS shell [25–28] and thereby efficient reduction of
nonradiative recombination after surface modifications
Among the two single shells, CuInS2 /ZnS contributes little bit higher intensity
compared to CuInS2 /MnS nanocrystals, but the latter exhibit higher intensities
than Mn2+ ion-doped core CuInS2 nanocrystals. The reason for slightly higher
intensity (Fig. 9b) for CuInS2 /ZnS compared to CuInS2 /MnS core/shell and CuInS2
core nanocrystals may be due to the reduced size of the core by cation exchange,
surface etching and partial interfacial alloying of CuInS2 core with ZnS, or surface
reconstruction during ZnS shell growth [21].
Growing multishells (paramagnetic MnS and diamagnetic ZnS) over CuInS2
core exhibits emission with exceptionally high intensity in the same NIR region
as that of CuInS2 , which not only suppresses the dangling bonds but also reduces
the strain and thereby eliminates surface trap states on the surface of CuInS2 core.
This is because the multishell formation ended with two well-defined green and
near-infrared (NIR) emissions at 543 and 752 nm with the excitation wavelength of
360 nm and 480 nm (Fig. 9a, b), respectively [29]; consequently it enhances the PL
intensity in an enormous nature.
The careful analysis of the influence of MnS on CuInS2 /MnS and CuInS2 /MnS
/ZnS depicts that there was no characteristic orange emission observed in both the
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 43

cases. The above observation reveals that the bandgap energy of CuInS2 is lower
than 4 T1 –6 A1 ligand-field Mn transition [30], which prescribes that the energy
transfer take place from the valence band (VB) of core nanocrystals to the Mn
centers, so that emission by core only is obtained [31]. This result also confirms
that MnS acts as a shell layer in the multishell core nanocrystals.
The full width at half maximum (FWHM) of the PL spectra suggests that
the PL spectra of CuInS2 /MnS and CuInS2 /ZnS nanocrystals were almost sim-
ilar but exhibit slightly higher emissions than the core CuInS2 , whereas the
CuInS2 /MnS/ZnS core-multishell nanocrystals have higher PL FWHM than all the
other nanocrystals. In addition, the PL FWHM (40 nm) of core-multishell nanocrys-
tals is apparently much wider than that of Mn2+ -doped CdS/ZnS nanocrystals
(20 nm) [32].

2.9.1 Influence of Shell Concentration on PL Properties


of CuInS2 /MnS/ZnS Core-Multishell Nanocrystals

The influence of shell concentration on PL property with varying shell concentra-


tions for MnS and ZnS from 0.1 to 0.5 mmol individually concludes that 0.1 mmol
by varying the concentration of one shell either MnS or ZnS at a time and keeping
the other parameters constant. Concentration of shell thickness (either inner MnS or
outer ZnS) would be the optimum level of shell growth thickness for the multishell
architecture as shown in Fig. 10.
Three main factors that affect the PL intensity in the present case are:
(i) Increase of MnS shell concentration may increase the amount of Mn2+ ions
in interfacial layer, providing more number of nonradiative recombination
sites (electron-hole pair) over the CuInS2 core. The formation of nonradiative
recombination sites over the surface of CuInS2 core would be affecting the PL
intensity of CuInS2 /MnS/ZnS nanocrystals [33].
(ii) PL quenching may occur, because of increase in the MnS shell concentration,
which may increase the Mn-Mn interactions due to high Mn2+ ion concentra-
tion; such interactions can cause a reduction in PL intensity [34, 35].
(iii) The decrease of PL intensity by increasing the outer ZnS shell concentration
may be due to the lattice mismatch between the ZnS outer shell and the
MnS inner shell that usually generate some structural defects, resulting in the
formation of nonradiative recombination sites [36–39].

2.10 X-Ray Photoelectron Spectroscopy (XPS) Analysis


of Core-Multishell CuInS2 /MnS/ZnS Nanocrystals

XPS analysis reveals that the splitting of Cu 2p core into Cu 2p3/2 and Cu 2p1/2
peaks, the binding energies located at 932.1 eV and 952.1 eV, respectively, confirms
44 V. Renuga and C. Neela Mohan

Fig. 10 PL spectra of CuInS2 /MnS/ZnS core-multishell nanocrystals with different shell concen-
trations (0.1–0.5 mmol) using excitation wavelength of 360 nm and 480 nm: (a, b) MnS shell
concentration and (c, d) ZnS shell concentration

that the oxidation state of Cu in the CuInS2 /MnS/ZnS nanocrystals is +1 instead


of +2, since the absence of satellite peak at 944.0 eV responsible for +2 states of
Cu2+ in the present case. The capping agent (organic thiols) is mainly responsible
for their In 3d state. The two peaks split at 444.9 eV (In 3d5/2) and 452.3 eV (In
3d5/2), which is consistent with a valence of +3. Further, the existence of two strong
peaks at 643.1 and 654.9 eV in the spectrum (Fig. 11) confirms the presence of Mn
in Mn 2p region with the corresponding binding energy of Mn 2p3/2 and Mn 2p1/2
[40]. In Mn 2p region (Fig. 11), the splitting width between 2p3/2 and 2p1/2 was
observed to be 11.8 eV.
The growth of ZnS over CuInS2/MnS core/shell nanocrystals caused decrease in
the signals of In 3d, Cu 2p, and Mn 2p, and Zn 2p peaks were divided into Zn 2p3/2
(1020.8 eV) and Zn 2p1/2 (1043.2 eV) peaks, which confirm the presence of Zn
atoms in the formation of ZnS shell over CuInS2/MnS nanocrystals. The presence
of three sulfur atoms (CuInS2, MnS, and ZnS) in the above nanocrystals is further
confirmed from the peak of S 2p centered at 161.40 eV.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 45

Fig. 11 XPS spectra of CuInS2 /MnS/ZnS core-multishell nanocrystals

2.11 Crystal Structure Analysis and the Influence of Shell


Concentration on Crystal Structure of Core-Multishell
CuInS2 /MnS/ZnS Nanocrystals

The four main diffraction peaks located at 2θ = 27.90, 31.83, 46.31, and 54.98◦
corresponding to the (111), (200), (220), and (311) planes reflect the presence of
zinc blende CuInS2 nanocrystals (Fig. 12a) [41]. The powder XRD pattern of the
nanocrystals product clearly matches with zinc blende CuInS2 (JCPDS Card No.
45-1304) [41].
The formation of ZnS shell over CuInS2 shifts the major diffraction peaks of core
CuInS2 nanocrystals to higher 2θ angle without changing its crystal structure (Fig.
12c). On the other hand, if MnS shell is passivated on the top of CuInS2 core, its
major diffraction peak is shifted to slightly lower 2θ angle with the same crystal
structure (Fig. 12b) [35]. These results suggest that either MnS or ZnS shell would
not affect the original crystal structure of CuInS2 core, indicating that almost similar
ionic radii of metal ions strongly diffuse into the core surface and not into the inner
part of the core, and hence it does not change the core’s crystal structure [19].
The XRD of CuInS2 /MnS/ZnS core-multishell nanocrystals (Fig. 12d) shows
shifts of diffraction peak positions of core-multishell nanocrystals to lower 2θ angle
when MnS shell is introduced between CuInS2 core and ZnS shell nanocrystal
compared to CuInS2 /ZnS. It revealed that certain amount of Mn2+ ions have been
introduced into the ZnS shell structure, which should be responsible for the slight
difference in XRD patterns [42]. This specific coating process does not affect the
crystal structure of host cubic CuInS2 structure and has the same zinc blende as
46 V. Renuga and C. Neela Mohan

Fig. 12 XRD patterns of synthesized nanocrystals: (a) CuInS2 core, (b) CuInS2 /MnS core/shell,
(c) CuInS2 /ZnS core/shell, and (d) CuInS2 /MnS/ZnS core-multishell nanocrystals

the core CuInS2. This indicates the formation of CuInS2/MnS/ZnS core-multishell


nanocrystals.
Varying the concentration (from 0.1 to 0.5 mmol) of either ZnS or MnS
shell over CuInS2 core does not alter the crystal structure of CuInS2 core and
CuInS2 /MnS/ZnS, as shown in Fig. 13, and exhibits the shift of diffraction peaks
to slightly higher 2θ with the same zinc blende structure. With increase of shell
concentrations from 0.1 to 0.5 mmol, the interplanar spacing decreased for both
MnS and ZnS shell structure, and the corresponding values are listed in Table 4 and
shown in Fig. 13a, b.

2.12 Morphological Analysis of CuInS2 /MnS/ZnS


Core-Multishell Nanocrystals

TEM images revealed that all the core, core/shell, and core-multishell nanocrystals
are close to the near-spherical shape and fairly monodispersed and are shown in
Fig. 14.
The appearance of interplanar spacing of 0.320 nm (Fig. 14) corresponding to
that of the (111) planes of zinc blende CuInS2 (0.3224 nm) is consistent with the
result of the X-ray diffraction pattern of (111) plane. Further Fig. 14 exhibit only the
lattice fringes without any well-defined interface between core and shell throughout
the entire crystal indicates the epitaxial growth of the shell over core nanocrystals
[36].
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 47

Fig. 13 XRD patterns of CuInS2 /MnS/ZnS nanocrystals with different shell concentrations (0.1–
0.5 mmol): (a) MnS shell and (b) ZnS shell

Table 4 Elemental composition of CuInS2 core, CuInS2 /MnS and CuInS2 /ZnS core/shell, and
CuInS2 /MnS/ZnS core-multishell of the synthesized CuInS2 nanocrystals
S.no Sample name Compositions (wt %)
Cu In Mn Zn S
1 CuInS2 28.20 20.77 – – 50.74
3 CuInS2 /MnS 28.99 21.34 2.72 – 46.84
2 CuInS2 /ZnS 26.45 18.61 – 13.87 39.96
4 CuInS2 /MnS/ZnS 24.90 11.62 18.83 16.04 28.61

Fig. 14 TEM and HR-TEM images of (a) CuInS2 core, (b) CuInS2 /ZnS, (c) CuInS2 /MnS
core/shell, and (d) CuInS2 /MnS/ZnS nanocrystals

The combined SAED and XRD patterns (Fig. 15) reveal the three lattice spacing
values for the diffraction rings, and the obtained values are 0.3206, 0.2042, and
0.1758 nm, which correspond to the d-spacing for lattice planes of (111), (220), and
(311) zinc blende of CuInS2 .
48 V. Renuga and C. Neela Mohan

Fig. 15 SAED pattern and XRD pattern (insert figure) and compositions of (a) CuInS2 , (b)
CuInS2 /ZnS, (c) CuInS2 /MnS, and (d) CuInS2 /MnS/ZnS nanocrystals

From EDX (Fig. 15), the presence of Cu-, In-, and S-related peaks indicates
the formation of CuInS2 nanocrystals, whereas in CuInS2 /ZnS and CuInS2 /MnS
nanocrystals, additionally Zn-, Mn-, and S-related peaks were also observed
(Fig. 15). These results confirmed the passivation of shells (MnS or ZnS) on CuInS2
nanocrystals.

2.13 UV-Vis Absorption Spectra of Cd2+ -Doped Core


and Core/Shell AgInS2 Nanocrystals

The second part of this chapter deals with core AgInS2 , core passivated by ZnS,
and core doped with Cd2+ ions and passivated by ZnS shell. Apart from that the
influence of Cd as Cd2+ dopant and as CdS shell on AuInS2 are also discussed.
Correspondingly their optical, crystal structure, phase, and morphology on the core
AgInS2 in single core/shell and core-multishell nanocrystals are discussed in a
detailed manner.
The optical properties of the doped and undoped AgInS2 core and AgInS2 /ZnS
core/shell nanocrystals were characterized by UV-Vis absorption, and the corre-
sponding PL emission spectroscopes and their bandgaps are also shown in Figs. 16
and 17 and Table 5.
In general, the ternary nanocrystals exhibit broad absorption without any sharp
excitation peak and do the same for AgInS2 -based nanocrystals also. So their onset
absorption peak edge was recorded and appeared as shown Fig. 16.
Doped AgInS2 (534 nm) with Cd2+ ion shifts its absorption to red shift
(602 nm) compared to ZnS-passivated AgInS2 (375 nm) (blue shift). Due to the
small lattice mismatch between orthorhombic AgInS2 and cubic CdS (Cd2+ ions
precursor), the alloying of Cd2+ ions with AgInS2 surface takes place [43]. The
ZnS-passivated AgInS2 core decreases the surface defects of AgInS2 and thereby
shifts its absorption toward blue-shift region.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 49

Fig. 16 UV-Vis absorption spectra of the doped and undoped AgInS2 -based core and core/shell
nanocrystals

Fig. 17 Photoluminescence spectra of doped and undoped AgInS2 -based core and core/shell
nanocrystals

But Cd2+ ion-doped ZnS shell over AgInS2 exhibits absorption (476 nm) toward
red shift compared to undoped ZnS over AgInS2 (375 nm) and toward blue shift
compared to Cd2+ ion-doped AgInS2 (602 nm). Similarly if Cd2+ ions were
doped on both core and shell, they exhibit onset absorption at 516 nm low blue-
50 V. Renuga and C. Neela Mohan

Table 5 Calculated onset absorption peak positions and bandgap of pure and doped AgInS2 -
based core and core/shell nanocrystals
S.no Sample name Onset absorption(nm) Bandgap (eV)
1 AgInS2 534 2.32
2 AgInS2 /ZnS 375 3.30
3 AgInS2 /ZnS:Cd2+ 476 2.60
4 AgInS2 :Cd2+ 602 2.06
5 AgInS2 :Cd2+ /ZnS 500 2.48
6 AgInS2 :Cd2+ /ZnS:Cd2+ 516 2.40

shift region compared to AgInS2 core. All the above results conclude that either
passivation or dopant will be the deciding factor to reduce the bandgap of the core;
among these two factors, dopant alone plays a major role in reducing the bandgap;
correspondingly their optical property also changes linearly. However, if doped
AgInS2 nanocrystals were passivated by ZnS shell, the bandgap is decreased to a
level in between that of undoped core/shell (AgInS2 /ZnS) and doped ZnS shell with
undoped AgInS2 (AgInS2 /ZnS:Cd2+ ).

2.14 Photoluminescence Properties of Cd2+ -Doped Core


and Core/Shell AgInS2 Nanocrystals

All the AgInS2 nanocrystals exhibit yellow emission at 570 nm with FWHM
~140 nm for AgInS2 , and for Cd2+ -doped AgInS2 , emission occurs at 580 nm
with FWHM ~120 nm with lower PL intensity with excitation at 385 nm. The
influence of completely filled d-orbital dopant (Cd2+ ) increases the size and leads
to nonradiative pathway; that results in very light red shift with low intensity.
The passivation by ZnS shell over AgInS2 may diffuse the size of AgInS2
nanocrystals and exhibit emission at 538 nm a blue shift with FWHM ~15 nm when
comparing the effect of both dopant and passivation simultaneously on AgInS2 core
passivation (AgInS2 :Cd2+ /ZnS) and undoped core with doped passivated ZnS shell
(AgInS2 /ZnS:Cd2+ ).
In the first case (AgInS2 :Cd2+ /ZnS) due to the dopant or impurity of Cd2+
atoms [44, 45], emission occurred at 512 nm with FWHM of 131 nm, a blue shift
with higher intensity compared to undoped AgInS2 and undoped AgInS2 core with
ZnS shell (AgInS2 /ZnS). In addition, the green emission from AgInS2 :Cd2+ /ZnS
core/shell nanocrystals is attributed either to direct band-to-band recombination
in CdS particles [46, 47] or recombination of Cd2+ ions as an impurity center
[48]. Another reason for the occurrence of green emission in the present case is
attributable to the direct recombination effect (E-De) of Cd2+ ions which is located
at the center between the core and shell [48]. Recombination is the direct evidence
of reduction of FWHM and green emission; the intensity of the AgInS2 :Cd2+ /ZnS
core/shell nanocrystals is higher compared to all other cases.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 51

When AgInS2 core surface is passivated by ZnS shell, PL emission occurs at


538 nm with the corresponding FWHM of 151 nm as shown in Fig. 17. After
passivating the surface of AgInS2 core nanocrystals, the emission results in blue
shift of the emission spectrum. Compared to doping of AgInS2 nanocrystals with
surface passivation, the slight red shift in the earlier case and blue shift in the latter
case lead to the conclusion that passivation by ZnS shell may diffuse the size of
AgInS2 nanocrystals, since there may be a chance of diffusion of a portion of ZnS
layer into the AgInS2 core.
In the second case (AgInS2 /ZnS:Cd2+ ), the emission occurred at 552 nm with
FWHM of 133 nm. The existence of slight blue shift from the core (AgInS2 ) and
doped core (AgInS2 :Cd2+ ), and red shift from the passivated core (AgInS2 /ZnS) and
doped core with ZnS shell (AgInS2 :Cd2+ /ZnS), warrants finding out whether the
dopant or passivation plays a vital role in predicting or influencing the PL property
of pure AgInS2 . The reason for the existing result in the present case is in view of
the lattice mismatch between the surface layer (ZnS shell) and inner core (AgInS2 )
[49] that may lead to the formation of trap states.
When Cd2+ ions were doped on both core and shell (AgInS2 :Cd2+ /ZnS:Cd2+ ),
the emission occurred at 525 nm, blue-shifted in all cases except Cd2+ ion-doped
AgInS2 core over ZnS shell (red shift) with FWHM of 148 nm. One important
observation here is both doped core passivated (AgInS2 :Cd2+ /ZnS) and doped core
with doped ZnS shell (AgInS2 :Cd2+ /ZnS:Cd2+ ) generally show higher intensity of
emission compared to all previous cases. Hence, the conclusion arrived here is that
surface modification plays a more vital role than dopant in increasing PL intensity
[50]. The main reason of increasing PL intensity in the latter case is change in
surface structure and crystal morphology that accompany the decrease in surface-
to-volume ratio of the present case when compared to previous one [51, 52]. The
nonradiative pathways are diminished by reducing the surface states either doping
of foreign ions on core or shell. This leads to increase in radioactive emission and
it resulted in increased PL intensity as shown in Fig. 17. This concept had been
discussed in many papers [53–57].

2.15 Crystal Structure Analysis of Cd2+ -Doped Core


and Core/Shell AgInS2 Nanocrystals

The diffraction peaks of all the AgInS2 -based nanocrystals in the XRD spectra
match well with the patterns of the orthorhombic phase and are shown in Fig. 18.
The existences of broader diffraction peaks in all cases are due to smaller size of the
AgInS2 -based nanocrystals.
AgInS2 :Cd2+ nanocrystals shift slightly to the lower angle. The gradual replace-
ment of both Ag+ and In3+ ions with two Cd2+ atoms to form Cd2+ -doped AgInS2
core in AgInS2 :Cd2+ nanocrystals (Fig. 18), which in turn is confirmed from EDX
analysis.
52 V. Renuga and C. Neela Mohan

ZnS shell-coated AgInS2 core exhibited the same orthorhombic phase structure
with peak positions shifted to the side of higher angle and located in the middle
position of the patterns of AgInS2 and the standard patterns of ZnS. The change in
peak positions suggests that some amount of Zn2+ ions diffuses or gets deposited
on the surface of the core AgInS2 nanocrystals [58, 59]. AgInS2 :Cd2+ /ZnS:Cd2+
core/shell nanocrystals also exhibit a similar orthorhombic crystal structure with
lower-angle 2θ shift which may be due to the dopant ions entering into both the core
and the shell.
The XRD patterns revealed that there was no change in crystal structure of Cd2+
ion-doped ZnS shell over AgInS2 core nanocrystals (AgInS2 /ZnS:Cd2+ ), com-
pared to undoped AgInS2 /ZnS. Both AgInS2 /ZnS and AgInS2 /ZnS:Cd2+ core/shell
nanocrystals have the same orthorhombic phases; similarly the XRD pattern of the
Cd2+ ion-doped AgInS2 nanocrystals passivated by ZnS shell was unchanged and
exhibited orthorhombic crystal structure similar to doped AgInS2 core nanocrystals.
In addition, the positions of diffraction peaks of undoped ZnS shell over Cd2+ ion-
doped AgInS2 nanocrystals were shifted to lower 2θ angles, which is an evidence of
the incorporation of Cd2+ ions into AgInS2 /ZnS core/shell structure.

Fig. 18 XRD patterns of undoped and Cd2+ -doped AgInS2 core and AgInS2 /ZnS core/shell
nanocrystals
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 53

2.16 Surface Morphology Analysis of Cd2+ -Doped Core


and Core/Shell AgInS2 Nanocrystals

The morphologies and sizes of AgInS2 /ZnS, AgInS2 :Cd2+ /ZnS, and AgInS2 :Cd2+
/ZnS:Cd2+ core/shell nanocrystals were characterized by the low-magnification
TEM images of pure AgInS2 /ZnS, and doped AgInS2 /ZnS core/shell nanocrystals
under the same reaction conditions revealed that both pure and doped core/shell
nanocrystals were irregular and sparingly monodispersed nanoparticles [60] as
shown in Fig. 19.
The crystallinity of the products examined by high-resolution TEM (HR-TEM)
and images is shown in Fig. 19. The interplanar spacing between each adjacent
plane is 0.304, 0.310, and 0.314 nm for AgInS2 /ZnS, AgInS2 :Cd2+ /ZnS, and
AgInS2 :Cd2+ /ZnS:Cd2+ core/shell nanocrystals, respectively, and corresponds with
the spacing of the (120) crystal planes of orthorhombic structure of AgInS2 , which
agrees well with the (120) d-spacing of the value in published literature (JCPDS
No.25-1328) [61].
The interplanar spacing of doped AgInS2 /ZnS core/shell nanocrystals is slightly
higher than pure AgInS2 /ZnS nanocrystals, which may be due to the incorporation
of Cd2+ ions into the core as well as both core and shell of AgInS2 /ZnS nanocrys-
tals.
The average sizes of pure and doped core/shell nanocrystals were found
to be 4.4 nm (AgInS2 /ZnS), 4.6 nm (AgInS2 :Cd2+ /ZnS), and 5.0 nm
(AgInS2 :Cd2+ /ZnS:Cd2+ ), suggesting that the doping composition of the precursor

Fig. 19 (a–c) TEM images, (d–f) HR-TEM images (interplanar spacing), and (g–i) SAED
patterns of AgInS2 /ZnS, AgInS2 :Cd2+ /ZnS, and AgInS2 :Cd2+ /ZnS:Cd2+ core/shell nanocrystals
and their elemental compositions
54 V. Renuga and C. Neela Mohan

did not greatly influence the size and size distribution of the resultant nanoparticles.
The selected area electron diffraction (SAED) patterns of the synthesized
nanocrystals are shown in Fig. 19. The estimated diffraction radii rings match
well with the (120), (200), (002), (121), (122), (320), (203), and (112) planes and
lattice planes of orthorhombic AgInS2 [61]. SAED results confirmed the data from
XRD measurements of the Cd2+ ion-doped AgInS2 nanocrystals.

2.17 Elemental Compositions of Cd2+ -Doped Core


and Core/Shell AgInS2 Nanocrystals

The chemical composition of core/shell and doped core/shell nanocrystals from


elemental dispersive X-ray (EDX) spectra is that AgInS2 core/shell nanocrystals
(Fig. 19) are composed of Ag, In, S, and Zn and doped core/shell are composed
of Ag, In, S, Zn, and Cd elements, respectively. Table 6 summarizes the atomic
composition of each element investigated using EDX. For pure AgInS2 /ZnS
core/shell nanocrystals, the atomic ratio of Ag and In was close to 1:1 stoichiometry.
In another case, cadmium-doped core only AgInS2 /ZnS nanocrystals show that
the atomic ratio of indium is slightly lower than pure core/shell nanocrystals and
implies the formation of possible vacancies at silver sites. It further denotes that,
apart from occupying the vacancies, the cadmium atoms would replace some of
the cationic elements; the replacement might have occurred primarily at the silver
sites compared to the indium ones. The change in composition of Ag+ ions in the
present nanocrystals might be due to cation exchange of Cd2+ ions into the Ag+
sites, because the ionic radius of cadmium ions (0.95 nm) is smaller than silver ions
(1.15 nm).

2.18 UV-Vis Studies of Core-Multishell and Alloyed


AgInS2 /CdS/ZnS Shell Nanocrystals

UV-Vis absorption spectra and bandgap of the AgInS2 core, type II AgInS2 /CdS,
type I AgInS2 /ZnS core/shell, type II/type I AgInS2 /CdS/ZnS, type I/reverse
type I AgInS2 /ZnS/CdS core-multishell, and their alloying shell system (type I

Table 6 Elemental composition of pure and Cd2+ -doped core/shell nanocrystals


S.no Samples Atomic%
Ag L In L Zn K Cd L SK
1 AgInS2 /ZnS 21.52 16.21 13.85 – 48.42
2 AgInS2 :Cd2+ /ZnS 19.21 15.87 13.43 03.11 48.38
3 AgInS2 :Cd2+ /ZnS:Cd2+ 19.16 15.73 11.70 05.72 47.69
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 55

AgInS2 /Zn1-x Cdx S) nanocrystals are characterized separately and are shown in
Fig. 20 and Table 7.
The quantum confinement effect and surface trap states of AgInS2 nanocrystals
exhibit onset absorption peak edge at 534 nm with a bandgap of 2.32 eV.
Passivation of CdS over AgInS2 resulted in type II core/shell architecture
with bathochromic shift (red shift), and they exhibit absorption around 564 nm,
whereas ZnS shell passivation over AgInS2 results in type I core/shell systems with
hypsochromic shift (blue shift) and shows absorption around 375 nm, due to their
inherent nature of band alignment.
Absorption spectra of AgInS2 /CdS/ZnS type II/type I and AgInS2 /ZnS/CdS type
I/reverse type I core-multishell nanocrystals exhibit blue shift for the former case
and red shift for the latter case. When ZnS were passivated over AgInS2 /CdS
core/shell nanocrystals, the absorption is blue-shifted from 564 nm (AgInS2 /CdS)
to 397 nm (AgInS2 /CdS/ZnS), and this may be due to the dangling bond present on
the surface of CdS which is suppressed by the higher-bandgap ZnS. Moreover, the

Fig. 20 UV-Vis spectra of core, core/shell, and core-multishell AgInS2 /CdS/ZnS nanocrystals

Table 7 Calculated onset absorption peak positions and bandgap of core, core/shell, core-
multishell, and core/alloyed shell AgInS2/CdS/ZnS nanocrystals
S.no Sample name Onset absorption (nm) Bandgap (eV)
1 AgInS2 534 2.32
2 AgInS2 /ZnS 375 3.30
3 AgInS2 /ZnS/CdS 421 2.92
4 AgInS2 /CdS 564 2.20
5 AgInS2 /CdS/ZnS 397 3.12
6 AgInS2 /Zn1-x Cdx S 460 2.70
56 V. Renuga and C. Neela Mohan

higher-bandgap ZnS compresses the lattice strain accumulated not only between
AgInS2 and CdS but also in AgInS2 /CdS/ZnS architecture. It leads to broader
absorption with higher bandgap energy, and this is responsible for blue shift in the
present case.
AgInS2 /ZnS/CdS nanocrystals exhibit red-shift absorption from 375 nm
(AgInS2 /ZnS) to 421 nm (AgInS2 /ZnS/CdS). This higher wavelength with lower
energy of absorption is due to the formation of type I/reverse type I architecture
formed during CdS over type I AgInS2 /ZnS. Here, the lattice strain and surface
trap states are greatly reduced. Passivation of reverse type I over type I material
will definitely lead to lower energy absorption [62]. Obviously, this may lead to
absorption in the red-shift region.
Schematic diagram 2 shows the band alignment and lattice mismatch between the
AgInS2 , ZnS, and CdS semiconductor nanocrystals. Here, AgInS2 /CdS core/shell
nanocrystals have less lattice mismatch (∼4.2%) than AgInS2 /ZnS (∼5.5%) and
CdS/ZnS (∼7.0%) core/shell nanocrystals. Therefore, layering the alloy Zn1-x Cdx S
shell over core AgInS2 nanocrystals provides better stability by decreasing inter-
facial strain between core and shell. This AgInS2 /Zn1-x Cdx S core/alloyed shell
systems exhibit blue-shift absorption (460 nm) compared to core AgInS2 nanocrys-
tals (534 nm).
Due to the low lattice mismatch between AgInS2 core and Zn1-x Cdx S alloyed
shell, the possibility of diffusion of metal ions present in alloy system reduces the
strain that accumulates at the interface of AgInS2 and Zn1-x Cdx S alloyed shell and
tends to influence the blue shift with broader absorption. Moreover, the alloyed shell
comprises a larger volume fraction (AgInS2 /Zn1-x Cdx S) than the core (AgInS2 ), so
as to cause interdiffusion effect mainly at the interface between AgInS2 core and
Zn1-x Cdx S alloyed shell which is also responsible for the existence of blue shift in
the core/alloyed shell systems.

2.19 Photoluminescence Studies of Core-Multishell


and Alloyed AgInS2 /CdS/ZnS Shell Nanocrystals

PL spectra of core, core/shell, core-multishell (ZnS/CdS and CdS/ZnS), and alloyed


shell-coated AgInS2 nanocrystals are shown in Fig. 21. The PL spectra of core
AgInS2 nanocrystals exhibit emission wavelength at 570 nm when excited at 385 nm
and have lower PL intensity with larger FWHM of 100 nm [63, 64]. The existence
of PL quenching is caused by an escape of carriers via a nonradiative recombination
path [65] due to the presence of dangling bonds on the surface which leads the
leakage of excitonic wave function outside the surface.
To increase the AgInS2 core emission intensity, two higher-bandgap inorganic
materials (CdS and ZnS) were passivated over AgInS2 either individually or in a
combined form, so as to remove the surface defects. It acts as a trap state for the
carriers and thereby minimizes the probability of undesired quenching emission via
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 57

Fig. 21 Comparison of PL spectra of AgInS2 /Zn1-x Cdx S core/alloyed shell nanocrystals with (a)
single shell (ZnS and CdS)-coated AgInS2 , (b) multishell (ZnS/CdS and CdS/ZnS)-coated AgInS2
nanocrystals, and (c) core AgInS2 nanocrystals

nonradiative recombination. The choice of these materials is based on the fact that
both ZnS and CdS provide tunneling effect for electrons and holes originating in the
AgInS2 core, thereby reducing the probabilities for the carriers to come out from
the surface of AgInS2 .
Passivation of AgInS2 core by ZnS shell, type I core/shell architecture, and CdS
resulted in type II core/shell architecture formation. Accordingly, their emissions
also vary in a surprising manner. When AgInS2 nanocrystals were passivated by
ZnS, the emission falls at 538 nm, a blue-shift region (from 570 nm for AgInS2 )
with higher intensity (compared to core), and its FWHM falls greater than 100 nm,
exhibiting larger stock shifts with broader band emissions [63]. On the other hand,
when CdS were passivated over AgInS2 , the emission fell at 585 nm with a
progressive red-shift region [49, 66] (from 570 nm for AgInS2 ) with comparatively
low intensity, and their corresponding FWHM fell at 140 nm.
One such reason for the existence of red shift of AgInS2 /CdS, compared to
AgInS2 /ZnS, is that the size of the cadmium is slightly higher than that of zinc atom.
During the passivation process, a small amount of strain may accumulate between
AgInS2 and CdS which leads to the dislocation of Cd2+ ions due to its photo-
ionization effect over AgInS2 core. It may decrease nonradiative recombination
center, so that it needs lower energy than Zn2+ ion, where minimum strain occurs.
Both type I (AgInS2 /ZnS) and type II (AgInS2 /CdS) nanocrystals were passi-
vated by higher-bandgap materials to form CSS architecture, so as to reduce the
problem faced by the single CS system. Here again both CdS and ZnS were used
58 V. Renuga and C. Neela Mohan

to passivate in the above cases to attain AgInS2 /ZnS/CdS a type I/reverse type I
and for AgInS2 /CdS/ZnS a type II/type I formation. By this arrangement, the size-
dependent band edge luminescence efficiency is improved reasonably, and thereby
the emission occurred at 572 nm (almost similar to core) for AgInS2 /ZnS/CdS
type I/reverse type I with higher intensity corresponding to their FWHM falling
at 180 nm.
At the same time, AgInS2 /ZnS were passivated by CdS a potential trapant (due to
larger radius) for electrons and holes or originate within the CdS shell and thereby
reduce the strain accumulated in the interface.
Passivation of AgInS2 /CdS type II systems by ZnS shell results in emissions
in the blue-shift region (527 nm) [67] compared to both AgInS2 (570 nm) and
AgInS2 /CdS (585 nm). The conjugations of ZnS over AgInS2/CdS take place in
a linear or homogenous manner due to their narrow size distribution with good
monodispersity and maintain a discrete form without aggregation.
It is noteworthy that alloying of Zn1-x Cdx S shell passivated over AgInS2
core nanocrystals provides emission at 532 nm toward red shift compared to
AgInS2 /CdS/ZnS and to blue shift for the other systems. It is due to equal
contribution of Zn and Cd ions relative to S precursor in the shell solution.
The formed AgInS2 /Zn1-x Cdx S core/alloyed shell would inevitably promote cation
exchange and diffusion between Zn2+ , Cd2+ , and the ions present in AgInS2 core.
This alloyed shell formation not only reduces the interfacial tensions but also
generates energy barrier for holes. The conduction band edge of Zn1-x Cdx S alloyed
shell material is located above that of AgInS2 and CdS, while the valance band
edge is deeper than CdS. This formed a larger potential trough in the valance band
edges as an additional blocking layer for holes [68]. This energy band structure can
further reduce the overlap of electrons and holes to decrease the strength of the first
excitation absorption peak and suppresses the reabsorption.

2.20 XPS Spectra of Core-Multishell and Alloyed


AgInS2 /CdS/ZnS Shell Nanocrystals

XPS analysis (Fig. 22) reveals that the existence of the 3d state of Cd exists in
two split states as Cd 3d5/2 and Cd 3d3/2 and their corresponding binding energies
are located at ~404 eV (3d5/2 ) and ~ 410 eV (3d3/2 ) with the peak separation of
~6.0 eV which is due to spin-orbit coupling and confirms the +2 oxidation state of
Cd. The existence of characteristic doublet peaks of Zn 2p at 1021 eV (Zn 2p3/2 )
and 1044 eV (Zn 2p1/2 ) with a splitting of ~23 eV is due to the presence of Zn(II),
and it establishes the formation of ZnS as a shell around the core nanocrystals. The
presence of three sulfur atoms in the coordinations65 of AgInS2 , Cd–S, and Zn–S is
proved by the appearance of binding energy of sulfur 2p electrons peak centered at
~160 eV.
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 59

Fig. 22 XPS spectra of core-multishell (AgInS2 /ZnS/CdS and AgInS2 /CdS/ZnS) and alloyed
shell (AgInS2 /Zn1-x Cdx S) nanocrystals

The core/alloyed shell nanocrystal’s total peak emission intensity is much


higher than that of multishell nanocrystals and is most probably due to the higher
contribution of the lattice sulfur atom from both the inner AgInS2 core and the
outer Zn1-x Cdx S shell. All the above results clearly confirmed that the valence of
the elements in the present case such as Ag, In, Zn, S, and Cd are found to be 1+,
3+, 2+, 2−, and 2+, respectively.

2.21 Crystal Structure Analysis of Core-Multishell and Alloyed


AgInS2 /CdS/ZnS Shell Nanocrystals

Powder X-ray diffraction patterns shown in Fig. 23 for AgInS2 -based core,
core/shell, and core/shell/shell nanocrystals confirmed the formation of
orthorhombic phase of AgInS2 (JCPDS Card No. 25-1328) with lattice parameters
of a = 0.7001 nm, b = 0.8258 nm, and c = 0.6704 nm. The existence of broad and
weak diffraction peak with the broad band at 2θ ∼ 21.8◦ is the characteristic nature
of the QDs present in the system [61]. The diffraction peaks at 2θ values of 24.93◦ ,
25.42◦ , 26.58◦ , 28.33◦ , 36.86◦ , 43.64◦ , 48.00◦ , and 52.68◦ matched well with the
(120), (200), (002), (121), (122), (040), (123), and (322) planes of orthorhombic
AgInS2 phase.
Deposition of a ZnS and CdS shell separately over AgInS2 does not alter the
general XRD pattern of the AgInS2 nanocrystals but results in a distinct shift of
60 V. Renuga and C. Neela Mohan

Fig. 23 XRD patterns of core (AgInS2 ), core/shell (AgInS2 /CdS and AgInS2 /ZnS), core-
multishell (AgInS2 /ZnS/CdS and AgInS2 /CdS/ZnS), and alloyed shell (AgInS2 /Zn1-x Cdx S)
nanocrystals

the reflections to larger angles for ZnS (Fig. 23), and slightly lower 2θ for CdS
indicates the incorporation of Zn2+ ions for ZnS and the lower lattice mismatch
(4.2%) between the AgInS2 and CdS for CdS. The multishell structures also exhibit
orthorhombic crystal structure. The alloyed Zn1-x Cdx S shell layer presented in
Fig. 23 exhibits a peak shift to a higher diffraction (2θ angle) with respect to that of
bare AgInS2 nanocrystals and retains the same orthorhombic crystal structure.

2.22 Morphological Studies of Core-Multishell and Alloyed


AgInS2 /CdS/ZnS Shell Nanocrystals

TEM image of AgInS2 /ZnS core/shell nanocrystal (Fig. 24) shows a reasonable
narrow size distribution with average diameters of 6.2 nm. When passivating
(ZnS/CdS) over AgInS2 nanocrystals (Fig. 24), the size of the formed core-
multishell nanocrystals (AgInS2 /ZnS/CdS) increased to 12.1 nm, but passivating by
CdS/ZnS shell over AgInS2 core nanocrystals, the size of final AgInS2 /CdS/ZnS is
reduced to 11.5 nm, smaller than AgInS2 /ZnS/CdS type I/reverse type I nanocrys-
tals. The images of the multishell systems are also irregular in shape with broad
distribution in sizes [61].
The morphology of the multishell nanocrystals was observed from the HR-
TEM image Fig. 24 and is found as pyramidal-like morphology for AgInS2/ZnS
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 61

Fig. 24 (a–l) HR-TEM images and SAED patterns and elemental compositions of (a) type
I AgInS2 /ZnS core/shell, (b) type I/reverse type I AgInS2 /ZnS/CdS, (c) type II/type I
AgInS2 /CdS/ZnS core-multishell, and (d) type I AgInS2 /Zn1-x Cdx S core/alloyed shell nanocrys-
tals

Table 8 Elemental composition of core/shell, core-multishell, and core/alloyed shell of AgInS2


nanocrystals
S.no Samples Atomic%
Ag L In L Zn K Cd L SK
1 AgInS2 /ZnS 21.52 16.21 13.85 – 48.42
2 AgInS2 /ZnS/CdS 14.98 14.92 15.40 06.96 47.74
3 AgInS2 /CdS/ZnS 14.74 11.66 13.07 12.42 48.11
4. AgInS2 /Zn1-x Cdx S 20.51 14.34 7.05 09.48 48.62

core/shell, plate-like morphology for AgInS2 /CdS/ZnS and AgInS2 /ZnS/CdS core-
multishell, and hexagonal-like structure for AgInS2 /Zn1-x Cdx S core/alloyed shell
nanocrystals.
The TEM selected area electron diffraction (SAED) pattern shows six distinct
diffraction rings. The calculated d-spacings match well with the major d-spacings
of orthorhombic phase AgInS2 (Fig. 24), which is consistent with the result of the
XRD measurements (Fig. 24).
Energy-dispersive X-ray spectroscopy (EDS) of AgInS2 /ZnS core/shell,
AgInS2 /CdS/ZnS, and AgInS2 /CdS/ZnS core-multishell nanocrystals, as shown
in Fig. 24, further suggested the presence of the elemental ratios of the Ag, In, Zn,
and S in core/shell nanocrystals (Fig. 24a) and additionally Cd in the core-multishell
nanocrystals (Fig. 24).
Figure 24 exhibits the EDX spectrum of the prepared AgInS2 /Zn1-x Cdx S
core/alloyed shell nanoparticles and found the ratio of Ag:In:Zn:S:Cd for the
obtained nanocrystal is about 13.14:12.97:22.07:44.53:7.29, respectively, which
justifies that the Cd and Zn had successfully grown into the obtained AgInS2
nanocrystal. The obtained elemental composition of core/shell and core-multishell
nanocrystals is listed in Table 8.
62 V. Renuga and C. Neela Mohan

3 Conclusions

This chapter deals with the utility of the full potential of ternary (CuInS2 and
AgInS2 ) chalcogenide as an alternative to the existing toxic nanocrystals systems
and to improve their optical property as well as intensity by architecting core/shell
and core-multishell nanocrystals.
To produce such highly crystalline nanocrystals with controllable size, hot-
injection method of synthesis is adopted, and the strategies adopted are surface
passivation, doping (high-luminescence Mn2+ and Cd2+ a d5 and d10 ionic sys-
tems), and alloying.
The dopant Cd2+ when doped both on core and shell in a combined manner
than individually on CuInS2 /ZnS single core/shell systems enhances optical and
luminance property considerably. Moreover the crystal structure of CuInS2 /ZnS
changes from zinc blende to wurtzite when Mn2+ were doped over CuInS2 /ZnS
system.
The multishell architecture of CuInS2 /MnS/ZnS exhibits dual emission both in
green and NIR region with high stability. The optimum thickness of the middle as
well as outer shell is preferably 0.1 mmol. The zinc blende structure is retained
throughout the entire growth process.
The core/shell QDs of AgInS2 are of high scientific and technological importance
in recent years. Cd2+ as dopant and ZnS as shell material in AgInS2 /ZnS exhibit
both optical absorption and PL intensity considerably when arriving the above
architecture compared to core AgInS2 with the same orthorhombic crystal structure.
Core-multishell nanocrystals of AgInS2 -based type II/type I and type I/reverse
type I (AgInS2 /CdS/ZnS and AgInS2 /ZnS/CdS) exhibit their absorption as well as
emission in the red region. Alloying Zn1-xCdxS shell over AgInS2 core leads to
blue-shift emission which is due to decrease of the first excitation absorption and
retained the same orthorhombic crystal structure with slight deviations in diffraction
peak positions compared to core AgInS2 . EDX and their elemental mapping images
confirmed Ag, In, Zn, Cd, and S are distributed uniformly.
These strategies enhance the fluorescent property of the core (CuInS2 and
AgInS2 ) chalcogenide materials but at the same time improved the scope of
developing into many energy-related applications, by reducing Cd2+ from its
toxicity in view of core location being coated by inner and outer shells. The
strategies adopted in the present work will be useful in designing new opto-electric
materials with high luminescent property and also in many biological fields.

References

1. Zhanga, Q., Cao, G.: Hierarchically structured photoelectrodes for dye-sensitized solar cells.
J. Mater. Chem. 21, 6769–6774 (2011)
2. Henglein, A.: Small-particle research: physicochemical properties of extremely small colloidal
metal and semiconductor particles. Chem. Rev. 89(8), 1861–1873 (1989)
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 63

3. Spanhel, L., Weller, H., Henglein: Photochemistry of semiconductor colloids. 22. Electron
ejection from illuminated cadmium sulfide into attached titanium and zinc oxide particles. J.
Am. Chem. Soc. 109(22), 6632–6635 (1987)
4. Radi, A., et al.: Nanoscale shape and size control of cubic, cuboctahedral, and octahedral
Cu−Cu2 O core−shell nanoparticles on Si(100) by one-step, templateless, capping-agent-free
electrodeposition. ACS Nano. 4(3), 1553–1560 (2010)
5. Zhao, H., Chaker, M., Ma, D.: Effect of CdS shell thickness on the optical properties of water-
soluble, amphiphilic polymer-encapsulated PbS/CdS core/shell quantum dots. J. Mater. Chem.
21, 17483–17491 (2011)
6. Lambert, K., Geyter, B.D., Moreels, I., Hens, Z.: PbTe|CdTe core|shell particles by cation
exchange, a HR-TEM study. Chem. Mater. 21, 778–780 (2009)
7. Peng, X., et al.: Epitaxial growth of highly luminescent CdSe/CdS core/shell nanocrystals with
photostability and electronic accessibility. J. Am. Chem. Soc. 119(30), 7019–7029 (1997)
8. Hines, M.A., Guyot-Sionnest, P.: Synthesis and characterization of strongly luminescing ZnS-
capped CdSe nanocrystals. J. Phys. Chem. 100, 468–471 (1996)
9. Vasudevan, D., et al.: Core-shell quantum dots: properties and applications. J. Alloys Compd.
636, 395–404 (2015)
10. EychmA Culler, A., Mews, A., Weller, H.: A quantum dot quantum well: CdS/HgS/CdS. Chem.
Phys. Lett. 208, 59–62 (1993)
11. Renuga, V., Neela Mohan, C., Mohamed Jaabir, M.S., Arul Prakash, P., Navaneethan, M.:
Synthesis and surface passivation of CuInS2 /MnS/ZnS Core-multishell nanocrystals, their
optical, structural, and morphological characterization, and their bioimaging applications. Ind.
Eng. Chem. Res. 57, 15703–11572 (2018)
12. Asgary, S., Mirabbaszadeh, K., Nayebi, P., Emadi, H.: Synthesis and investigation of optical
properties of TOPO-capped CuInS2 semiconductor nanocrystal in the presence of different
solvent. Mater. Res. Bull. 51, 411 (2014)
13. Nag, A., Sarma, D.D.: White light from Mn2+ -doped CdS nanocrystals. J. Phys. Chem. C. 111,
641 (2007)
14. Castro, S.L., Bailey, S.G., Raffaelle, R.P., Banger, K.K., Hepp, A.F.: Synthesis and character-
ization of colloidal CuInS2 nanoparticles from a molecular single-source precursor. J. Phys.
Chem. B. 108, 12429 (2004)
15. Yu, X.F., Peng, X.N., Chen, Z.Q., Lian, C., Su, X.R., Li, J.B., Li, M., Liu, B.L., Wang,
Q.Q.: High temperature sensitivity of manganese-assisted excitonic photoluminescence from
inverted core/shell ZnSe:Mn/CdSe nanocrystals. Appl. Phys. Lett. 96, 123104 (2010)
16. Vlaskin, V.A., Janssen, N., van Rijssel, J., Beaulac, R., Gamelin, D.R.: Tunable dual emission
in doped semiconductor nanocrystals. Nano Lett. 10, 3670 (2010)
17. Quan, Z.N., Wang, Z.L., Yang, D.P., Lin, J., Fang, J.Y.: Synthesis and characterization of
high quality ZnS, ZnS:Mn2+ , and ZnS:Mn2+ /ZnS (core/shell) luminescent nanocrystals. Inorg.
Chem. 46, 1354 (2007)
18. Nam, D.E., Song, W.S., Yang, H.: Facile, air-insensitive solvothermal synthesis of emission-
tunable CuInS2 /ZnS quantum dots with high quantum yields. J. Mater. Chem. 21, 18220
(2011)
19. Hua, J., Zhang, Y., Yuan, X., Cheng, H., Meng, X., Zhao, J., Li, H.: Photoluminescence
properties of Cu–Mn–In–S/ZnS core/shell quantum dots. Superlattice. Microst. 73, 214 (2014)
20. Li, M., Zhao, Q., Yi, X., Zhong, X., Song, G., Chai, Z., Liu, Z., Yang, K.: Au@MnS@ZnS
core/shell/shell nanoparticles for magnetic resonance imaging and enhanced cancer radiation
therapy. ACS Appl. Mater. Interfaces. 8, 9557 (2016)
21. Zhang, R., Yang, P., Wang, Y.: Facile synthesis of CuInS2 /ZnS quantum dots with highly near-
infrared photoluminescence via phosphor-free process. J. Nanopart. Res. 15, 1910 (2013)
22. Sitbon, G., Bouccara, S., Tasso, M., Francois, A., Bezdetnaya, L., Marchal, F., Beaumonte,
M., Pons, T.: Multimodal Mn-doped I–III–VI quantum dots for near infrared fluorescence and
magnetic resonance imaging: from synthesis to in vivo application. Nanoscale. 6, 9264 (2014)
23. Park, J., Kim, S.W.: CuInS2 /ZnS core/shell quantum dots by cation exchange and their blue-
shifted photoluminescence. J. Mater. Chem. 21, 3745 (2011)
64 V. Renuga and C. Neela Mohan

24. Xing, H., Wei, T., Lin, X., Dai, Z.: Near-infrared MnCuInS/ZnS@BSA and urchin-like Au
nanoparticle as a novel donor-acceptor pair for enhanced FRET biosensing. Anal. Chim. Acta.
1042, 71 (2018)
25. Liu, W., Zhang, Y., Zhai, W., Wang, Y., Zhang, T., Gu, P., Chu, H., Zhang, H., Cui, T., Wang,
Y., Zhao, J., Yu, W.W.: Temperature-dependent photoluminescence of ZnCuInS/ZnSe/ZnS
quantum dots. J. Phys. Chem. C. 117, 19288 (2013)
26. Sun, J., Ikezawa, M., Wang, X., Jing, P., Li, H., Zhao, J., Masumoto, Y.: Photocarrier
recombination dynamics in ternary chalcogenide CuInS2 quantum dots. Phys. Chem. Chem.
Phys. 17, 11981 (2015)
27. Jo, D.Y., Yang, H.: Synthesis of highly white-fluorescent CuGaS quantum dots for solid-state
lighting devices. Chem. Commun. 52, 709 (2016)
28. Jo, D.Y., Kim, D., Kim, J.H., Chae, H., Seo, H.J., Do, Y.Y., Yang, H.: Tunable white fluorescent
copper gallium sulfide quantum dots enabled by Mn doping. ACS Appl. Mater. Interfaces. 8,
12291 (2016)
29. Jiang, T., Zhang, W., Song, J., Yang, M., Wang, H., Xia, R., Ye, X., Zhu, L., Wang, H., Xu, X.:
Aqueous synthesis of color tunable Cu doped Zn–In–S/ZnS nanoparticles in the whole visible
region for cellular imaging. J. Mater. Chem. B. 3, 2402 (2015)
30. Langer, D., Richter, H.: Zero-phonon lines and phonon coupling of ZnSe: Mn and CdS: Mn.
Phys. Rev. 146, 554 (1966)
31. Beaulac, R., Archer, P.I., Liu, X., Lee, S., Salley, G.M., Dobrowolska, M., Furdyna, J.K.,
Gamelin, D.R.: Spin-polarizable excitonic luminescence in colloidal Mn2+ -doped CdSe
quantum dots. Nano Lett. 8, 1197 (2008)
32. Hazarika, A.: Ultra narrow and widely tunable Mn2+ -induced photoluminescence from single
Mn-doped nanocrystals of ZnS-CdS alloys. Phys. Rev. Lett. 110, 267401 (2013)
33. Yang, Y., Chen, O., Angerhofer, A., Cao, Y.C.: On doping CdS/ZnS core/shell nanocrystals
with Mn. J. Am. Chem. Soc. 130, 15649 (2008)
34. Zhang, J., Xie, R., Yang, W.: A simple route for highly luminescent quaternary Cu-Zn-In-S
nanocrystal emitters. Chem. Mater. 23, 3357 (2011)
35. Zheng, J., Ji, W., Wang, X., Ikezawa, M., Jing, P., Liu, X., Li, H., Zhao, J., Masumoto, Y.:
Improved photoluminescence of MnS/ZnS core/shell nanocrystals by controlling diffusion of
Mn ions into the ZnS shell. J. Phys. Chem. C. 114, 15331 (2010)
36. Dabbousi, B.O., Rodriguez-Viejo, J., Mikulec, F.V., Heine, J.R., Mattoussi, H., Ober, R.,
Jensen, K.F., Bawendi, M.G.: (CdSe)ZnS core−shell quantum dots: synthesis and character-
ization of a size series of highly luminescent nanocrystallites. J. Phys. Chem. B. 101, 9463
(1997)
37. Steckel, J.S., Zimmer, J.P., Coe-Sullivan, S., Stott, N.E., Bulovic, V., Bawendi, M.G.: Blue
luminescence from (CdS)ZnS core-shell nanocrystals. Angew. Chem. Int. Ed. 43, 2154 (2004)
38. Eychmuller, A.: Structure and photophysics of semiconductor nanocrystals. J. Phys. Chem. B.
104, 6514 (2000)
39. Lad, A.D., Mahamuni, S.: Effect of ZnS shell formation on the confined energy levels of ZnSe
quantum dots. Phys. Rev. B. 78, 125421 (2008)
40. Lee, J.W., Hall, A.S., Kim, J.D., Mallouk, T.E.: A facile and template-free hydrothermal
synthesis of Mn3 O4 nanorods on graphene sheets for supercapacitor electrodes with long cycle
stability. Chem. Mater. 24, 1158 (2012)
41. Liu, Q., Deng, R., Ji, X., Pan, D.: Alloyed Mn–Cu–In–S nanocrystals: a new type of diluted
magnetic semiconductor quantum dots. Nanotechnology. 23, 255706 (2012)
42. Cao, S., Zheng, J., Zhao, J., Wang, L., Gao, F., Wei, G., Zeng, R., Tian, L., Yang, W.: Highly
efficient and well-resolved Mn2+ ion emission in MnS/ZnS/CdS quantum dots. J. Mater. Chem.
C. 1, 2540 (2013)
43. Zhang, G., Monllor-Satoca, D., Choi, W.: Band energy levels and compositions of CdS-based
solid solution and their relation with photocatalytic activities. Cat. Sci. Technol. 3, 1790 (2013)
44. Feng, Z., Dai, P., Ma, X., Zhan, J., Lin, Z.: Monodispersed cation-disordered cubic AgInS2
nanocrystals with enhanced fluorescence. Appl. Phys. Lett. 96, 013104 (2010)
Design, Synthesis, and Properties of I-III-VI2 Chalcogenide-Based Core-. . . 65

45. Giribabu, K., Suresh, R., Manigandan, R., Vijayaraj, A., Prabu, R., Narayanan, V.: Cadmium
sulphide nanorods: synthesis, characterization and their photocatalytic activity. Bull. Kor.
Chem. Soc. 33, 9 (2012)
46. Orii, T., Kaito, S., Matsuishi, K., Onari, S., Arai, T.: Photoluminescence of CdS nanoparticles
suspended in vacuum and its temperature increase by laser irradiation. J. Phys. Condens.
Matter. 14, 9743 (2002)
47. Yao, J., Zhao, G., Wang, D., Han, G.: Solvothermal synthesis and characterization of CdS
nanowires/PVA composite films. Mater. Lett. 59, 3652 (2005)
48. Yang, P., Lu, M., Xu, D., Yuan, D., Zhou, G.: Synthesis and photoluminescence characteristics
of doped ZnS nanoparticles. Appl. Phys. A Mater. Sci. Process. 73, 455 (2001)
49. Lo, S.S., Mirkovic, T., Chuang, C.H., Burda, C., Scholes, G.D.: Emergent properties resulting
from type-II band alignment in semiconductor nanoheterostructures. Adv. Mater. 23, 180
(2011)
50. Sivasubramanian, V., Arora, A.K., Premila, M., Sundar, C.S., Sastry, V.S.: Optical properties
of CdS nanoparticles upon annealing. Phys. E. 31, 93 (2006)
51. Chen, M., Kim, Y.N., Li, C., Cho, S.O.: Controlled synthesis of hyperbranched cadmium
sulfide micro/nanocrystals. Cryst. Growth Des. 8, 629 (2008)
52. Yao, W.T., Yu, S.H., Liu, S.J., Chen, J.P., Liu, X.M., Li, F.Q.: Architectural control syntheses
of CdS and CdSe nanoflowers, branched nanowires, and nanotrees via a solvothermal approach
in a mixed solution and their photocatalytic property. J. Phys. Chem. B. 110, 11704 (2006)
53. Chen, S., Zaeimian, M.S., Jorge, H.S.K., Monteiro, Zhao, J., Mamalis, A.G., A. de B-Dias,
Zhu, X.: Mn doped AIZS/ZnS nanocrystals: synthesis and optical properties. J. Alloys Compd.
725, 1077 (2017)
54. Wang, X., Xie, C., Zhong, J., Liang, X., Xiang, W.: Synthesis and temporal evolution of Zn-
doped AgInS2 quantum dots. J. Alloys Compd. 648, 127 (2015)
55. Podgurska, I., Rachkov, A., Borkovska, L.: Effect of Pb2+ ions on photoluminescence of ZnS-
coated AgInS2 nanocrystals. Phys. Status Solidi A. 215, 1700450 (2017)
56. Zeng, Z., Wang, A., Ping, L., Yang, J., Wang, Q.: Encapsulation of lanthanides in ternary I–
III–VI AgInS2 nanocrystals and their physical properties. Mater. Lett. 141, 225 (2015)
57. Mao, B., Chuang, C.H., McCleese, C., Zhu, J., Burda, C.: Near-infrared emitting AgInS2 /ZnS
nanocrystals. J. Phys. Chem. C. 118, 13883 (2014)
58. Powder Diffraction File; The JCPDS International Centre for Diffraction Data: Swarthmore,
PA, 1990; No. 251330
59. Hamanaka, Y., Yukitoki, D., Kuzuya, T.: Structural transformation and photoluminescence
modification of AgInS2 nanoparticles induced by ZnS shell formation. Appl. Phys. Express. 8,
095001 (2015)
60. Liao, S., Huang, Y., Zhang, Y., Shan, X., Yan, Z., Shen, W.: Highly enhanced photolumines-
cence of AgInS2 /ZnS quantum dots by hot-injection method. Mater. Res. Express. 2, 015901
(2015)
61. Tan, L., Liu, S., Li, X., Chronakis, I.S., Shen, Y.: A new strategy for synthesizing AgInS2
quantum dots emitting brightly in near-infrared window for in vivo imaging. Colloids Surf. B:
Biointerfaces. 125, 222 (2015)
62. Ivanov, S.A., Piryatinski, A., Nanda, J., Tretiak, S., Zavadil, K.R., Wallace, W.O., Werder, D.,
Klimov, V.I.: Type-II core/shell CdS/ZnSe nanocrystals: synthesis, electronic structures, and
spectroscopic properties. J. Am. Chem. Soc. 129, 11708 (2007)
63. Zhong, H., Bai, Z., Zou, B.: Tuning the luminescence properties of colloidal I–III–VI
semiconductor nanocrystals for optoelectronics and biotechnology applications. J. Phys. Chem.
Lett. 3, 3167 (2012)
64. Park, Y.J., Oh, J.H., Han, N.S.: Photoluminescence of band gap states in AgInS2 nanoparticles.
J. Phys. Chem. C. 118, 25677 (2014)
65. Terai, Y., Kuroda, S., Takita, K., Okuno, T., Masumoto, Y.: Zero-dimensional excitonic
properties of self-organized quantum dots of CdTe grown by molecular beam epitaxy. Appl.
Phys. Lett. 73, 3757 (1998)
66 V. Renuga and C. Neela Mohan

66. Gabka, G., Bujak, P., Kotwica, K., Ostrowski, A., Lisowski, W., Sobczakb, J.W., Prona,
A.: Luminophores of tunable colors from ternary Ag–In–S and quaternary Ag–In–Zn–S
nanocrystals covering the visible to near-infrared spectral range. Phys. Chem. Chem. Phys.
19, 1217 (2017)
67. Wang, S., Li, J.J., Lv, Y., Wu, R., Xing, M., Shen, H., Wang, H., Li, L.S., Chen, X.: Synthesis
of reabsorption-suppressed type-II/type-I ZnSe/CdS/ZnS core/shell quantum dots and their
application for immunosorbent assay. Nanoscale Res. Lett. 12, 380 (2017)
68. Boldt, K., Schwarz, K.N., Kirkwood, N., Smith, T.A., Mulvaney, P.: Electronic structure
engineering in ZnSe/CdS type-II nanoparticles by interface alloying. J. Phys. Chem. C. 118,
13276 (2014)
Unique Luminescent Properties
of Composition-/Size-Selected Aqueous
Ag-In-S and Core/Shell Ag-In-S/ZnS
Quantum Dots

Oleksandr Stroyuk, Oleksandra Raievska, and Dietrich R. T. Zahn

Abstract Ternary metal-chalcogenide quantum dots (QDs), including copper and


silver-indium-sulfide QDs and related core/shell QDs, combine broad tunability
of optical properties through extensive variations in the QD composition and size
with a unique photoluminescence (PL) emission behavior. In particular, the PL
of nonstoichiometric Ag-In-S (AIS) QDs and core/shell AIS/ZnS QDs reveals a
combination of properties typical for excitonic emission (high quantum yields,
distinct size dependence of spectral PL parameters, characteristic temperature
dependences) while simultaneously revealing a large spectral width of emission
bands typically observed for defect-related radiative recombination. In this view,
the origins and mechanisms of the broadband PL of such ternary QDs remain a
topic of vivid discussions.
This chapter is an account of our recent efforts in aqueous syntheses of
composition- and size-selected AIS and AIS/ZnS QDs with bright and multicolor
PL. We provide an overview of our methods of stabilizing AIS and AIS/ZnS QDs
in aqueous colloids by small multifunctional molecular ligands using the example
of biocompatible glutathione (GSH) and discuss dependences of absorption and
PL characteristics of the nonstoichiometric AIS and core/shell AIS/ZnS QDs on
their composition and size as well as temperature. We focus on the origins of the

O. Stroyuk ()
Forschungszentrum Jülich GmbH, Helmholtz-Institut Erlangen Nürnberg für Erneuerbare
Energien (HI ERN), Erlangen, Germany
e-mail: ostroyuk@fz-juelich.de
O. Raievska
Semiconductor Physics, Chemnitz University of Technology, Chemnitz, Germany
L.V. Pysarzhevsky Institute of Physical Chemistry, National Academy of Sciences of Ukraine,
Kyiv, Ukraine
e-mail: oleksandra.raievska@physik.tu-chemnitz.de
D. R. T. Zahn
Semiconductor Physics, Chemnitz University of Technology, Chemnitz, Germany
e-mail: zahn@physik.tu-chemnitz.de

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 67
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_3
68 O. Stroyuk et al.

broadband PL of AIS and AIS/ZnS QDs discussing the structure of the QD PL


bands of different size, their time-resolved PL behavior, temperature dependences
in a broad 10–300 K range, as well as our recent results from PL studies of single
AIS/ZnS QDs.

Keywords Luminescent properties · Aqueous colloidal quantum dots ·


Glutathione-capped AgInS2 quantum dots · AgInS2/ZnS core/shell quantum
dots · Temperature-dependent optical properties

1 Introduction

Ternary metal-chalcogenide quantum dots (QDs), such as CuInS2 (CIS) and AgInS2
(AIS), reveal an unrivaled tunability of optical and other physicochemical properties
due to the broad variations in QD composition and size as well as tolerance to
abundant doping [1–7]. This variability of application-relevant features combined
with relatively high photoluminescence (PL) quantum yields (QYs) reaching 60–
80% raised strong interest in using these ternary In-based QDs in optoelectronic
systems, such as photodetectors and light-emitting devices, in bioimaging and
biosensing systems, as well as in photocatalytic systems and solar cells [1, 3–7].
At the same time, the nature of the excited states of CIS and AIS QDs and the
mechanisms of light absorption and PL emission remain topics of vivid discussions
[1, 4, 8].
Ternary CIS and AIS QDs as well as more frequently studied and more stable
core/shell CIS/ZnS and AIS/ZnS QDs reveal many special features in absorption
and PL spectra. These include broad absorption bands with relatively intense sub-
bandgap (Urbach) tails that can even mask the fundamental absorption band edges
of CIS and AIS QDs [9, 10], broad PL bands with spectral widths of 200–400 meV,
and an inhomogeneity of the PL lifetimes across the PL band [1, 4, 8, 9, 11–14].
The broadband character of PL admittedly compromises to some extent the
perspectives of such QDs in applications requiring a high purity of the emitted light
[3, 15, 16]. At the same time, a large spectral width (or full width at half maximum –
FWHM) of the PL bands results in spectacular visual changes of the emission
color even at comparatively tiny variations of the PL band position, allowing
for multicolored and bright luminophores attractive for biomedical labeling and
multicolored cell tracking [2, 3, 5, 16].
By analogy with well-studied binary II–VI and IV–VI QDs, such as CdS, CdSe,
and PbS, the PL properties of ternary CIS and AIS as well as related core/shell QDs
are often rationalized in terms of a donor-acceptor (D-A) PL model [1, 4, 9, 12]
with contributions from size and chemical inhomogeneities on the single dot basis.
According to the D-A model, the photogenerated charge carriers are rapidly trapped
by lattice defects such as interstitial atoms, cation/anion vacancies, as well as surface
states, assumed to be abundant in ternary QDs especially in nonstoichiometric
ones, that can introduce local states within the bandgap. The trapped carriers can
then recombine radiatively or non-radiatively with free/trapped carriers of opposite
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 69

charge. Alternatively, they can be detrapped by thermal activation or recombine via


a tunneling mechanism. The radiative recombination of these electron-hole pairs
results in the emission of photons with a distribution of energies, broadened by
inhomogeneities in QD size and chemical composition as well as the distribution in
depths/energy position of the trap states and the distances between trapped electrons
and holes within a single particle [17–19].
However, several experimental observations cannot or can only hardly be
explained by the D-A model. In particular, numerous studies showed the spectral
width of the PL bands of CIS and AIS QDs to be unaffected by the average QD
size and size distribution [11, 14, 20], temperature [11, 12, 20, 21], and excitation
intensity variations [12, 20], despite the fact that all these factors are expected
to alter the distribution of the emissive centers formed upon light absorption. In
particular, we reported recently the aqueous syntheses of nonstoichiometric CIS
and AIS QDs with a relatively narrow size distribution (2–4 nm) stabilized by
mercaptoacetic acid (MAA) and GSH [14] revealing PL QY values as high as
40–45% (MAA-capped AIS/ZnS QDs) and 65–70% (GSH-capped AIS/ZnS QDs).
Such high PL QYs can hardly originate from an emission mediated solely by QD
lattice defects. For example, in binary metal-sulfide QDs, the participation of defect
states in the non-radiative recombination of the charge carriers leads to a relatively
weak broad PL with a PL QY usually below 5% [17–19].
An alternative explanation of the unique PL properties of ternary QDs can be
found in the model of self-trapped excitons (STE). This model assumes one of
the photogenerated charge carriers (typically the hole) to be localized at a certain
lattice site resulting in a considerable lattice distortion and strong electron-phonon
interaction [1, 8, 17, 22]. In this context, the broad PL band is interpreted as a
series of phonon replicas of the pure excitonic (zero-phonon) emission line (ZPL)
similarly to the molecular fluorescence with a distinct vibrational structure [23].
Gamelin et al. applied the STE model to the broadband emission of copper- and
silver-doped cadmium chalcogenide QDs and argued that this model can also
describe the PL features of CIS and AIS QDs as ultimate cases of Cu and Ag
doping [8, 22]. Indeed, Cu- and Ag-doped CdS and CdSe QDs reveal PL properties
resembling those of CIS and AIS QDs. We showed recently that the PL band
parameters of Cu- and Ag-doped binary CdS and CdSe/CdS and ternary AIS and
AIS/ZnS QDs synthesized using the same or similar ligands can be attributed to the
same strong electron-phonon interaction and can be modeled using the general STE
model and corresponding vibrational parameters of bulk semiconductors [10, 12,
14].
The STE model allows the broadband PL of ternary QDs to be described
without assumptions about the exact nature, density, and energies of lattice defects.
This model has two important consequences: (1) the broadband PL is assumed to
originate from the QD lattice as a whole, not from specific defects, and, therefore,
by a careful design of the lattice, ternary QDs can be realized with extremely high
PL QYs as already reached for binary metal-chalcogenide QDs [17–19]; (2) the
broadband PL is an inherent property of each and every single ternary QD and can
only be very slightly modulated by variations of QD ensemble properties.
70 O. Stroyuk et al.

Currently however, reports of experimental evidence supporting the STE for


ternary QDs are quite scattered and need to be extracted from many studies to obtain
a single and consistent picture of the PL mechanism. In this view, we aim in the
present chapter to provide a comprehensive overview of our recent experimental
findings supporting the STE interpretation of the PL phenomena in aqueous ternary
QDs. Our works performed in the last 5 years (2015–2019) encompass a number of
aqueous ternary QDs, including MAA-capped CIS QDs [24, 25], MAA-stabilized
AIS QDs [13, 24], copper-doped AIS QDs [26], GSH-capped AIS QDs [14, 27],
and corresponding core/shell QDs with ZnS shells.
In this chapter, we focus mainly on core AIS and core/shell AIS/ZnS QDs
capped with surface complexes with GSH that reveal the highest stability and
reproducibility of optical properties. In particular, we discuss the aqueous syntheses
of composition- and size-selected AIS and AIS/ZnS QDs with bright multicolor PL
and dependences of absorption and PL characteristics of the nonstoichiometric AIS
and core/shell AIS/ZnS QDs on their composition and size as well as temperature.
We focus on the origins of the broadband PL of AIS and AIS/ZnS QDs, the
structure of the PL bands of QDs of different sizes, their time-resolved PL behavior,
temperature PL dependences in a broad temperature range of 10–300 K, as well as
our recent results from the PL studies of single AIS/ZnS QDs.
The absorption and emission spectra of QD ensembles typically produced by
colloidal synthesis always reflect the collective properties of many single particles
in the QD ensemble. In terms of the D-A model of the radiative electron-hole
recombination, the distribution of the PL quantum energy EPL can be assessed using
Eq. (1) [8, 9, 12, 20]:

e2
EPL = Eex − (ED + EA ) + (1)
εR
Here Eex is the optical bandgap, ED and EA are the binding energies of the donor
and acceptor sites that trap a hole and an electron, e and ε are the electron charge
and dielectric constant of the semiconductor, and R is the distance between a donor
and an acceptor state participating in the generation of PL.
Equation (1) implies that several distributions and sources of inhomogeneity
can affect the EPL energies of the emitted photons and, thus, the spectral width
FWHM of the resulting PL band. In the case of QDs with sizes below the exciton
diameter that exhibit spatial confinement effects, the energy of the interband electron
transitions depends on the QD size d, thus introducing a distribution in Eg (d)
[17]. The relative contribution of this factor to the FWHM of the PL band can be
evaluated by exploring optical properties of size-selected series of QDs provided
that other factors like the QD chemical composition, ligand shell, temperature, and
PL registration conditions are maintained unchanged. In the present chapter, we
focus on the methods of size-selective precipitation that, when applied to ternary
AIS and AIS/ZnS QDs, allows series of QDs of different size to be produced from
the same maternal QD ensemble for any given composition of the ensemble. Studies
of the optical properties of such size- and composition-selected series allow for a
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 71

reliable evaluation of size effect on the PL emission from ternary QDs not interfered
by other possible influences.
The photogenerated electrons and holes are trapped by acceptor and donor
states introduced by various lattice defects, with the diversity of the possible
defects yielding a distribution of the “energy depth” of the electron and hole trap
states or EA and ED levels. Under the influence of the thermal lattice energy,
the trapped charge carriers can “hop” between trap states with closely matching
energies, which also contributes to the FWHM of the PL band [17, 18, 28, 29].
The relative contribution of this factor to the total PL band broadening can be
evaluated by a partial elimination of the defect states either reversibly by decreasing
the temperature and freezing out the thermally activated detrapping of the charge
carriers or irreversibly by annealing of the QDs or by modifying the QD surface
with a passivating shell. The contribution of such temperature-dependent variation
to the PL properties of AIS and AIS/ZnS QD ensembles is discussed in detail.
The third contribution to PL broadening arises from the different distances R
between the trapped electrons and holes, as follows from the last term of Eq. (1).
The recombination of closer lying charge carrier pairs results in a larger Coulombic
contribution to the PL energy. The relative contribution of this factor is the most
challenging one to be evaluated directly. This phenomenon is affected by both the
QD size variation, with a narrower distribution of R expected for smaller QDs, and
by parameters like temperature and excitation energy.
A collection of experimental data discussed later in this chapter shows that
the D-A model of PL emission, when applied to ternary composition- and size-
selected AIS and CIS QDs (and corresponding core/shell QDs), reveals a number of
inconsistencies and, therefore, a more realistic approach for the description of PL
phenomena in ternary QDs is required.

2 Aqueous Synthesis and Optical Properties of AIS and


AIS/ZnS QDs

The most brightly emitting ternary QDs are typically produced by heating-up
and hot-injection protocols elaborated earlier in great detail for binary metal-
chalcogenide QDs. These syntheses utilize high-boiling-point organic solvents
(such as octadecene or trioctylphosphine) and long-chain surface-capping ligands
(1-dodecanethiol or oleylamine) to precisely control the size and surface perfection
of the forming QDs [2, 5, 30–32]. As the biosensing applications require the QDs
to be water-soluble and ready for the conjugation to various biomolecules, the
as-prepared luminescent QDs need to be subjected to a ligand exchange with water-
soluble agents and phase transfer from the original organic oils to water [5, 16,
30–33]. The most efficient phase transfer and stabilization in aqueous solutions
is typically achieved with multifunctional transfer agents such as mercaptoacids
(MAA or mercaptopropionic acid), cysteine, or GSH that can bind strongly to the
72 O. Stroyuk et al.

QD surface by a mercapto-group and simultaneously prohibit the QD aggregation


in water due to the electrostatic repulsion between the carboxyl anions [32, 33].
As compared to other water-soluble multifunctional ligands (mercapto-acids,
cysteine, etc.), GSH exhibits an excellent resistivity toward oxidation and hydrolysis
[34] and provides GSH-capped QDs with a high stability against aggregation. GSH
was used to transfer PbS [34] or CIS [35] QDs from organic media to water via
ligand exchange. This approach is very popular for binary II–VI metal-chalcogenide
QDs; however, in the case of ternary I-III-VI compounds, it results in a considerable
loss of the PL efficiency [34, 36]. For example, the phase transfer of CIS QDs
originally capped with oleylamine with a PL QY of 70% to water mediated by GSH
yields aqueous QDs with a PL QY of only 34% [36].
In this view, attempts of direct syntheses of luminescent GSH-stabilized QDs in
aqueous media have constantly been conducted. In particular, successful syntheses
were reported for GSH-capped ZnS [37], CdS [38, 39], PbS [40], doped ZnS [41],
Cdx Zn1–x S [42], HgS [43], CdSe [44], ZnSe [45–48] QDs, alloyed ZnSex Te1-x [49],
Cdx Zn1–x Se [48], and Cdx Zn1–x Te [50] QDs, as well as Ag2 Se [51] and CdTe [52–
57] QDs. Glutathione can act not only as a surface-binding ligand but also as a
precursor for the in situ formation of a ZnS shell [46, 53] as well as a reducing
agent for Se precursors in the syntheses of metal selenide QDs [51].
Recently, GSH was shown to serve as an efficient stabilizer for direct and aqueous
syntheses of composition- and size-selected ternary I-III-VI QDs, such as CIS [58–
61], AIS [36, 62], and ZAIS [63]. Luminescent AIS QDs were reported to form
when treating bovine serum albumin-stabilized Ag2 S QDs with an In(III) complex
with GSH [64]. The treatment results in a simultaneous In(III) incorporation into the
QD structure and a surface ligand exchange. The capability of GSH to form stable
complexes and even coordination polymers with metal cations can be used to tune
finely the size and size distribution of the final metal-chalcogenide QDs, as reported
for CdS QDs [39].
The capping with the functionality-rich GSH allows for the post-synthesis
conjugation of QDs to biomolecules while preserving the PL efficiency. Such QD-
biomolecule conjugates were used as luminescent contrast agents for living cell
tracking and labeling [39, 51, 52, 56, 59–64]. By capping QDs selectively with L or
D optical isomers of GSH, chiral QDs can be produced for focused cell toxicity and
sensing studies [54].

2.1 Optimization of the Synthesis of Aqueous AIS


and AIS/ZnS QD Ensemble Colloids

The composition of AIS QDs can be expressed as the molar ratio xAg :xIn :xS
of the corresponding components. We varied systematically the xAg , xIn , and xS
parameters, the concentration of the capping agent, the temperature and the duration
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 73

Fig. 1 Absorption (a) and PL spectra (b) of colloidal AIS QDs produced with different silver
contents (as indicated in the figures); In:S = 4:5. (c) Bandgap (curve 1), PL band maximum (curve
2), and integral PL intensity (curve 3) of AIS QDs as a function of molar Ag/In ratio

of the thermal treatment, as well as the mode and the conditions of the ZnS shell
formation to achieve the highest possible PL QY.
We found that an increase in the Ag/In ratio results in an increase of the integral
absorbance of the colloidal AIS-GSH solutions as well as in a steady shift of
the absorption band edge to lower energies (Fig. 1a). In contrast to AIS colloids
stabilized by mercaptoacids, the GSH-based systems demonstrate quite distinct
absorption band thresholds allowing for a reliable estimation of the bandgap Eg
by replotting the spectra in Tauc coordinates for direct transitions and deducing
the intercepts with the abscissa. The bandgap was found to decrease monotonically
from around 2.8 eV to ≈2.1 eV as the Ag/In ratio was increased from 0.05 to 0.5
(Fig. 1c, curve 1). The Ag 4d orbitals have a strong contribution to the valence band
(VB) “top” near the edge [8, 15], and, therefore, an increase in the silver content
expectedly results in an increase in the absorption and a bandgap narrowing.
The AIS-GSH QDs emit PL in broad bands with the position and intensity being
strongly dependent on the silver content (Fig. 1b). As the Ag/In ratio is increased,
the PL band peak shifts to lower energies from ≈2.1 eV to 1.8 eV (Fig. 1c, curve
2), while the PL intensity increases, reaching a maximum at an Ag/In ratio around
0.25 and then decreasing for Ag-richer QDs (Fig. 1c, curve 3).
Similarly, the spectral parameters of the PL band depend considerably on the
S/Ag ratio, while the Ag/In ratio is kept constant (Fig. 2a). The PL intensity
increases with increasing S/Ag ratio, displays a maximum at S/Ag ≈ 5, and then
decreases for lower silver contents (Fig. 2b). When plotted as a topographic surface,
both dependences of the PL intensity on the sulfur and silver contents produce a
maximum corresponding to Ag/S ratios from 1:4 to 1:5 (Fig. 2c). In this view,
we selected QDs with a composition corresponding to xAg :xIn :xS = 1:4:5 for the
following studies.
The AIS QDs form in a reaction between GSH complexes with Ag(I) and In(III)
and hydrosulfide ions at 96–98 ◦ C. The duration of such a thermal treatment affects
the optical properties of the QDs in quite a decisive manner. In particular, the
absorption band edge of AIS QDs (Ag:In:S = 1:4:5) shows a steady shift to lower
74 O. Stroyuk et al.

Fig. 2 (a) PL spectra of colloidal AIS QDs produced with different S/Ag ratios (as indicated in the
figure); In/Ag = 4. (b) Integral PL intensity as a function of the ratio of sulfur to silver. (c) 2D map
of the variation of PL intensity of AIS QDs with xAg and xS at a constant xIn = 4; Ag:In:S = 1:4:5

Fig. 3 (a, b) Absorption (a) and PL spectra (b) of colloidal AIS QDs produced at different
durations of heating (as indicated in the figures); Ag:In:S = 1:4:5. (c) Bandgap (curve 1), PL
band maximum (curve 2), and integral PL intensity (curve 3) of AIS QDs as a function of the
duration of the thermal treatment

energies as the treatment duration is increased (Fig. 3a). The bandgap shrinks from
2.85 eV before heating to ≈2.3 eV after an 80-min treatment (Fig. 3c, curve 1),
which is most probably related to a growth of the QDs. The PL band maximum EPL
follows this trend showing a shift from ≈2.1 eV to 1.85 eV (Fig. 3c, curve 2).
While the PL shift is moderate, the PL intensity increases markedly with heating
time (Fig. 3b), almost doubling after the 80-min treatment (Fig. 3c, curve 3). This
observation shows that, besides increasing the mean QD size, the heating results in
an increase of the lattice ordering of the AIS-GSH QDs leading to the elimination
of defects responsible for non-radiative electron-hole recombination.
Typically, heating of a colloidal QD ensemble results in the increase of the overall
PL intensity and a shift of the PL maximum toward lower energies. The former
is associated with the elimination of the sites responsible for the non-radiative
recombination as a result of the thermally induced lattice reconstruction [17].
Simultaneously with the lattice reconstruction, Ostwald ripening of the colloidal
QD ensemble occurs, leading to an increase of the average QD size, which is
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 75

commonly reflected by a red shift of the PL maximum. Both effects are observed
for AIS QDs (Fig. 3b, c). For the conventional D-A model of the broadband
PL originating from lattice defects, the heating-induced partial elimination of
sub-bandgap states associated with lattice defects should result in a considerable
narrowing of the PL band. However, this expectation contradicts the experimental
results presented in Fig. 3b.
As the heating of the colloidal QD ensemble proceeds, the integral PL intensity
increases by almost an order of magnitude, thereby confirming the suppression of
non-radiative recombination pathways. At the same time, we found that the FWHM
of the PL band is reduced from 0.43 eV to around 0.40 eV in the first 10 min of
the thermal treatment but remains almost unchanged during the rest of the heating
procedure.
Apparently more than half of the thermally induced increase in PL intensity
between 10 min and 80 min of the heating phase occurs without changes in PL
FWHM. Later in this chapter, we show that the size variation of the AIS QDs
does not affect appreciably the FWHM of the emission band and, therefore, all
variations of the FWHM during heating should be attributed to changes in the defect
state density. As no changes in FWHM are observed, we can conclude that any
possible reduction of the density of defect states responsible for the non-radiative
recombination processes is not reflected in the spectral width of the PL band.
Covering of AIS QDs with a ZnS shell was found to be an efficient way of
enhancing the PL emission of AIS-MAA QDs [13]. This approach is also valid for
the AIS-GSH system, the enhancement factor reaching ≈2 (Fig. 4a). The formation
of a ZnS shell results in a small shift of EPL to higher energies (Fig. 4a) as a result of
the well-reported diffusion of Zn2+ ions from the shell into the AIS lattice [15, 30].
The latter trend becomes more pronounced with an increased Zn/In ratio (Fig. 4b);
however, a saturation of the PL enhancement is achieved already at Zn:In = 3:1.
Similar to the previously discussed heat treatment, the formation of the ZnS shell
affects the PL FWHM only in the case of very small amount of ZnS (Fig. 4b). For
higher Zn-to-Ag ratios, the FWHM remains almost constant although the integral
PL intensity increases monotonously as the Zn/Ag ratio is elevated from 1 to 25–
30. This observation indicates that the ZnS deposition eliminates the lattice defects
both on the AIS QD core surface and in the QD volume by compensating the anion
vacancies and by inclusion of the Zn2+ into the lattice cation vacancies, thereby
blocking non-radiative decay channels, but obviously these lattice defects do not
contribute to the FWHM of the PL band.

2.2 Size Selection of AIS and AIS/ZnS QDs

The nature of the broadband PL of ternary CIS and AIS still remains a field of
discussion, because these QDs seem to reveal different electron-hole recombination
mechanisms as compared to both the corresponding bulk ternary semiconductors
and the broadly studied II–VI binary QDs [8, 15, 29, 65]. In particular, a generally
76 O. Stroyuk et al.

Fig. 4 (a) Absorption spectra (curves 1,2) and PL spectra (curves 1 ,2 ) of AIS QDs (curves 1)
and AIS/ZnS QDs (curves 2) produced at the same Ag:In:S:Zn ratio of 1:4:5:0 and 1:4:5:16,
respectively. (b) PL spectra of colloidal AIS/ZnS QDs produced with different zinc-to-indium
ratios (as indicated in the figure); Ag:In:S = 1:4:5

accepted model of trap-state-mediated radiative recombination typically evoked to


account for the large FWHM and Stokes shifts of the PL bands is currently ques-
tioned and revised in terms of its applicability to the ternary QDs [8, 15]. A deeper
insight into the mechanisms of PL emission requires to probe separately the QD
composition, size, doping, and surface capping, which is a challenge still to be met
in the case of ternary and quaternary In-based QDs. Of great help in this direction
can be synthetic approaches allowing one of the abovementioned factors to be varied
while leaving others untouched, such as size-selective precipitation/redispersion.
Size-selected series of various semiconductor QDs can be produced by their
selective precipitation from original polydisperse colloids. The precipitation is
induced by a measured addition of a “poor” solvent that mixes freely with the col-
loidal solution but decreases the aggregation stability of the QDs. The aggregation
involves the largest QDs in the present ensemble that flocculate and precipitate,
while smaller QDs remain dispersed in the colloidal solution. The QDs precipitate
and can be collected and redispersed and the entire procedure repeated many times
to produce a series of size-selected QDs. Such a method became “classical” for
the size-selective separation of cadmium chalcogenide QDs [19, 66–78], but it also
revealed general applicability and was successfully used for other semiconductors,
such as Si [79], InP [71, 80], InAs [77], CIS [81–83], CuGaS2 [84], AgIn(Ga)S2
[85], and Cu2 ZnSnS4 nanoparticle colloids [86, 87]. The reported examples of the
size-selective precipitation of ternary QDs are largely limited to colloids produced
by the hot-injection/heating-up methods.
We demonstrated the feasibility of the size-selection method for the deconvolu-
tion of an original colloidal ensemble of aqueous GSH-capped AIS QDs into a broad
series of fractions containing QDs with different average sizes and considerably
focused size distributions as compared to the parental ensemble [13, 14]. The
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 77

QDs were found to have similar compositions allowing the influence of size on
the spectral QD characteristics to be discriminated. In this section we focus on
the discussion of the nature of the broadband PL revealed by such selected QDs
with the FWHM being almost independent on the QD size and size distribution. We
argue that this inherent property of the ternary In-based QDs should be interpreted in
terms of a self-trapped exciton model and strong electron-phonon interaction, rather
than in the frame of the conventional trap-mediated donor-acceptor recombination
model.
The addition of an ample (volumetric) excess of 2-propanol results in the
instant coagulation of colloidal GSH-capped AIS and AIS/ZnS QDs, as well as
their mercaptoacetate-stabilized analogues. This procedure can be used for batch
separation of all AIS QDs from the parental solution for purification purposes
as the coagulate can be readily redissolved in deionized water yielding again
stable colloidal solutions. If the 2-propanol addition is performed in portions, a
selective precipitation of the largest QDs in the colloidal ensemble can be achieved.
By separating the coagulate and repeating this procedure, numerous fractions of
QDs differing in the optical properties, and, therefore, in the average size, can be
produced from the parental colloid.
Figure 5a exemplifies the absorption spectra of eight fractions produced from
“standard” parental AIS and AIS/ZnS QD colloids (Ag:In:S = 1:4:5, 60-min
heating). As the fraction number increases from 1 to 8, the absorption onset of the
AIS QDs shifts to lower wavelengths (higher energies) clearly indicating that the
size of selectively precipitated QDs decreases resulting in an enhancement of the
spatial exciton confinement and a widening of the bandgap.
The bandgap of size-selected AIS QDs is found to increase in a seemingly linear
manner with the fraction number from 2.3 eV (fraction 1) to almost 3.0 eV for the
smallest QDs in the fraction 8 (Fig. 5a, inset). The same trend can be observed
for AIS/ZnS QDs with the bandgap increasing from 2.3 eV to 3.1 eV for the last
fraction 8 (Fig. 5b). A larger Eg for AIS/ZnS QDs as compared to the “core” AIS

Fig. 5 Absorption spectra of size-selected AIS QDs (a) and AIS/ZnS QDs (b) in fractions 1
through 8. Inserts in (a, b): bandgap Eg of QDs as a function of the fraction number
78 O. Stroyuk et al.

Fig. 6 Relative amounts of QDs in different fractions produced from AIS (a) and AIS/ZnS
colloids (b). Insets: Photographs of size-selected AIS (a) and AIS/ZnS (b) colloids taken under
UV illumination (365 nm); Ag:In:S:Zn = 1:4:5:8. (Ô) XRD patterns for AIS precipitates later
redispersed to fractions 1 and 7; Ag:In:S = 1:4:5

QDs in the respective fraction numbers (starting from 4–5 and higher) indicates that
the incorporation of Zn2+ is obviously more efficient for smaller QDs than for larger
ones, in accordance with the XPS data discussed below.
To evaluate the population of QDs in different fractions, we took the optical
density of the size-selected QDs far from the absorption band edge (at 300 nm) as
a quantitative measure of the molar AIS concentration (Fig. 6). This evaluation is
regarded as a first approximation, because the strong quantum confinement in small
AIS QDs can affect not only the band-edge absorption but also the selection rules
and the molecular unit-weighted oscillator strengths of all electron transitions in the
QDs.
The first and second fractions of AIS QDs are the most populated ones in terms of
the AIS concentration, while the amount of QDs in the following fractions declines
rapidly being ≈20 times lower for fraction 8 as compared with the most populated
fraction 2 (Fig. 6a). In the case of AIS/ZnS, again, fraction 2 is the most populated,
while the concentration of the QDs decreases in a much steeper way than for AIS
QDs (Fig. 6b), and the last fractions 7 and 8 contain two orders of magnitude less
AIS QDs than the most populated fraction 2.
The size-selected AIS QDs reveal X-ray diffraction (XRD) patterns typical
for chalcopyrite Ag-In-S with the reflection intensity gradually decreasing with
increasing fraction number (Fig. 6c). The latter fact indicates a gradual disordering
of the QD lattice with decreasing size, most probably due to the lattice relaxation
to accommodate the growing surface-to-volume ratio. The estimations based on the
Scherrer equation show the AIS QDs in fractions 7–8 being smaller than 2 nm, while
larger QDs (2.5–3.0 nm) are found in the first two fractions.
A basic issue for the nonstoichiometric AIS QDs is to discriminate the evolution
of the optical properties caused by the QD size variation from variations of
the QD composition in different fractions. We subjected a series of ten size-
selected fractions of AIS/ZnS QDs to an X-ray photoelectron spectroscopic (XPS)
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 79

Fig. 7 (a) High-resolution X-ray photoelectron spectra of AIS/ZnS QDs in the range of the
binding energies of Ag3d, In3d, Zn2p, and S2p electrons. Gray lines, experimental spectra; blue
lines, background; red lines, fitting results; Ag:In:S:Zn = 1:4:5:8 [14]. (b) Atomic ratios of
components of size-selected AIS/ZnS QDs as a function of the bandgap (corresponding fraction
numbers are given in the figure) as determined by XPS (Ag:In:S:Zn = 1:5:10:10, ten fractions)

study aiming at the evaluation of possible compositional deviations. The high-


resolution photoelectron spectra in the ranges of Zn2p, Ag3d, and In3d core level
emissions (Fig. 7a) show corresponding doublets at the positions typical for Zn2+
(1020/1043 eV), Ag+ (367/373 eV), and In3+ (444/451 eV) [88, 89].
The S2p range reveals two doublets at 161.4/162.6 eV and 163.4/164.6 eV
characteristic for the lattice sulfide and the mercapto-group bound to the QD surface,
respectively. The relative contribution of the latter component increases somewhat
(by around 30%) from fraction 1 to fraction 10 in line with the increase of the total
QD surface/volume ratio, but no steady trend throughout the entire series could be
obtained due to the scatter in data.
Figure 7b shows a set of atomic In/Ag (blue bars), Zn/In (green bars), and S/In
(red bars) ratios calculated from the corresponding areas in the high-resolution XPS
spectra for ten fractions of core/shell AIS/ZnS-GSH QDs. The indium-to-silver ratio
which most strongly affects the optical properties of the size-selected QDs is found
to vary between 2.4 and 3.0 and, with regard to the experimental error, can be
assumed to be almost constant throughout the series studied.
Inspecting the data we can safely assume that the observed changes in the
absorption spectra (as well as the changes in the PL spectra discussed below)
originate from the QD size variation between the fractions and not from variations of
the silver-to-indium ratio. As this ratio is smaller than the one set in the synthesis of
the parental colloid (Ag:In = 1:4), we can also assume that In(III) is partially bound
in a complex with GSH, which remains unprecipitated in the parental solution.
The ratio of zinc to indium is found to increase steadily from around 1–1.5 for
the first fractions to 2.1–2.2 for the last fractions, most probably due to the presence
of Zn(II)-GSH complexes on the QD surface and the incorporation of Zn2+ ions
80 O. Stroyuk et al.

Fig. 8 PL spectra of size-selected AIS QDs (a) and core/shell AIS/ZnS QDs (b) in fractions 1
through 8. (c) PL QY of colloidal AIS QDs (red squares 1) and AIS/ZnS QDs (blue circles 2) as a
function of the QD bandgap; Ag:In:S:Zn = 1:4:5:8 (eight fractions)

into the QD lattice. Both factors are expected to increase with decreasing QD size
resulting in the observed trend. The S-to-In ratio is found to increase considerably,
from around 4 for the first fractions to ≈7 for the smallest QDs, which can also be
accounted for by the increase in the total QD surface area.

2.3 Static PL of the Size-Selected AIS and AIS/ZnS QDs

The size-selective precipitation produces a series of colloidal solutions differing


considerably by their PL characteristics. The size-selected AIS and AIS/ZnS QDs
emit in broad bands (Fig. 8a, b), similar to those of the original parental solutions,
but the PL band position and intensity (normalized to the QD concentration in each
fraction) depend on the fraction number. In the case of AIS QDs, the PL band
maximum shifts from ≈660 nm for the first fraction to ≈560 nm for the last fraction
in the series, while the PL intensity is reduced by about an order of magnitude
(Fig. 8a).
A similar trend is observed for AIS/ZnS QDs exhibiting PL maxima at ≈620 nm
for fraction 1 and ≈520 nm for fraction 8 and a similar ≈sixfold reduction in the PL
intensity (Fig. 8b). The PL bands are characterized by a similar FWHM of around
0.35–0.40 eV for all fractions. Visually the PL emission color changes from orange-
red to green in the series of size-selected AIS QDs (Fig. 6a, inset) and from orange
to bluish-green for AIS/ZnS QDs (Fig. 6b, inset).
We measured the absolute PL quantum yields (QYs) for all fractions using two
different setups equipped with integrating spheres to assure a high reproducibility.
It is found that the highest PL QYs of 32% and 60% are observed for the most
populated fractions 2 of AIS and AIS/ZnS QDs, respectively (Fig. 8c). The twice
as large PL QY of the core/shell AIS/ZnS QDs as compared to the bare AIS clearly
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 81

indicates the passivating effect of the ZnS shell, most probably arising from both
the elimination of surface defects and possible lattice imperfections due to the
Zn2+ diffusion into the AIS core. As the AIS QD size decreases (and the bandgap
increases), the PL QY decreases down to only a few percent for the least populated
fractions 7 and 8 (Fig. 8c). The core/shell AIS/ZnS QDs show a similar trend, but
in this case the fractions with the smallest QDs reveal distinctly higher PL QYs of
around 10%, showing that the effect of passivation with ZnS is stronger for smaller
QDs, in accordance with the conclusions derived from the absorption spectra and
the XPS results.
As compared to the GSH-capped AIS and AIS/ZnS QDs, the stabilization of
AIS-based QDs with MAA results in a broader size distribution allowing a much
higher number of fractions to be separated by the size-selective precipitation. Figure
9a, b demonstrates PL spectra and emission colors for the record 17 size-selected
fractions of the core/shell MAA-capped AIS/ZnS QDs. As the fraction number
increases, the PL maximum shifts to higher energies indicating an increased exciton
confinement in the smaller QDs. Remarkably, the PL maximum energy changes in
an almost linear manner with the fraction number (red squares in Fig. 9c) revealing
a continuous variation of the size and bandgap of the QDs in the size-selected
fractions. The PL FWHM, however, remains almost unchanged varying between
0.38 and 0.40 eV with an average width of 0.39 eV very close to the PL FWHM
of the initial QD ensemble (Fig. 9c). The fact shows that a possible variation in the
QD size distribution in the different QD fractions does not contribute to the spectral
width of the PL band of AIS/ZnS QDs.
According to the assumptions reported in Ref. [90] for core/shell CIS/ZnS QDs,
the hole localization center in these QDs is characterized by a radius of around
0.3 nm, which is considerably smaller than the size of about 12–13 nm of the studied
QDs. This can favor a scatter of possible hole positions from the QD center up to the
QD surface. In the case of our 2–4 nm-sized AIS/ZnS QDs, this distribution should

Fig. 9 PL spectra (a) and emission colors (b) of size-fractionated MAA-capped AIS/ZnS QDs
(fraction numbers are given in figures). Higher fraction numbers correspond to smaller QDs in
the fractions. The photograph (b) was taken under UV illumination (365 nm). (c) Energy of PL
maximum (red squares) and FWHM (blue circles) of the PL bands of size-selected AIS/ZnS QDs
82 O. Stroyuk et al.

be considerably narrower. For the smallest AIS/ZnS QDs with a size of around
2 nm, the QD core size becomes comparable with the reported hole localization
diameter of 0.6 nm. Thus, the distribution of Coulombic energies of the interacting
free electron and trapped hole is also expected to be considerably reduced. However,
the smallest AIS/ZnS QDs still reveal almost the same FWHM as the largest ones
(Fig. 9c), thereby underlining the absence of a FWHM dependence on a possible
distribution of electron-hole distances. Similar observations were already made for
small CdS QDs in 1986 by Brus et al. and rationalized by the assumption that the
minimal D-A distance corresponds to the length of the Cd-S bond or that the QD
surface acts as a collective defect state for hole trapping [91]. Both assumptions
imply that the broadband PL is an inherent property of the AIS/ZnS QDs rather
than a collective property of a QD ensemble.
The nature of the broadband PL of ternary CIS and AIS QDs, similar to copper-
and silver-doped binary cadmium chalcogenide QDs, still remains a subject of
discussion [8, 15], and the proposed interpretations are quite controversial. Most
frequently, the broadband PL of CIS/AIS QDs is interpreted in a “conventional”
way by applying a model of donor-acceptor (D-A) recombination [8, 9, 12, 15,
92–94] expressed by Eq. (1) as discussed shortly in Sect. 1. This model implies
that the PL emission originates from the radiative recombination between trapped
electrons and holes and the spectral width of the PL band is a function of at least
three distributions: (i) a distribution of sizes; (ii) D and A trap level distributions
by “depth” (i.e., the energy gap between the corresponding band edge and the local
trap levels in the bandgap), and (iii) a distribution of distances separating the trapped
electrons and the holes within each single QD, the latter distance being limited by
the maximum QD size in the ensemble, i.e., (iii) can partially correlate with (i).
A size distribution of the QDs in a colloidal ensemble results in a distribution
of Eex , while an inhomogeneity of the QD lattice and many possible origins of the
electron and hole traps contribute to the distribution of the ED and EA energies.
Finally, trapped carriers being closer to each other are characterized by a higher
Coulombic interaction when recombining, thus generating a PL with a higher energy
than those pairs situated further apart. All three factors are believed to contribute
to the formation of broad PL bands [8, 9, 12, 15, 92–94]. As the closer carrier
pairs recombine faster, the D-A-emitting QDs typically exhibit a dependence of the
PL lifetime on the PL observation wavelength (energy) – the lower the PL energy
the longer the PL lifetime. Such dependences are usually taken as an unambiguous
proof of the D-A-type PL both for binary and ternary metal-chalcogenide QDs [15,
65]. However, when applied to ternary AIS and CIS QDs, especially those subjected
to size selection, the D-A model manifests certain inconsistencies that, combined
together, show the need for a more realistic approach for the description of PL
phenomena.
The recent progress in the synthesis of CIS and AIS QDs with strictly tailored
morphology, phase, and composition allows producing colloidal QDs with low
dispersities [2, 5, 30–33, 95] showing, nevertheless, broad PL bands like their
polydisperse analogues. The size selection applied in the present work is also
expected to considerably narrow the size distribution separating the original parental
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 83

ensemble into numerous fractions with different average sizes. However, the size-
selected AIS and AIS/ZnS QDs reveal the same or even somewhat larger FWHMs of
the PL bands than the parental colloids, showing no indications of band narrowing
(Figs. 8 and 9). Finally, recent studies on single CIS and AIS QDs [96, 97] as well
as those reported in the present chapter showed them to emit in equally broad bands
with the band maximum position changing randomly but within the PL range of the
whole QD ensemble. All these observations allow the size dependence of Eex (Eg )
to be excluded as a major contribution to the FWHM of ternary QDs.
The idea of a broad distribution of trap states by depth also seems to be
impaired by various static PL measurements. A broad depth distribution implies
a large number of available defects, but the defects inherently can participate both
in radiative and non-radiative recombination. The highest PL QY of 60% in the
present work is very close to the best results typically reported for AIS/ZnS QDs
produced by heating-up/hot-injection syntheses [9, 11, 93]. Even higher PL QYs
were reported for larger (5–6 nm) alloyed ZAIS QDs reaching up to 79% [98]
indicating that we still have room for improvement of our aqueous synthesis.
In the frame of the D-A model, the broad PL is assumed to originate from a broad
distribution of energies and distances of the trapped carriers. Hence, we should
expect the spectral parameters of the PL bands of the AIS QDs to be strongly
dependent on temperature. For an ensemble of traps, there exist states separated
in energy close to kT (k is the Boltzmann constant), and therefore a repopulation of
traps at room temperature is expected with the charge carriers coming from deeper
traps to shallower traps under the influence of the thermal lattice energy [29, 99]. As
the temperature is lowered, the detrapping becomes slower, resulting in a “red” shift
and a narrowing of the trap-state-mediated PL bands. However, the reported studies
of temperature dependences of ternary CIS and AIS QDs [11, 12] showed no effects
of this sort to occur with the spectral width of the PL band remaining essentially
the same in the whole temperature range studied. Our studies of the temperature
dependence of the PL of the size-selected AIS/ZnS QDs (see next sections) also
showed the PL band width being essentially the same in the broad T range from 310
to 10 K where “freezing” of the traps can be expected.
A possible contribution of the distribution of the trapped carriers by distance
within single QDs can also be questioned in the light of the reports on the
dependence of the PL emission of AIS QDs on the excitation power [9, 12, 20]
as well as by our studies on size-selected AIS QDs. Indeed, as the size of AIS
QDs becomes smaller, we should expect a narrowing of the spectrum of possible
distances between the trapped electrons and holes. This is not the case for the size-
selected QDs showing the same FWHM of the PL band in the entire range of sizes.
Moreover, the smallest QDs (d < 2 nm) have only a few lattice planes, and a broad
distribution of D-A distances is not realistic.
Additionally, if a distribution of distances between the trapped electrons and
holes is assumed, an effect of the PL excitation intensity on the population of
electronic states (corresponding to different e-h distances) is expected, i.e., a
narrowing of the PL band would be expected as the excitation intensity is reduced.
No such trend was found for AIS QDs with the PL excitation power varied by several
84 O. Stroyuk et al.

orders of magnitude [9, 12, 20] indicating that the concept of D-A recombination
with differently separated e-h pairs needs to be revised at least for the present case
of ternary In-based chalcogenide QDs [8, 15, 92].
D. Gamelin et al. proposed to describe the PL emission of copper-doped
cadmium chalcogenide QDs as well as ternary CIS QDs by the STE model [8],
originally introduced for silver halide QDs and then suggested by L. Brus et al.
to account for a specific PL behavior of ultrasmall CdS QDs [91]. The model is
based on the assumption of a strong electron-phonon interaction and participation of
vibrational modes of the QD lattice in the fate of the photogenerated charge carriers.
One of the carriers, typically the valence band hole, is assumed to be bound to a
local state (Cu+ dopant in copper-doped QDs or any M+ ion in silver halides, CIS,
or AIS QDs [8]) resulting in a considerable local distortion of the lattice stabilizing
the trapped carrier. Due to strong coupling to the lattice vibrations, a portion of the
excitation energy is dissipated as phonons, typically as longitudinal optical (LO)
phonons, because a trapping site is assumed to consist of a central metal ion in
the tetrahedral “shell” of four anions and the trapped charge carrier is coupled to
metal-anion stretching vibrations [8, 9, 12]. The radiative recombination of a free
electron with a trapped hole results, therefore, in a series of phonon replicas of the
zero-phonon line EZPL generally observed as a broad PL band for silver halide QDs,
Cu-doped CdS (CdSe), CIS, and AIS QDs [8]. The energy En and relative intensity
In of each phonon replica forming the PL band can be estimated in the frame of a
configuration coordinate model [9, 12] as

En = EZPL − nω, (2)

S n e−S
In = (3)
n!
where n is the number of emitted phonons, -hω is the phonon energy, and S is
the Huang-Rhys factor equal to the average number of phonons emitted with PL.
The case of n = S corresponds to the peak energy EPL of the PL band, that is,
EPL = EZPL – S-hω.
Hamanaka et al. were first to apply this model to simulate the PL spectrum of
2.6 nm AIS QDs [12]. Their fitting yielded a realistic value of the AIS LO phonon
energy, a large Huang-Rhys factor of 23, and a EZPL value corresponding to a low-
intensity excitonic feature in the absorption spectrum. These observations indicate
that the experimental PL spectra can be modeled using empirically determined
values – EZPL equal to the bandgap and the LO phonon energy derived from Raman
spectra, while the Huang-Rhys factor S can be estimated from the Stokes shift
between the Eg and EPL normalized to the phonon energy.
The present size-selected AIS-GSH QDs (a series of ten fractions) were found
to exhibit different Stokes shifts varying from 330 meV for fraction 1 to 630 meV
for fraction 10 showing that the average number of phonons emitted along with
PL is twice as large for the smallest QDs as for the largest QDs. In terms of the
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 85

STE model, these observations imply that the vibrational relaxation of the excitation
energy is around two times more efficient in the smallest QDs.
As smaller QDs have a larger surface-to-volume ratio, the above observation
may indicate that the hole can be trapped on the interface between the QD core and
the surface shell of the ligands. Indeed, earlier we reported on the ultrasmall (1.8–
2.0 nm) CdS QDs stabilized by surface Cd(II) complexes with polyethyleneimine
(PEI) [100] or mixed complexes with mercaptoacetate (MA) anions and ammonia
[101], exhibiting similar broadband and intense PL. We found that such QDs are
crystalline and very stable indicating a high lattice quality. A moderate heating
of such CdS colloids resulted in a strong PL quenching, but the PL intensity was
completely restored after cooling. We interpreted this effect as a result of the
thermally activated dissociation of the surface Cd(II) complex species participating
in the radiative recombination and the formation of aqua-complexes involved in
non-radiative energy dissipation.
A similar situation can take place in the present case of GSH-stabilized AIS
QDs because this ligand provides the same combination of mercapto- and amino-
groups as in the case of CdS-MA-NH3 QDs. Our experiments showed that the PL
intensity of AIS-GSH QDs indeed drops by an order of magnitude after heating of
colloidal solution from room temperature (RT) to 60–70 ◦ C (more details in the
next sections), and this effect is completely reversible upon cooling. Therefore,
we can assume that the layer of surface-capping GSH complexes can act as a
“collective” trap for the photogenerated hole and the vibrational relaxation becomes
expectedly faster for the smaller QDs with a larger surface-to-volume ratio. A strong
contribution of both direct exciton-ligand coupling and indirect surface charge
trapping events are currently discussed as possible mechanisms of a vibrational loss
of the excitation energy [65].
We tried to reconstruct the PL spectra of two fractions of size-selected AIS QDs
(1 and 10) using the Eg values derived from the corresponding absorption onset
(Eg 1 = 2.40 eV and Eg 10 = 2.92 eV; Fig. 10a, b; superscript index is the fraction
number), -hω = 38 meV (the energy corresponding to the strongest phonon peak in
the Raman spectra [13]), and a Huang-Rhys factor of 9 (fraction 1) and 17 (fraction
10) calculated from the corresponding Stokes shifts. The modeled spectra (bars
in Fig. 10a, b) appeared to be much narrower than the experimental PL spectra,
with the difference between the modeling and experiment being much larger for
fraction 1.
The differences between the model and the observed PL indicate the existence
of an additional factor contributing to the PL band broadening. To investigate
the possible nature of this factor, we modeled the PL spectrum of previously
studied CdS-PEI QDs characterized by a very narrow and sharply resolved excitonic
absorption feature [100] indicating a very narrow QD size distribution (Fig. 11a).
The modeling using the CdS LO phonon energy, EZPL = 3.50 eV, and S = 27
(derived from the Stokes shift) showed a perfect match with the experimental PL
spectrum indicating a good applicability of the model and the chosen parameters
for the case of QD ensembles with low polydispersity. Therefore, the mismatch
between the modeling results and the PL spectra of AIS-GSH QDs can arise from a
86 O. Stroyuk et al.

Fig. 10 Normalized absorption and PL spectra (solid curves) of size-selected colloidal AIS QDs
from fractions 1 (a, c) and 10 (b, d) combined with the results of PL spectra modeling (bars)
without (a, b) and with (c, d) the correction on the QD size distribution; Ag:In:S:Zn = 1:4:5:8

broader size distribution in the corresponding QD fractions compared to CdS-PEI.


This assumption also agrees with a lower mismatch for the smaller QDs in fraction
10 that is obviously characterized by a narrower size distribution as compared to the
first fraction of QDs.
To account for a possible size distribution of the QDs in size-selected fractions,
we adopted a method of the determination of an average Eg from the first derivative
of the absorption spectrum near the band edge proposed by Pesika et al. for ZnO
QDs [102]. This method allows a more realistic average bandgap to be obtained
for an ensemble of QDs in the regime of strong spatial exciton confinement as
compared with the tangent method, the latter providing a lower limit of Eg . The
average bandgap values for fractions 1 and 10 determined by the derivative method
(see example in Fig. 11b) are found to be higher (Eg 1 = 2.65 eV, Eg 10 = 3.29 eV)
providing larger Stokes shifts (Fig. 10c, d) and the Huang-Rhys factors of S1 = 15
and S10 = 27. The reconstruction of the PL spectra using these corrected values
results in a much better match between the modeled and the experimental spectra
(Fig. 10c, d) demonstrating a good applicability of the STE model to the size-
selected AIS QDs.
These results also attest that the position and shape of the PL bands of AIS
QDs do not originate from trap depth distributions but rather from the dynamics
of the electron-phonon interaction and the vibrational relaxation in the ultrasmall
QDs coupled to ligand molecules. In this view, the discussed QDs differ strongly
from larger “regular” nanoparticles and bulk semiconductors where the broadband
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 87

Fig. 11 (a) Normalized absorption and PL spectra (solid curves) of PEI-stabilized CdS QDs [100]
combined with the results of PL spectra modeling (bars). (b) The first derivative of the absorption
spectrum of size-selected AIS QDs from fraction 1

or multiband emission is associated with different defects and recombination


mechanisms. The possibilities of a drastic difference in the origin of the PL between
bulk CuInS2 and CIS QDs and a size dependence of the stability of self-trapped
excitons were discussed by Gamelin et al. for Cu-doped QDs [8]. It was suggested
that this model becomes even more appropriate as the size of ternary QDs is
decreased.

2.4 Time-Resolved Photoluminescence of the Size-Selected


GSH-Stabilized AIS and Core/Shell AIS/ZnS QDs

The PL decays of size-selected GSH-stabilized AIS (AIS/ZnS) QDs reveal a distinct


multi-exponential character (Fig. 12a) typical for ternary metal-chalcogenides [9,
11, 15, 36, 62, 93]. The PL decay curves of CIS and AIS QDs are typically
fitted with linear combinations of two–three single-exponential functions to derive
a weighted average PL lifetime [9, 11, 36, 62, 93].
The average PL lifetime <τ> of AIS and AIS/ZnS QDs depends on the QD
bandgap (Fig. 12b) decreasing from around 380 ns for the largest QDs in fraction 1
(Eg 1 = 2.32 eV) to ≈100 ns for the smallest QDs in fraction 10 (Eg 10 = 2.95 eV).
Covering of the AIS cores with a ZnS shell did not affect the PL lifetime in fractions
1–6, while increasing <τ> for the last two fractions of the smallest QDs. The
latter effect can be accounted for by a more massive penetration of Zn2+ ions
into the lattice of the smallest QDs, in accordance with the results of absorption
spectroscopy and XPS.
88 O. Stroyuk et al.

Fig. 12 (a) PL decay curves of size-selected colloidal AIS/ZnS QDs (fraction numbers are given
in the figure) registered in the maxima of corresponding PL bands. (b) Average radiative lifetime
of colloidal AIS QDs (squares 1) and AIS/ZnS QDs (squares 2) as a function of the QD bandgap;
Ag:In:S:Zn = 1:4:5:8

By combining stationary measurements of the PL QY with the kinetic data on


<τ>, we can extract values of the rate constants of the radiative recombination kr and
the non-radiative recombination knr as a function of Eg . The radiative QD lifetime
is inversely proportional to the sum of the rate constants of all recombination routes
(Eq. 4), while the PL QY represents the fraction of the radiative channels in the
entire set of recombination events (Eq. 5):

1
< τ >= (4)
kr + knr

kr
PLQY = (5)
kr + knr

By combining Eqs. (4) and (5), we can evaluate the radiative recombination rate
constant
PLQY
kr = (6)
<τ >
and then calculate the rate constant of the non-radiative recombination processes.
The radiative recombination rate is found to follow the PL QY, being maximal
in fractions 2 of the AIS and the AIS/ZnS QDs (Fig. 13a), and then decreasing for
smaller QDs. As for the PL QY, the incorporation of Zn2+ retards the kr decrease
for the fractions 7 and 8 of the smallest AIS/ZnS QDs. One can see that the PL
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 89

Fig. 13 Radiative recombination rate constant (a) and non-radiative recombination rate constant
(b) of colloidal AIS QDs (squares 1) and core/shell AIS/ZnS QDs (squares 2) as a function of the
QD bandgap; Ag:In:S:Zn = 1:4:5:8

emission rate is generally by 60–70% higher for the core/shell AIS/ZnS QDs as
compared to bare AIS QDs in the corresponding fractions. In absolute values, the
radiative recombination rate constant derived here for the size-selected AIS and
AIS/ZnS QDs falls into the typical range of (0.1–1) μs−1 reported for such ternary
QDs [9, 11, 93].
Surprisingly, the rate constant of the non-radiative recombination was found to be
only slightly lower for the core/shell AIS/ZnS QDs than for their non-passivated AIS
analogues in the first fractions with the highest PL QY (Fig. 13b). As the fraction
number and Eg increase, the knr difference between the core AIS and core/shell
AIS/ZnS QDs becomes more evident, and for the smallest QDs, the non-radiative
recombination occurs almost two times slower after the QD passivation with the
ZnS shell. These observations indicate that the PL enhancement effect of the ZnS
shell may be interpreted as a result of both an electron-hole interaction enhancement
in the QD cores by the higher-bandgap outer ZnS layer and the formation of
additional sites for the radiative recombination [11], the latter effect becoming more
pronounced for the smallest and Zn-enriched AIS/ZnS QDs.
The PL lifetime is found to depend considerably on the PL registration wave-
length, that is, on the PL energy. The emission rate increases by a factor of around
two as the PL energy grows from around 1.95 eV to 2.35 eV and from 2.15 eV
to 2.65 eV for the first (No. 1) and the last (No. 10) fractions studied, respectively
(Fig. 14a). Conventionally, such dependences are interpreted in terms of a higher
probability of the radiative recombination between closer electron-hole (D-A) pairs
[8, 9, 12, 15]. The distance-dependent Coulomb interaction between trapped charge
carriers is supposed to contribute to the energy of the emitted PL quanta being
higher for closer pairs. Based on the above discussion about the nature of the
90 O. Stroyuk et al.

Fig. 14 (a) Radiative lifetime of size-selected colloidal AIS QDs from fractions 1, 6, and 10 as a
function of the observation wavelength (PL energy). (b) Average time of single phonon emission
tsp (blue bars 1) and rate of the energy loss hv/τ (red bars 2) determined for the size-selected
AIS QDs as functions of the QD bandgap

PL emission in AIS QDs, we suggest an alternative interpretation of the “τ–hv”


dependences which is consistent with the self-trapped exciton model and assumes
the continuous dissipation of the excitation energy in the form of vibrational quanta
until the event of radiative e-h recombination occurs. The larger portion of energy
has to be dissipated as phonons, the lower is the probability of such an event, and the
longer a QD remains in the excited state. This tendency results in lower-energy PL
quanta to be emitted at longer times. In this interpretation, the PL energy loss divided
by the difference in two radiative lifetimes between which the loss is observed, that
is, the slope of a “<τ> – hv” plot, can be taken as a quantitative measure of the rate
of phonon emission.
We found the rate (i.e., the probability) of vibrational relaxation to be size-
dependent increasing by a factor of almost two as the bandgap of AIS QDs increases
from 2.4 to 2.92 eV (Fig. 14b, bars 2). The nominal time tsp is also found to decrease
from 18 ns for the largest AIS QDs to 9 ns for the smallest QDs (Fig. 14b, bars
1). These observations indicate that the phonon emission in the smallest AIS QDs
occurs about two times faster than in the largest QDs in the studied series resulting
in a larger relative PL energy loss during the period when the QDs remain in the
excited state. In this view, the spectral parameters of the PL bands of AIS QDs are
functions of two size-dependent parameters – the bandgap as a zero-phonon line of
the PL spectrum and the rate of vibrational relaxation of a trapped charge carrier
determining the maximum position and FWHM of the PL band.
Hence, by using two different approaches, namely, the analysis of the absorp-
tion/PL spectra in terms of the STE model and from the kinetic PL measurements,
we come to the same conclusion that the rate of vibrational relaxation in AIS QDs
is size-dependent with its probability increasing approximately by a factor of two in
the studied size range.
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 91

2.5 Single-Particle PL Measurements

The measurement of single-particle PL spectra can provide the most straightforward


and convincing argument for the inherent nature of the broadband PL of AIS/ZnS
QDs [17, 97]. We found that the core/shell AIS/ZnS QDs derived from aqueous
solutions are not suitable for the preparation of samples for single-particle studies
by conventional spin-coating on various substrates including glass and mica because
of the strong aggregation of the QDs on the surface upon drying. To prepare well-
separated QDs for the intended study of single QDs, we developed a phase transfer
protocol allowing the AIS/ZnS QDs to be transferred from water to toluene without
any appreciable variation in their absorption and PL properties. The protocol
includes the precipitation of the AIS/ZnS QDs with ethanol, the redispersion of the
viscous precipitate in water, and the extraction of the QDs into toluene containing
oleylamine and oleic acid [27]. The phase-transferred oleylamine- and oleic acid-
stabilized AIS/ZnS colloids are very stable in toluene and exhibit the same PL
intensity as the initial AIS/ZnS colloid (Fig. 15a, b). The absorption spectra of the
AIS/ZnS QDs in water and toluene are identical, while the PL maximum of AIS/ZnS
QDs in toluene is slightly blue shifted (Fig. 15a), reflecting the reduced stabilization
of the charge-separated excited state of the QDs in nonpolar toluene compared to
polar water.
Atomic force microscopic (AFM) images of core/shell AIS/ZnS QDs deposited
by spin-coating of a very diluted toluene dispersion on atomically flat mica (Fig.
16a) show evenly dispersed QDs with a height of 2–4 nm, with the height
distribution being centered at about 3 nm (Fig. 16b). The QD height distribution
matches almost perfectly the size distribution of aqueous AIS/ZnS QDs derived
from transmission electron microscopy (TEM) images (Fig. 16c).
We simultaneously took AFM images and confocal microscope images of the QD
samples to ensure that only photons originating from QDs not bigger than 3–4 nm

Fig. 15 (a) Absorption (right part) and PL spectra (left part) of colloidal AIS/ZnS QD ensemble
in water (red curves) and toluene (blue curves). (b) Photographs of size-selected GSH-capped
AIS/ZnS QDs in water and toluene under UV illumination (350–390 nm). (Photograph of AIS/ZnS
QDs in toluene – courtesy of L. Dhamo and Prof. U. Resch-Genger (BAM Berlin, 2019))
92 O. Stroyuk et al.

Fig. 16 (a) AFM image of AIS/ZnS QDs deposited on mica. (b) Size (height) distribution of
AIS/ZnS QDs derived from an AFM image. (c) Size distribution of AIS/ZnS QDS derived from
TEM

Fig. 17 (a) Optical confocal microscope image of AIS/ZnS QDs deposited on a glass substrate.
(b) Trace recorded on the spot marked by a circle in left panel

(i.e., single QDs) were collected. For each selected dot, we simultaneously detected
a PL spectrum and time traces (Fig. 17) that reveal the typical emission intermittency
where a QD is in an emissive “bright” state and in a non-emitting “dark” state
for random time lapses. The characteristic blinking behavior provides additionally
proof of probing single QDs [17]. This was performed for ten representative sample
spots.
The observed variations of the PL intensity in the bright state can originate from
the fast blinking of QDs within each single binning period of the recorded time trace
yielding intermediate integral PL intensities. Also, the single AIS/ZnS QDs show
broad PL bands within the energy range of the PL band of the parent QD ensemble
which can be fitted with a single Gaussian curve. Figure 18a shows a representative
PL spectrum of a single AIS/ZnS QD including its fit and the corresponding QD
ensemble spectrum for comparison. Gamelin et al. reported similar results for single
CuInS2 QDs that revealed large PL FWHMs with a PL maximum fluctuating from
QD to QD within the wavelength range covered by the PL band of the respective
QD ensemble [96].
The position of the PL band of the single AIS/ZnS QDs varies from ≈2.0 to
≈2.1 eV (Fig. 18b). Its maximum closely matches the PL band maximum of the
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 93

Fig. 18 (a) PL spectra of the parent ensemble of AIS/ZnS QDs (curve 1) and a single AIS/ZnS
QD (2), curve 3 obtained by fitting the single-particle spectrum 2 with a single Gauss profile. (b,
c) PL band maxima (b) and PL band FWHM (c) of ten separate single AIS/ZnS QDs; black lines
show the corresponding parameters of the colloidal ensemble and blue lines the parameters of the
most prominent size-selected fraction 2 of AIS/ZnS QDs, respectively

most prominent size-selected fraction extracted from the respective QD ensemble.


The probability of measuring signals from AIS/ZnS QDs from this fraction is the
highest, and therefore, it is expected that the average PL maximum position of single
QDs is closer to the PL maximum of this fraction, rather than to the position of
the PL maximum of the whole as-synthesized QD ensemble. A similar behavior
is observed for the PL FWHM of single AIS/ZnS QDs (Fig. 18c). The FWHM
values vary from 0.24 to 0.36 eV, with the average value being much closer to the
FWHM of the size-selected AIS/ZnS QDs from the most populated fraction 2 than
the spectral width of the original colloidal QD ensemble.
The fact that all single QDs revealed broad PL spectra indicates that this feature is
inherent to every single AIS/ZnS QD. A similar behavior was observed by Gamelin
et al. for single CIS QDs with FWHM values varying ≈0.20 to ≈0.28 eV [96]. The
results of the present study agree well with findings of Sharma et al. [103] reporting
broad PL spectra for single ZAIS QDs with a composition-dependent PL maximum
position and a homogeneous broadening of the PL band resulting in FWHMs as
high as 230–270 meV. This report also showed single ZAIS QDs to manifest a
phenomenon of spectral diffusion, the PL position and FWHMs changing randomly
during the spectral measurements. The spectral diffusion effect can also be expected
to contribute to the spectral width of the PL band of the present AIS/ZnS QD as
illustrated by a random distribution of the PL parameters of ten representative QDs
shown in Fig. 18b, c. According to Sharma et al. [103], the spectral diffusion can
account for up to one third of the FWHM of the single ZAIS QDs; however, the PL
band of such QDs is still much larger than the FWHMs typical for single cadmium
chalcogenide QDs [sp15].
Contrary, Klimov et al. reported PL bands as narrow as ≈0.13 eV for single
CIS/ZnS QD [90], even though these data were collected for core/shell QDs with
94 O. Stroyuk et al.

very thick ZnS shells that revealed almost no blinking. This could suggest that the
PL mechanism and the spectral parameters of such thick-shell CIS/ZnS QDs can
vary from those of the AIS/ZnS QDs discussed here and the ternary QDs studied
in Refs. 98 and 105 having only thin surface passivation shells. This assumption
is further corroborated by an unusually large scatter of the FWHM values reported
by Klimov et al. [90], ranging from 50 to 210 meV, while a much narrower FWHM
distribution was observed by Gamelin et al. for the single CIS QDs and the AIS/ZnS
QDs studied by our group. Later, Klimov et al. showed that the PL behavior of CIS
QDs including the mode of charge trapping and even the recombination mechanism
can differ drastically for stoichiometric and nonstoichiometric samples [104]. This
may also account for the observed differences in the single-particle PL behavior of
the strongly nonstoichiometric AIS/ZnS QDs.
The comparison of the FWHM values of individual AIS/ZnS QDs and the
QD ensemble (Fig. 18) shows that possible FWHM variations caused by a size-
dependent distribution of the PL maximum position and changes of the PL
properties of between individual QDs and induced by the respective QD microen-
vironment/dispersive medium including the spectral diffusion effect [103] can
increase the FWHM of single QDs in the AIS/ZnS ensemble, but only to a
limited extent, adding not more than 15–20%. Overall, the single QD measurements
confirmed our assumption of the inherent broadband character of the PL band of
AIS/ZnS QDs derived from the PL features of both the initial QD ensemble and the
size-selected QD fractions.

3 Temperature-Dependent Luminescence of AIS


and AIS/ZnS QDs

Semiconductor QDs emitting bright size- and composition-dependent PL reached


an advanced stage of application in light-emitting technologies, including LEDs
and optical sensors, but still require an in-depth understanding of many aspects
of the mechanisms and dynamics of their light emission, in particular temperature
dependences of the efficiency and spectral parameters of PL.
Investigations of the PL temperature dependence were primarily focused on
cadmium CdX and lead chalcogenide PbX QDs (X = S, Se). These studies provided
ample information on the origins and mechanisms of both radiative and non-
radiative processes of electron-hole recombination and charge transfer, being of
interest, particularly, for bioimaging and biosensing systems [33, 105, 106]. High
thermal sensitivity and photostability of QDs as compared to molecular fluorescent
markers [107], engineered biocompatibility, and size-dependent optical properties
of QDs can be used for multiplexed sensing and open broad perspectives of using
luminescent QDs as temperature sensors at the nanoscale [106, 107].
The applications in biosensing require the QDs to be affordable and “green,” that
is, to avoid highly toxic metals such as cadmium, lead, or mercury [16, 32, 33, 106].
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 95

In recent years, ternary metal-chalcogenides such as CIS and AIS have been broadly
recognized as highly promising luminophores with the spectral emission parameters
tailorable in a broad range by variations in the QD composition, size, and doping [4,
12, 16, 32, 33, 108–110].
In the present section, we discuss temperature dependences of ternary AIS and
core/shell AIS/ZnS QDs stabilized in aqueous solutions by surface GSH complexes
and probed in two regimes – in colloids at 10–80 ◦ C and as drop-casted dried films
in a broader temperature range of 10–310 K.

3.1 Temperature Range of 10–80 ◦ C

Recently reported aqueous syntheses [4, 16, 32, 33] allow highly stable and
brightly emitting ternary QDs to be obtained in aqueous solutions by simple,
reliable, and reproducible procedures. The surface of such QDs can be modified
by multifunctional ligands ready for the post-synthesis conjugation to various target
biomolecules. In particular, we reported on the aqueous syntheses of ternary AIS and
core/shell AIS/ZnS QDs stabilized by surface complexes with polyethyleneimine
(PEI) [111], MAA [13, 112], and GSH (see Sect. 2) [14]. Our approaches to aqueous
ternary QDs base on previous experience with ultrasmall (the size d < 2 nm) CdS
QDs capped with Cd(II) complexes with PEI [100, 113] or mixed complexes of
Cd(II) with ammonia and MA anions [10, 101], as well as ultrasmall core/shell
CdSe/CdS QDs [10, 114]. We found that luminescent properties of aqueous
ultrasmall CdX QDs and ternary AIS (CIS) QDs are very similar, both QD types
emitting relatively intense visible PL characterized by large spectral widths and
Stokes shifts [10, 13, 14, 100, 101, 111, 113].
The ultrasmall CdS QDs stabilized by Cd(II) complexes with PEI or NH3 /MA
revealed unusually strong temperature dependences of both absorption edge and PL
intensity when probed in aqueous colloidal solutions [101, 113], showing a peculiar
PL quenching mechanism differing from that typical for “regular” CdX QDs capped
by covalently bound ligands [115–121]. In this subsection we discuss temperature
dependences of the PL parameters of GSH-capped AIS (AIS/ZnS) QDs in the
temperature range of 10–80 ◦ C in comparison with previously studied ultrasmall
CdX QDs capped with Cd(II) complexes. The thermal sensitivity of these QDs is
governed by similar mechanisms involving dissociation/association equilibria on the
QD surface.
Heating of aqueous colloidal solutions of AIS and core/shell AIS/ZnS QDs up
in the temperature range between freezing and boiling points of water results in
a decrease of the total PL intensity, in a shift of the PL band maximum to lower
energies, as well as in an increase of the rate of the PL decay, all these parameters
changing proportionally to the temperature of the solutions. The observed changes
of PL behavior are completely reversible upon cooling of the colloidal solutions to
the original temperature. The changes of the PL intensity, energy, and lifetime are,
therefore, cyclic and can be repeated in many consecutive heating/cooling cycles
96 O. Stroyuk et al.

Fig. 19 (a) PL spectra of aqueous GSH-capped AIS QDs at different temperatures varied from 10
to 80 ◦ C in steps of 5 ◦ C. (b) PL QY of aqueous colloidal AIS QDs with temperature changing in a
cyclic manner from 10 to 80 ◦ C and back (five consecutive cycles). (c) Spectral width of PL band
of AIS-GSH QDs (circles 1) and CdS QDs capped with Cd(II)-NH3 -MA complexes (rectangles
2) measured upon the temperature increase from 10 ◦ C (283 K) to 80 ◦ C (353 K, red symbols)
followed by the decrease to the original value (blue symbols)

without any deterioration of the QD stability. In the following discussion, we address


separately the effects of heating/cooling on the PL intensity, PL energy, and PL
decay rate.

3.1.1 Variations of PL Intensity

An increase of temperature of aqueous colloidal AIS QD solutions results in a


PL quenching proportional to the temperature increment (Fig. 19a). In particular,
heating from 10 to 80 ◦ C induces a threefold decrease of the PL QY – from 36%
to ≈12% (Fig. 19b). The reverse cooling of the solution results in the complete
restoration of the PL QY to the original value, and such “heating/cooling” cycle
of the PL intensity change can be repeated many times (five consecutive cycles
presented in Fig. 19b) without any noticeable signs of aggregation of AIS QDs
or degradation of their emissive capacity. Also, the PL FWHM remains constant
during the heating-/cooling-induced variations of the PL intensity (Fig. 19c, circles
1). These observations indicate that the heating-induced PL quenching originates
not from some irreversible deterioration of the QDs, but from a reversible process
of activation/deactivation of some non-radiative relaxation pathways.
The behavior of core/shell AIS/ZnS QDs under the heating/cooling cycles is very
similar to that of the core-only AIS QDs. A series of six size-selected fractions of
aqueous colloidal AIS-GSH and AIS/ZnS-GSH QDs produced by using the size-
selected precipitation procedure with 2-propanol as a non-solvent [14] showed the
same character of the temperature dependences of the PL QY (Fig. 20a) as the
original colloidal QD ensemble, with some variations in the slopes of the “PLQY-T”
dependences discussed below in more detail.
Earlier we reported on the synthesis and PL properties of ultrasmall (the size
d < 2 nm) CdS QDs stabilized by Cd(II) complexes with PEI [100, 113] or by
a combination of MA and ammonia [10, 101]. Both types of CdS QDs revealed
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 97

Fig. 20 Variation of PL QY (a) and PL maximum energy (b) with temperature in a cycle of
heating/cooling for a series of fractions of size-selected GSH-capped aqueous core/shell AIS/ZnS
QDs (fraction numbers are provided in the figure)

very similar reversible variations of the PL intensity when heated up and cooled
down in aqueous colloidal solutions. Similar to AIS QDs, the heating-/cooling-
induced absorption and PL changes of ultrasmall CdS QDs are reversible and
could be repeated many times without any deterioration of the stability of the CdS
QDs. By analyzing the temperature dependences of absorption and PL properties
of CdS-PEI QDs, we concluded [113] that the heating-induced PL quenching of
colloidal QDs originates from the thermally activated dissociation of surface Cd(II)-
PEI complexes and the formation of Cd(II)-water complexes acting as centers of
non-radiative vibrational relaxation. Cooling of the colloidal solution results in the
reestablishment of Cd(II)-PEI complexes on the QD surface and restoration of the
original PL efficiency. We further assumed that the heating-induced PL quenching
of GSH-capped AIS QDs also arises from the thermally activated dissociation of
metal-GSH complexes on the surface of the AIS QDs and introduction of water
molecules into the inner coordination sphere favoring a non-radiative vibrational
relaxation of the excitation energy.
If the effects of PL quenching and de-quenching under heating/cooling cycles
originate from the activation/deactivation of additional pathways of the non-
radiative recombination, we can take the PL intensity as a quantitative parameter
proportional to the rate constant of the non-radiative recombination and evaluate an
apparent activation energy Ea of the quenching/de-quenching from the temperature
dependences of PL intensity IPL . These dependences were found to be fairly linear
in the coordinates of the Arrhenius equation “ln(IPL ) − 1000/T” (Fig. 21a) both
for the core AIS QDs (blue rectangles 1) and for the core/shell AIS/ZnS QDs (red
rectangles 2). Linear fits of these plots yielded Ea = (130 ± 10) meV for AIS QDs
and (90 ± 10) meV for AIS/ZnS QDs, respectively.
These values need to be compared with the enthalpies of formation of the GSH
complexes with corresponding metal ions – Ag(I) and In(III) in the case of AIS
and Zn(II) in the case of core/shell AIS ZnS QDs – to evaluate the feasibility of
98 O. Stroyuk et al.

Fig. 21 (a) Temperature dependences of PL QY of AIS QDs (curve 1) and AIS/ZnS QDs (2)
presented as ln(PL QY) versus 1000/T. (b) Apparent activation energy Ea of thermal PL quenching
of AIS QDs in different fractions (red bars) as well as CdS-PEI and CdS-NH3 -MA QDs (blue bars)

the thermally activated dissociation of such surface complexes as a possible reason


for the PL quenching. Among these metals we could only find reliable data on the
thermodynamics of the formation of Zn(II) complexes with GSH analogues relevant
for bio-applications. In particular, for Zn(II) complexes with glutathione disulfide
(dimer of GSH), the formation enthalpy was reported to be 15 KJmol−1 assuming
a reaction between Zn2+ and the L3− ligand [122], corresponding to 150 meV for
a single complex molecule. This value is on the same order of magnitude as the Ea
energies calculated from the above-discussed temperature dependences for AIS/ZnS
QDs supporting our assumption that the PL quenching results from the dissociation
of the Zn(II)-GSH complexes on the surface of AIS/ZnS QDs. As mentioned, no
report on the enthalpies of formation for GSH complexes with Ag(I) or In(III)
can be found to be compared with the Ea calculated in the present work, but these
complexes with GSH are expected to be more stable than those with Zn(II), resulting
in a higher activation energy for the case of the PL quenching of core-only AIS QDs,
in accordance with the experimental results.
We also evaluated the Ea values for a series of size-selected AIS QDs and found
this parameter to be size-dependent, increasing from 120 meV for fraction 2 of the
largest AIS QDs to ≈215 meV for fractions 6–7 containing the smallest AIS QDs
(Fig. 21b). These data show that we should expect a stronger PL dependence on
temperature for smaller AIS QDs, which is also in accordance with our experimental
findings.
As discussed above, the temperature dependences of the intensity and spectral
parameters of aqueous AIS (and AIS/ZnS) QDs stabilized by GSH complexes are
very similar to those of aqueous CdS QDs stabilized by Cd(II)-PEI complexes [100,
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 99

113] and Cd(II)-NH3 -MA complexes [10, 101]. All three types of QDs indeed show
linear Arrhenius-like dependences of the PL intensity on temperature with apparent
PL quenching activation energies of 2 nm CdS-PEI and 2 nm CdS-NH3 -MA QDs
estimated to be (310 ± 10) meV and (250 ± 10) meV, respectively (Fig. 21b, blue
bars). As discussed in Ref. [113], the Ea of CdS-PEI QDs is very close to the
dissociation energy of Cd(II) complexes with ammonia or ethylenediamine which
can be taken as models of PEI showing that the thermally induced PL quenching
of CdS-PEI and AIS-GSH QDs has the same origin and depends on the nature of
the capping ligand. The Ea value determined for CdS-NH3 -MA QDs is closer to the
Ea range of the smallest of the size-selected AIS QDs (see Fig. 21a), than for the
case of CdS-PEI. This can be expected because GSH and the NH3 /MA combination
formally provide the same functionalities to be bound to the QD surface, while PEI
can bind only via the amino-groups.

3.1.2 Variations of the PL Band Maximum Position

Simultaneously with the PL intensity decrease, heating of aqueous AIS QD


solutions also results in a “red” shift of the PL band maximum (Fig. 22a, red
rectangles 2) – from 2.01 eV at 10 ◦ C (283 K) to 1.95 eV at 80 ◦ C (353 K). The same
behavior was observed for all size-selected series of AIS QDs (Fig. 20b) indicating
a general character of this phenomenon for all the AIS QDs present in the ensemble
with no specific size selectivity. The shift displays the same reversible behavior with

Fig. 22 (a) Temperature dependences of PL maximum energy of AIS QDs registered for QDs
dried on a glass plate (rectangles 1) with T increasing from 220 to 310 K and for aqueous colloidal
AIS QDs with T increasing from 283 K to 353 K (rectangles 2) and decreasing back from 353 K
to 283 K (circles 3). Arrows guide the eye in the direction of the T change. (b) Temperature
dependences of the PL band maximum energy decrement (curves 1–3) of a QD ensemble dried
on a glass plate (1) and in aqueous colloid (2), as well as of a fraction of the smallest colloidal
AIS QDs produced from the ensemble (3). Curve 4 shows the decrement of the first absorption
maximum energy of aqueous PEI-capped CdS QDs upon heating. (c) Temperature coefficient of
the PL band maximum energy of size-selected colloidal AIS QDs for different fractions
100 O. Stroyuk et al.

the PL band peak position restoring to its original value as the colloidal QD solution
is being cooled again (Fig. 22a, blue circles 3). Figure 22a also shows a section
of the temperature dependence of the PL band peak of AIS QDs (curve 1) probed
in a far larger T window of 10–300 K. In this experiment (see details in the next
section), the AIS QDs were dried on a glass substrate before the measurements, and
the comparison of the datasets gives a clear impression about the effect of water on
the temperature dependences of the PL properties of AIS QDs.
It can be seen that the slopes of the curves 1 and those of the curves 2 and 3 are
considerably different, i.e., it is steeper for the cycle of heating/cooling of aqueous
colloidal AIS QD solution than for the same QDs deposited on glass and dried for
the cryogenic measurements. This observation clearly indicates that the mechanism
of the temperature-induced changes of the PL parameters of AIS QDs is different
for aqueous colloids and in the case of dried QDs. Earlier, we observed a similar
thermally induced “red” shift of the excitonic maximum in absorption spectra of
the colloidal CdS-PEI QDs. This shift is completely reversible, and the position of
the absorption maximum restores completely after the solution is cooled back to the
original temperature [113].
Another typical feature common for the temperature dependences of PL param-
eters of aqueous AIS and CdS QDs is a weak or no temperature dependence of the
PL FWHM in the T range studied (Fig. 19c). The FWHM value of both CdS-PEI
and CdS-NH3 -MA QDs is the same and shows a very slight trend to increase from
≈570 meV to ≈600 meV as the temperature is elevated from 10 to 80 ◦ C with a
perfectly reversible character of this change upon cooling of the colloidal solutions
(Fig. 19c, rectangles). The aqueous colloidal AIS QDs also demonstrate a stability
of the FWHM values in this T range with a very slight declining trend from 400 meV
at 10 ◦ C to ≈390 meV at 80 ◦ C and the same cyclic character of the FWHM changes
in the cooling/heating cycles (Fig. 19c, circles).
Figure 22b shows a relative decrement (energy values at 10 ◦ C taken as zero)
of the PL band energy achieved by decreasing temperature of dried AIS-GSH QDs
on the glass plate (curve 1) as well as of aqueous colloidal AIS-GSH (curves 2,3)
and CdS-PEI (curve 4) QD solutions. The data show that the shift of the PL band
maximum is much larger for aqueous colloidal AIS-GSH solutions as compared to
the dried samples, but nevertheless smaller than the shift of the excitonic absorption
feature detected for CdS-PEI QDs (compare curves 2 and 4). At the same time, these
colloidal samples are characterized by a different QD size, namely, ≈3 nm for the
AIS-GSH colloid and ≈2 nm for the CdS-PEI QD solution. In view of this fact, we
compared CdS-PEI QDs with fractions 6–7 of the smallest (d ≤ 2 nm) size-selected
AIS-GSH QDs and found that the heating-/cooling-induced PL maximum shift for
such QDs is indeed larger than for the AIS QD ensemble and very close to the
shift of the excitonic absorption maximum of 2 nm CdS-PEI QDs (compare curves
3 and 4, Fig. 22b). We can conclude from these observations that the temperature
dependences of the spectral PL parameters of aqueous colloidal AIS QDs depend
on the QD size and for the case of 2 nm AIS QDs are almost the same as for 2 nm
CdS QDs despite some differences in the structure of capping ligands.
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 101

The stabilization of previously studied PEI-capped CdS QDs originates from


the formation of complexes of undercoordinated or adsorbed Cd(II) cations on the
QD surface with amino-groups of PEI [100, 113]. The macromolecular PEI can be
substituted by a combination of Cd(II) complexes with ammonia and MA forming
in situ during the synthesis of CdS QDs from the mixture of Cd(II)-NH3 and Cd(II)-
MA complexes [10, 101]. Both types of CdS QDs capped either by Cd(II)-PEI
or by Cd(II)-NH3 -MA complexes were found to have almost identical absorption
and PL properties indicating a very similar stabilization mechanism. The AIS QDs
are produced from a mixture of Ag(I) and In(III) complexes with GSH molecules
[14], the latter containing mercapto-groups, amino-groups, and carboxyl-groups,
and binding to the AIS QD surface in a similar way as PEI or Cd(II)-NH3 -MA
complexes bound to the surface of CdS QDs. It can be anticipated, therefore, that for
the CdS and AIS QDs stabilized by the same mechanism, that is, by the formation
of surface complexes, the temperature dependences of the optical properties will
be similar, provided that the QD size is the same. The results discussed above and
shown in Fig. 22b definitively speak in favor of this assumption.
The reversible shift of the excitonic absorption of CdS-PEI QDs was previously
explained by a decrease of the spatial exciton confinement resulting from the
thermally activated dissociation of the surface Cd(II)-PEI complexes [113]. The
amino-groups of PEI possess free electron pairs that form a layer of increased
electron density on the QD surface and contribute to the confinement of photo-
generated electrons. The thermal dissociation of Cd(II)-PEI complexes eliminates
this additional factor of electron confinement resulting in a decrease of the exciton
energy without any actual change in the QD size. Similar effects were reported to
arise from the substitution of electron-donating to electron-accepting ligands on the
surface of “magic-size” CdX clusters [123, 124]. We assume that such effects can
be (at least partially) responsible for the reversible shifts of the PL band maximum
for the case of AIS-GSH QDs.
In contrast to the ultrasmall CdS QDs, both absorption and PL excitation spectra
of AIS and AIS/ZnS QDs reveal no distinct excitonic features and the presence
of broad sub-bandgap “tails” [13, 14, 26]. Such structure of the absorption spectra
makes the assessment of the bandgap quite tricky and prohibits the evaluation of
the thermally induced shifts in the absorption/PL excitation spectra of ternary QDs.
At the same time, the PL band position of AIS QDs is directly dependent on the
bandgap energy, and, therefore, the changes of the PL band maximum energy can
be taken as directly reflecting the temperature-induced changes of the bandgap of
AIS QDs.
As the PL band maximum energy depends in an almost linear manner on
temperature of the colloidal solutions both for the AIS QD ensemble (Fig. 22a)
and size-selected AIS QDs (Fig. 20b), we can characterize different QDs by
a linear temperature coefficient α T . The AIS QD ensemble is characterized by
α T = (0.85 ± 0.05) meVK−1 , which is quite close to the α T value obtained
earlier from the analysis of the temperature dependence of the exciton absorption
maximum of colloidal CdS-PEI QDs – (0.95 ± 0.05) meVK−1 [113]. This fact
indicates a general mechanism of the spectral shifts in both cases as discussed
102 O. Stroyuk et al.

above. Typically, much lower α values on the order of 0.3–0.5 meVK−1 are observed
for CdS [125–127] and CdSe [128] QDs stabilized by conventional covalently
bound surface ligands even for the cases of ultrasmall QDs [126, 128]. Also, the
α T coefficient for oleylamine-capped ZCIS/ZnSe/ZnS was found to be only 0.23
meVK−1 in the T range of 50–373 K [110]. Therefore, the present results show that
water-dispersed metal-chalcogenide QDs stabilized with surface metal complexes
are much more sensitive to temperature variations due to the labile thermally
activated dissociation/association behavior of such complexes in aqueous solutions.
The size-selected colloidal AIS QDs revealed a dependence of the temperature
coefficient α T on the fraction number (Fig. 22c), that is, on the average QD size. The
largest QDs from the original ensemble (fraction 2) are characterized by α T = 0.55
meVK−1 . This value increases to 1.25–1.40 meVK−1 for smaller QDs in fractions
6 and 7. The size dependence of the α T value can be easily understood, as the
influence of surface ligands on the electronic system of the entire QD is expected to
be stronger for smaller QDs.
A unique combination of quite strong PL quenching of AIS QDs induced by
heating with a distinctly detectable “red” shift of the PL band maximum position
opens possibilities to utilize these broadband-emitting QDs as a ratiometric sensor
accurately reporting temperature changes irrespectively of the absolute value of the
PL intensity of the emitting QDs.
This approach was developed in Ref. [129] for broadband-emitting CdSe QDs
and is based on the fact that the PL quenching is accompanied by a shift of the PL
band, which is even more pronounced in the present case of AIS QDs than in the
case for CdSe QDs reported in Ref. [129]. In this method, the PL emission of QD
reporters is detected independently in two channels arbitrarily chosen on the left
“wing” of the PL band (channel S1) and on the right “wing” of the PL band (S2), as
depicted in Fig. 23a for the case of AIS QDs.

Fig. 23 Temperature-dependent variation of the PL spectrum of aqueous AIS QDs in a range of


10–80 ◦ C. (a) Bars mark two PL ranges (S1 range, 500–525 nm, and S2 range, 600–625 nm) used
for the self-calibration procedure. (b) Self-calibrated PL intensity versus temperature (circles).
Solid red line corresponds to a linear fit of the experimental data
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 103

The sensor output is determined as R = (S1 –S2 )/(S1 + S2 ), where S1 and S2 are
the areas under the spectral curves in two selected channels (Fig. 23a). This value
changes with the temperature because of the spectral shift of the PL band induced
by heating/cooling. The procedure of the simultaneous detection of PL signals in
two independent channels and online calculation of the R ratio was reported to be
feasible by using a two-channel portable spectrometer and corresponding software
[129]. R changes linearly with temperature for different spectral S1 /S2 scopes and
different sizes of CdSe QDs. Here, we found that the R ratio is also linear for
aqueous GSH-stabilized AIS QDs (as well as for the core/shell AIS/ZnS QDs,
both tested as unfractionated colloidal ensembles) in the entire probed temperature
range of 10–80 ◦ C (Fig. 23b). This feature of aqueous AIS and AIS/ZnS QDs, in
combination with high PL QYs, QD stability in many heating/cooling cycles, as
well as a reasonable stability of QDs toward photobleaching in the conditions of
conventional PL measurement experiments, opens perspectives for the application
of GSH-capped AIS/ZnS QDs in biosensing systems.

3.1.3 Variations of the PL Decay Rate

The PL decay curves of aqueous colloidal AIS-GSH QDs reveal a non-


monoexponential character (Fig. 24a) typical for such ternary QDs [4]. An increase
of the temperature of AIS (and AIS/ZnS) colloids results in an acceleration of the
PL decay indicating a considerable redistribution of the efficiencies of radiative
and non-radiative secondary processes in QDs. As the colloids are cooled back,
the PL decay is retarded, and the decay curves registered at the same temperatures

Fig. 24 (a) Kinetic curves of PL decay of colloidal AIS QDs detected at different temperatures
(gray lines) approximated by linear combinations of three monoexponential functions (red lines).
(b) PL lifetime as a function of temperature of colloidal AIS QD solution in a regime of T increase
(red squares) and consecutive T decrease (blue squares)
104 O. Stroyuk et al.

during the heating and cooling stages almost coincide showing a perfectly reversible
character of the PL decay rate changes with no appreciable hysteresis effects.
We applied two models for the description of the PL decay curves. The first is
a conventional description of PL decay curves with linear combinations of three
single-exponential functions [14, 26, 130]. It allows the average PL lifetime to
be evaluated and the representative values of rate constants of radiative and non-
radiative recombination (provided that PL QYs are available) to be calculated
without any assumptions on the exact emission mechanism. The second model is
based on a Kohlrausch-type stretched exponent model describing the PL emission
as a purely radiative process perturbed by non-radiative energy transfer processes
with the participation of the vibrational states of the surface ligands [131]. This
model appears to be less favorable for the qualitative description of the PL decay
dynamics because of high error margins of the fitting parameters, but it allows the
contribution of the interactions between the colloidal QDs, their ligand “shell,” and
the surrounding dispersive medium to the dynamics of the radiative recombination
to be evaluated.
Fitting of the PL decay curves with linear combination of monoexponential
functions. The PL decay curves can be satisfactorily fitted (Fig. 24a, red solid lines)
with linear combinations of three single-exponential components and the average
radiative lifetime calculated from Eq. (4). An increase in temperature of aqueous
AIS QD colloids was found to result in an acceleration of the PL decay, the average
PL lifetime decreasing from 420 ns at 10 ◦ C to ≈250 ns at 70 ◦ C in an almost linear
manner (Fig. 24b, red rectangles). As the solution is cooled again, the average PL
lifetime increases monotonously reaching the starting value of 420 ns at 10 ◦ C (Fig.
24b, blue rectangles) and showing a perfectly cyclic character of the temperature-
induced changes of the PL decay rate in many heating/cooling cycles (data not
shown). A qualitatively similar picture was found for the core/shell AIS/ZnS QDs,
and the PL lifetime varies from 495 ns at 15 ◦ C to 370 ns at 70 ◦ C and back
in a cyclic manner without changes in the stability of QD solution in multiple
heating/cooling cycles.
The collected array of PL QY and PL lifetime values allows the rate constants
of the radiative (kr ) and non-radiative (knr ) recombination in AIS (AIS/ZnS) QDs
at different temperatures to be calculated and the activation energy of the PL
quenching to be assessed from the kinetic data. The rate constants of the radiative
recombination were found to be higher for core/shell AIS/ZnS QDs than for core
AIS QDs, both types of QDs showing a monotonous decrease of kr with an increase
in temperature (Fig. 25a). At that, the core AIS QDs reveal a stronger dependence of
kr on temperature (around 40% reduction of kr with temperature elevated from 10 to
70 ◦ C for AIS QDs versus ≈20% reduction for AIS/ZnS QDs). This fact indicates a
higher activation energy of the PL quenching of core AIS QDs, in accordance with
the above-discussed Ea estimations made from the stationary PL measurements. The
rate constant of the non-radiative recombination also showed a strong dependence
on temperature, the knr increasing by a factor of 2.3 (AIS QDs) and 2.0 (AIS/ZnS
QDs) as T is elevated from 10 to 70 ◦ C (Fig. 25b).
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 105

Fig. 25 Rate constants of radiative (a) and non-radiative (b) recombination in AIS QDs (curves
1) and core/shell AIS/ZnS QDs (curves 2) at different temperature (in heating regime). (c) Plot of
ln(knr ) versus 1000/T for colloidal AIS QDs (curve 1) and AIS/ZnS QDs (2)

Figure 25c shows that the temperature dependences of knr are reasonably linear
in the coordinates of the Arrhenius equation (rectangles 1 for AIS and 2 for
AIS/ZnS QDs). The dependences yield activation energies for the non-radiative
recombination Ea = (120 ± 5) meV for AIS QDs and (90 ± 5) meV for AIS/ZnS
QDs (the same Ea values were obtained for the “heating” and “cooling” branches
of the cyclic temperature dependence of knr ). Both values show a perfect agreement
with the values derived from the analysis of the temperature dependences of the
stationary PL intensity, the latter one being also close to the dissociation enthalpy
reported for Zn(II)-GSH complexes. Therefore, the kinetic data unambiguously
show that the thermally activated non-radiative recombination processes can be
considered as a main reason for the PL quenching. The numerical values of Ea allow
these non-radiative processes to be associated with the reversible dissociation of the
surface metal-GSH complexes and introduction of water into the first coordination
sphere of QDs.
Fitting of the PL decay curves with stretched exponential functions. Alternatively
to the well-reported examples of fitting the PL decay curves of metal-chalcogenide
QDs with linear combinations of two or three monoexponential functions, the PL
decay can be described as a purely radiative process competing with a channel of
electronic energy transfer to vibrational states using a Kohlrausch-type stretched
exponential function
   
t t β
I (t) = I0 exp − −α , (7)
2τb 2τb

where τ b is the lifetime of the “bright” state (purely radiative lifetime) and α and
β are constants characterizing the energy transfer mechanism and the strength of
interactions between the energy donor and acceptor [131].
106 O. Stroyuk et al.

In particular, this model was applied to fit the PL decay curves of a series of size-
selected CdSe QDs capped with trioctylphosphine oxide (TOPO) molecules [131].
The authors of this report assumed that the deviation of the PL decay curves of CdSe
QDs from the single-exponential character arises from the energy transfer between
the electronically excited donor (CdSe QD) and vibrational states of acceptor (C-H
vibrations of TOPO ligand). They found the parameter β to be close to 0.5 typical
for the Förster-type resonant energy transfer from a donor to acceptors randomly
distributed in a three-dimensional medium. The parameter α was found to depend
on CdSe QD size and interpreted as an overlapping factor of the electronic states of
the donor and vibrational states of the acceptor. In this view, the α can be viewed as
a qualitative parameter proportional to the probability of such electronic-vibrational
energy transfer.
We found that the PL decay curves of colloidal AIS QDs detected at different
temperatures can be well fitted with the Kohlrausch stretched exponential function
allowing τ b , α, and β to be calculated for different temperatures (Fig. 26). As in
the case of TOPO-capped CdSe QDs [131], the parameter β varies in the range of
0.5–0.6 indicating the same energy transfer mechanism as discussed in Ref. [131].
The characteristic PL lifetimes produced by fitting the PL decay curves with
the Kohlrausch function were found to be rather constant as well, varying around
400 ns in the whole T range studied both during the heating and cooling parts of
the cyclic T changes (Fig. 26a). Typical error margins for the τ b values were around
±10–15% (much higher than for the average lifetimes determined using the above
three-exponential functions) which makes it difficult to discern any dependences
between τ b and temperature in the rather narrow T range studied here (10–80 ◦ C).
At the same time, the average τ b , 400 ± 50 ns, is close to the <τ > values determined
by the three-exponential fitting at the lowest temperature of 10 ◦ C, (420 ± 5) ns,
indicating a good correspondence between the two fitting models.

Fig. 26 Fitting parameters τ b (a), α (b), and β (c) derived from Kohlrausch stretched exponential
functions at different temperature in the forward heating (red squares) and backward cooling (blue
circles) regimes. The arrows guide the eye to the temperature change direction
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 107

The parameter α depends quite strongly on the temperature of the colloidal AIS
QD solution increasing from around 0.8 at 10 ◦ C to around 2.6 at 70 ◦ C (Fig. 26b,
red rectangles). This change was found to be completely reversible, the α decreasing
upon cooling and attaining the original value as the temperature is lowered back to
10 ◦ C (Fig. 26b, blue circles). This change can be reproduced many times in a cyclic
manner.
As discussed in Ref. [131], the parameter α is proportional to an overlap of the
emission spectrum maximum of the QDs ν max and √
vibrational absorption spectrum
ε
of the surface ligands εvib (ν max ), i.e., α ∝ ν 2 vib . Therefore, this parameter is
max
expected to increase as ν max lowers with an increase of the QD size. In accordance
with this expectation, the TOPO-capped CdSe QDs showed an increase of α from
0.20 to 0.33 as the average QD size increased from 2.5 nm to 5.0 nm [131]. We
should note that the α values derived for aqueous AIS QDs are much higher than
those reported for CdSe-TOPO QDs even at the lowest T studied. Thus, the GSH-
capped aqueous AIS QDs demonstrate a much stronger interaction with the ligand
shell and the solvent environment, even for roughly the same average QD size, as
compared with CdSe-TOPO QDs reported in Ref. [131].
As discussed earlier, the heating of AIS QDs results in a “red” shift of the PL
band maximum. This shift means a decrease of νmax and can, therefore, contribute
to the observed increase of the parameter α with temperature. However, even rough
estimations assuming a constant εvib , but varying ν max from 16,210 cm−1 (2.01 eV)
to 15,720 cm−1 (1.95 eV), that is, in the range of PL band maximum energies
presented in Fig. 22a, show that such “red” shift of the PL band can result only in
7–8% increase of the parameter α. Therefore, the main contribution to the heating-
induced increase of α should originate from the changes of the εvib spectrum.
The parameter α should grow with an increase in the energy of the anharmonic
oscillators participating in the energy transfer. In this view, the increase of α can
be interpreted as a result of the dissociation of the surface QD-GSH complexes
and incorporation of water into the first coordination sphere of the AIS QDs, as
discussed before based on the stationary and time-resolved PL measurements. The
surface water complexes can supply high-energy anharmonic oscillations of OH
groups to contribute to the energy transfer increasing in this way the factor α
proportionally to the temperature of the colloidal solution. This interpretation is in
line with the above assumptions on the complex equilibria on the AIS QD surface
responsible for the PL changes and provides a physically relevant background for
the analysis of the time-resolved PL traces.

3.2 Temperature Range of 10–300 K

The studies of PL temperature dependences in a broad temperature range descend-


ing from room temperature to a few Kelvins provide ample information on the
origins and mechanisms of both radiative and non-radiative processes of electron-
108 O. Stroyuk et al.

hole recombination and charge transfer in single semiconductor QDs [116, 132]
and QD ensembles [133–137], as well as of the interfaces between QDs and various
matrices [115, 125, 138]. The studies of the PL temperature dependences of CdX
[115–121, 133, 135, 137] and PbX QDs [139, 140] allowed the structure and
energies of emitting and “dark” states, the size dependences of electron-phonon
interaction, the contribution of lattice vibrations to the relaxation of charge carriers,
and other vital parameters to be assessed. Furthermore, the relative contributions of
ordered QD core and disordered QD surface areas to the distribution of PL energies
and lifetimes were estimated from temperature-dependent PL evolutions [118, 133,
141].
A general trend to sustainable semiconductor nanomaterials drew wide attention
to multinary (ternary, quaternary, and more complex) metal-chalcogenide composi-
tions based on indium, gallium, tin, etc. as potential alternatives of binary CdX and
PbX QD absorbers and emitters in light-harvesting and PL applications [4, 16, 32,
142, 143]. Recently our group reported on the aqueous syntheses of ternary CIS,
AIS, and Cu-doped AIS QDs as well as corresponding core/shell QDs with a zinc
sulfide shell stabilized by surface metal complexes with a variety of ligands [13, 14,
25, 26, 111, 112, 144]. Such ternary QDs demonstrate relatively strong PL (with
PL quantum yields of up to 55–60%) emitted in broad symmetrical bands strongly
red-shifted relative to the corresponding absorption band edges.
By analyzing the spectral PL parameters of core AIS and core/shell AIS/ZnS
QDs subjected to the size-selection procedure [14] and comparing the PL properties
of single AIS/ZnS QDs and QD ensembles [27], we gained evidence that the
broadband PL of such QDs should be described in terms of a self-trapped exciton
(STE; see Sect. 2) model that assumes the emission band to be a series of phonon
replicas of a zero-phonon line, rather than by the conventional donor-acceptor
pair recombination model involving energetically and spatially distributed deep
intra-bandgap charge traps [142, 145, 146]. Indeed, the PL properties of ternary
Ag(Cu)-In-S QDs differ fundamentally from the PL properties of larger (d ≥ 3 nm)
CdX QDs abundantly reported in the literature and studied by us earlier [147–151].
Such QDs are often referred to as “regular” QDs and typically show a “classical” PL
pattern with two components – an excitonic PL (EPL) band near the absorption edge
and a defect-related PL (DPL) band strongly broadened and red-shifted relatively
to the bandgap [149–151]. The CdS and CdSe QDs were reported to exhibit clear
differences in the temperature-dependent evolution of EPL and DPL features [118,
125, 133, 135–137, 152], and, therefore, a comparison of the temperature-dependent
PL properties of “regular” CdX QDs emitting both EPL and DPL with those of
ternary AIS and CIS QDs can provide unique information on the PL mechanisms
inherent to the latter complex objects. In this section we discuss variations of PL
properties of aqueous AIS and core/shell AIS/ZnS QDs capped with GSH and
dropcasted on glass in a broad temperature window of 10–310 K.
The PL bands of AIS and AIS/ZnS QDs can be fitted with a single Gauss profile
and do not show any multiple components, as in the case of the “regular” CdX
QDs. Though being apparently similar to the DPL bands of CdX QDs by spectral
parameters (large FWHMs and Stokes shifts), the PL of AIS and AIS/ZnS QDs
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 109

is obviously different, showing very high efficiency and no size- or composition-


related variations in FWHM [13, 14, 27]. Here, we show that, despite its apparent
spectral resemblance of DPL, the nature of emission of the ternary AIS-based QDs
indeed resembles EPL and definitely differs from the DPL of CdX QDs.
The PL band shape of AIS (AIS/ZnS) QDs remains basically the same in the
entire temperature range of 10–310 K (Fig. 27). Similar observations indicating a
small or no variation of the PL band shape with decreasing temperature were also
reported for Zn-doped 5 nm AISe QDs [108] and 2.3–2.7 nm CIS/ZnSe/ZnS QDs
[110].
The PL intensity of AIS QDs was found to increase as temperature is lowered
reaching some saturation in the range of T = 170–220 K and then gradually decrease
at T < 150 K almost to the same values at 50 K as those observed at 290–300 K (Fig.
28a, blue rectangles). The core/shell AIS/ZnS QDs show the same tendency, the PL
intensity increasing from T = 300 K to 160–170 K, reaching a certain saturation
level at T = 120–150 K and decreasing at lower T – first slightly at T = 70–120 K
and then abruptly, as T becomes lower than 70 K (Fig. 28a, red rectangles).
The temperature dependences of the PL intensity of aqueous GSH-capped
AIS and AIS/ZnS QDs differ considerably from those reported earlier for similar

Fig. 27 PL spectra of
AIS/ZnS QDs on glass at
different temperatures

Fig. 28 PL intensity (a), PL band maximum position (b), and PL band FWHM (c) for core AIS
and core/shell AIS/ZnS QDs as functions of temperature
110 O. Stroyuk et al.

compound QDs. In particular, 1-dodecanethiol-/oleylamine-capped CIS and Zn-


doped CIS (ZCIS) QDs [153–155], 5 nm ZAISe QDs [108], and 2.3–2.7 nm
ZCIS/ZnSe/ZnS QDs [110] showed a non-monotonous increase in the entire
temperature range studied. At the same time, a decrease in the PL intensity in
the temperature range of T < 50–70 K was reported earlier for Si QDs [156], CdS
QDs [125], CdSe QDs [141], PbS QDs [140], and 2.6 nm dodecanethiolate-capped
AIS QDs [12] and attributed mainly to a thermally activated transition from the
delocalized excitonic state into the STE state [125, 140, 141, 156], which becomes
strongly retarded at lower T, thus resulting in the decrease of the PL quantum yield.
As suggested in Refs. 133, 135, 136, 157, the broadband and “red”-shifted
emission of CdSe QDs can be attributed to the STE recombination in the surface
QD layer. Depending on the QD size and ligand nature, the radiative recombination
can proceed entirely through this channel resulting in a single broad PL band, as is
the case for the reported ultrasmall CdX QDs [123, 126], Si QDs [158], and PbS
QDs [140] and for the aqueous GSH-capped AIS and AIS/ZnS QDs studied by our
group. The differences in the shape of temperature dependences of the PL intensity
between AIS and AIS/ZnS QDs (Fig. 28a) may be related to different activation
energies of the free-to-self-trapped excitonic transition. Besides, the AIS QDs under
discussion are synthesized at a Ag:In ratio of 1:4 that is far from the stoichiometry,
and thus the lattice of such QDs is expected to be abundant with silver vacancies
that can be partially filled with Zn after the deposition of a ZnS shell [13, 14]. The
shell formation can, therefore, result in an increase of the lattice rigidity and affect
strongly the temperature dependence of the PL intensity. A similar phenomenon
of the ZnS shell influence on the PL intensity was reported for CdSe QDs [141].
The deposition of a ZnS shell on nonstoichiometric CIS QDs was found to suppress
considerably the PL quenching induced by the temperature increase from 6 to 300 K
[155]. We also reported a strong suppression of optical phonon decay, after the CdSe
QD-polymer interface was replaced by a CdSe/ZnS one [151, 159].
The shape of T dependences of PL intensity of AIS-based samples shows that we
definitively deal with a different PL mechanism as compared to binary CdX QDs.
At the same time, temperature dependences of the PL band maximum of both AIS
and AIS/ZnS QDs behave in a “classical” way, that is, these parameters increase
monotonically (Fig. 28b) as could be expected for the EPL band of “regular” binary
CdX QDs [33, 116, 117, 121]. Only, some drop of the PL band energy can be
observed at T < 20 K for the core/shell AIS/ZnS QDs (Fig. 28b, red rectangles)
which is unique for these QDs. Other reports on ternary chalcogenide QDs also
show them to demonstrate “classical” temperature dependences of the PL band
maximum typical for the EPL of conventional binary QDs. This behavior was
reported for 5 nm ZAISe QDs [108] and 2.3–2.7 nm ZCIS/ZnSe/ZnS QDs [110].
A comparison of the dependences presented in Fig. 28b shows that core-only
AIS QDs are more susceptible to temperature-dependent changes in the electronic
structure than the core/shell AIS/ZnS QDs. In particular, the overall increment of
the PL maximum energy achieved by decreasing T from 313 to 10 K amounts to
0.042 eV, which is only 2.1% of the starting EPL value at T = 313 K for AIS/ZnS
QDs, and to 0.083 eV, that is, 4.4% of the starting EPL value for the AIS QDs. We
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 111

can conclude from these observations that the lattice of AIS/ZnS QDs, which is
more stiff due to filling of Ag vacancies by Zn ions from the shell, is less prone to
changes with temperature. This phenomenon can also be the reason for differences
in the shape of the T dependences of PL intensity between AIS and AIS/ZnS QDs
(Fig. 28a).
The PL FWHM of core AIS QDs also shows a “classical” temperature depen-
dence expected for EPL, decreasing slightly from 0.30 eV to ≈0.29 eV as the
temperature is lowered from 300 to 50 K (Fig. 28c, blue rectangles). The core/shell
AIS/ZnS QDs show, on the contrary, a certain FWHM increment, from ≈0.37 eV
to 0.39 eV, upon lowering T down to 40–50 K and a slight backward decrease
to 0.38 eV at T < 40 K (Fig. 28c, red rectangles). This behavior has no analogy
in the reports on binary QDs and is specific for the presently studied ternary
compounds. For example, a decrease of FWHM by more than 150 meV was reported
for 1-dodecanethiol-capped 5 nm ZAISe QDs with the temperature lowered from
273 to 10 K [108], which can be accounted for by a QD size distribution in the
reported samples. Overlap of two or more emission bands with different temperature
behavior (e.g., two sorts of STE with slower and faster T dependence) may be
responsible, for instance, for the observed anomalous broadening of the AIS/ZnS
PL with lowering T.
Summarizing the discussion of the PL parameters of AIS and AIS/ZnS QDs
studied in a broad temperature range, we note that these QDs reveal certain signs
of the STE behavior reported earlier for a broad range of bulk and nanometer
compounds. The STE model was applied to explain the broad quasi-white PL of
single-crystal TiO2 [160], various metal halides [161–163], and hybrid perovskites
[164–167]. Self-trapping of the exciton was also suggested for explaining the PL
of SiO2 [168], silicon QDs [116, 169], carbon dots [170, 171], and SiC QDs [172]
indicating a very general character of this phenomenon. The exciton self-trapped at
a Te atom was invoked to explain the temperature dependence of the PL intensity
of ZnSe1-x Tex [173], particularly a transfer of the exciton between the trapped and
free-exciton state. This range of compounds can also be extended by the reports on
Si, CdS, CdSe, PbS, and ternary QDs already discussed above [125, 140, 141, 156].
At the same time, the shape of the temperature dependences of the PL intensity
and FWHM vary quite strongly among the reports on the STE emission, clearly
indicating important contributions from the specifics associated with different core
materials, crystal sizes, and ligand environments.
Based on the results presented here and in other works mentioned above, we
may assume that the radiative recombination in QDs can choose between at least
three major mechanisms/routes: (1) direct interband recombination (EPL), (2)
trap-mediated recombination (DPL), and (3) strongly phonon-coupled self-trapped
exciton emission (STE). The domination of one channel or coexistence of several
recombination routes in the given QD sample is the result of numerous factors,
such as QD size, QD crystallinity and stoichiometry, concentration of radiative
and non-radiative traps, properties of the ligand and medium, etc. Overlap of two
or more emission bands with different temperature behavior may be responsible,
for instance, for the observed anomalous broadening of the AIS/ZnS PL with
112 O. Stroyuk et al.

lowering T. Moreover, in the case of very small QDs of 2–3 nm and less, the above-
discussed channels can converge and do not exist in a pure form anymore, because
of the strong electronic confinement, large portion of surface atoms, and perceptible
electronic coupling between the QD interior core and its ligand/environment.

4 Summary

The principal aim of this chapter is to show that relatively “green” aqueous
syntheses of core AIS and core/shell AIS/ZnS QDs combined with the method
of size-selective precipitation constitute a powerful instrument for the preparation
of light-absorbing and light-emitting ternary QD-based materials for potential
applications in photovoltaics, photocatalysis, sensorics, and LED technologies. The
approach is extremely versatile allowing the composition, size, ligand shell, core and
shell material, dopant nature and content, and many other parameters to be varied in
a broad range.
We showed that the structure of the broad PL bands of AIS and AIS/ZnS QDs
can be described by the model of self-trapped exciton implying that the PL band
consists of a sequence of phonon replica of a zero-phonon line resulting from a
strong electron-phonon interaction and a partial conversion of the excitation energy
into lattice vibrations. These results also attest that the position and shape of the PL
bands of AIS QDs originate not from energy factors (depth and distribution of trap
states, etc.), but rather from the dynamics of the electron-phonon interaction and the
vibrational relaxation in the QDs.
By analyzing parameters affecting the PL features like the post-synthesis anneal-
ing and the deposition of a ZnS surface passivation shell, we could confirm the lack
of a correlation between the FWHM of the PL band of AIS and AIS/ZnS QDs and
the density of the lattice defects and corresponding sub-bandgap states. Moreover,
the measurements of the PL of AIS/ZnS QDs in a broad temperature window from
310 to 10 K demonstrated that the FWHM of the PL band does not decrease
with decreasing temperature as expected for an emission arising from thermally
activated detrapping processes as assumed in the conventional donor-acceptor PL
model. Also, we show that the model of the self-trapped exciton can be applied to
reconstruct the shape of PL spectra of size-selected AIS/ZnS QDs and can account
for the effects typically attributed to variations in defect state energy. Measurements
of the PL properties of single AIS/ZnS QDs highlighted the inherent broadband
nature of the emission of individual QDs.
The analysis of the PL spectra of a series of size-selected AIS/ZnS QDs revealed
that the PL FWHM of fractionated QDs does not depend on the QD size and size
distribution, despite the fact that the latter is expected to be narrower than that of the
parent QD ensemble. All these observations indicate that the size, size distribution,
the density, and energetic distribution of possible lattice defects do not affect in
a considerable way the spectral parameters of the PL band, in particular the PL
FWHM, as predicted by the donor-acceptor PL model. These findings also show
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 113

that, despite the inherently broad distribution of the PL energies, the emission of
ternary QDs most probably does not originate from lattice defects, but involves the
QD lattice as a whole and, therefore, by tailoring the QD structure potentially PL
efficiencies as high as those reported for binary cadmium or lead chalcogenide QDs
can be eventually reached.
The rate of the conversion of the electronic excitation into vibrational modes in
AIS QDs was found to be size-dependent, increasing almost twice from the largest
to the smallest QDs in the studied size-selected series. This fact may indicate that
the exciton trapping and coupling to the lattice vibrations results in a much higher
distortion of the lattice of smaller QDs or, alternatively, that the trapping occurs on
the QD surface periphery with the participation of the ligand shell which is expected
to be of a higher efficiency for smaller QDs.
We found that aqueous colloidal AIS (AIS/ZnS) QDs show quite strong PL
quenching and a “red” shift of the PL peak energy when the colloids are heated from
10 to 80 ◦ C. The PL intensity and PL band maximum return to the original values
upon cooling of the solutions and the total reversibility of the spectral PL parameters
can be observed in many cycles of heating/cooling without any losses in the stability
and emissivity of the colloidal QDs. This behavior resembles strongly the previously
reported PL changes of ultrasmall CdS QDs stabilized by Cd(II) complexes and
originates from the reversible thermally activated dissociation of complexes formed
by ligands with the metal cations on the QD surface and introduction of water
into the QD coordination sphere, both processes being completely reversible upon
cooling of the colloidal solution. The proposed mechanism was supported by the
analysis of the temperature-dependent PL decay curves showing that the activation
energy of the non-radiative relaxation processes in AIS (AIS/ZnS) QDs is close to
the dissociation energy of the GSH complexes of the QD surface. The broadband
character of the PL emission combined with both thermal quenching and a “red”
shift of the PL band maximum makes ratiometric temperature determination feasible
in aqueous media using colloidal AIS QDs. The Kohlrausch stretched exponential
function was applied to describe the kinetic PL decay traces of colloidal AIS-
GSH QDs. The physical model behind this function implies that the PL emission
is accompanied by a resonant Förster-type energy transfer between electronically
excited QDs and vibrational states of the surface ligands.
The temperature dependences of the PL peak energy and PL band FWHM of the
ternary AIS and AIS/ZnS QDs were found to differ from those found for defect-
related PL of CdSe and CdSe/ZnS QDs. In contrast, they resemble temperature
dependences typical for excitonic PL and are similar to the data collected for
the direct interband radiative recombination of CdSe and CdSe/ZnS QDs. In
combination with the previously reported results, these temperature-dependent PL
signatures serve as an additional argument in favor of the STE mechanism of the
broadband PL emission of ternary AIS (AIS/ZnS) QDs.
114 O. Stroyuk et al.

References

1. Aldakov, D., Lefrançois, A., Reiss, P.: Ternary and quaternary metal chalcogenide nanocrys-
tals: synthesis, properties and applications. J. Mater. Chem. C. 1, 3756–3776 (2013)
2. Kolny-Olesiak, J., Weller, H.: Synthesis and application of colloidal CuInS2 semiconductor
nanocrystals. ACS Appl. Mater. Interfaces. 5, 12221–12237 (2013)
3. Torimoto, T., Kameyama, T., Kuwabata, S.: Photofunctional materials fabricated with
chalcopyrite-type semiconductor nanoparticles composed of AgInS2 and its solid solutions.
J. Phys. Chem. Lett. 5, 336–347 (2014)
4. Stroyuk, O., Raevskaya, A., Gaponik, N.: Solar light harvesting with multinary metal
chalcogenide nanocrystals. Chem. Soc. Rev. 47, 5354–5422 (2018)
5. Thomas, S.R., Chen, C.W., Date, M., Wang, Y.C., Tsai, H.W., Wang, Z.M., Chueh, Y.L.:
Recent developments in the synthesis of nanostructured chalcopyrite materials and their
applications: a review. RSC Adv. 6, 60643–60656 (2016)
6. Regulacio, M.D., Han, M.Y.: Multinary I-III-VI2 and I2 -II-IV-VI4 semiconductor nanostruc-
tures for photocatalytic applications: Acc. Chem. Res. 49, 511–519 (2016)
7. Azimi, H., Hou, Y., Brabec, C.J.: Towards low-cost. Environmentally Friendly Printed
Chalcopyrite and Kesterite Solar Cells. Energy Environ. Sci. 7, 1829–1849 (2014)
8. Knowles, K.E., Hartstein, K.H., Kilburn, T.B., Marchioro, A., Nelson, H.D., Whitham,
P.J., Gamelin, D.R.: Luminescent colloidal semiconductor nanocrystals containing copper:
synthesis, Photophysics, and applications. Chem. Rev. 116, 10820–10851 (2016)
9. Hamanaka, Y., Ozawa, K., Kuzuya, T.: Enhancement of donor−acceptor pair emissions in
colloidal AgInS2 quantum dots with high concentrations of defects. J. Phys. Chem. C. 118,
14562–14568 (2014)
10. Stroyuk, O., Raevskaya, A., Gaponik, N., Selyshchev, O., Dzhagan, V., Schulze, S., Zahn,
D.R.T.: Origin of the broadband photoluminescence of pristine and Cu+ /Ag+ –doped Ultra-
small CdS and CdSe/CdS quantum dots. J. Phys. Chem. C. 122, 10267–10277 (2018)
11. Sharma, D.K., Hirata, S., Bujak, L., Biju, V., Kameyama, T., Kishi, M., Torimoto, T., Vacha,
M.: Influence of Zn on the photoluminescence of colloidal (AgIn)x Zn2(1-x) S2 nanocrystals.
Phys. Chem. Chem. Phys. 19, 3963–3969 (2017)
12. Hamanaka, Y., Ogawa, T., Tsuzuki, M.: Photoluminescence properties and its origin of
AgInS2 quantum dots with chalcopyrite structure. J. Phys. Chem. C. 115, 1786–1792 (2011)
13. Raevskaya, A., Lesnyak, V., Haubold, D., Dzhagan, V., Stroyuk, O., Gaponik, N., Zahn,
D.R.T., Eychmüller, A.: A fine size selection of brightly luminescent water-soluble Ag-In-
S and Ag-In-S/ZnS quantum dots. J. Phys. Chem. C. 121, 9032–9042 (2017)
14. Stroyuk, O., Raevskaya, A., Spranger, F., Selyshchev, O., Dzhagan, V., Schulze, S., Zahn,
D.R.T., Eychmüller, A.: Origin and dynamics of highly efficient broadband photolumines-
cence of aqueous glutathione-capped size-selected Ag−In−S quantum dots. J. Phys. Chem.
C. 122, 13648–13658 (2018)
15. Sandroni, M., Wegner, K.D., Aldakov, D., Reiss, P.: Prospects of Chalcopyrite-type Nanocrys-
tals for energy Applications. ACS Energy Lett. 2, 1076–1088 (2017)
16. Zu, G., Zeng, S., Zhang, B., Swihart, M.T., Yong, K.T., Prasad, P.N.: New genera-
tion cadmium-free quantum dots for biophotonics and nanomedicine. Chem. Rev. 116,
12234–12327 (2016)
17. Rogach, A. (ed.): Semiconductor Nanocrystal Quantum Dots: Synthesis, Assembly. Spec-
troscopy and Applications. Springer, Wien (2008)
18. Eychmüller, A.: Structure and Photophysics of semiconductor nanocrystals. J. Phys. Chem.
B. 104, 6514–6528 (2000)
19. Lesnyak, V., Gaponik, N., Eychmüller, A.: Colloidal semiconductor nanocrystals: the aqueous
approach. Chem. Soc. Rev. 42, 2905–2929 (2013)
20. Hamanaka, Y., Ogawa, T., Tsuzuki, M., Ozawa, K., Kuzuya, T.: Luminescence properties of
chalcopyrite AgInS2 nanocrystals: their origin and related electronic states. J. Lumin. 133,
121–124 (2013)
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 115

21. Ding, Q., Zhang, X., Li, L., Lou, X., Xu, J., Zhou, P., Yan, M.: Temperature dependent
photoluminescence of composition tunable Znx AgInSe quantum dots and temperature sensor
application. Opt. Express. 25, 19065 (2017)
22. Knowles, K.E., Nelson, H.D., Kilburn, T.B., Gamelin, D.R.: Singlet−triplet Splittings in the
luminescent excited states of colloidal Cu+ :CdSe, Cu+ :InP, and CuInS2 nanocrystals: charge
transfer configurations and self-trapped excitons. J. Am. Chem. Soc. 137, 13138–13147
(2015)
23. Valeur, B.: Molecular fluorescence: principles and applications. Wiley-VCH Verlag GmbH
(2001)
24. Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y.A.: Ag-In-S and Cu-In-S nanoparticles in
aqueous media: preparation and spectral-luminescent properties. Theoret. Experim. Chem.
53, 315–325 (2017)
25. Raevskaya, A., Rosovik, O., Kozytskiy, A., Stroyuk, O., Dzhagan, V., Zahn, D.R.T.: Non-
stoichiometric Cu-In-S@ZnS nanoparticles produced in aqueous solutions as light harvesters
for liquid-junction Photoelectrochemical solar cells. RSC Adv. 6, 100145–100157 (2016)
26. Raevskaya, A., Rozovik, O., Novikova, A., Selyshchev, O., Stroyuk, O., Dzhagan, V.,
Goryacheva, I., Gaponik, N., Zahn, D.R.T., Eychmüller, A.: Luminescence and Photoelec-
trochemical properties of size-selected aqueous copper-doped Ag–In–S quantum dots. RSC
Adv. 8, 7550–7557 (2018)
27. Stroyuk, O., Weigert, F., Raevskaya, A., Spranger, F., Würth, C., Resch-Genger, U., Gaponik,
N., Zahn, D.R.T.: Inherently broadband photoluminescence in Ag-In-S/ZnS quantum dots
observed in ensemble and single particle studies. J. Phys. Chem. C. 123, 2632–2641 (2019)
28. Eychmüller, A., Hässelbarth, A., Katsikas, L., Weller, H.: Fluorescence mechanism of highly
monodisperse Q-sized CdS colloids. J. Lumin. 48-49, 745–749 (1991)
29. Eychmüller, A., Hässelbarth, A., Katsikas, L., Weller, H.: Photochemistry of semiconductor
colloids. 36. Fluorescence investigations on the nature of electron and hole traps in Q-sized
colloidal CdS particles. Ber. Bunsenges. Phys. Chem. 95, 79–84 (1991)
30. Coughlan, C., Ibañez, M., Dobrozhan, O., Singh, A., Cabot, A., Ryan, K.M.: Compound
Copper Chalcogenide Nanocrystals. Chem. Rev. 117, 5865–6109 (2017)
31. Kershaw, S.V., Susha, A.S., Rogach, A.L.: Narrow bandgap colloidal metal chalcogenide
quantum dots: synthetic methods, heterostructures, assemblies, electronic and infrared optical
properties. Chem. Soc. Rev. 42, 3033–3087 (2013)
32. Reiss, P., Carrière, M., Lincheneau, C., Vaure, L., Tamang, S.: Synthesis of semicon-
ductor nanocrystals, focusing on nontoxic and earth-abundant materials. Chem. Rev. 116,
10731–10819 (2016)
33. Jing, L., Kershaw, S.V., Li, Y., Huang, X., Li, Y., Rogach, A.L., Gao, M.: Aqueous Based
Semiconductor Nanocrystals. Chem. Rev. 116, 10623–10730 (2016)
34. Deng, D., Xia, J., Cao, J., Qu, L., Tain, J., Qian, Z., Gu, Y., Gu, Z.: Forming highly fluorescent
near-infrared emitting PbS quantum dots in water using glutathione as surface-modifying
molecule. J. Colloid Interface Sci. 367, 234–240 (2012)
35. Zhao, C., Bai, Z., Liu, X., Zhang, Y., Zou, B., Zhong, H.: Small GSH-capped CuInS2 quantum
dots: MPA-assisted aqueous phase transfer and bioimaging applications. ACS Appl. Mater.
Interfaces. 7, 17623–17629 (2015)
36. Luo, Z., Zhang, H., Huang, J., Zhong, X.: One-step synthesis of water-soluble AgInS2 and
ZnS–AgInS2 composite nanocrystals and their photocatalytic activities. J. Colloid Interface
Sci. 377, 27–33 (2012)
37. Singhal, M., Sharma, J.K., Kumar, S.: Effect of biocompatible glutathione capping on Core–
Shell ZnS quantum dots. J. Mater. Sci. Mater. Electron. 23, 1387–1392 (2012)
38. Zou, L., Fang, Z., Gu, Z., Zhong, X.: Aqueous phase synthesis of biostabilizer capped CdS
nanocrystals with bright emission. J. Lumin. 129, 536–540 (2009)
39. Huang, P., Jiang, Q., Yu, P., Yang, L., Mao, L.: Alkaline post-treatment of cd(II)−glutathione
coordination polymers: toward green synthesis of water-soluble and Cytocompatible CdS
quantum dots with tunable optical properties. ACS Appl. Mater. Interfaces. 5, 5239–5246
(2013)
116 O. Stroyuk et al.

40. Nakane, Y., Tsukasaki, Y., Sakata, T., Yasuda, H., Jin, T.: Aqueous synthesis of glutathione-
coated PbS quantum dots with tunable emission for non-invasive fluorescence imaging in the
second near-infrared biological window (1000–1400 nm). Chem. Commun. 49, 7584–7586
(2013)
41. Kolmykov, O., Coulon, J., Lalevée, J., Alem, H., Medjahdi, G., Schneider, R.: Aqueous
synthesis of highly luminescent glutathione-capped Mn2+ -doped ZnS quantum dots. Mater.
Sci. Eng. C. 44, 17–23 (2014)
42. Chen, Y., Huang, L., Li, S., Pan, D.: Aqueous synthesis of glutathione-capped Cu+ and Ag+ -
doped Znx Cd1–x S quantum dots with full color emission. J. Mater. Chem. C. 1, 751–756
(2013)
43. Yang, J., Zhang, W.H., Hu, Y.P., Yu, J.S.: Aqueous synthesis and characterization of
glutathione-stabilized β-HgS nanocrystals with near-infrared photoluminescence. J. Colloid
Interface Sci. 379, 8–13 (2012)
44. Wang, L.L., Jiang, J.S.: Optical performance evolutions of reductive glutathione coated CdSe
quantum dots in different environments. J. Nanopart. Res. 13, 1301–1309 (2011)
45. Ding, Y., Shen, S.Z., Sun, H., Sun, K., Liu, F.: Synthesis of L-glutathione-capped ZnSe
quantum dots for the sensitive and selective determination of copper ion in aqueous solutions.
Sensors Actuators B. 203, 35–43 (2014)
46. Zhang, J., Li, J., Zhang, J., Xie, R., Yang, W.: Aqueous synthesis of ZnSe nanocrystals by
using glutathione as ligand: the pH-mediated coordination of Zn2+ with glutathione. J. Phys.
Chem. C. 114, 11087–11091 (2010)
47. Fang, Z., Li, Y., Zhang, H., Zhong, X., Zhu, L.: Facile synthesis of highly luminescent UV-
blue-emitting ZnSe/ZnS Core/Shell nanocrystals in aqueous media. J. Phys. Chem. C. 113,
14145–14150 (2009)
48. Zheng, Y., Yang, Z., Ying, J.Y.: Aqueous synthesis of glutathione-capped ZnSe and
Zn1–x Cdx Se alloyed quantum dots. Adv. Mater. 19, 1475–1479 (2007)
49. Lesnyak, V., Dubavik, A., Plotnikov, A., Gaponik, N., Eychmüller, A.: One-step aqueous
synthesis of blue-emitting glutathione-capped ZnSe1–x Tex alloyed nanocrystals. Chem. Com-
mun. 46, 886–888 (2010)
50. Du, J., Li, X., Wang, S., Wu, Y., Hao, X., Xu, C., Zhao, X.: Microwave-assisted synthesis
of highly luminescent glutathione-capped Zn1–x Cdx Te alloyed quantum dots with excellent
biocompatibility. J. Mater. Chem. 22, 11390–11395 (2012)
51. Gu, Y.P., Cui, R., Zhang, Z.L., Xie, Z.X., Pang, D.W.: Ultrasmall near-infrared Ag2 Se
quantum dots with tunable fluorescence for in vivo imaging. J. Am. Chem.Soc. 134, 79–82
(2012)
52. Tian, J., Liu, R., Zhao, Y., Xu, Q., Zhao, S.: Controllable synthesis and cell-imaging studies on
CdTe quantum dots together capped by glutathione and Thioglycolic acid. J. Colloid Interface
Sci. 336, 504–509 (2009)
53. Liu, Y.F., Yu, J.S.: In situ synthesis of highly luminescent glutathione-capped CdTe/ZnS
quantum dots with biocompatibility. J. Colloid Interface Sci. 351, 1–9 (2010)
54. Li, Y., Zhou, Y., Wang, H.Y., Perrett, S., Zhao, Y., Tang, Z., Nie, G.: Chirality of glutathione
surface coating affects the cytotoxicity of quantum dots. Angew. Chem. Int. Ed. 50, 5860–
5864 (2011)
55. Qian, H., Dong, C., Weng, J., Ren, J.: Facile one-pot synthesis of luminescent, water-soluble,
and biocompatible glutathione-coated CdTe nanocrystals. Small. 2, 747–751 (2006)
56. Zheng, Y., Gao, S., Ying, J.Y.: Synthesis and cell-imaging applications of glutathione-capped
CdTe quantum dots. Adv. Mater. 19, 376–380 (2007)
57. Samanta, A., Deng, Z., Liu, Y.: Aqueous synthesis of glutathione-capped CdTe/CdS/ZnS and
CdTe/CdSe/ZnS Core/Shell/Shell nanocrystal Heterostructures. Langmuir. 28, 8205–8215
(2012)
58. Liu, H., Gu, C., Xiong, W., Zhang, M.: A sensitive hydrogen peroxide biosensor using ultra-
small CuInS2 nanocrystals as peroxidase mimics. Sensors Actuators B. 209, 670–676 (2015)
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 117

59. Gedda, G., Chen, G.R., Yao, Y.Y., Girma, W.M., Li, J.D., Yen, C.L., Ling, Y.C., Chang, J.Y.:
Aqueous synthesis of dual-targeting Gd-doped CuInS2 /ZnS quantum dots for cancer-specific
bi-modal imaging. New J. Chem. 41, 14161–14170 (2017)
60. Chang, J.Y., Chen, G.R., Li, J.D.: Synthesis of Magnetofluorescent Gd-doped CuInS2 /ZnS
quantum dots with enhanced longitudinal Relaxivity. Phys. Chem. Chem. Phys. 18, 7132–
7140 (2016)
61. Xiong, W.W., Yang, G.H., Wu, X.C., Zhu, J.J.: Aqueous synthesis of color-tunable
CuInS2 /ZnS nanocrystals for the detection of human interleukin 6. ACS Appl. Mater.
Interfaces. 5, 8210–8216 (2013)
62. Xiong, W.W., Yang, G.H., Wu, X.C., Zhu, J.J.: Microwave-assisted synthesis of highly
luminescent AgInS2 /ZnS nanocrystals for dynamic intracellular cu(II) detection. J. Mater.
Chem. B. 1, 4160–4165 (2013)
63. Deng, D., Cao, J., Qu, L., Achilefu, S., Gu, Y.: Highly luminescent water-soluble quaternary
Zn–Ag–In–S quantum dots for tumor cell-targeted imaging. Phys. Chem. Chem. Phys. 15,
5078–5083 (2013)
64. Song, J., Ma, C., Zhang, W., Li, X., Zhang, W., Wu, R., Cheng, X., Ali, A., Yang, M., Zhu, L.,
Xia, R., Xu, X.: Bandgap and structure engineering via cation exchange: from binary Ag2 S
to ternary AgInS2 . Quaternary AgZnInS alloy and AgZnInS/ZnS Core/Shell Fluorescent
Nanocrystals for Bioimaging. ACS Appl. Mater. Interfaces. 8, 24826–24836 (2016)
65. Leach, A.D.P., Macdonald, J.E.: Optoelectronic properties of CuInS2 nanocrystals and their
origin. J. Phys. Chem. Lett. 7, 572–583 (2016)
66. Chemseddine, A., Weller, H.: Highly monodisperse quantum sized CdS particles by size
selective precipitation. Ber. Bunsenges. Phys. Chem. 97, 636–638 (1993)
67. Murray, C.B., Norris, D.J., Bawendi, M.G.: Synthesis and characterization of nearly monodis-
perse CdE (E = S, se, Te) semiconductor Nanocrystallites. J. Am. Chem. Soc. 115,
8706–8715 (1993)
68. Wang, C.L., Fang, M., Xu, S.H., Cui, Y.P.: Salts-based size-selective precipitation: toward
mass precipitation of aqueous nanoparticles. Langmuir. 26, 633–638 (2010)
69. Anand, M., Odom, L.A., Roberts, C.B.: Finely controlled size-selective precipitation and sep-
aration of CdSe/ZnS semiconductor nanocrystals using CO2 -gas-expanded liquids. Langmuir.
23, 7338–7343 (2007)
70. Rogach, A.L., Kornowski, A., Gao, M., Eychmüller, A., Weller, H.: Synthesis and Character-
ization of a size series of extremely small thiol-stabilized CdSe nanocrystals. J. Phys. Chem.
B. 103, 3065–3069 (1999)
71. Rogach, A.L., Talapin, D.V., Shevchenko, E.V., Kornowski, A., Haase, M., Weller, H.:
Organization of Matter on different size scales: monodisperse nanocrystals and their super-
structures. Adv. Funct. Mater. 12, 653–664 (2002)
72. Kudera, S., Zanella, M., Giannini, C., Rizzo, A., Li, Y., Gigli, G., Cingolani, R., Ciccarella,
G., Spahl, W., Parak, W.J., Manna, L.: Sequential growth of magic-size CdSe nanocrystals.
Adv. Mater. 19, 548–552 (2007)
73. Talapin, D.V., Shevchenko, E.V., Kornowski, A., Gaponik, N., Haase, M., Rogach, A.L.,
Weller, H.: A new approach to crystallization of CdSe nanoparticles into ordered three-
dimensional Superlattices. Adv. Mater. 13, 1868–1871 (2001)
74. Holmes, M.A., Townsend, T.K., Osterloh, F.E.: Quantum confinement controlled photocat-
alytic water splitting by suspended CdSe nanocrystals. Chem. Commun. 48, 371–373 (2012)
75. Bao, H., Wang, E., Dong, S.: One-pot synthesis of CdTe nanocrystals and shape control of
luminescent CdTe–Cystine nanocomposites. Small. 2, 476–480 (2006)
76. Gaponik, N., Talapin, D.V., Rogach, A.L., Hoppe, K., Shevchenko, E.V., Kornowski,
A., Eychmüller, A., Weller, H.: Thiol-capping of CdTe nanocrystals: an alternative to
organometallic synthetic routes. J. Phys. Chem. B. 106, 7177–7185 (2002)
77. Talapin, D.V., Rogach, A.L., Shevchenko, E.V., Kornowski, A., Haase, M., Weller, H.:
Dynamic distribution of growth rates within the ensembles of colloidal II-VI and III-V
semiconductor nanocrystals as a factor governing their photoluminescence efficiency. J. Am.
Chem. Soc. 124, 5782–5790 (2002)
118 O. Stroyuk et al.

78. Karabudak, E., Brookes, E., Lesnyak, V., Gaponik, N., Eychmüller, A., Walter, J., Segets, D.,
Peukert, W., Wohlleben, W., Demeler, B., Cölfen, H.: Simultaneous identification of spectral
properties and sizes of multiple particles in solution with subnanometer resolution. Angew.
Chem. Int. Ed. 55, 11770–11774 (2016)
79. Mastronardi, M.L., Maier-Flaig, F., Faulkner, D., Henderson, E.J., Kübel, C., Lemmer, U.,
Ozin, G.A.: Size-dependent absolute quantum yields for size-separated Colloidally-stable
silicon nanocrystals. Nano Lett. 12, 337–342 (2012)
80. Guzelian, A.A., Katari, J.E.B., Kadavanich, A.V., Banin, U., Hamad, K., Juban, E., Alivisatos,
A.P., Wolters, R.H., Arnold, C.C., Heath, J.R.: Synthesis of size-selected, surface-passivated
InP nanocrystals. J. Phys. Chem. 100, 7212–7219 (1996)
81. Nam, D.E., Song, W.S., Yang, H.: Facile, air-insensitive Solvothermal synthesis of emission-
tunable CuInS2 /ZnS quantum dots with high quantum yields. J. Mater. Chem. 21,
18220–18226 (2011)
82. Xie, R., Rutherford, M., Peng, X.: Formation of high-quality I-III-VI semiconductor
nanocrystals by tuning relative reactivity of cationic precursors. J. Am. Chem. Soc. 131,
5691–5697 (2009)
83. Akdas, T., Walter, J., Segets, D., Distaso, M., Winter, B., Birajar, B., Spieckler, E., Peukert,
W.: Investigation of the size–property relationship in CuInS2 quantum dots. Nanoscale. 7,
18105–18118 (2015)
84. Chang, S.H., Chiu, B.C., Gao, T.L., Jeng, S.L., Tuan, H.Y.: Selective synthesis of copper
gallium sulfide (CuGaS2 ) nanostructures of different sizes, crystal phases, and morphologies.
CrystEngComm. 16, 3323–3330 (2014)
85. Uematsu, T., Doi, T., Torimoto, T., Kuwabata, S.: Preparation of luminescent AgInS2 -AgGaS2
solid solution nanoparticles and their optical properties. J. Phys. Chem. Lett. 1, 3283–3287
(2010)
86. Collord, A.D., Hillhouse, H.W.: Composition control and formation pathway of CZTS and
CZTGS nanocrystal inks for Kesterite solar cells. Chem. Mater. 27, 1855–1862 (2015)
87. Singh, A., Geaney, H., Laffir, F., Ryan, K.M.: Colloidal synthesis of Wurtzite Cu2 ZnSnS4
Nanorods and their perpendicular assembly. J. Am. Chem. Soc. 134, 2910–2913 (2012)
88. Moulder, J.F., Stickle, W.F., Sobol, P.E., Bomben, K.D.: Handbook of X ray photoelectron
spectroscopy: a reference book of standard spectra for identification and interpretation of Xps
data. Phys. Electron. (1995)
89. Naumkin, A.V., Kraut-Vass, A., Gaarenstroom, S.W.: NIST X-ray photoelectron spectroscopy
database. NIST, Stand. Ref. Database 20, Version 4.1 (2012)
90. Zang, H., Li, H., Makarov, N.S., Velizhanin, K.A., Wu, K., Park, Y.S., Klimov, V.I.: Thick-
Shell CuInS2 /ZnS quantum dots with suppressed “blinking” and narrow single-particle
emission line widths. Nano Lett. 17, 1787–1795 (2017)
91. Chestnoy, N., Harris, T.D., Hull, R., Brus, L.: Luminescence and photophysics of CdS
semiconductor clusters: the nature of the emitting electronic state. J. Phys. Chem. 90,
3393–3399 (1986)
92. Liu, S., Su, X.: The synthesis and application of I–III–VI type quantum dots. RSC Adv. 4,
43415–43428 (2014)
93. Chevallier, T., Benayad, A., Le Blevennec, G., Chandezon, F.: Method to determine radiative
and non-radiative defects applied to AgInS2 –ZnS luminescent nanocrystals. Phys. Chem.
Chem. Phys. 19, 2359–2363 (2017)
94. Chevallier, T., Le Blevennec, G., Chandezon, F.: Photoluminescence properties of AgInS2 –
ZnS nanocrystals: the critical role of the surface. Nanoscale. 8, 7612–7620 (2016)
95. Pietryga, J.M., Park, Y.S., Lim, J., Fidler, A.F., Bae, W.K., Brovelli, S., Klimov, V.I.: Spec-
troscopic and device aspects of nanocrystal quantum dots. Chem. Rev. 116, 10513–10622
(2016)
96. Whitham, P.J., Marchioro, A., Knowles, K.E., Kilburn, T.B., Reid, P.J., Gamelin, D.R.:
Single-particle photoluminescence spectra, blinking, and delayed luminescence of colloidal
CuInS2 nanocrystals. J. Phys. Chem. C. 120, 17136–17142 (2016)
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 119

97. Cichy, B., Rich, R., Olejniczak, A., Gryczynski, Z., Strek, W.: Two blinking mechanisms
in highly confined AgInS2 and AgInS2 /ZnS quantum dots evaluated by single particle
spectroscopy. Nanoscale. 8, 4151–4159 (2016)
98. Kameyama, T., Takahashi, T., Machida, T., Kamiya, Y., Yamamoto, T., Kuwabata, S.,
Torimoto, T.: Controlling the electronic energy structure of ZnS−AgInS2 solid solution
nanocrystals for photoluminescence and photocatalytic hydrogen evolution. J. Phys. Chem.
C. 119, 24740–24749 (2015)
99. Gaponenko, S.V.: Optical properties of semiconductor nanocrystals. Cambridge University
Press, Cambridge (1998)
100. Rayevskaya, A.E.: Grodzyuk, G.Ya., Dzhagan, V.M., Stroyuk, O.L., Kuchmiy, S.Ya., Plyus-
nin, V.F., Grivin, V.P., Valakh, M.Ya.: synthesis and characterization of white-emitting CdS
quantum dots stabilized with Polyethyleneimine. J. Phys. Chem. C. 114, 22478–22486 (2010)
101. Raevskaya, A.E., Stroyuk, O.L., Solonenko, D.I., Dzhagan, V.M., Lehmann, D., Kuchmiy,
S.Y.A., Plyusnin, V.F., Zahn, D.R.T.: Synthesis and Luminescent Properties of Ultrasmall
Colloidal CdS Nanoparticles Stabilized by Cd(II) Complexes With Ammonia and Mercap-
toacetate. J. Nanopart. Res. 16, 2650-1-10 (2014)
102. Pesika, N.S., Stebe, K.J., Searson, P.C.: Relationship between absorbance spectra and particle
size distributions for quantum-sized nanocrystals. J. Phys. Chem. B. 107, 10412–10415
(2003)
103. Sharma, D.K., Hirata, S., Bujak, L., Biju, V., Kameyama, T., Kishi, M., Torimoto, T., Vacha,
M.: Single-particle spectroscopy of I–III–VI semiconductor nanocrystals: spectral diffusion
and suppression of blinking by two-color excitation. Nanoscale. 8, 13687–13694 (2018)
104. Fuhr, A.S., Yun, H.J., Makarov, N.S., Li, H., McDaniel, H., Klimov, V.I.: Light emission
mechanisms in CuInS2 quantum dots evaluated by spectral electrochemistry. ACS Photonics.
4, 2425–2435 (2017)
105. Michalet, X., Pinaud, F.F., Bentolila, L.A., Tsay, J.M., Doose, S., Li, J.J., Sundaresan, G., Wu,
A.M., Gambhir, S.S., Weiss, S.: Quantum dots for live cells, in vivo imaging, and diagnostics.
Science. 307, 538–544 (2005)
106. Quintanilla, M., Liz-Marzan, L.M.: Guiding rules for selecting a Nanothermometer. Nano
Today. 19, 126–145 (2018)
107. Resch-Genger, U., Grabolle, M., Cavaliere-Jaricot, S., Nitschke, R., Nann, T.: Quantum dots
versus organic dyes as fluorescent labels. Nat. Methods. 5, 763–775 (2008)
108. Ding, Q., Zhang, X., Li, L., Lou, X., Xu, J., Zhou, P., Yan, M.: Temperature dependent
photoluminescence of composition tunable Znx AgInSe quantum dots and temperature sensor
application. Opt. Express. 25, 298033 (2017)
109. Matsuda, Y., Torimoto, T., Kameya, T., Kameyama, T., Kuwabata, S., Yamaguchi, H., Niimi,
T.: ZnS–AgInS2 nanoparticles as a temperature sensor. Sens. Actuators B. 176, 505–508
(2013)
110. Liu, W., Zhang, Y., Zhai, W., Wang, Y., Zhang, T., Gu, P., Chu, H., Zhang, H., Cui, T., Wang,
Y., Zhao, J., Yu, W.W.: Temperature-dependent photoluminescence of ZnCuInS/ZnSe/ZnS
quantum dots. J. Phys. Chem. C. 117, 19288–19294 (2013)
111. Raevskaya, A.E., Ivanchenko, M.V., Skoryk, M.A., Stroyuk, O.: Brightly luminescent col-
loidal Ag-In-S nanoparticles stabilized in aqueous solutions by branched polyethyleneimine.
J. Lumin. 178, 295–300 (2016)
112. Raevskaya, A.E., Ivanchenko, M.V., Stroyuk, O.L., Kuchmiy, S.Y.A., Plyusnin, V.F.: Lumi-
nescent Ag-doped In2 S3 Nanoparticles ntabilized by mercaptoacetate in water and glycerol.
J. Nanoparticle Res. 17, 135 (2015)
113. Raevskaya, A.E., Grodzyuk, G.Y.A., Stroyuk, O.L., Kuchmiy, S.Y.A., Plyusnin, V.F.: Effect
of temperature on the optical properties of polyethylenimine-stabilized CdS nanoparticles.
Theoret. Experim. Chem. 48, 95–101 (2012)
114. Raevskaya, A., Stroyuk, O., Panasiuk, Y.A., Dzhagan, V., Solonenko, D., Schulze, S., Zahn,
D.R.T.: A new route to very stable water-soluble ultra-small core/shell CdSe/CdS quantum
dots. Nano Struct. Nano Objects. 13, 146–154 (2018)
120 O. Stroyuk et al.

115. Valerini, D., Cretí, A., Lomascolo, M., Manna, L., Cingolani, R., Anni, M.: Temperature
dependence of the photoluminescence properties of colloidal CdSe/ZnS core/shell quantum
dots embedded in a polystyrene matrix. Phys. Rev. B. 71, 235409 (2005)
116. Labeau, O., Tamarat, P., Lounis, B.: Temperature dependence of the luminescence lifetime of
single CdSe/ZnS quantum dots. Phys. Rev. Lett. 90, 257404 (2003)
117. de Mello Donegá, C., Bode, M., Meijerink, A.: Size- and temperature-dependence of exciton
lifetimes in CdSe quantum dots. Phys. Rev. B. 74, 085320 (2006)
118. Nirmal, M., Murray, C.B., Bawendi, M.G.: Fluorescence-line narrowing in CdSe quantum
dots: surface localization of the photogenerated exciton. Phys. Rev. B. 50, 2293–2300 (1994)
119. Crooker, S.A., Barrick, T., Hollingsworth, J.A., Klimov, V.I.: Multiple temperature regimes of
radiative decay in CdSe nanocrystal quantum dots: intrinsic limits to the dark-exciton lifetime.
Appl. Phys. Lett. 82, 2793–2795 (2003)
120. Liptay, T.J., Ram, R.J.: Temperature dependence of the exciton transition in semiconductor
quantum dots. Appl. Phys. Lett. 89, 223132 (2006)
121. Liu, T.C., Huang, Z.L., Wang, H.Q., Wang, J.H., Li, X.Q., Zhao, Y.D., Luo, Q.M.:
Temperature-dependent photoluminescence of water-soluble quantum dots for a bioprobe.
Anal. Chim. Acta. 559, 120–123 (2006)
122. Kr˛ezel, A., Wójcik, J.W., Maciejczyk, M., Bal, W.: Zn(II) complexes of glutathione disulfide:
structural basis of elevated stabilities. Inorg. Chem. 50, 72–85 (2011)
123. Jethi, L., Mack, T.G., Kambhampati, P.: Extending semiconductor nanocrystals from the
quantum dot regime to the molecular cluster regime. J. Phys. Chem. C. 121, 26102–26107
(2017)
124. Voznyy, O., Mokkath, J.H., Jain, A., Sargent, E.H., Schwingenschlögl, U.: J. Phys. Chem. C.
120, 10015–10019 (2016)
125. Santhi, S., Bernstein, E., Paille, F.: Temperature dependence of trap luminescence of CdS
doped glasses. J. Lumin. 117, 101–112 (2006)
126. Vossmeyer, T., Katsikas, L., Giersig, M., Popovic, I.G., Diesner, K., Chemseddine, A.,
Eychmüller, A., Weller, H.: CdS nanoclusters: synthesis, characterization, size dependent
oscillator strength, temperature shift of the Excitonic transition energy, and reversible
absorbance shift. J. Phys. Chem. 98, 7665–7673 (1994)
127. Dai, Q., Song, Y., Li, D., Chen, H., Kan, S., Zou, B., Wang, Y., Deng, Y., Hou, Y., Yu, S.,
Chen, L., Zou, G.: Temperature dependence of band gap in CdSe nanocrystals. Chem. Phys.
Lett. 439, 65–68 (2007)
128. Kamisaka, H., Kilina, S.V., Yamashita, K., Prezhdo, O.V.: Ab initio study of temperature and
pressure dependence of energy and phonon-induced dephasing of electronic excitations in
CdSe and PbSe quantum dots. J. Phys. Chem. C. 112, 7800–7808 (2008)
129. Jorge, P.A.S., Mayeh, M., Benrashid, R., Caldas, P., Santos, J.L., Farahi, F.: Quantum dots as
self-referenced optical fibre temperature probes for luminescent chemical sensors. Meas. Sci.
Technol. 17, 1032–1038 (2006)
130. Wuister, S.F., van Houselt, A., de Mello Donega, C., Vanmaekelbergh, D., Meijerink, A.:
Temperature antiquenching of the luminescence from capped CdSe quantum dots. Angew.
Chem. Int. Ed. 43, 3029–3033 (2004)
131. Bodunov, E.N., Danilov, V.V., Panfutova, A.S., Simoes Gamboa, A.L.: Room-temperature
luminescence decay of colloidal semiconductor quantum dots: nonexponentiality revisited.
Ann. Phys. (Berlin). 528(272–277), 272 (2016)
132. Li, S., Zhang, K., Yang, J.M., Lin, L., Yang, H.: Single quantum dots as local temperature
markers. Nano Lett. 7, 3102–3105 (2007)
133. Kambhampati, P.: On the kinetics and thermodynamics of excitons at the surface of
semiconductor nanocrystals: are there surface excitons? Chem. Phys. 446, 92–107 (2015)
134. Biju, V., Makita, Y., Sonoda, A., Yokoyama, H., Baba, Y., Ishikawa, M.: Temperature-
sensitive photoluminescence of CdSe quantum dot clusters. J. Phys. Chem. B. 109, 13899–
13905 (2005)
135. Mooney, J., Krause, M.M., Saari, J.I., Kambhampati, P.: A microscopic picture of surface
charge trapping in semiconductor nanocrystals. J. Chem. Phys. 138, 204705 (2013)
Unique Luminescent Properties of Composition-/Size-Selected Aqueous. . . 121

136. Woodall, D.L., Tobias, A.K., Jones, M.: Resolving carrier recombination in CdS quantum
dots: a time-resolved fluorescence study. Chem. Phys. 471, 2–10 (2016)
137. Jethi, L., Mack, T.G., Krause, M.M., Drake, S., Kambhampati, P.: The effect of exciton-
delocalizing thiols on intrinsic dual emitting semiconductor nanocrystals. ChemPhysChem.
17, 665–669 (2016)
138. Walker, G.W., Sundar, V.C., Rudzinski, C.M., Wun, A.W., Bawendi, M.G., Nocera, D.G.:
Quantum-dot optical temperature probes. Appl. Phys. Lett. 83, 3555–3557 (2003)
139. Olkhovets, A., Hsu, R.C., Lipovskii, A., Wise, F.W.: Size-dependent temperature variation of
the energy gap in Lead-salt quantum dots. Phys. Rev. Lett. 81, 3539–3542 (1998)
140. Turyanska, L., Patane, A., Henini, M., Hennequin, B., Thomas, N.R.: Temperature depen-
dence of the photoluminescence emission from thiol-capped PbS quantum dots. Appl. Phys.
Lett. 90, 101913 (2007)
141. Mooney, J., Krause, M.M., Saari, J.I., Kambhampati, P.: Challenge to the deep-trap model of
the surface in semiconductor nanocrystals. Phys. Rev. B. 87, 081201 (2013)
142. Berends, A.C., Mangnus, M.J.J., Xia, C., Rabouw, F.T., de Mello Donega, C.: Optoelectronic
properties of ternary I-III-VI2 semiconductor nanocrystals: bright prospects with elusive
origins. J. Phys. Chem. Lett. 10, 1600–1616 (2019)
143. Bai, X., Purcell-Milton, F., Gun’ko, Y.K.: Optical properties, synthesis, and potential appli-
cations of cu-based ternary or quaternary anisotropic quantum dots, polytypic nanocrystals,
and Core/Shell heterostructures. Nano. 9, 85 (2019)
144. Stroyuk, O., Raevskaya, A., Spranger, F., Gaponik, N., Zahn, D.R.T.: Temperature-dependent
photoluminescence of silver-indium-sulfide nanocrystals in aqueous colloidal solutions.
ChemPhysChem. 20, 1640–1648 (2019)
145. V.I. Klimov (Ed.): Nanocrystal quantum dots. 2nd edn. CRC Press (2010)
146. Konstantatos, G., Sargent, E.H. (eds.): Colloidal quantum dot optoelectronics and photo-
voltaics. Cambridge, Cambridge University Press (2013)
147. Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y.A.: Preparation of colloidal CdSe and
CdS/CdSe nanoparticles from sodium selenosulfate in aqueous polymers solution. J. Colloid
Interface Sci. 302, 133–141 (2006)
148. Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y.A., Azhnyuk, Y.M., Dzhagan, V.M., Yukhim-
chuk, V.O., Valakh, M.Y.A.: Growth and spectroscopic characterization of CdSe nanoparticles
synthesized from CdCl2 and Na2 SeSO3 in aqueous gelatine soslutions. Colloids Surf. A. 290,
304–309 (2006)
149. Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y., Dzhagan, V.M., Valakh, M.Y., Zahn, D.R.T.:
Optical study of CdS- and ZnS-passivated CdSe nanocrystals in gelatin films. J. Phys.
Condens. Matter. 19, 386237 (2007)
150. Dzhagan, V.M., Valakh, M.Y., Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y., Zahn, D.R.T.:
Characterization of semiconductor core-shell nanoparticles by resonant Raman scattering and
photoluminescence spectroscopy. Appl. Surf. Sci. 255, 725–727 (2008)
151. Dzhagan, V.M., Valakh, M.Y., Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y., Zahn, D.R.T.:
The influence of shell parameters on phonons in core-shell nanoparticles: a resonant Raman
study. Nanotechnology. 20, 365704 (2009)
152. Walsh, B.R., Saari, J.I., Krause, M.M., Nick, R., Coe-Sullivan, S., Kambhampati, P.: Surface
and interface effects on non-radiative exciton recombination and relaxation dynamics in
CdSe/CdZnS nanocrystals. Chem. Phys. 471, 11–17 (2016)
153. Seo, J., Raut, S., Abdel-Fattah, M., Rice, Q., Tabibi, B., Rich, R., Fudala, R., Gryczynski, I.,
Kim, W.J., Jung, S., Hyun, R.: Time-resolved and temperature-dependent photoluminescence
of ternary and quaternary nanocrystals of CuInS2 with ZnS capping and cation exchange. J.
Appl. Phys. 114, 094310 (2013)
154. Shi, A., Wang, X., Meng, X., Liu, X., Li, H., Zhao, J.: Temperature-dependent photolumines-
cence of CuInS2 quantum dots. J. Lumin. 132, 1819–1823 (2012)
155. Kim, Y.K., Ahn, S.H., Chung, K., Cho, Y.S., Choi, C.J.: The photoluminescence of CuInS2
nanocrystals: effect of non-stoichiometry and surface modification. J. Mater. Chem. 22, 1516–
1520 (2012)
122 O. Stroyuk et al.

156. Kobitski, A.Y., Zhuravlev, K.S., Wagner, H.P., Zahn, D.R.T.: Self-trapped exciton recombi-
nation in silicon nanocrystals. Phys. Rev. B. 63, 115423 (2001)
157. Mack, T.G., Jethi, L., Kambhampati, P.: Temperature dependence of emission line widths
from semiconductor nanocrystals reveals Vibronic contributions to line broadening processes.
J. Phys. Chem. C. 121, 28537–28545 (2017)
158. Kamenev, B.V., Nassiopoulou, A.G.: Self-trapped excitons in silicon nanocrystals with sizes
below 1.5 Nm in Si/SiO2 multilayers. J. Appl. Phys. 90, 5735–5740 (2001)
159. Dzhagan, V.M., Valakh, M.Y., Raevskaya, A.E., Stroyuk, O.L., Kuchmiy, S.Y., Zahn, D.R.T.:
Temperature-dependent resonant Raman scattering study of core/shell nanocrystals. J. Phys.:
Conf. Series. 92, 012045 (2007)
160. Tang, H., Berger, H., Schmid, P.E., Levy, F., Burri, G.: Photoluminescence in TiO2 anatase
single crystals. Solid State Commun. 87, 847–850 (1993)
161. Purdy, A.E., Murray, R.B., Song, K.S., Stoneham, A.M.: Studies of self-trapped exciton
luminescence in KCl. Phys. Rev. B. 15, 2170–2176 (1977)
162. Williams, R.T., Song, K.S.: The self-trapped exciton. J. Phys. Chem. Solids. 51, 679–716
(1990)
163. Cramer, L.P., Cumby, T.D., Leraas, J.A., Langford, S.C., Dickinson, J.T.: Effect of surface
treatments on self-trapped exciton luminescence in single-crystal CaF2 . J. Appl. Phys. 97,
103533 (2005)
164. Smith, M.D., Watson, B.L., Dauskardt, R.H., Karunadasa, H.I.: Broadband emission with a
massive stokes shift from sulfonium Pb-Br hybrids. Chem. Mater. 29, 7083–7087 (2017)
165. Yangui, A., Garrot, D., Lauret, J.S., Lusson, A., Bouchez, G., Deleporte, E., Pillet, S., Bendeif,
E.E., Castro, M., Triki, S., Abid, Y., Boukheddaden, K.: Optical investigation of broadband
white-light emission in self-assembled organic-inorganic perovskite (C6 H11 NH3 )2 PbBr4 . J.
Phys. Chem. C. 119, 23638–23647 (2015)
166. McCall, K.M., Stoumpos, C.C., Kostina, S.S., Kanatzidis, M.G., Wessels, B.W.: Strong
Electron-phonon coupling and self-trapped excitons in the defect halide perovskites A3 M2 I9
(A = Cs, Rb; M = Bi, Sb). Chem. Mater. 29, 4129–4145 (2017)
167. Hu, T., Smith, M.D., Dohner, E.R., Sher, M.J., Wu, X., Trinh, M.T., Fisher, A., Corbett,
J., Zhu, X.Y., Karunadasa, H.I., Lindenberg, A.M.: Mechanism for broadband white-light
emission from two-dimensional (110) hybrid perovskites. J. Phys. Chem. Lett. 7, 2258–2263
(2016)
168. Itoh, C., Tanimura, K., Itoh, N., Itoh, M.: Threshold energy for photogeneration of self-
trapped excitons in SiO2 . Phys. Rev. B. 39, 11183–11186 (1989)
169. De La Torre, J., Bremond, G., Souifi, A., Guillot, G., Buffet, N., Mur, P.: Simultaneous
observation of “self trapped exciton” and Q-confined exciton luminescence emission in
silicon nanocrystals. Opt. Mater. (Amst). 27, 1004–1007 (2005)
170. Yu, P., Wen, X., Toh, Y.R., Tang, J.: Temperature-dependent fluorescence in carbon dots. J.
Phys. Chem. C. 116, 25552–25557 (2012)
171. Fu, M., Ehrat, F., Wang, Y., Milowska, K.Z., Reckmeier, C., Rogach, A.L., Stolarczyk, J.K.,
Urban, A.S., Feldmann, J.: Carbon dots: a unique fluorescent cocktail of polycyclic aromatic
hydrocarbons. Nano Lett. 15, 6030–6035 (2015)
172. Benin, B.M., Dirin, D.N., Morad, V., Wörle, M., Yakunin, S., Rainò, G., Nazarenko, O.,
Fischer, M., Infante, I., Kovalenko, M.V.: Highly emissive self-trapped excitons in fully
inorganic zero-dimensional tin halides. Angew. Chemie Int. Ed. 57, 11329–11333 (2018)
173. Chang, S.K., Lee, C.D., Park, H.L., Chung, C.H.: Exciton transfer processes in ZnSe1-x Tex .
J. Cryst. Growth. 117, 793–796 (1992)
Electronic and Optical Characteristics
of Core/Shell Quantum Dots

D. A. Baghdasaryan, H. T. Ghaltaghchyan, D. B. Hayrapetyan,


E. M. Kazaryan, and H. A. Sarkisyan

Abstract This chapter presents the results of a theoretical study of single-electron,


two-electron, and impurity states as well as interband optical absorption and single-
electron current in spherical and cylindrical core/shell quantum dots. Analytically
exact solvable models of the confining potentials of core/shell quantum dots are
discussed.
The dependences of various physical characteristics of these structures on the
geometric parameters of quantum dots are studied. The controllability of the
energy spectrum of charge carriers in core/shell quantum dots is demonstrated, and
ability to control the optical and current characteristics of spherical and cylindrical
core/shell quantum dots is shown.

Keywords Core/shell quantum dot · Impurity states · Absorption coefficient ·


One-electron current

1 Introduction

Quantum dots (QDs) are unique objects where it is possible to investigate many
fundamental quantum mechanical principles of atoms and molecules experimentally
[1–3]. Being quantum systems of gigantic sizes, these structures make it relatively
easy to identify the features of the discrete spectrum of particles localized in
them, as well as to flexibly manipulate the energy levels. The manipulation of

D. A. Baghdasaryan · H. T. Ghaltaghchyan
Russian-Armenian University, Yerevan, Armenia
e-mail: david199108@mail.ru
D. B. Hayrapetyan · E. M. Kazaryan · H. A. Sarkisyan ()
Russian-Armenian University, Yerevan, Armenia
Peter the Great Saint-Petersburg Polytechnic University, St. Petersburg, Russia
e-mail: eduard.ghazaryan@rau.am; shayk@ysu.am

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 123
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_4
124 D. A. Baghdasaryan et al.

physical properties of QDs are attractive because of its potential application in


the development of semiconductor optoelectronic devices and quantum computers.
The electronic structure and optical properties of semiconductor QD have been
intensively studied in recent years thanks to the rapid development of fabrication
technology. To date due to availability of various methods of the QD growth such
as molecular beam epitaxy [4–6], MOCVD [7–9], liquid-phase epitaxy [10–
12], the Stranski-Krastanov method [13–15], etc., spherical [16–18], cylindrical
[19, 20], ellipsoidal [21, 22], lens-shaped [23, 24], pyramidal [25, 26], and
conical QDs [27] have been fabricated. The uniqueness of these systems lies
in the fact that they made it possible to experimentally verify many of quantum
mechanics principles, which were originally formulated purely theoretically and
were more likely academic and abstract ideas. A striking example is a parabolic
QD, which made it possible to test the fundamental properties of two- and three-
dimensional quantum oscillators experimentally. In particular, a generalization of
Kohn’s theorem [28] in parabolic QDs [29–33] proved the feasibility of these
models applied to quantum nanostructures. On the other hand, using ring-shaped
QDs, an experimental verification of the Aharonov-Bohm effect for bound states
became possible [34].
Among zero-dimensional systems, a special class is made up of spherical and
cylindrical layered structures consisting of a core and a shell – so-called core/shell
QDs. Such structures are experimentally implemented and investigated on the basis
of various methods. The problem of synthesis and optical properties of such systems
is addressed in many works. Thereby in [35], the authors report about the synthesis
of luminescent (CdSe) ZnS-composite QDs with CdSe nuclei by considering
optical properties and determining how the structure of the ZnS shell affects the
properties of photoluminescence. QDs with a core/shell structure were developed
in [36]. The authors in this paper showed that the final shell of ZnSe is used
to stabilize the dots for applications in aqueous environments and the first shell
of I nP leads to redshift and an increase in the quantum yield. The third-order
nonlinear optical properties of a core-shell QD using the Z-scanning method are
studied in [37]. It was demonstrated that measured values of the effective nonlinear
absorption coefficient and nonlinear refractive index of CdS − Ag QDs are several
orders of magnitude larger than those of uncoated CdS QDs. The authors of
[38] fabricated structured colloidal QD of GaAs/ZnSe and I nx Ga1−x As/ZnSe
using an indium and gallium acetylacetonate complexes. They also demonstrate
that shell protects the core and improves quantum yield. In a review [39] present
semiconductor nanocrystals, which describes the use of core-shell QDs. Core-shell
QDs can increase the photoluminescence efficiency of single QDs. Luminescent
properties of Ge/Si core/shell QD were realized in a paper [40]. Si-QD with a
Ge core is used in light-emitting devices that combine with ultra-large integrated
processing. Authors of [41] have chemically synthesized CdT e/CdSe type II
core/shell QDs. It was found that compared to CdT e cores, the CdT e/CdSe type
II core/shell QDs have much longer PL decay times, up to about 100 ns.
Depending on the position of the valence and conduction band and the essential
energy gap between them in the semiconductors, different categories of core/shell
Electronic and Optical Characteristics of Core/Shell Quantum Dots 125

Fig. 1 Types of core/shell QD

QD exist. The core could have a larger band gap than the shell (type I) or vice versa
(inverse type I), or the conduction band or valence band of the core could be located
within the band gap of the shell (type II) (Fig. 1).
Each of the types has their advantages which define their application in different
optoelectronic devices of new generation. In this monography we will consider only
type I QDs.
From a theoretical point of view, the description of the core/shell QD is
associated with the construction of a realistic Hamiltonian of a studied system. One
of the important stages is the presentation of a mathematical model of the confining
potential of the system. The layered geometry of the core/shell QD determines the
presence of boundary conditions imposed on the wave function, both at the outer
and inner boundary [42]. For example, in the case of a spherical core/shell QD, the
radial motion of particles inside the layer is limited by both the inner R1 and outer
R2 radii [43].
In various works, different models of the confining potential of the core/shell QD
were proposed: a rectangular impenetrable well [44, 45] (Fig. 2), a shifted oscillator
[46, 47] (Fig. 2), Winternitz-Smorodinsky potential or a volcanic potential [48, 64]
(Fig. 3), and the molecular potential of Kratzer [49, 50] (Fig. 4).
Some of the potentials presented for single-particle states are analytically
exactly solvable, which allows us to give a detailed description of various physical
parameters of studied structures. An important circumstance of the core/shell QD
theoretical description is the generalizing nature of the results. Indeed, by means of
the corresponding limit transitions, one can pass from the core/shell QD geometry
to quantum nanostructures with a simpler geometry. For example, in the model of
a spherical core/shell QD with a fixed external radius R2 by making the transition,
then as a result we arrive at a case of spherical QD of radius R2 . On the other hand,
if without changing the thickness of the layer R2 − R1 = L = const at the same
126 D. A. Baghdasaryan et al.

Fig. 2 Rectangular and


shifted parabolic confinement
potentials

Fig. 3
Winternitz-Smorodinsky
confinement potential

Fig. 4 Kratzer confinement


potential
Electronic and Optical Characteristics of Core/Shell Quantum Dots 127

Fig. 5 Limiting cases of spherical core/shell QD

time realizing the transitions {R1 → ∞, R2 → ∞}, then we get a quantum well of
thickness L (Fig. 5).
Physical processes in core/shell QDs were theoretically investigated in many
works. The authors of [51] studied the electronic states and wave functions of
both type I and type II core/shell CdSe/CdS and CdSe/CdT e QDs. The optical
transitions and Stark shift in spherical core/shell QD are studied in the “strong”
quantization regime in the paper [52]. The author discusses the influence of
quantum-confined Stark effect on the interband and intersubband optical transitions
in such system. In the paper [53] in the frame of four-band envelope function
formalism, the electronic structure of P bSe/P bS core-shell QDs has been studied.
The propagation of exciton polaritons in near-field-coupled core/shell QD chains
was studied in [54]. The authors used the density-matrix formalism and showed
that at least for low temperature, it is possible to use electronically controlled
switching by the quantum-confined Stark effect. In [55] the authors investigate
the simultaneous influence of the magnetic field combined to the hydrostatic
pressure and the geometrical confinement on the behavior of an impurity confined
in GaN/I nGaN core/shell QD and the photoionization effect in such system. It
was shown that the size of the core/shell, magnetic field, and hydrostatic pressure
affect not only the binding energy but also the magnitude of the resonance peaks of
the photoionization and their shift toward the higher energies.
Fast development of the fabrication technology of core/shell QD bring to a
huge number of applications in many research fields, such as sensing, electronics,
bio-imaging, and medicine. For example, the core/shell QD conjugated with
different enzymes can serve as FRET probes to sense glucose [56] and has also
been demonstrated to be effective pH sensors [57]. Another application of such
structures is cell separation where the core/shell acted as a shell in the polymer-
coated magnetite core [58]. Core/shell QDs show emitting behavior and high
128 D. A. Baghdasaryan et al.

quantum yield which is important for LED applications [59, 60]. The core/shell
structures play an important role in technology of infrared devices both as infrared
light emitters (e.g., in light-emitting diodes, biological imaging, etc.) and infrared
absorbers (e.g., in photovoltaics, solar fuels, photon up-conversion, etc.) of new
generation [61].
Thus, the study of the physical properties of core/shell QDs continues to
remain in the center of attention of specialists, since the results obtained can
find important practical applications. In this chapter, we present some results of
studying the electronic, optical, and spin characteristics of spherical and cylindrical
core/shell QDs.

2 Theory

2.1 One Particle States in Spherical Core/Shell QD

As it was discussed in the introduction, the confining potential of the core/shell


QD should consider availability of both inner and outer interfaces. In this case,
a mathematical model of such potential is formed both by geometry, physical,
and chemical characteristics of QD core, shell, and the environment. For relatively
simple spherical and cylindrical geometry, it is possible to introduce model confin-
ing potentials allowing to give analytical solutions to the one-particle Schrodinger
equation. In this paragraph we discuss two such potentials. Let us start from simplest
confinement potential model of infinite potential well:

0, R1 < r < R2
Vconf (r) = , (1)
∞, r ≤ R1 , r ≥ R2

where R1 and R2 are inner and outer radii, respectively Fig. 6.


In order to find the energy spectrum and envelope wave function in the framework
of effective mass approximation, we consider the Schrodinger equation with the
following Hamiltonian:

h̄2
Ĥ = − Δ + Vconf (r) . (2)

In dimensionless quantities (lengths are measured in effective Bohr radii aB =


h̄2 d /μe2 , and energies in effective Rydberg constant ER = h̄2 /2μaB 2 ), the
Schrodinger equation inside the layer region has the form:

− ΔΨnr , ,m (r, θ, ϕ) = EΨnr , ,m (r, θ, ϕ) , (3)


Electronic and Optical Characteristics of Core/Shell Quantum Dots 129

Fig. 6 Schematic view of the


spherical core/shell QD

where Δ (r, θ, ϕ) is the Laplace operator in spherical coordinates, E is the total


energy of the electron, Ψnr , ,m (r, θ, ϕ) is the wave function, and nr is the radial, =
0, 1, . . . is the orbital, and m = 0, ±1, ±2 . . . is the azimuthal quantum numbers.
Since the system has spherical symmetry, its wave function can be represented
as the product of the radial and angular parts:

Ψnr , ,m (r, θ, ϕ) = fnr , (r)Y ,m (θ, ϕ), (4)

where Y ,m (θ, ϕ) are spherical harmonics [62].


For the radial equation, we get:

d 2 fnr , (r) 2 dfnr , (r) ( + 1)


+ − fnr , (r) + k 2 fnr , (r) = 0, (5)
dr 2 r dr r2

where knr , ≡ k = E. The solution to this equation is the linear combination of
spherical Bessel functions [62]:

fnr , (r) = C1 J (kr) + C2 Y (kr), (6)

where C1 and C2 are normalization constants and J (kr) and Y (kr) are Bessel
spherical functions of the first and second kind, respectively.
Since we are considering an infinitely deep potential, the wave function at the
boundary of the spherical layer should vanish of the two boundary conditions:

Ψnr , (R1 ) = Ψnr , (R2 ) = 0. (7)


130 D. A. Baghdasaryan et al.

Fig. 7 Core/shell QD with (a)


infinitely deep rectangular
potential energy spectrum
dependence on its outer (a)
and inner (b) radii

(b)

The system of equations has nontrivial solutions if the determinant of the


coefficients for C1 and C2 is equal to zero.

J (kR1 ) Y (kR1 )

det = = 0. (8)
J (kR2 ) Y (kR2 )

This transcendental equation is solved numerically. As a result, it is possible to


obtain energy spectrum dependencies on inner and outer radii.
As one can see from Fig. 7, the dependencies E (R1 ) and E (R2 ) have the oppo-
site behavior. This fact can be easily explained based on Heisenberg’s uncertainty
relation. Indeed, with increasing R1 for a fixed value of R2 , the layer thickness
decreases. Accordingly, the localization area of the charge carrier decreases. As a
Electronic and Optical Characteristics of Core/Shell Quantum Dots 131

Fig. 8 (a) The density of the


radial distribution of the
probability of an electron in a
spherical layer for 1S and 2S
states at fixed R1 = 1.5aB
and R2 = 2aB . (b) The
density distribution in
cross-sectional region of an
electron in a spherical layer
for 1S and 1P states at fixed
R1 = 1.5aB and R2 = 2aB

result, its momentum increases and hence its energy. If the inner radius remains
unchanged and the outer radius grows, then the opposite picture arises.
Note that the coefficients C1 and C2 depend on each other, so just define one of
them. Then the radial wave function can be represented as follows:

fnr , (r) = A (J (kr) + CY (kr)) , (9)

 − 1

2
R 2
where A = {J (kr) + CY (kr)}2 r 2 dr normalization factor and C =
R1
−J (kR1 )/Y (kR1 ).
The radial probability densities of the electron inside the spherical layer are
presented in Fig. 8.
As it can be seen from Fig. 8a, the radial probability distribution for the ground
state has a maximum near the shell central region, which means that the electron is
localized in between inner and outer radii region. For the case when radial quantum
number is excited, we can see distribution with two maximums and minimum
located in between. In Fig. 8b the density distribution is plotted in cross-sectional
132 D. A. Baghdasaryan et al.

region. It is obvious that in the ground state, the wave function depends only on
radial coordinate.
Along with infinitely deep walls, the confining potential can be successfully
modeled using a three-dimensional Winternitz-Smorodinsky potential, or it is
otherwise called a three-dimensional volcano potential. Winternitz-Smorodinsky
potential not only well describes the presence of the inner and the outer boundaries
of quantum ring but also allows one to take into account the asymmetric behavior
of the potential function with respect to the minimum point which can play an
important role in studying the optical characteristics of such structures [63, 64]. This
confinement potential in the spherical coordinate system has the following form:

A √
Vconf (r) = 2
+ Br 2 − 2 AB, (10)
r
where A and B are constants characterizing the nanostructure described by the
height of the potential leap and width of the confinement potential on the QD
interface. The possible schematic presentation of the confinement potential fitted
to the energy profile of the Ga1−x Alx As/GaAs core/shell structure is shown in
Fig. 9. Here x is concentrations of Al, in core region. In the direct gap range,
the dependence of band gap on x concentration is determined by Eg (eV ) =
1.424 + 1.247x. Effective mass of charge carriers can be considered the same in
all structure of core/shell for small concentrations of Al.
Note that the confinement potential parameters A and B are connected with the
outer and inner radii by the following relations:

BRi2 + A/Ri2 − 2 AB = Ui , i = 1, 2 (11)

where Ui is defined by a relation of Ui = 211.9Qi xi , where Qe = 0.63 for electron


and Qh = 0.37 for hole. It is obvious that solutions of the system for electrons and
holes are connected with each other (Ah , Bh ) = Qh /Qe (Ae , Be ) , and solving this

Fig. 9 Winternitz-
Smorodinsky confinement
potential and energy profile of
core/shell QD
Electronic and Optical Characteristics of Core/Shell Quantum Dots 133

system of equations for the Winternitz-Smorodinsky parameters, one will get:



R1 2 R2 2 R2 2 U1 + R1 2 U2 + 2R1 R2 U1 U2
Ae = 2 2 , (12)
R1 − R2 2

R1 2 U1 + R2 2 U2 + 2R1 R2 U1 U2
Be = 2 2 . (13)
R1 − R2 2

Let us consider the strong size quantization regime, when we can neglect
electron-hole interaction because inter-particle interaction energy is much smaller
than the energy caused by size quantization. In this regime, the issue separates to
the one particle problem. The radial part of the Schrodinger equation for the electron
in dimensionless quantities will have the following form:
 
d 2 fnr , (r) 2 d fnr , (r) ( + 1) Ae
+ − + E − 2 − Be r fnr , (r) = 0,
2
(14)
dr 2 r dr r2 r

here fn√r , (r) is radial wave function, μ is electron (hole) effective mass, and E =
E + 2 AB, where E is the energy of the system and orbital quantum number.
Making following notations

2s (2s + 1) = ( + 1) + A, (15)
E
√ = 4 (nr + s) + 3. (16)
B

From which we can write


 
1
s= −1 + (2 + 1)2 + 4A , (17)
4
 
1 E
nr = −2 + 4A + (1 + 2 )2 + √ , (18)
4 B

where nr is radial quantum number. Spectrum and the solution of the equation are
known and given by the following expressions [62]:
 
√ √
Enr , = B 4nr + 2 + (2 + 1) + 4A − 2 AB, nr = 0, 1, ..
2
(19)

Note that the spectrum (19) for the A = 0 precisely comes to the expression for
spherical oscillator:
 
√ √ 3
osc
EN = B {4nr + 2 + 3} = 2 B N + , (20)
2
134 D. A. Baghdasaryan et al.

Table 1 The Winternitz-Smorodinsky parameters and ground-state energy value for different
values of inner and outer radii. Here the values of Al concentrations x1 = x2 = 0.4

R1 R2 Ae , ER aB2 Br , ER /aB2 E0,0 , ER


1 1.5 362.464 174.657 26.60
1 2 156.696 44.824 13.52
1 3 85.097 11.574 6.89

where N = 2nr + is principal quantum number. The corresponding values of the


ground-state energies for different inner and outer radii are presented in Table 1.
In its turn, the wave function will have the form of the product angular and radial
parts:

Ψnr , ,m (r, θ, ϕ) = Y ,m (θ, ϕ) fnr , (r) , (21)

where Y ,m (θ, ϕ) again are spherical harmonics, and for fnr , (r) we have [62]:
 
−ξ/2 s 3 √
fnr , (r) = e ξ 1 F1 −nr , 2s + ; ξ , ξ = Br 2 (22)
2
here 1 F1 (a, b; x) is confluent hyper-geometric function of the first kind.

2.2 Impurity States in Spherical Core/Shell QD

Along with single-particle states in spherical core/shell QD, Coulomb systems such
as hydrogen-like impurities and excitons are intensively studied. An interesting task
is the problem of the influence of the impurity center on the electronic properties of
layered nanostructures. In recent years this problem has been studied by several
authors. Thus, the linear and nonlinear optical absorption associated with the
transition between 1s and 2s states corresponding to the electron-donor-impurity
complex in GaAs/Ga1−x Alx As three-dimensional concentric double quantum
rings is investigated in [65]. The linear and nonlinear intraband optical absorption
coefficients are considered in GaAs three-dimensional single quantum rings [66].
The donor impurity in the elliptic quantum ring [67, 68] in the presence of magnetic
field was observed in [67]. Aharonov-Bohm oscillations were found caused by both
magnetic field and eccentricity.
Consider a core/shell spherical QD of GaAs in the center of which the donor
impurity is located. Just note that we do not take into account polarization effects
and we expect that the permittivity d is the same within the QD and in the
environment.
It can be seen from Fig. 4 that the parameters are related with the height of
confinement potential and the potential minimum point r0 of Vconf (r) potential
according to the relation:
Electronic and Optical Characteristics of Core/Shell Quantum Dots 135

α = U0 r02 , β = 2U0 r0 . (23)

In the spherical coordinates, the Schrodinger equation takes the form:


   
h̄2 1 e2
− Δr + 2 Δθ,ϕ + Vconf (r) − Ψ ( r ) = EΨ ( r ) , (24)
2μ r d r

where d is the dielectric constant of QD and

α β
Vconf (r) = − + U0 . (25)
r2 r
Since the field is spherically symmetric, taking into account the spin of the
electron, the wave function can be written as an eigenfunction of the total momen-
 2  
tum Jˆ2 = L ˆ + s ˆ and its projection Jˆz = L̂z + ŝz ŝz = h̄ σz [69], where
2
 
1 0
σz = is Pauli matrix:
0 −1
 
−i ∂ϕ

+ 1
0
Jˆz = h̄ 2
, (26)
0 −i ∂ϕ

− 1
2

 2 3 2 
ˆ L̂ + 4 h̄ + h̄L̂z h̄L̂+
J =2
, (27)
h̄L̂− L̂2 + 34 h̄2 − h̄L̂z

where L̂± = L̂x ± i L̂y . Electron wave function can be represented as a product of
angular and radial functions:
 
j +1/2 fnr , (r) + mj + 1/2 · Y ,mj −1/2 (θ, ϕ)
Ψnr , ,mj (r, θ, ϕ) = √  , (28)
2 + 1 − − mj + 1/2 · Y ,mj +1/2 (θ, ϕ)
 
j −1/2 fnr , (r) − mj + 1/2 · Y ,mj −1/2 (θ, ϕ)
Ψnr , ,mj (r, θ, ϕ) = √  . (29)
2 + 1 + mj + 1/2 · Y ,mj +1/2 (θ, ϕ)

For the radial wave function, we obtain the equation:


 
d 2 2 d ( +1) 2μα 1 2μβ 1 2μe2 1 2μ (E−U0 )
+ − − 2 2+ 2 + + fnr , (r) =0
dr 2 r dr r2 h̄ r h̄ r d h̄2 r h̄2
(30)
136 D. A. Baghdasaryan et al.

making notations:
  
2μα 

 1  2 8μα e2
, e =β +
2
( +1) + 2 = + 1 , = −1+ ( +1) + 2
h̄ 2 h̄ d
(31)
2μ (E − U0 ) e 2μ
= −γ 2 , κ = , (32)
h̄2 γ h̄2

as well as introducing new wave function χnr , (r)

χnr , (r)
fnr , (r) = , (33)
r
we obtain the following equation for χnr , (r)
 
 2γ κ   + 1
χnr , (r) + −γ + − 2
χnr , (r) = 0. (34)
r r2

The solution of this equation satisfying the standard conditions is the following
expression [69]:
  
fnr ,l (r) = Cr e−γ r 1 F1  + 1 − κ, 2  + 2; 2γ r , (35)

where −nr =  + 1 − κ, nr = 0, 1, 2, . . . .
If we now introduce a non-integer quantum number n defined as
  
1 8μα
n = nr + 1 + −1 + (2 + 1)2 + 2 , (36)
2 h̄

then we finally obtain for the energy of the system:


 2
e2
2 β+ d μ
En3D
r ,
= U0 − . (37)
h̄2 (2n)2

The binding energy of the impurity is defined as the difference between the levels
of the electron without and with impurity:

Ebind = Ewithout − Ewith . (38)

Figure 10 shows the dependence of the probability density distribution of


electron localization in the layer on the r radial variable. It can be seen that the
shift of peaks occurs toward the outer radius for = 1 states. This is due to the fact
that angular energy increases with , and therefore, the maximum peak shifts toward
the outer boundary of the QD.
Electronic and Optical Characteristics of Core/Shell Quantum Dots 137

Fig. 10 Dependence of the


probability density of the
electron localized in a
spherical nanolayer in the
presence of the impurity on
the radial variable for
different states

Fig. 11 Dependence of the


electron energy levels on the
parameter r0

At a fixed U0 value with the increase of r0 , impurity electron begins to remove


from the geometric center of the core/shell QD. As a consequence, the Coulomb
interaction energy decreases (in absolute value). Figure 1a shows that the well width
increases with r0 , and therefore the size quantization energy decreases. As a result,
the total energy of the system begins to decrease monotonically in the case of r0
increase (Fig. 11).
The contrary situation takes place when the r0 remains constant; however, the
potential jump U0 changes (Fig. 12). In this case, the quantum well becomes
narrower, and U0 increases at the outer edge. As a result, there is a monotonic
increase of impurity energy levels (Fig. 12).
Finally, let’s pay attention to the binding energy of our system. As it can be
seen from Fig. 13, the impurity binding energy initially increases with the decrease
of r0 . This is due to the fact that the impurity electron approaches to the center
and, accordingly, the Coulomb energy increases. However, because of the finite
height of confinement potential at the outer boundary, the electron emission occurs
from core/shell spherical QD into the external environment for small values of r0 .
138 D. A. Baghdasaryan et al.

Fig. 12 Dependence of the electron energy levels on the parameter U0

Fig. 13 Dependence of
electron binding energy on
the parameter r0 for fixed
values of the well depth U0

Electron is bound with impurity when it is in the environment. Localization radius


increases and the module of Coulomb energy decreases.

2.3 Two-Electron Sates in Thin Spherical Core/Shell QD: Two


Electrons on a Sphere

It should be noted that for the case of sufficiently thin nanolayer, the radial
localization of charge carriers may become so narrow that the quantum state of
the particle can effectively be described by a motion of a spherical surface. In this
situation the effective mechanism for theoretical description of particle states in a
Electronic and Optical Characteristics of Core/Shell Quantum Dots 139

nanolayer is the adiabatic method when the system can be separated into “fast”
and “slow” subsystems. The given approximation for the case of one particle states
allows to apply the model of two-dimensional and spherical rotators and obtain
energy spectrum and wave functions of the charge carrier. Now, if we consider
two-electron system then, as it was shown in Ref. [70] on the example of quantum
ring, due to the strong radial quantization in the first approximation, we can discuss
the Coulomb interaction dependent only upon relative angle between electrons. In
the mentioned paper [70] instead of exact potential for two electrons positioned
on the circle and interacting according to the Coulomb law, it has been inserted
an approximated potential of interaction dependent on the inter-electron angle ϕ1
and ϕ2 :
 √ 
approx e2 1 2−1
VCoul (ϕ1 − ϕ2 ) = √ + cos(ϕ1 − ϕ2 ) . (39)
d Reff 2 2

This allowed to reduce the solution of the two-dimensional problem to the


Mathieu equation and to find analytical expressions for two-electron spectra
expressed through the well-known tabulated solutions of Mathieu equation. It was
demonstrated that the given system can perform both harmonic and an harmonic
oscillations as well as librational motion. It is clear that in the case of the thin
spherical nanolayer containing two electrons, the Coulomb interaction force also
depends on the relative angle between electrons [69, 71]. Indeed, as it was shown
in papers [72–75], it is possible to reduce two particle Hamiltonian to the one
particle by inserting inter-electron angle. However, the problems of spherium
and harmonium (Hookium atom) were studied by the authors of the mentioned
papers initially discussing the electrons on the surface of the sphere and without
connecting the problems to nanostructures. Alavi et al. [76–79] have investigated
two electrons interacting with a Coulomb potential in a sphere of radius R. Loos
and Gill along with two electrons in a sphere [80] have also studied two-electron
states on the surface of a sphere [81–84]. This is so-called spherium model. It
should be mentioned that the spherium model was studied by Berry at al. [85–
88] to understand both weakly and strongly correlated systems and to suggest
an “alternating” version of Hund’s rule [89]. Various models of repulsive force
between particles have been discussed such as exact Coulomb, Gauss, and delta
potentials. Energy spectrum as well as eigenfunctions is obtained for electron on a
sphere and also on concentric spheres too. Seidl et al. [90, 91] studied the spherium
system in the context of density functional theory (DFT) to develop new correlation
functionals within the adiabatic connection. In the current paragraph, the adiabatic
method is used in order to study two-electron states in a spherical nanolayer of small
thickness. It is shown that the Coulomb interaction between electrons here also can
be approximated by potential (39) and, thereby, the two-electron problem can be
reduced to the one-electron case.
140 D. A. Baghdasaryan et al.

Let us consider the two-electron states in a spherical QD with confining potential:



0, R1 < r < R2
Vconf (r) = , d = R2 − R1  {R1 , R2 } , (40)
∞, r ≤ R1 , r ≥ R2

where d is the thickness of the shell. From the uncertainty principle, it directly
follows that the quantization energy in radial direction is very strong compared to
angular and Coulomb interaction energies. Indeed,

P2 h̄2
∼ , (41)
2μ 2μ (R2 − R1 )2

h̄2
therefore, the radial energy is proportional to , whereas the Coulomb
2μ(R2 −R1 )2
one, for the model we discuss (electrons are localized in diametrically opposite
sides of the sphere because of the Coulomb repulsive force), is proportional to
e2
r1 +r2 , where r 1 , r 2 are radius vectors of electrons. Hence, when d is small enough,
the radial quantization energy becomes very large. This makes possible to apply
stationary adiabatic approximation, i.e., we can consider that variables of radial
and angular parts can be separated as “fast” and “slow” subsystem coordinates. In
its turn the stronger is the quantization, the further are the energy levels of radial
part from each other. Hence, we consider that in radial direction electrons are in
one-electron ground states. In such case the most probable radius (maximum of
radial wave function) which describes the location of electrons is approximately
Reff = (R2 + R1 )/2 . Hence, we can consider that the electrons are located on
a sphere with Reff effective radius. Therefore, the problem of two electrons in
spherical nanolayer we reduce to the problem of two electrons on a sphere. The
latest leads the Coulomb interaction to be dependent on angular variables only.
The Hamiltonian of the system has the following form:

h̄2
Ĥ (1, 2) = − Δr1 ,Ω1 +Δr2 ,Ω2 +Vconf ( r1 )+Vconf ( r2 )+VCoul (r1 , r2 , Ω1 , Ω2 ) ,

(42)
where Ωi = {θi , ϕi }.
Using adiabatic approximation and taking into account our assumption that
the Coulomb interaction potential depends only from angular variables, one can
reduce two-electron Schrodinger equation (which was not analytically solvable) into
the two-dimensional Schrodinger equation for angular part of the quantum ring with
the effective potential. For the radial wave function, we will have [71]:

π k0
f0 (r) = C1 J1/2 (k0 r) + C2 J−1/2 (k0 r) , (43)
2r
Electronic and Optical Characteristics of Core/Shell Quantum Dots 141


2μE0rad
where Jυ (x) is the Bessel function of the first kind and k0 = , E0rad is radial
h̄2
ground-state energy.
In order to calculate the ground-state energy of radial quantization, we come to
the equation below if we consider the boundary conditions f0 (R1 ) = f0 (R2 ) = 0:

J1/2 (k0 R1 ) J−1/2 (k0 R1 )

det = = 0. (44)
J1/2 (k0 R2 ) J−1/2 (k0 R2 )

Solving of the system, one can obtain the energy of the ground state:

π 2 h̄2
E0rad = . (45)
2μ (R2 − R1 )2

In the introduction it was already mentioned that different authors have discussed
various models of interaction between electrons such as exact Coulomb, Gauss,
delta, and many other potentials. Here, in this paragraph, instead of exact Coulomb
interaction potential of electrons, we are going to use the same approximated
potential which, for harmonic approximation, also leads to an analytically solvable
equation. The reason why we discuss the Coulomb potential within π2 and π angles
is that in our model we consider that electrons don’t come closer then π2 . Because
of the repulsive force between electrons, they tend to be located in opposite sides
on diagonal of the sphere. Note that angle θ is related with the angles {θ1 , ϕ1 } and
{θ2 , ϕ2 } according to the following expression:

( r1 r 2 )
cos θ = = cos θ1 cos θ2 + sin θ1 sin θ2 cos (ϕ1 − ϕ2 ) . (46)
R2
With direct calculations we can show that the Schrodinger equation takes the
following form if we transfer from the variables {θ1 , ϕ1 } and {θ2 , ϕ2 } to variable θ
[72, 73]:
   √ 
h̄2 d2 d e2 1 2−1
− + cot θ Φ (θ ) + √ + cos θ Φ (θ )
μReff dθ 2 dθ d Reff 2 2
 
π 2 h̄2
= E− Φ (θ ) . (47)
μ(R2 − R1 )2

In general, the given equation is not integrable; therefore let’s discuss the case
when the electrons are positioned on diametrically opposite sides and perform small
oscillations around the corresponding points of equilibrium. In this case θ ∼ π ;
hence we perform change of variables:

φ = π − θ. (48)
142 D. A. Baghdasaryan et al.

The condition θ ∼ π in terms of the new variable will have the form φ ∼ 0, and
equation (47) will have the following form:
   √ 
h̄2 d2 d e2 1 2−1
− + cotφ Φ (φ) + √ − cosφ Φ (φ)
μReff dφ 2 dφ d Reff 2 2
 
π 2 h̄2
= E− Φ (φ) . (49)
μ (R2 − R1 )2

By the analogy of papers of Refs. [73–75], considering that the angle is small
enough, we can expand functions cot φ and cos φ to the series of Taylor. After some
transformations for the definition of Φ (φ), we will come to an equation which in its
form coincides with the equation of cyclic oscillator with magnetic quantum number
equal to zero:
 
h̄2 d2 1 d μ ωeff
2 ξ2
− + Φ (ξ ) + Φ (ξ ) = EΦ (φ) , (50)
μ dξ 2 ξ dξ 2

where the following notations are used:


√ 1
 μ 2−1 e2 2
μ = , ωeff = , (51)
2 2 d μ Reff
3

π 2 h̄2 e2
E =E− − , (52)
μ (R2 − R1 )2 2 d Reff

ξ = Reff φ. (53)

For Φ (ξ ) and E, one can write directly:

ξ2
 
Φn (ξ ) = Cn e−λ 2 1 F1 −n; 1; λξ 2 , (54)

μ ωeff
where λ = h̄ and Cn is normalization coefficient,

En = h̄ωeff (2n + 1) , (55)

where n = 0, 1, . . . is oscillator quantum number.


It is necessary to mention that according to expressions (51), (52), and (53),
cyclic frequency ωeff in expression (55) is proportional to Reff −3/2 . Therefore, by
increasing Reff cyclic frequency, ωeff decreases, and inter-level distance decreases
too. This is an expected result as far as by increasing Reff , within the frameworks
of our model, the distance between electrons increases and, hence, interaction
Electronic and Optical Characteristics of Core/Shell Quantum Dots 143

potential decreases. Therefore, the oscillation conditioned by relative location of


electrons becomes weak. Also let us note that for ωeff it was obtained an analogical
dependence on Reff in Ref. [73]. Finally, in the frameworks of considered model
for energy spectrum of two-electron system in the harmonic approximation, we can
write:

π 2 h̄2 e2
E = h̄ωeff (2n + 1) + + . (56)
μ(R2 − R1 )2 2 d Reff

In its turn the wave function of the system will have the following form:
 2  
 π k0
Ψ (r1 , r2 , θ ) = C1 J1/2 (k0 ri ) +C2 J−1/2 (k0 ri ) ×
2ri
i=1
ξ2
 
×Cn e−λ 2 1 F1 −n; 1; λξ 2 . (57)

For comparison of our results with other results, we discuss also the first order
correction energy in the frameworks of our model. The first order correction of
energy can be written by using perturbation theory:

∞
En(1) = 2π Φn (ξ )V̂ Φn (ξ ) ξ dξ, (58)
0

where

h̄2 φ d 2 − 1 e2 φ 4
V̂ = 2 3 dφ
− . (59)
μReff 2 d Reff 24

Perturbation operator V̂ is easy to obtain by considering next terms of Taylor


expansion of cot φ and cos φ. For the ground state, the perturbation energy is
calculated and gives the following expression:
 √ 
(1) 1 h̄ 2−1 e2
E0 =− 2 + 3
. (60)
3Reff μ 8 λ2 d Reff

Figure 14 represents the value of correction energy dependence on Reff for


GaAs/Gax Al1−x As/GaAs structure, for which d ≈ 13. As we can see in
the figure, the modulus of energy correction of relative motion decreases with
increasing the effective radius. This is an expected result as far as by increasing
the effective radius (according to our approximation the relative angle θ is closed to
π ) the distance between electrons increases. Therefore, the interaction of electrons
conditioned by the relative motion should become weaker. Also, it is interesting
144 D. A. Baghdasaryan et al.

Fig. 14 Dependence of
perturbation energy on
effective radius

to discuss the ratio of first-order correction energy and harmonic approximation


energy given by expression (55). For the ground state when Reff = 20 nm, the
ratio is approximately equal to −0.56, for Reff = 40 nm is equal to −0.39, and for
Reff = 60 nm is equal to −0.32. One can see the ratio becomes one or more orders
smaller from harmonic approximation energy for enough big radii. This is explained
by the fact that we have discussed the Coulomb interaction as perturbation; hence,
the perturbation should be small (and will act as a small correction) only for big
radii, where Coulomb interaction is weak enough.

2.4 Interband Optical Absorption in Spherical Core/Shell QD

We focus our attention on calculation of different spectral characteristics of the


spherical core/shell QDs in previous paragraphs. Let us consider concrete physical
parameters of core/shell QDs.
In this paragraph, we will study the interband optical absorption in the spherical
core/shell QD with the Winternitz-Smorodinsky confining potential. On the basis of
the results obtained in the first paragraph for single QD, it is possible to calculate
absorption coefficient of QD ensemble.
Let us consider the direct interband optical absorption of the ensemble of
core/shell QDs with the dispersion of their sizes in the regime of strong size
quantization. Such dispersion will lead to broadening of spectral lines. According
to [68] for light absorption coefficient for the strong regime, we have following
expression:
   
2
K (ω) = K0 J PLS (u) δ h̄ω − Eg − Eνee − Eνhh du, (61)
νe ,νh
Electronic and Optical Characteristics of Core/Shell Quantum Dots 145

Fig. 15 The dependence of


the absorption edge on the
outer radius of the core/shell
structure for different values
of the Al concentrations in
shell region

where PLS is the distribution function for QD geometrical parameter, Eg is the


energy gap of the semiconductor,
 ω is frequency of the incident light, and νe =
nre , e , me and νh = nrh , h , mh are the sets of quantum numbers of the
electron and the hole, respectively. K0 is a quantity proportional to the square of
modulus of the matrix element of the dipole moment taken over the Bloch functions.
The dimensionless oscillator strength is defined in the following way:

J = Fνe ,νh /F0 = Ψe ( re ) Ψh ( rh ) dV , (62)

where F0 is the oscillator strength of the bulk material. From the argument of delta
function, one can easily define the absorption edge h̄ω000 , the dependence of which
is shown in Fig. 15.
As we can see from Fig. 15, the absorption edge decreases with the increase of
outer radius. The curves corresponding to the bigger values of the Al concentrations
in shell region are located above.
The oscillator strength in the strong regime of size quantization in this system in
the explicit form can be presented as the following expression [68]:

 ∞
J = Y e ,me (θ, ϕ) Y h ,mh (θ, ϕ) dΩ fne , e (r) fnh , h (r) r 2 dr, (63)
Ω=4π 0

where dΩ = sin θ dθ dϕ the element of solid angle. The first integral in (63) can be
easily calculated by using the orthogonality of spherical harmonics [62]:

Y e ,me (θ, ϕ) Y h ,mh (θ, ϕ) dΩ = δ e , h δme ,−mh . (64)
Ω=4π
146 D. A. Baghdasaryan et al.

Table 2 Values of the


nre , nrh Fnre ,nrh /F0
oscillator strength for
different radial quantum 0, 0 0.914
numbers. Here the values of 1, 1 0.744
Al concentrations
1, 0 0.141
x1 = x2 = 0.4,
R1 = 1, R2 = 2 2, 0 0.348
2, 1 0.341

From (64), the selection rules for the orbital and magnetic quantum numbers are
immediately obtained:

e = h , (65)
me = −mh . (66)

The radial part of the expression cannot be calculated analytically; thus it must
be determined numerically. Note that for nre = nrh , integral is not equal to zero;
however, it is small for this transition (see Table 2).
In order to describe the distribution of the geometrical parameter around the mean
value, we will use Gaussian distribution. Forms of these functions are given below:
 
1 (u − 1)2
PLS (u) = √ exp − . (67)
σ 2π 2σ 2

In (67) the variable u is the ratio of dispersive parameters to their average


R2 −R1
values u = R 2 −R1 
and σ is the standard deviation. After taking into account this
distribution function, one can obtain the dependence of the absorption coefficient on
the incident photon energy (see Fig. 16).
Figure 16 shows that there are sets of the peaks corresponding to the quantum
transitions between energy levels. The first set of peaks corresponds to the transition
nrh = 0 → nre = 0. The peaks in this set correspond to the transitions with
different values of orbital quantum number . The second set of peaks origins from
the overlay of transitions nrh = 1 → nre = 0 and nrh = 0 → nre = 1. It is
interesting that the intensity of this peak is not equal to zero. This is due to the
asymmetry of the Winternitz-Smorodinsky confinement potential. The intensity of
the third set of peaks is less than the first one. The same picture observes for the
second absorption curve corresponding to the smaller value of outer radii, with the
exception that curve for the hole has a blue shift.
The photoluminescence spectra are calculated using the Roosbroeck-Shockley
relation [92, 93] (see in Fig. 17):

Pc (1 − Pv )
RP L (h̄ω) = R0 h̄ω K (h̄ω) (68)
Pv − Pc
Electronic and Optical Characteristics of Core/Shell Quantum Dots 147

Fig. 16 The absorption


coefficient dependence on the
incident photon energy

Fig. 17 The dependence of


the PL intensity on the
frequency of the incident
light in arbitrary units

where R0 is proportional to the square modulus of the dipole moment matrix


element taken over the Bloch functions, Pc and 1 − Pv are probabilities of the
conduction band states being occupied, and the valance band states being empty,
respectively. For the high temperatures, the term PPc (1−P
v −Pc
v)
in Roosbroeck-Shockley
relation transforms to the Boltzmann-like form. Note that the PL curve has been
calculated for the room temperature and the exponential term vanishes the peaks in
the high-energy region.
148 D. A. Baghdasaryan et al.

2.5 Quadrupole Moment Created by Impurity Electron


in the Core/Shell QD

As it has been mentioned in the introduction, QDs have a wide range of applications.
They can be used in such devices as heterostructure lasers, sensors, solar cells, and
single-electron transistors [94–96]. In the case of single-electron transistors, it is
necessary to study the Coulomb blockade of the electron when it enters the QD in the
simplest case containing one electron. The problem of describing the electrostatic
interaction between two electrons becomes very significant. It is clear that electron
having wave properties can create both dipole and quadrupole fields. Therefore, it
is important to take into account these multipoles. On the other hand, the presence
of the static impurity also leads to the creation of an electrostatic field. Thus, the
resulting field is the superposition of these two fields. One of the remarkable features
of QD is the possibility of manipulating the energy levels of the impurity electron,
as well as its wave function by changing its geometric dimensions. As a result, we
can control the electrostatic multipoles (dipole and quadrupole moments) produced
by electrons.
Here it is worth giving a picture of a spherical nanolayer with an electron inside.
We note right away that for S states the potential will possess exclusively central
symmetry.
For states with = 0 arises angular distribution different from the centrally
symmetric. Therefore, nonzero values arise for electrostatic multipoles. Now we
note that the average dipole moment is zero:


Di  = Ψn , ,m Di dV , (69)
r j

Di = exi , (70)

Dx  = Ψn∗r , ,mj {r sin θ cos ϕ}Ψnr , ,mj dV , (71)


!
Dy = Ψn∗r , ,mj {r sin θ sin ϕ}Ψnr , ,mj dV , (72)


Dz  = Ψn∗r , ,mj {r cos θ }Ψnr , ,mj dV . (73)

By direct calculation, we can verify that all three integrals are equal to zero.
Let us turn to the calculation of the quadrupole moment of the system. The
quadrupole moment tensor is determined according to the relation:

Qik = 3xi xk − r 2 δik . (74)


Electronic and Optical Characteristics of Core/Shell Quantum Dots 149

Qik is a symmetric Qik = Qki and its trace is zero:

Sp {Qik } = 0. (75)

For average value of the Qik , we can write:

∞ 
2
Qik  = Ψn (r, θ, ϕ) Qik r 2 drdΩ. (76)
r , ,mj

0 Ω=4π

On the basis of the expression for wave function Ψnr , ,mj after direct calculations,
one can find:
! !
Qxy = Qxz  = Qyz = 0, (77)

! 1
Qxz  = Qyy = − Qzz  , (78)
2
where for Qzz we have
 
1 3m2j
Qzz  = 1− Inr , , (79)
2 j (j + 1)

where
    
d nr F  + 5/2,  + 3, 2  + 2, A2 /B 2

(1 − h)2 +2 B 2 +5

1 Γ (2  + 5) dhnr
Inr , =     
h=0
.
(2γ )2 Γ (2  + 3) d nr F  + 3/2,  + 2, 2  + 2, A2 /B 2

(1 − h)2 +2 B 2 +3

dhnr
h=0
(80)

Since the quadrupole moment is equal to


⎛ ⎞
1
 − 0 
0
3m2j ⎜ 2 ⎟
1 ⎜ ⎟
Q = 1− Inr , ⎜ 0 − 1 0⎟ . (81)
2 j (j + 1) ⎝ 2 ⎠
0 0 1

Table 3 shows when j = 1/2 the


spherical symmetry have both S and P
states. With increasing values of mj elongated shape of the electron, probability
distribution becomes flattened.
150 D. A. Baghdasaryan et al.

Table 3 The values of the ration of the quadrupole corrections to the point charge Coulomb
potential
Coefficient for nr = 0 − 0.32, for nr = 1 − 0.47, for nr = 2 − 0.67
States J mj = ±1/2 mj = ±3/2 mj = ±5/2 mj = ±7/2
S1/2 , P1/2 1/2 0 − − −
P3/2 , D3/2 3/2 +2/5 −2/5 − −
D5/2 , F5/2 5/2 +16/35 +4/35 −20/35 −
F7/2 , G7/2 7/2 +10/21 +6/21 −2/21 −14/21

2.6 Orbital and Spin Magnetic Moment Current


in the Cylindrical Core/Shell QD
2.6.1 Cylindrical Nanolayer

In the previous paragraphs, we considered the core/shell QD of spherical symmetry.


On the other hand, from a theoretical point of view, it is interesting to study physical
processes in core/shell QDs with cylindrical symmetry. In such ring-shaped systems,
persistent single-particle currents can arise [97, 98], a detailed study of which can
be of practical importance. The problem of charge current in ring-like structures
was discussed in many papers. For example, in Ref. [98] there has been studied the
effect of electron-electron interaction on the magnetic moment (associated with the
persistent current) of electrons in a quantum ring. There was introduced a model
where the electron makes a circular motion in a parabolic confinement simulating
a quantum ring which is subjected to a perpendicular magnetic field. The electron
states in such a ring with and without the Coulomb interaction are then investigated.
There also explored the limits of narrow and wide rings. In Ref. [99] it was
demonstrated the theoretical possibility of obtaining a pure spin current in a 1D
ring with spin-orbit interaction by irradiation with a non-adiabatic, two-component
terahertz laser pulse, whose spatial asymmetry is reflected by an internal phase
difference. In Ref. [100] the persistent current in two vertically coupled quantum
rings containing few electrons is studied. It was shown that the Coulomb interaction
between the rings in the absence of tunneling affects the persistent current in each
ring and the ground-state configurations. Quantum tunneling between the rings
alters significantly the ground state and the persistent current in the system. Also this
problem is discussed in Refs. [97, 101, 102]. In general, the quantum mechanical
expression for one-electron current in presence of magnetic field with consideration
of the spin of electron consists of two components. The first one characterizes the
orbital current j Orb , and it is connected with orbital motion of the electron [62]. The
second one is caused by the magnetic moment of electron, and it is called density of
spin magnetic moment current j SMM . As far as this current is not conditioned with
directed motion of the electron, therefore, its divergence equals to zero:

div j SMM = 0. (82)


Electronic and Optical Characteristics of Core/Shell Quantum Dots 151

This current is specific and its existence is caused by the presence of the spin
magnetic moment of the electron. In Ref. [103] the peculiarities of spin magnetic
moment current were discussed, particularly when hydrogen atom electron is in S
states. In these states, the total current is formed exclusively by the spin magnetic
moment current of the electron. In this paragraph the investigation of orbital and spin
magnetic moment current densities for an electron located in cylindrical core/shell
QD are presented in presence of axial magnetic field.

Energy Spectrum and Wave Functions

Let us consider the confining potential of the cylindrical core/shell QD in the


following form:

0, R1 < ρ < R2 , |z| < L2
Vconf (ρ, z) = , (83)
∞, ρ ≤ R1 , ρ ≥ R2 , |z| ≥ L
2

where L is the height of the cylindrical core/shell QD and R1 and R2 are,


respectively, the inner and outer radii (Fig. 18).


Let us also consider that the system is in an axial H homogenous magnetic field
with the gauge:
( )

A = Aρ = Az = 0, Aϕ = 2 . (84)

For the gauge chosen above, the variables in Schrodinger equation can be
separated. By taking into account the spin of the electron, one can obtain the three-
dimensional equation:

Fig. 18 Schematic view of


cylindrical core/shell QD
152 D. A. Baghdasaryan et al.

   
2
h̄2 1 ∂ ∂ ∂2 1 ∂ 2 i h̄ωH ∂ μωH ∗
− ρ + 2+ 2 2− + ρ +Vconf (ρ, z) −μ H σz ×
2
2μ ρ ∂ρ ∂ρ ∂z ρ ∂ϕ 2 ∂ϕ 8
×ψ (ρ, ϕ) g (z) χsz =
= Enz ,nρ ,m,sz ψ (ρ, ϕ) g (z) χsz . (85)

with the following boundary conditions:

Ψ |z=± L = Ψ |ρ=R1 = Ψ |ρ=R2 = 0. (86)


2

where μ∗ is the effective magnetic moment of electron (for GaAs μ∗ = gL μB ,


μB = 2m eh̄
ec
, gL = 0, 44); σz is the z component of Pauli matrices; ωH = eH
μc is the
cyclotron frequency, where μ = 0, 067me and me is electron rest mass; and χs is
the spin part of the wave function. For the z representation of Pauli matrices, one
can write:
   
1 0 1
σz = , χsz = 1 = . (87)
0 −1 2 0

Equation (85) has analytical solutions. After corresponding calculations for the
wave function, one can obtain the following expression:
   
n ,s 1 2 sin πLnz z (nz = 2k)
1
Ψnρz,mz (ρ, ϕ, z, sz ) = √ eimϕ fnρ ,m (ρ) .
2π L cos πLnz z (nz = 2k + 1)
0
(88)
where m = 0, ±1, ±2, . . . is the magnetic quantum number, nρ the radial quantum
number, and nz the axial quantum number.
After substitution of (88) into (85), we obtain an equation for the radial wave
E−E π 2 h̄2 nz 2
function. After inserting the following notations β = h̄ωHnz − m2 , Enz = ,
2μL2
aH = ωHh̄ μ – magnetic length the solutions for that radial equation can be written
as follows:

ρ2

fnρ ,m (ρ) = ρ |m| e 4aH ×
2

       
( |m| +1 ρ2 |m| +1 ρ2 )
× C11 F1 − β− , m+1, 2
+C2 U − β− , m+1, 2
,
2 2aH 2 2aH
(89)

where 1 F1 and U are confluent hypergeometric functions of first and second kinds
[62]. In Fig. 19 the distribution of electron probability density is depicted. As it
follows from Fig. 19, the maximal probability of localizing of electron is in the
Electronic and Optical Characteristics of Core/Shell Quantum Dots 153

Fig. 19 Distribution of
electron
probability density,

H = 104 Oe

center of the nanolayer in radial direction. In order to find energy spectrum for the
ground state, it is necessary to consider the radial boundary conditions (86) and
solve the obtained transcendental equation below:
       
|m|+1 R12 |m|+1 R12
1 F1 − β − , m + 1, U − β− , m + 1,
2 2aH2 2 2aH2
        = 0.

F − β− R22
, m + 1, 22
2
|m|+1 |m|+1 R
1 1 2 , m + 1, 2 U − β− 2
2aH 2aH
(90)
As a result of those operations, we will get:

 m  π 2 h̄2 nz 2
Enρ ,m,nz = h̄ωH β + + . (91)
2 2μL2

where β can be obtained by numerical calculations.

Orbital Current

On the basis of the results obtained above, one can calculate orbital and spin
magnetic moment current densities for the discussed system [62]:

j Tot = j Orb + j SMM . (92)

where the first term represents the orbital and the last one the spin magnetic moment
current.
According to the theory of quantum mechanics, the expression of orbital current
density of a charge carrier in the magnetic field has the following form:
154 D. A. Baghdasaryan et al.

ieh̄  ∗  e2
j Orb =
Ψ ∇Ψ − Ψ ∗ ∇Ψ − |2 .
A|Ψ (93)
2μ μc

and the spin magnetic moment current density as follows [62]:



j SMM = μ∗ c rot Ψ ∗ σ Ψ . (94)

By direct calculations it could be shown that


   
j Orb = 0, j Orb = 0,
 ρ  z  (95)
e2 H
jOrb = μρ − 2μc ρ |ψ(ρ, ϕ)|2 |g (z)|2 ,
eh̄m
ϕ

where ψ(ρ, ϕ) and g (z) are the two components of Ψ wave function dependent on
correspondingly ρ, ϕ, and z variables. The orbital current has two components. If
we consider that h̄m is the orbital momentum characterizing the state with quantum
number m, then it is clear that the first term in formula (95) characterizes the current
conditioned by the orbital motion. In its turn eH μc is the cyclotron frequency of the
electron. The multiplication of cyclotron frequency and ρ characterizes the linear
velocity of cyclotron movement of the electron. Therefore, the second term is the
cyclotron frequency current and describes the contribution of the magnetic field
in j Orb .
2H √
As we can see from (95), when eμρ h̄m
= e2μc ρ correspondingly ρm = 2maH and
when radial coordinate of the electron has this value, the orbital current in nanolayer
becomes zero. Let us discuss it in details. As far as we consider electron, let us
write this equation as − |e|μρ
h̄m 2H
= e2μc ρ. For this equation one of the following two
conditions should be satisfied:

→ −

1. m > 0 and H ↑↓ Oz, which means that H has the opposite direction with
respect to Oz axis


2. m < 0 and H ↑↑ Oz, which, respectively, means that has the same direction as
Oz axis
It is important to notice that for each fixed value − of m, there is its own fixed

value of ρm . For a given value of magnetic field H on distances equal to ρm =

2maH , orbital current equals to zero, as it was mentioned later, and when the
radial coordinate of the electron is larger than this value, we have oppositely directed
orbital current. Physically this phenomenon could be explained as follows. With
increasing ρ the contribution of energy caused by angular momentum decreases,
but at the same time, contribution of energy of cyclotron rotation increases. For the
radiuses larger than ρm , the direction of rotation is mainly conditioned by magnetic


field H . In Tables 4 and 5, there are shown several values of ρm for different values


of H and m.
Electronic and Optical Characteristics of Core/Shell Quantum Dots 155

Table 4 Several values of ◦


H (Oe) ρ m (A )
ρm for fixed m = 1 and
different values of H −1.1 · 105 109.43
−1.5 · 105 93.71
−1.9 · 105 83.26
−2.3 · 105 75.68
−2.65 · 105 70.50

Table 5 Several values of ◦


m ρm (A)
m for fixed
ρ
1 70.50
H = 2.65 ∗ 105 Oe and
2 99.71
different values of m
3 122.12
4 141.01
5 157.65

In order to calculate orbital and spin magnetic moment current densities, first we
should find integration constants; therefore, we should use the boundary conditions
(86). Thus, for those coefficients we obtain the following expressions:
   
|m|+1 R1 2
1 F1 − β − 2 , m + 1, 2
2aH
C2 = −     C1 . (96)
U − β − |m|+1
2 , m + 1, R1 2
2 2aH

from where, by using the normalization condition, we find that

C1 =

1
* ⎡ ⎛ ⎞⎤2 .
+  
+ |m| + 1
+ R2 ⎢ ρ2   −β, m+1,
R1 2  
+
⎢ |m| − 4aH2 ⎜ 1 F1
2 2 ⎟⎥
⎜1 F 1
2aH
+ |m|+1
− β, m + 1, ρ2
−   |m|+1
− β, m + 1, ρ2 ⎟⎥ ρdρ
, ⎣ρ e ⎝ 2 2
2aH
U |m|+1
−β, m+1,
R1 2
U 2 2
2aH ⎠⎦
R1 2 2
2aH

(97)

In Fig. 20 the dependence of orbital current density on radial coordinate for the
ground state of electron (nρ = 1, m = 0) is illustrated. Let us discuss the behavior
of these current densities in detail when L = 500Å, R1 = 70Å, and R2 = 110Å.
In Fig. 20 it can be seen that the modulus of orbital current density fully repeats the
radial distribution of the electron in the nanolayer and that is natural. As far as we
consider the state when m = 0, the orbital current is negative, and its profile has the
inverted form of density of probability of radial distribution.
156 D. A. Baghdasaryan et al.

Fig. 20 Dependence of
orbital current density on
radial coordinate for quantum
cylindrical
nanolayer,

H = 104 Oe

Spin Magnetic Moment Current

Now let us turn to the investigation of spin magnetic moment current. As for the
spin magnetic moment current density calculation, first it is necessary to calculate
the components of the following vector:

σ ¯ = μ∗ Ψ ∗ σ ˆ Ψ. (98)

By taking into consideration the forms of Pauli matrices and χs wave function, with
direct calculations it can be obtained that
   
σ ¯ = σ ¯ = 0. (99)
ρ ϕ

and
 
σ ¯ = 2μ∗ sz |ψ|2 |g|2 . (100)
z
2 2
|ψ| |g| multiplication is only dependent on ρ and z coordinates; therefore
different from zero will be only the ϕ component of rot σ ¯ . For ρ, ϕ, and z
components of spin magnetic moment current density, one can obtain:
    ∂
j SMM = 0, j SMM = −2μ∗ csz |ψ|2 |g|2 ,
ρ ϕ ∂ρ
 
j SMM = 0. (101)
z
Electronic and Optical Characteristics of Core/Shell Quantum Dots 157

Finally, for the total current density, it can be written as:


   
eh̄m e2 H ∂
j tot = − ρ |ψ|2 |g|2 − 2μ∗ csz |ψ|2 |g|2 . (102)
ϕ μρ 2μc ∂ρ

These results are quite expected because in ρ and z directions, charge carrier’s
motion is limited. On the other hand, we have periodical rotation of particle around
Oz axis which creates cyclical current.
It should be noted that in contrast with the case of quantum wire, the considera-
tion of the quantization along the Oz axis leads to the vanishing of the orbital current
in the direction of magnetic field. On the other hand, both in orbital and in spin
magnetic moment current expressions, the presence of Oz direction is expressed
by the presence of |g|2 factor. Therefore, dependent on which z = const plain is
discussed the current, for the same ρ, it will have
 different values. It is important to
note that in the expression of total current j tot , the factor |g|2 characterizes the
ϕ
quantization along Oz axis. Here with

2 π nz

L z (nz − even)
2
L sin
|g| =
2
(103)
2 2 π nz z (n − odd)
L cos L z

Hence, along Oz axis of cylindrical nanolayer, there are plains where total
current equals to zero (see Fig. 21). These plains match to |g|2 factor zeros. Let
us consider two cases: one for an even value of and another for an odd one. For
example, when nz = 6, we should consider |g|2 = L2 sin2 πLnz z . Then there
will be seven flats where total current will be equal to zero. z coordinates of
these plains could be calculated from L sin L z = 0 condition. For the case
2 2 6π

of nz = 7 (the odd case), zero plains are obtained by the same calculations for
|g|2 = L2 cos2 πLnz z . These results could be illustrated as follows:
One more important thing that should be noted is by fixing m in expression
(102) one can choose the magnetic vector H so that in the total current density,
the contribution would have only the spin magnetic moment current:
 
  ∂|ψ|2
j tot = −2μ∗ csz |g|2 . (104)
ϕ ∂ρ

which is conditioned by the gradient of electron radial distribution.


In Figs. 22 and 23, spin magnetic moment and total current densities are depicted,
respectively. As we can see in Fig. 22, spin magnetic moment has a non-monotonic
μ ∗ , and
character. Such behavior is specific for the layered geometries. As far as ∇,

jSMM form a right-handed orthogonal trio, from (94) it follows that when ∇|ψ| 2

changes its direction to opposite, jSMM should change its direction to opposite.
In fact, the spin magnetic moment current is a pseudo current. Here under the spin
magnetic moment current, it must be understood imaginary currents would create
158 D. A. Baghdasaryan et al.

Fig. 21 Illustration of plane where total current is equal to zero

Fig. 22 Dependence of spin


magnetic moment current
density on radial coordinate
for quantum cylindrical


nanolayer, H = 104 Oe

Fig. 23 Dependence of total


current density on radial
coordinate for quantum
cylindrical
nanolayer,

H = 104 Oe
Electronic and Optical Characteristics of Core/Shell Quantum Dots 159

the same magnetic field distribution which is created by the spin magnetic moment
of the electron. From the physical point of view, it is obvious that for the cylindrical
nanolayer, the spin magnetic moment should create a magnetic field parallel to the
axis of the cylindrical nanolayer as the spin was directed by the axis originally. A
magnetic field parallel to Oz axis can be caused by oppositely directed currents in
a plain perpendicular to the Oz axis.

3 Conclusion

In the present chapter, we have presented some properties of core/shell QDs.


Our goal was to show the possibility of flexible manipulation of spectral optical
and spin characteristics of spherical and cylindrical core/shell QD. The results of
the theoretical investigation show that it is possible to test fundamental quantum
mechanical principles using core/shell QDs. Particular features of optical properties
such as high quantum yield of such systems allow us to consider them as promising
candidates on the role of an active element base for optoelectronic devices of the
new generation. It is clear from the main technological application of core/shell
quantum nanostructures. On the other hand, such structures allow the experimental
implementation of quantum systems that were commonly considered abstract and
aroused academic interest. An example of the above is the problem of the behavior
of the two-electron system in a thin spherical nanolayer. As it is shown in the
chapter, this system is analogous to the system of two electrons localized on the
spherical surface. It is clear that in the simplest one-particle case, this problem
is reduced to the problem of a spherical rotator. Finally, in cylindrical core/shell
QD, under certain conditions, damped single-electron currents can be generated.
Moreover, taking into account the electron spin in such systems leads to the
necessity of taking into account the specific current due to the spin magnetic
moment of the electron. At the same time in the case of orbital current absence,
the magnetic field in such a system is associated with the spin magnetic moment of
the electron.
Thus, spherical and cylindrical core/shell QDs have a wide range of applications
starting from purely applied to academic problems. Such systems continue to be in
the center of attention of specialists since they provide important information on
quantum processes in layered nanostructures.

Acknowledgments The work was performed within the framework of the state basic program
“Investigation of physical properties of quantum nanostructures with a complex geometry and
different confining potentials.”
160 D. A. Baghdasaryan et al.

References

1. Jacak, L., Wojs, A., Hawrylak, P.: Quantum Dots. Springer, Berlin (1998)
2. Chakraborty, T.: Quantum Dots. Elsevier, Amsterdam (1999)
3. Wang, Z.: Self-Assembled Quantum Dots. Springer, Berlin (2008)
4. Singha, R.K., Manna, S., Bar, R., Das, S., Ray, S.K.: Surface potential, charging and local
current transport of individual Ge quantum dots grown by molecular beam epitaxy. Appl.
Surf. Sci. 407, 418–426 (2017)
5. Blumenthal, S., Reuter, D., As, D.J.: Optical properties of Cubic GaN quantum dots grown
by molecular beam epitaxy. Phys. Stat. Sol. B 255, 1700457 (2018)
6. Chevuntulak, C., Rakpaises, T., Sridumrongsak, N., Thainoi, S., Kiravittaya, S., Nuntawong,
N., Sopitpan, S., Yordsri, V., Thanachayanont, C., Kanjanachuchai, S., Ratanathammaphan,
S., Tandaechanurat, A., Panyakeow, S.: Molecular beam epitaxial growth of interdigitated
quantum dots for heterojunction solar cells. J. Cryst. Growth 512, 159–163 (2019)
7. Liu, H., Wang, Q., Chen, J., Lie, K., Ren, X.: MOCVD growth and characterization of multi-
stacked InAs/GaAs quantum dots on misoriented Si (100) emitting near 1.3 μm. J. Cryst.
Growth 455, 168–171 (2016)
8. Yan, X., Tang, F., Wu, Y., Li, B., Zhang, X., Ren, X.: Growth of isolated InAs quantum dots on
core-shell GaAs/InP nanowire sidewalls by MOCVD. J. Cryst. Growth 468, 185–187 (2017)
9. Aleshkin, V.Y., Baidus, N.V., Dubinov, A.A., Kudryavtsev, K.E., Nekorkin, S.M., Kruglov,
A.V., Reunov, D.G.: Submonolayer InGaAs/GaAs quantum dots grown by MOCVD. Semi-
conductors 53, 1138–1142 (2019)
10. Moiseev, K.D., Parkhomenko, Y.A., Ankudinov, A.V., Gushchina, E.V., Mikhaĭlova, M.P.,
Titkov, A.N., Yakovlev, Y.P.: InSb/InAs quantum dots grown by liquid phase epitaxy. Tech.
Phys. Lett. 33, 295–298 (2008)
11. Maronchuk, I.I., Sanikovich, D.D., Velchenko, A.A.: Photoluminescence of gallium
phosphide-based nanostructures with germanium quantum dots, grown by liquid-phase
epitaxy. J. Appl. Spectrosc. 84, 880–883 (2017)
12. Wang, Y., Hu, S., Xie, H., Lin, H., Wang, C., Sun, Y., Dai, N.: Photoluminescence
investigation of type -II GaSb/GaAs quantum dots grown by liquid phase epitaxy. Infrared
Phys. Technol. 91, 68–71 (2018)
13. Prieto, J.E., Markov, I.: Second-layer nucleation in coherent Stranski-Krastanov growth of
quantum dots. Phys. Rev. B 84, 195417 (2011)
14. Soto Rodriguez, P.E., Aseev, P., Gomez, V.J., Kumar, P., UI Hassan Alvi, N., Calleja, E.,
Manuel, J.M., Morales, F.M., Jiménez, J.J., Garcia, R., Senichev, A., Lienau, C., Nötzel, R.:
Stranski-Krastanov InN/InGaN quantum dots grown directly on Si (111). Appl. Phys. Lett.
106, 023105 (2015)
15. Prieto, J.E., Markov, I.: Stranski-Krastanov mechanism of growth and the effect of misfit sign
on quantum dots nucleation. Surf. Sci. 664, 172–184 (2017)
16. Liao, Y.C., Lin, S.Y., Lee, S.C., Chia, C.T.: Spherical SiGe quantum dots prepared by thermal
evaporation. Appl. Phys. Lett. 77, 4328 (2000)
17. Surawijaya, A., Mizuta, H., Oda, S.: Observation and analysis of tunneling properties of single
spherical nanocrystalline silicon quantum dot. Jpn. J. Appl. Phys. 45, 3638 (2006)
18. Aroutiounian, V.M., Gambaryan, K.M., Harutyunyan, V.G., Soukiassian, P.G., Boeck, T.,
Schmidtbauer, J., Bansen, R.: The Ostwald ripening at nanoengineering of InAsSbP spherical
and ellipsoidal quantum dots on InAs (100) surface. J. Contemp. Phys. 48, 37–42 (2013)
19. Bayer, M., Schilling, O., Forchel, A., Reinecke, T.L., Knipp, P.A., Pagnod-Rossiaux, P.,
Goldstein, L.: Splitting of electronic levels with positive and negative angular momenta in
In0.53 Ga0.47 As/InP quantum dots by a magnetic field. Phys. Rev. B 53, 15810 (1996)
20. Huang, S., Dai, Z., Qu, F., Zhang, L., Zhu, X.: Self-assembled large-scale and cylindrical
CuInSe2 quantum dots on indium tin oxide films. Nanotechnology 13, 691–694 (2002)
Electronic and Optical Characteristics of Core/Shell Quantum Dots 161

21. Gambaryan, K.M., Aroutiounian, V.M., Boeck, T., Schulze, M., Soukiassian, P.: Strain-
induced InAsSbP islands and quantum dots grown by liquid phase epitaxy on a InAs (1 0
0) substrate. J. Phys. D: Appl. Phys. 41, 162004 (2008)
22. Gambaryan, K.M., Aroutiounian, V.M., Harutyunyan, V.G.: Nucleation features and energy
levels of type-II InAsSbP quantum dots grown on InAs(100) substrate. Appl. Phys. Lett. 101,
093103 (2012)
23. Tonkikh, A., Werner, P.: Surfactant-mediated Stranski-Krastanov islands. Phys. Stat. Sol. B
250, 1795–1798 (2013)
24. Balagula, R.M., Sofronov, A.N., Vorobjev, L.E., Firsov, D.A., Tonkikh, A.A.: Temperature
evolution of the photoexcited charge carriers dynamics in Ge/Si quantum dots. Phys E: Low-
Dimens. Syst. Nanostruct. 106, 85–89 (2019)
25. Grundmann, M., Christen, J., Ledentsov, N.N., Böhrer, J., Bimberg, D., Ruvimov, S.S.,
Werner, P., Richter, U., Gösele, U., Heydenreich, J., Ustinov, V.M., Egorov, A.Y., Zhukov,
A.E., Kopev, P.S., Alferovet Z.I.: Ultranarrow luminescence lines from single quantum dots.
Phys. Rev. Lett. 74, 4043 (1995)
26. Moroni, S.T., Chung, T.H., Juska, G., Gocalinska, A., Pelucchi, E.: Statistical study of
stacked/coupled site-controlled pyramidal quantum dots and their excitonic properties. Appl.
Phys. Lett. 111, 083103 (2017)
27. Pickering, S., Kshirsagar, A., Ruzyllo, J., Xu, J.: Patterned mist deposition of tri-colour
CdSe/ZnS quantum dot films toward RGB LED devices. Opto-Electron. Rev. 20, 148–152
(2012)
28. Kohn, W.: Cyclotron resonance and de Haas-van Alphen oscillations of an interacting electron
gas. Phys. Rev. 123, 1242–1244 (1961)
29. Maksym P.A., Chakraborty T.: Quantum dots in a magnetic field: role of electron-electron
interactions. Phys. Rev. Lett. 65, 108–111 (1990)
30. Peeters, F.M.: Magneto-optics in parabolic quantum dots. Phys. Rev. B 42, 1486–1487 (1990)
31. Govorov, A.O., Chaplik, A.V.: Magnetoabsorption at quantum points. JETP Lett. 52, 31–33
(1990)
32. Gudmundsson, V., Gerhardts, R.: Self-consistent model of magnetoplasmons in quantum dots
with nearly parabolic confinement potentials. Phys. Rev. B 43, 12098–12101 (1991)
33. Sarkisyan, H., Hayrapetyan, D., Petrosyan, L., Kazaryan, E., Sofronov, A., Balagula, R.,
Firsov, D., Vorobjev, L., Tonkikh, A.: Realization of the Kohn’s theorem in Ge/Si quantum
dots with hole gas: theory and experiment. Nanomaterials 9, 56 (2019)
34. Lorke, A., Luyken, R.J., Govorov, A.O., Kotthaus, J.P., Garcia, J.M., Petroff, P.M.: Spec-
troscopy of nanoscopic semiconductor rings. Phys. Rev. Lett. 84, 2223–2226 (2000)
35. Dabbousi, B.O., Rodriguez-Viejo, J., Mikulec, F.V., Heine, J.R., Mattoussi, H., Ober, R.,
Jensen, K.F., Bawendi, M.G.: (CdSe) ZnS core-shell quantum dots: synthesis and charac-
terization of a size series of highly luminescent nanocrystallites. J. Phys. Chem. B 101,
9463–9475 (1997)
36. Kim, S.W., Zimmer, J.P., Ohnishi, S., Tracy, J.B., Frangioni, J.V., Bawendi, M.G.: Engineer-
ing InAsxP1-x/InP/ZnSe III-V alloyed core/shell quantum dots for the near-infrared. J. Am.
Chem. Soc. 127, 10526–10532 (2005)
37. Gong, H.M., Wang, X.H., Du, Y.M., Wang, Q.Q.: Optical nonlinear absorption and refraction
of CdS and CdS-Ag core-shell quantum dots. J. Chem. Phys. 125, 024707 (2006)
38. Park, J.P., Lee, J.J., Kim, S.W.: Fabrication of GaAs, InxGa1-xAs and their ZnSe core/shell
colloidal quantum dots. J. Am. Chem. Soc. 138, 16568–16571 (2016)
39. Vasudevan, D., Gaddam, R.R., Trinchi, A., Cole, I.: Core–shell quantum dots: properties and
applications. J. Alloys. Compd. 636, 395–404 (2015)
40. Makihara, K., Ikeda, M., Fujimura, N., Yamada, K., Ohta, A., Miyazaki, S.: Electrolumi-
nescence of superatom-like Ge-core/Si-shell quantum dots by alternate field-effect-induced
carrier injection. Appl. Phys. Exp. 11, 011305 (2018)
41. Hien, N.T., Chi, T.T.K., Vinh, N.D., Van, H.T., Thanh, L.D., Do, P.V., Tuyen, V.P., Ca, N.X.:
Synthesis, characterization and the photoinduced electron-transfer energetics of CdTe/CdSe
type-II core/shell quantum dots. J. Lumin. 217, 116822 (2020)
162 D. A. Baghdasaryan et al.

42. Feng, X., Xiong, G., Zhang, X., Gao, H.: Third-order nonlinear optical susceptibilities
associated withintersubband transitions in CdSe/ZnS core–shell quantum dots. Phys. B
Condens. Matter 383, 207–212 (2006)
43. Zoheir, M., Manaselyan, A.K., Sarkisyan, H.A.: Electronic states and the Stark shift in narrow
band InSb quantum spherical layer. Phys. E: Low-Dimens. Syst. Nanostruct. 40, 2945–2949
(2008)
44. Zoheir, M., Manaselyan, A.K., Sarkisyan, H.A.: Magneto- and electroabsorption in narrow-
gap InSb cylindrical layer quantum dot. Phys. E: Low-Dimens. Syst. Nanostruct. 41, 1583–
1590 (2009)
45. Kirakosyan, A.A., Kazaryan, E.M., Mughnetsyan, V.N., Sarkisyan, H.A.: Tunability of
absorption threshold frequencies and Stark shift in the InSb narrow gap spherical quantum
layer. Semicond. Sci. Technol. 27, 085003 (2012)
46. Halonen, V., Pietiläinen, P., Chakraborty, T.: Optical-absorption spectra of quantum dots and
rings with a repulsive scattering centre. EPL 33, 377–382 (1996)
47. Niemelä, K., Pietiläinen, P., Hyvönen, P., Chakraborty, T.: Fractional oscillations of electronic
states in a quantum ring. EPL 36, 533–538 (1996)
48. Simonin, J., Proetto, C.R., Barticevic, Z., Fuster, G.: Single-particle electronic spectra of
quantum rings: a comparative study. Phys. Rev. B 70, 205305 (2004)
49. Hayrapetyan, D.B., Kazaryan, E.M., Petrosyan, L.S., Sarkisyan, H.A.: Core/shell/shell
spherical quantum dot with Kratzer confining potential: Impurity states and electrostatic
multipoles. Phys. E: Low-Dimens. Syst. Nanostruct. 66, 7–12 (2015)
50. Hayrapetyan, D.B., Amirkhanyan, S.M., Kazaryan, E.M., Sarkisyan, H.A.: Effect of hydro-
static pressure on diamagnetic susceptibility of hydrogenic donor impurity in core/shell/shell
spherical quantum dot with Kratzer confining potential. Phys. E: Low-Dimens. Syst. Nanos-
truct. 84, 367–371 (2016)
51. Li, J., Wang, L.W.: First principle study of core/shell structure quantum dots. Appl. Phys.
Lett. 84, 3648–3650 (2004)
52. Harutyunyan, V.A.: Optical transitions in semiconductor nanospherical layer under the
presence of perturbating electrical field. Phys. E: Low-Dimens. Syst. Nanostruct. 39, 37–49
(2007)
53. Bartnik, A.C., Wise, F.W., Kigel, A., Lifshitz, E.: Electronic structure of PbSe/PbS core-shell
quantum dots. Phys. Rev. B 75, 245424 (2007)
54. Holmstrom, P., Thylen, L.: Electro-optic switch based on near-field-coupled quantum dots.
Appl. Phys. A 115, 1093–1101(2014)
55. El Aouami, A., Feddi, E., Talbi, A., Dujardin, F., Duque, C.A.: Electronic state and
photoionization cross section of a single dopant in GaN/InGaN core/shell quantum dot under
magnetic field and hydrostatic pressure. Appl. Phys. A 124, 442 (2018)
56. Gill, R., Bahshi, L., Freeman, R., Willner, I.: Optical detection of glucose and acetylcholine
esterase inhibitors by H2 O2 -sensitive CdSe/ZnS quantum dots. Angewandte Chemie. 120,
1700–1703 (2008)
57. Tomasulo, M., Yildiz, I., Kaanumalle, S.L., Raymo, F.M.: PH-sensitive ligand for luminescent
quantum dots. Langmuir 22(24), 10284–10290 (2006)
58. Hu, S.H., Liu, T.Y., Huang, H.Y., Liu, D.M., Chen, S.Y.: Magnetic-sensitive silica
nanospheres for controlled drug release. Langmuir 24(1), 239–244 (2008)
59. Surana, K., Salisu, I.T., Mehra, R.M., Bhattacharya, B.: A simple synthesis route of low-
temperature CdSe-CdS core-shell quantum dots and its application in the solar cell. Opt.
Mater. 82, 135–140 (2018)
60. Kumar, B.G., Sadeghi, S., Melikov, R., Aria, M.M., Jalali, H.B., Ow-Yang, C.W., Nizamoglu,
S.: Structural control of InP/ZnS core/shell quantum dots enables high-quality white LEDs.
Nanotechnology 29(34), 345605 (2018)
61. Haipeng, L., Carroll, G.M., Neale, N.R., Beard, M.C.: Infrared quantum dots: progress,
challenges, and opportunities. ACS Nano 13(2), 939–953 (2019)
62. Landau, L.D., Lifshitz E.M.: Quantum Mechanics, Non–Relativistic Theory, vol. 3. Pergamon
Press, London (1965)
Electronic and Optical Characteristics of Core/Shell Quantum Dots 163

63. Atayan, A.K., Kazaryan, E.M., Meliksetyan, A.V., Sarkisyan, H.A.: Magnetoexcitonic states
in a quantum ring with the Winternitz-Smorodinsky confinement potential. J. Contemp. Phys.
45(3), 126–131 (2010)
64. Atayan, A.K., Kazaryan, E.M., Meliksetyan, A.V., Sarkisyan, H.A.: Interband magnetoab-
sorption in cylindrical quantum layer with Smorodinsky-Winternitz confinement potential. J.
Comput. Theor. Nanosci. 7(6), 1165–1171 (2010)
65. Baghramyan, H.M., Barseghyan, M.G., Kirakosyan, A.A., Restrepo, R.L., Mora-Ramos,
M.E., Duque, C.A.: Donor impurity-related linear and nonlinear optical absorption coeffi-
cients in concentric double quantum rings: effects of geometry, hydrostatic pressure, and
aluminum concentration. J. Lumin. 145, 676–683 (2014)
66. Liu, G., Guo, K.: Linear and nonlinear intersubband optical absorption coefficients and
refractive index changes with the ring-shaped non-spherical oscillator potential. Superlattice.
Microst. 52(5), 997–1009 (2012)
67. Li, B., Liu, Y.-H., Liu, J.-J.: Donor impurity states in elliptical quantum rings subjected to a
magnetic field. Phys. Lett. A 375(8), 1205–1208 (2011)
68. Baghdasaryan, D.A., Hayrapetyan, D.B., Kazaryan, E.M.: Optical properties of narrow band
prolate ellipsoidal quantum layers ensemble. J. Nanophotonics 10(3), 033508–033508 (2016)
69. Flugge, S.: Practical Quantum Mechanics: Part 1. Springer, Berlin (1971)
70. Aghekyan, N.G., Kazaryan, E.M., Petrosyan, L.S., Sarkisyan, H.A.: Two electronic states in
a quantum ring: Mathieu equation approach. J. Phys. Conf. Ser. 248, 012048 (2010)
71. Galitskii, V.M., Karnakov, B.M., Kogan, V.I.: Problems in Quantum Mechanics. Nauka,
Moscow (1981)
72. Seidl, M.: Adiabatic connection in density-functional theory: two electrons on the surface of
a sphere. Phys. Rev. A 75, 062506 (2007)
73. Loos, P., Gill, P.: Ground state of two electrons on a sphere. Phys. Rev. A 79, 062517 (2009)
74. Loos, P., Gill, P.: Correlation energy of two electrons in the high-density limit. J. Chem. Phys.
131, 241101 (2009)
75. Loos, P.: Hooke’s law correlation in two-electron systems. Phys. Rev. A. 81, 032510 (2010)
76. Alavi, A.: Two interacting electrons in a box: an exact diagonalization study. J. Chem. Phys.
113, 7735 (2000)
77. Thompson, D.C., Alavi, A.: Two interacting electrons in a spherical box: an exact diagonal-
ization study. Phys. Rev. B 66, 235118 (2002)
78. Thompson, D.C., Alavi, A.: A comparison of Hartree-Fock and exact diagonalization
solutions for a model two-electron system. J. Chem. Phys. 122, 124107 (2005)
79. 29 Thompson, D.C., Alavi, A.: Electron correlation in a hard spherical external potential:
Wigner molecule formation and hybridization. Phys. Rev. B 69, 201302/1–201302/4 (2004)
80. Loos, P., Gill, P.: Correlation energy of two electrons in a ball. J. Chem. Phys. 132, 234111/1–
234111/6 (2010)
81. Loos, P., Gill, P.: Ground state of two electrons on concentric spheres. Phys. Rev. A 81,
052510/1–052510/9 (2010)
82. Loos, P., Gill, P.: Two electrons on a hypersphere: a quasiexactly solvable model. Phys. Rev.
Lett. 103, 123008/1–123008/4 (2009)
83. Loos, P., Gill, P.: Excited states of spherium. Mol. Phys. 108, 2527–2532 (2010)
84. Loos, P., Gill, P.: A tale of two electrons: correlation at high density. Chem. Phys. Lett. 500,
1–8 (2010)
85. Ezra, G.S., Berry, R.S.: Correlation of two particles on a sphere. Phys. Rev. A 25, 1513–1527
(1982)
86. Ezra, G.S., Berry, R.S.: The quantum states of two particles on concentric spheres. Phys. Rev.
A 28, 1989–2000 (1983)
87. Ojha, P.C., Berry, R.S.: Angular correlation of two electrons on a sphere. Phys. Rev. A 36,
1575–1585 (1987)
88. Hinde, R.J., Berry, R.S.: Correlation of two weakly attractive particles on a sphere. Phys. Rev.
A 42, 2259–2266 (1990)
89. Warner, J.W., Berry, R.S.: Hund’s rule. Nature 313, 160 (1985)
164 D. A. Baghdasaryan et al.

90. Seidl, M., Perdew, J.P., Kurth, S.: Simulation of all-order density-functional perturbation
theory, using the second order and the strong-correlation limit. Phys. Rev. Lett. 84, 5070–
5073 (2000)
91. Seidl, M.: Strong-interaction limit of density-functional theory. Phys. Rev. A 60, 4387–4395
(1999)
92. van Roosbroeck, W., Shockley, W.: Photon-radiative recombination of electrons and holes in
germanium. Phys. Rev. 94(6), 1558–1560 (1954)
93. Bhattacharya, R., Pal, B., Bansal, B.: On conversion of luminescence into absorption and the
van Roosbroeck-Shockley relation. Appl. Phys. Lett. 100(22), 222103 (2012)
94. Miralaie, M., Leilaeioun, M., Abbasian, K., Hasani, M.: Modeling and analysis of room-
temperature silicon quantum dot-based single-electron transistor logic gates. J. Comput.
Theor. Nanosci. 11, 15–24 (2014)
95. Yuan, M., Pan, F., Yang, Z., Gilheart, T.J., Chen, F., Savage, D.E., Lagally, M.G., Eriksson,
M.A., Rimberg, A.J.: Si/SiGe quantum dot with superconducting single-electron transistor
charge sensor. Appl. Phys. Lett. 98, 142104 (2011)
96. Lewis, E.A., Page, R.C., Binks, D.J., Pennycook, T.J., O’Brien, P., Haigh, S.J.: Probing the
core-shell-shell structure of CdSe/CdTe/CdS type II quantum dots for solar cell applications.
J. Phys.: Conf. Ser. 522, 012069 (2014)
97. Pietiläinen, P., Chakraborty, T.: Interacting-electron states and the persistent current in a
quantum ring. Solid State Commun. 87, 809–812 (1993)
98. Chakraborty, T., Pietilainen, P.: Electron-electron interaction and the persistent current in a
quantum ring. Phys. Rev. B 50, 8460–8468 (1994)
99. Niţă, M., Marinescu, D.C., Manolescu, A., Ostahie, B., Gudmundsson, V.: Persistent oscilla-
tory currents in a 1D ring with Rashba and Dresselhaus spin-orbit interactions excited by a
terahertz pulse. Phys. E: Low-Dimens. Syst. Nanostruct. 46, 12–20 (2012)
100. Castelano, L.K., Hai, G.Q., Partoens, B., Peeters, F.M.: Control of the persistent currents in
two interacting quantum rings through the Coulomb interaction and interring tunneling. Phys.
Rev. B 77, 235314 (2008)
101. Niţă, M., Marinescu, D.C., Manolescu, A., Gudmundsson, V.: Nonadiabatic generation of a
pure spin current in a one-dimensional quantum ring with spin-orbit interaction. Phys. Rev. B
83, 155427 (2011)
102. Bellucci, S., Onorato, P.: Quantum rings with tunnel barriers in a threading magnetic field:
spectra, persistent current and ballistic conductance. Phys. E: Low-Dimens. Syst. Nanostruct.
41, 1393–1402 (2009)
103. Mita, K.: Virtual probability current associated with the spin. Am. J. Phys. 68, 259–264 (2000)
Exciton–Phonon Interactions
and Temperature Behavior of Optical
Spectra in Core/Shell InP/ZnS Quantum
Dots

Sergey Savchenko, Alexander Vokhmintsev, and Ilya Weinstein

Abstract The chapter summarizes research results on temperature evolution of


optical absorption spectra in a wide temperature range of 6.5–296 K for colloidal
core/shell InP/ZnS quantum dot ensembles with a different size distribution.
Exciton–optical transition energies in the InP core and ZnS shell were determined
by the second-order derivative spectrophotometry. Numerical analysis within the
frame of linear exciton–phonon coupling showed that the shifts of the observed
spectral maxima with temperature were due to the interaction of excited states with
longitudinal acoustic vibrations. It was established that the half-width of the spectral
bands remained unchanged, which indicated the inhomogeneous broadening effect
and a high degree of static disorder in the ensembles under study. The semi-
phenomenological model was proposed that took into account the effects of the
exciton–phonon interaction and allowed one to analyze the influence of the static
and dynamic types of atomic disorder on behavior of optical absorption spectra in
InP/ZnS quantum dots under temperature variation.

Keywords Quantum dot · Core/shell · Optical absorption · Derivative


spectrophotometry · Half-width · Ensemble · Temperature dependence · Size
distribution · Semiconductor nanocrystal · Static and dynamic disorder ·
Inhomogeneous broadening · Exciton-phonon coupling

1 Introduction

Colloidal quantum dots (QDs) or semiconductor nanocrystals provide an interesting


opportunity to observe the evolution of material properties depending on dimen-
sional factors [1–3]. As a rule, the phenomenon mentioned above is caused by the
quantum confinement of elementary excitations over all three spatial dimensions in

S. Savchenko · A. Vokhmintsev · I. Weinstein ()


NANOTECH Centre, Ural Federal University, Ekaterinburg, Russia
e-mail: s.s.savchenko@urfu.ru; a.s.vokhmintsev@urfu.ru; i.a.weinstein@urfu.ru

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 165
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_5
166 S. Savchenko et al.

zero-dimensional objects. In this regard, QDs have a size-tunable optical bandgap,


which, in turn, enables to dramatically enhance the efficiency of luminescent
processes, as compared to the bulk state. In addition, QDs are promising candidates
in all applications related to absorption, emission, and conversion of visible light.
Over a short historical period, a number of research works have demonstrated QDs’
prospects in the development of modern light-emitting diodes, lasers, displays,
phosphors, sensors, solar cells, photocatalysts, luminescent labels, etc. [4–14].
QDs based on compounds of III–V groups are of particular interest because they
do not contain hazardous chemical elements such as Cd, Pb, and Hg. Therefore,
they can be widely used as biosensors and biomarkers and in various biocompatible
applications [15]. InP-based nanocrystals are regarded as an alternative to Cd-
based QDs, which are a reference in the context of optical characteristics. InP/ZnS
QDs are nontoxic and, after surface passivation with a ZnS shell, possess effective
luminescence. The latter can be adjusted in the entire visible range by nanocrystal
resizing [4, 16, 17]. One of the key parameters is also the half-width of the optical
bands, which largely determines the possibility of using QDs.
Despite intensive investigations in the field of the synthesis of InP/ZnS-based
nanocrystals with specified characteristics, the nature and mechanisms of the
formation of their optical properties are poorly understood [18–20]. In general,
in zero-dimensional structures, inhomogeneities and effects inherent in molecular
systems have a significant effect. This results in broadening the bands of discrete
exciton transitions relative to a homogeneous half-width value and their overlapping
and complicating spectra [21, 22]. Among the indicated inhomogeneities are size,
shape, stoichiometry, imperfection, local surrounding, charge distribution, etc. of
nanocrystals in the ensemble [21–27]. These factors affect the transition energy in
individual nanocrystals of the ensemble, causing an inhomogeneous broadening of
the observed experimental optical absorption (OA) and luminescence bands. Thus,
an analysis of the homogeneous and inhomogeneous contributions to the observed
broadening requires a deeper insight into the relationship between the structural
characteristics and the optical properties of such systems. Studying the temperature
behavior of experimental spectral features gains fundamental quantitative informa-
tion on the parameters of the exciton–phonon interaction.
As of today, there are research works on the temperature change in the pho-
toluminescence (PL) spectra of semiconductor InP/ZnS nanocrystals [28–30]. This
chapter deals with the behavior of the optical absorption (OA) spectra for ensembles
of InP/ZnS core/shell quantum dots with different average particle sizes in the
temperature range of 6.5–296 K.

2 Nanocrystal Absorption and Derivative Spectrophotometry

The electronic structure of semiconductor nanocrystals is manifested in their


optical absorption spectra. We have explored two types of InP/ZnS nanocrystals
with different average particle sizes in ensembles, which in what follows will be
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 167

Fig. 1 The investigated


solutions of QD-1 and QD-2
InP/ZnS under room light

Table 1 Characteristics of Parameter QD-1 QD-2


QD-1 and QD-2 InP/ZnS
ensembles Concentration (g/l) 40 48.4
Average radius (nm) 2.1 2.3
PL quantum yield (%) 27 10
PL band maximum (nm) 534 578
PL band half-width (nm) 65 95

designated as QD-1 and QD-2. They were produced by the Research Institute of
Applied Acoustics in Dubna. As the manufacturer claims, the initial nanocrystal
samples as aqueous colloidal solutions had a three-layer structure: an InP core,
a ZnS shell, and a modified polyacrylic acid coating. Figure 1 displays their
appearance under room light. The characteristics of the ensembles are given in
Table 1.
The InP/ZnS OA spectra measurements were taken at room temperature using a
Shimadzu UV-2450 spectrophotometer that operates under the traditional double-
beam scheme of measuring the optical density D = log (I0 /I), where I0 and I are
the intensities of optical radiation incident on and passing through the sample,
respectively. During the measurements, the spectral width of the slit was 2 nm, and
the sampling interval was 0.5 nm. Absorption was studied for QDs in the form of
solutions and films. The solutions were placed in a quartz cuvette with an optical
path length of 1 cm. Concentration series were prepared by successively diluting the
solutions with distilled water in the 1:1 volume ratio in the range 0.04–40 g/l. The
films were made of colloidal solutions deposited physically on a 1-mm-thick quartz
substrate at room temperature.
Figure 2 shows the OA spectra of QD-1 (a) and QD-2 (b) quantum dots for
the colloidal solutions of various concentrations (curves 1–7) and films (curves
8, 9). As can be seen in the figure, the optical density drops monotonically with
decreasing concentration, and various parts of the spectrum fall into the range of
168 S. Savchenko et al.

Fig. 2 OA spectra of QD solutions of various concentrations and films precipitated. (a) QD-1.
Solution: 1–40 g/l, 2–10 g/l, 3–5 g/l, 4–2.5 g/l, 5–0.63 g/l, 6–0.16 g/l, 7–0.04 g/l. Film: 8–40 g/l,
9–20 g/l. (b) QD-2. Solution: 1–40 g/l, 2–20 g/l, 3–10 g/l,4–5 g/l, 5–2.5 g/l, 6–0.31 g/l, 7–0.04 g/l.
Film: 8–48.4 g/l

infallible values measured by the instrument, D = 0.1–3. This allows investigating


the samples’ absorption in a wide range of wavelengths. The optical absorption of
the QD-1 and QD-2 samples grows in different spectral ranges. At a concentration of
40 g/l, the QD-1 and QD-2 solutions have an optical density of 0.1 at a wavelength of
560 nm and 600 nm, respectively. As the QD-1 concentration changes, the spectrum
exhibits a maximum at 450–480 nm (curves 2–4), while a shoulder can be revealed
in the range of 240–270 nm (curves 5–7). In the case of QD-2, shoulders can be
seen in the intervals of 490–570 nm, 290–330 nm, and 250–280 nm (curves 1–3).
In Fig. 2a, curves 8 and 9 correspond to the absorption of QD-1 films produced
using the solutions with concentrations of 40 and 20 g/l, respectively. Curve 8 in
Fig. 2b corresponds to the absorption of a QD-2 film obtained by precipitating
the solution of the initial concentration. The beginning of the absorption range for
the films demonstrates a feature in the same wavelength intervals and so does for
the solutions. However, with a decrease in the wavelength, the picture differs from
absorption of the solutions. For the QD-1 and QD-2 films, one spectral shoulder is
observed in the wavelength range of 260–340 nm.
It should be underscored that the shape of the long-wavelength part of the spec-
trum where the growth of light absorption begins is significantly different for the
samples. For the QD-1, a band with a pronounced maximum is observed, while the
QD-2 spectra have a shoulder in this region. The second case appears to be explained
by either a wider particle-size distribution or the implementation of various quantum
confinement regimes in the samples. The concentration dependencies of D for fixed
wavelengths are linear. This indicates no deviations from the Beer–Lambert law in
the above concentration range and evidences the stability of the solutions made.
An idealized theoretical description predicts a discrete, atomic-like absorption
spectrum of quantum dots. It is formed by transitions between the quantized energy
levels of the nanocrystals. The absorption of the real samples is a continuous
spectrum with maxima and/or shoulders. This is due to the fact that the absorption
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 169

bands have a certain width; they mutually overlap and are distorted by the presence
of background. In the general case, the latter factor is a certain monotonically
increasing function and can be related to the contribution of the corresponding
bulk density of states, scattering processes, high-energy transitions, etc. [16, 22,
31]. In this regard, analytical approaches of the derivative spectrophotometry can
be utilized for characterizing optical transitions and determining the positions of
absorption bands [32].
Derivative spectrophotometry is an effective method widely used for resolving
slightly marked spectral features. It makes it possible to explicitly highlight hidden
absorption bands by reducing their width and eliminating the unwanted influence of
the background signal. The method consists in calculating various-order derivatives
of the absorption spectra using hardware or numerical methods. The latter are
currently dominating owing to the instrument reading digitalization. In the case of
an ideal analytical band, the zeros of the first and all odd derivatives, as well as
the positions of the extrema on the second and all even derivatives, correspond to
the location of the absorption bands on the initial spectral curve. The regularities
specified above underlie the analysis of absorption spectra conditional upon the
mutual overlapping of the analytical signals and superimposition of the background.
Some noise level is inherent in real spectra. This raises the need to smooth
the experimental data before calculating the derivative since the signal-to-noise
ratio decreases inversely proportional to Hn as the order of the derivative n
increments [33]. To decline the influence of the noise, the Savitzky–Golay algorithm
was applied for calculating the derivatives of the spectrum [34, 35]. It includes
constructing an approximating polynomial in the neighborhood of each measure-
ment point using the least-squares method. This approach enables the data to be
smoothed for computing various-order derivatives without distorting the shape of
the spectrum. For calculating the derivative, the size of the smoothing window is
set maximum to save all the spectral features of the experimental absorption data.
A too large window’s value may distort the initial shape of the spectrum and, as a
consequence, the absorption band location can be inadequately estimated. As a rule,
increasing the averaging interval enhances the smoothing, but the peaks broaden
and diminish in height. Otherwise speaking, to preserve the fine structure of the
spectrum, a compromise value of the smoothing window should provide the desired
level of high-frequency noise filtering. The OA spectra measured for the InP/ZnS
nanocrystals were processed using the Savitzky–Golay second-order differentiating
filter with an approximating polynomial of the second degree. The minima of the
spectrum derivatives, obtained in such a way, corresponded to the position of the
absorption bands.
Figure 3 presents the OA spectra of films and solutions of the indicated
concentrations, as well as their second derivatives d2 D/dE2 . The estimates of the
energies of optical transitions at room temperature are shown by arrows. The values
are summarized in Table 2. It is worth emphasizing that the predicted energy
values do not change for colloidal solutions of QDs of various concentrations. This
statement is another confirmation of the stability of the concentration series at hand
and stability of the spectral processing technique applied. The spectra of film-shaped
170 S. Savchenko et al.

Fig. 3 OA spectra (solid lines) and their second derivatives (dashed lines) for solutions and films
of QD-1 (a) and QD-2 (b). The arrows denote the energies of optical transitions that correspond to
the minima of the d2 D/dE2 curves

Table 2 The energies of the Energy (eV)


InP/ZnS optical absorption
Sample E1 E2 E3
bands at room temperature
QD-1 Solution 2.60 4.70
Film 2.60 4.12
QD-2 Solution 2.38 4.06 4.68
Film 2.37 4.17

QDs demonstrate the same energy value for the first transition E1 and so do for
the solutions. In the higher-energy range, the transition with an energy of 4.12 eV
is typical of the QD-1 samples, whereas the transition with an energy of 4.17 eV
is observed for the QD-2 samples (see Fig. 3). The difference in the transition
energies in the solution and on the substrate can be associated with the effects of
the interaction of nanocrystals since a change in the distance between them during
deposition can lead to a modification of their energy structure [36].
The E1 transition is associated with the first exciton absorption band of indium
phosphide. For QDs, it is also regarded as either the bandgap Eg or the transition
between states corresponding to the top of the valence band and the bottom of
the conduction band. The paper [37] theoretically calculates the spectra of exciton
transitions for InP QDs with diameters of 2.02, 2.37, 2.80, and 3.48 nm within
the atomistic pseudopotential approach. The first exciton transition (s → s) for
the indicated sizes is 2.51, 2.34, 2.20, and 1.98 eV, respectively. The next energy
transition corresponds to the excitation of an s-like state in the valence band into an
s-like state of the minimum of the conduction band. The s-like state in the valence
band is caused by spin–orbit splitting. Between these transitions, there is a size-
independent energy interval of ~0.11 eV. Higher-energy transitions separated from
Eg by energies up to 0.7 eV and corresponding to p-like states are also calculated.
Moreover, there is an energy gap between states with different symmetries. It
increases with decreasing the size from 0.3 eV to 0.4 eV. Given the fact that the
model offers no oscillations and exchange interactions in the system, the calculated
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 171

findings are in good agreement with the experimentally observed absorption features
for the QD-1 and QD-2.
Energy E1 has a significant blue shift relative to the bandgap Eg (bulk) = 1.34 eV
for bulk InP at T = 290 K [38]. This is due to the influence of the quantum
confinement effect. The position of the first exciton absorption band of InP can
be exploited for evaluating the size of nanocrystals. To date, a number of theoretical
models have been proposed to describe the dependence of the bandgap of QDs on
the size of Eg (a) and to account for various quantum-mechanical approaches and
corrections [2, 3, 21, 39–43]. We used the well-known ratio that was derived in the
approximation of the effective mass and can be written in the following analytical
form [21]:
 a 2 aB
B
Eg (a) = Eg (bulk) + π2 Ry − 1.786 Ry − 0.248Ry, (1)
a a
where aB = 10.1 nm is the Bohr exciton radius, Ry = 5.8 meV is the Rydberg
exciton energy for InP, and a is the nanocrystal radius, nm. The average particle
radius calculated in this way was 2.1 nm for QD-1 and 2.3 nm for QD-2.
The values of E2 and E3 exceed the value of Eg for bulk ZnS, equal to 3.6 eV at
T = 290 K [38]. Accordingly, the above transitions can have higher energy due to
quantum effects in the shell of InP nanocrystals. This assumption is confirmed by
published data on the synthesis of nanopowders and ZnS colloids, where Eg values
are reported in the energy ranges of 3.71–4.3 [44, 45] and 4.47–4.82 [46] eV. In
addition, in QDs, the oscillator strength for the first optical absorption transitions
that are characterized by the lowest energy and s-like symmetry has the greatest
value [26, 31, 47]. In this case, the observed features are located noticeably higher
in energy than the first exciton transitions for InP. Thus, the absorption bands under
discussion can be referred to the ZnS shell.

3 Temperature Dependence of the Energy Gap

One of the main constants of semiconductor materials is the bandgap Eg . With a


decrease in dimension, this characteristic stops uniquely identifying the material and
becomes size-dependent. A top-down theoretical analysis of the energy structure
of nanocrystals is carried out within the effective mass approximation, and the
elementary excitation of their electronic subsystem is called an exciton. However,
excitons in both QDs and bulk semiconductors are not identical to each other. In the
case of a weak quantum confinement, when a >> aB , quantization of the motion of
the center of mass of the exciton occurs. The energy spectrum of such a quasiparticle
resembles a hydrogen-like one but does not coincide with it. In the limit of the strong
quantum confinement when aB >> a, it is implied that an electron and a hole never
form bound hydrogen-like states. Therefore, their motion in the first approximation
is separately analyzed [21]. Thus, in spite of regarding elementary excitation at a
172 S. Savchenko et al.

quantum dot as an exciton, the latter itself does not reflect the term accepted for bulk
semiconductors [48]. In this connection, the bandgap for QDs corresponds to the
first-transition energy in the exciton absorption spectrum. It should be stressed that
the terms such as electron–phonon interaction and exciton–phonon interaction are
used with respect to QDs as interchangeable ones due to the specifics of elementary
excitations in zero-dimensional objects.
At quantum dots, along with the size, the temperature factor affects the bandgap.
In this case, an important characteristic is the temperature coefficient β = dEg /dT.
For most materials, the latter is thought to be a constant with a negative value [49].
Let us look into the existing relations to describe the dependence Eg (T).
The simplest case describes the temperature change in the bandgap using a linear
model:

Eg (T ) = Eg (0) − βT . (2)

Within this representation, the magnitude of β is determined by the slope of the


linear part of the experimental temperature dependence Eg (T). However, at low
temperatures, it can be essentially nonlinear.
The linear–quadratic relation proposed by Varshni and widely used for describing
Eg (T) appears as [50]:

α1 T 2
Eg (T ) = Eg (0) − , (3)
α2 + T

where α 1 and α 2 are empirical parameters that have no specific physical meaning.
The α 2 constant having the dimension [α 2 ] = [T] is assumed to be close in mag-
nitude to the Debye temperature. In the limit of high temperatures, when T >> α 2 ,
it follows from Eq. (3) that α 1 ≈ β. In some cases, the α 1 and α 2 coefficients are
negative. Therefore, the physical interpretation of the recorded dependencies meets
certain difficulties. Nevertheless, despite a shortage of information extracted, Eq.
(3) quite satisfactorily describes the experimentally observed shape of the Eg (T)
temperature characteristic, which was tested on a large number of different objects
[49].
Within the single-phonon approximation and second-order perturbation theory,
the temperature dependence of the bandgap can be represented as the relation [51]:
 −1
Eg (T ) = Eg (0) − AF n , where n = exp (ω/kT ) − 1 . (4)

Here AF is the Fan parameter, depending on the microscopic properties of the


material, eV; n is the Bose–Einstein factor for phonons with an average energy
ω; k is the Boltzmann constant, eV/K. Earlier, it was shown in [49] that Eq. (4)
reduces to the form of Eq. (2) in the limit of high temperatures (kT >> -hω). In the
process, the temperature coefficient can be written as
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 173

k
β∞ = AF . (5)

It should be pointed out that Eq. (4) explicitly accounts for no contribution of
thermal dilation of the lattice. As shown in [51, 52], this contribution to the total
temperature change in the bandgap amounts to a magnitude of the order of 20%, and
it can be neglected in the first approximation. In addition, at higher temperatures,
the contribution of thermal expansion to the shift in energy levels is expected to be
also proportional to n [52]. In this case, the calculated value of the AF parameter
takes into account both internal (electron–phonon interaction) and external (thermal
expansion) contributions to the Eg (T) dependence.
Note that the AF parameter has an energy dimension and coincides in magnitude
with a change in the bandgap at a temperature when the average number of phonons
responsible for the displacement of the energy levels of the band edges is unity. The
microscopic expression for the Fan parameter was written in [53] in the following
way:
⎡ 3 3 ⎤
e2 13 1  1   1
1 ⎢ me 2
 1
mh 2⎥
AF = √ (m0 ω) 2 − ⎣ + ⎦. (6)
2 4πε ε∞ ε0 m0 m0

Here e is the electron charge; ε is the dielectric constant; ε0 and ε∞ are the static
and high-frequency permittivities, respectively; m0 is the mass of a free electron; me
and mh are the effective masses of an electron and a hole, respectively. Equation (6),
in which all quantities are tabular, can be used for estimating the fan parameter in
the case of the lack of experimental data on the dependence Eg (T).
The paper [49] demonstrated the relationship between the Varshni relation (3)
and the Fan expression (4). Expanding in a series the right-hand side of Eq. (4) in
the limit of kT >> -hω to the quadratic terms in temperature, we can arrive at an
expression identical to Eq. (3) with the following coefficients:

k ω
α1 = AF · , α2 = , AF = 2α1 α2 . (7)
ω 2k
Consequently, if the condition T >> 2α 2 holds, the Varshni coefficients α 1 and α 2
should contain information on the effective phonon energy.
In [54], a semiempirical relation was proposed, the terms of which explicitly take
into account effects of both thermal expansion and electron–phonon interaction:
4   5

Eg (T ) = Eg (0) − U1 T U2
− U3 ω coth −1 , (8)
2kT

where U1 , U2 , and U3 are temperature-independent parameters. The second term


on the right-hand side of Eq. (8) represents thermal expansion; the third term is
responsible for electron–phonon interaction. It is easy to see that the second and
174 S. Savchenko et al.

third terms in the right-hand sides of the expressions (4) and (8), respectively,
coincide, with AF = 2U3 -hω.
To approximate experimental data, the authors of [55] resorted to the relation:

Eg (T ) = a − b (1 + 2 n) , (9)

where a–b = Eg (0) and b is a parameter characterizing the force of the electron–
phonon interaction [56]. After comparing expressions (4) and (9), it can be written
that AF = 2b. The authors of [57] proposed to describe the Eg temperature
dependence of semiconductors as follows:

Eg (T ) = Eg (0) − 2Sω n , (10)

where S is the Huang–Rhys factor proportional to the force of the electron–phonon


interaction [28, 29]. Comparing Eqs. (4) and (10), we can see that AF = 2Sω. Thus, it
is obvious that expressions (4), (8), (9), and (10) are similar to each other and allow
describing the Eg temperature dependencies for the materials to obtain quantitative
information on the effective phonon energy.
The above models of the Eg (T) dependence are traditionally used to describe
the properties of both bulk materials and low-dimensional objects. For example,
papers [28, 29] applied relation (10) for delineating the temperature influence on
the maxima of the photoluminescence bands of various-size InP/ZnS nanocrystals.
Paper [56] analyzed the temperature dependencies of the energies of various optical
transitions in indium phosphide nanowires using the PL excitation spectra and
Eq. (9).
In general case, there are four temperature-dependent factors that affect the
positions of energy levels in quantum dots: dilation of the lattice, thermal expansion
of the envelope function, mechanical strain, and electron–phonon coupling [58,
59], with the electron–phonon contribution dominating both for quantum dots [58]
and for bulk materials [49]. In this regard, it seems more justified to use models
that explicitly take into account phonon statistics and make it possible to extract
information about the effective energy of phonons, the interactions with which
determine the observed shifts of energy levels.

4 The Influence of Exciton–Phonon Interaction


on the Energy of Optical Transitions

The temperature dependencies of the absorption of nanocrystals were studied for


InP/ZnS films. The OA spectra were measured at T = 6.5–296 K using a setup based
on a Shimadzu UV-2450 spectrophotometer and a Janis model CCS-100/204N
closed-cycle helium cryostat. The temperature was adjusted by a LakeShore Model
335 controller equipped with a DT-670B-CU diode temperature sensor. The vacuum
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 175

Fig. 4 The temperature dependencies of the OA of the QD-1 (a and b) and QD-2 (c and d) films
in various spectral regions corresponding to transitions E1 and E2 . The dotted line indicates the
direction of shift of the spectral features

inside the cryostat was produced using a HiCube 80 Eco turbo pump station and
maintained during measurements at a level of less than 7·10−5 mbar. The OA spectra
were recorded at fixed temperatures: 6.5 K, in the range of 10–100 K with a step of
10 K, and in the range of 100–296 K with a step of 20 K.
Figure 4 sketches the OA spectra of the QD-1 and QD-2 InP/ZnS films,
measured at various temperatures. To explore the features of the QD-1 absorption
spectrum, corresponding to the E1 and E2 transitions, two samples were prepared by
precipitating solutions with concentrations of 40 and 20 g/l. This made it possible
to tune the optical density range on the desired spectral region. Curves 8 and 9
in Fig. 2a refer to the absorption of these samples at room temperature. Figure
4a, b displays the temperature dependencies of the QD-1 absorption for the above
spectral regions. To study the QD-2 ensemble, we used a sample obtained by
precipitating a solution of the initial concentration. The resulting absorption of the
QD-2 ensemble is depicted by curve 8 in Fig. 2b. This film was characterized by
a rather low optical density D < 0.1 in the region of the first InP exciton band. To
enhance the optical density in this range, we made several attempts to raise the initial
concentration. This was done in two ways: either by evaporating or by repeatedly
depositing the solution on a substrate. However, the foregoing procedure led to an
176 S. Savchenko et al.

Fig. 5 The OA second-derivative spectra for determining the optical transition energies E1 (a), E2
(b) of QD-1 and E1 (c), E2 (d) of QD-2 at the indicated temperatures

increase in scattering only. Figure 4c, d shows the temperature evolution of the QD-
2 absorption curves in different wavelength ranges. As can be understood from the
figure, the QD-1 and QD-2 ensembles, when cooled, demonstrate spectral shifts
of the maxima and shoulders towards higher energies. Besides, the corresponding
optical density grows. The inset in Fig. 4a illustrates these changes in the first QD-1
exciton absorption band in more detail.
For determining the E1 and E2 energies of the optical transitions in QD-1
and QD-2 at various temperatures, the derivative spectrophotometry method was
used. In Fig. 5, the second-derivative spectra for absorption curves at different
temperatures serve as an example; they are designated as dotted lines. The minima
in the derivative spectra allow accurately tracking the quantitative changes in the
spectral parameters of the QD absorption bands as temperature varies [60].
In Fig. 6, square-shaped symbols represent the temperature dependencies for
the E1 and E2 optical transition energies in QDs at hand. It can be seen that
the magnitudes of E1 and E2 increase with decreasing T. Their behavior is
characteristic of the temperature-induced change in the width of the optical gap
in bulk semiconductor crystals [38, 49 and references therein]. At low temperatures,
a range of unchangeable energies is observed. As becomes clear from the figure,
temperatures of reaching this range for QD-1 are lower than those for QD-2. The
solid lines show the approximation of the experimental data using Fan relation
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 177

Fig. 6 The temperature dependencies of the E1 and E2 energies for QD-1 (a and b) and QD-
2 (c and d). Comparison of E1 shift for the QDs and bulk InP (e). Square-shaped symbols are
the experimental estimates; solid lines denote the approximation according to Fan expression (4);
dashed lines are the approximation within the linear model (2); short dashed lines indicate the
temperature dependence; dash-dotted lines denote the limit level β ∞ according to Eq. (5); round
symbols present the value of temperature coefficient β when approximated by a linear model

(4). The dashed lines represent the description within the linear model (2). As a
comparison, the obtained parameters and independent literature data are tabulated
in Table 3.
178 S. Savchenko et al.

Table 3 Parameters of the E1 (T) and E2 (T) temperature dependencies for QD-1 and QD-2
E(0) (eV) AF , eV h̄ω (meV) S β (10−4 eV/K) β ∞ (10−4 eV/K)
QD-1 E1 2.72 0.09 15 2.98 4.76 (90–296 K)5.13
0.06* [28] 13 [28] 3.83* [28]
0.14* [29] 23 [29] 5.20* [29]
E2 4.21 0.07 14 2.37 3.84 (120–330 K) 4.08
QD-2 E1 2.50 1.06 59 9.03 9.97 (260–330 K) 15.56
E2 4.19 0.27 37 3.76 5.31 (200–330 K) 6.47
InP 1.50 0.09* 21 2.20* 3.78*
nanowires [56]
Bulk InP 1.42 [38] 0.05* [64] 14* [64] 1.78* 4.6 [38] 3.06* [64]
*Our estimate using data from the cited work

Temperature behavior of the first exciton absorption band position in QD-1 is


governed by the interaction with effective phonons of 15-meV energy corresponding
to the energy of longitudinal acoustic vibrations (LA) in bulk InP [61, 62]. Overall,
the values of the -hω effective phonon energy, the AF Fan and S = AF /2-hω Huang–
Rhys parameters (see above) characterizing the E1 and E2 transitions, respectively,
are consistent with the values for InP nanowires [56] (the length is of about 1 μm,
and the diameter is of about 100 nm accordingly to the TEM image, but the authors
do not give their values) and InP/ZnS quantum dots with 2.3-nm [28] and 2.1-nm
[29] diameters, see Table 3.
In QD-2, the shift of the E1 and E2 exciton transitions is due to the interaction
with phonon modes having different effective energies [63]. The Huang–Rhys factor
S for the first InP exciton absorption band is 2.4 times larger than that for the E2
transition. This means stronger exciton–phonon interaction for the E1 band. Figure
6e compares the data as a shift E1 (T) = E1 (T) − E1 (0) for the E1 transition of the
QD-1 and QD-2 ensembles with the findings for the temperature behavior of exciton
absorption levels in InP single crystals [64].
In Fig. 6, the short dotted line shows the dependence of dE/dT for the cor-
responding optical transitions. It characterizes the change in the β-coefficient
with temperature. The dash-dotted line meets the limit value of the coefficient
calculated using Eq. (5) and designated as β ∞ . The value of β indicated by the red
circle was computed by approximating the experimental temperature dependencies
of E1 and E2 within the linear model (2). In the general case, this coefficient
depends on the interval T used for the linear approximation. This is because the
dependencies involved exhibit a nonlinear character in the low-temperature region.
The appropriate temperature ranges for the data satisfactorily described by the β-
constant coefficient are listed in Table 3. The coefficient value for QD-1 is consistent
with the reference data for bulk InP and significantly exceeds them in the case of
QD-2. The relation β ≈ β ∞ being fulfilled for QD-1, it can be argued that the
considered temperature ranges for the bands under study completely satisfy the
high-temperature condition kT >> -hω.
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 179

5 The Half-Width of the First Exciton Absorption Band

Change in the half-width of the characteristic bands reflects the temperature


influence on the optical absorption spectra of condensed matter, as well. Figure 7
illustrates the normalized OA spectra of QD-1 and QD-2 at various temperatures.
It should be noted that, with decreasing T, the low-energy edge of the exciton
absorption band shifts towards higher energies without changing the slope. In other
words, the shape of the band is temperature-independent. The insets in Fig. 7
confirm the above statement. The spectra are shifted along the energy axis for visual
clarity of their identical shape in the low-energy part at different temperatures.
In Fig. 8a, b, solid lines represent the experimental dependencies of the optical
density on the photon energy for QD-1 at various temperatures in the range studied.
It is worth pointing out that, along with the shift of the first exciton absorption

Fig. 7 Normalized OA spectra of QD-1 (a) and QD-2 (b) for different T. The insets illustrate the
identity of the shape of the first exciton absorption band in QDs. The dashed line is a Gaussian
approximation for estimating HG

Fig. 8 Temperature behavior of the first exciton absorption band of QD-1. An estimate of the Hvis
half-width (a) and HG (b) is illustrated. The dashed line shows the approximation by a Gaussian
band. The insets illustrate the changes in optical density at the maximum of the exciton band [65]
180 S. Savchenko et al.

band of the nanocrystal ensemble towards higher energies, an increase in the visible
maximum Dmax of optical density is observed when the temperature drops from
296 K to 6.5 K (see the inset in Fig. 8a). In this case, the shape of the band has no
changes. The estimation of the H half-widths was made in two ways: (1) visual Hvis
was estimated by the energy position of Dmax and (2) HG by using the Gaussian
approximation of the low-energy part of the spectrum.
The calculation of the Hvis value was carried out over the low-energy part of
the spectrum. In doing so, the width at half height was graphically determined and
multiplied by two conditionals upon that the band was said to be symmetric (see Fig.
8a). The Hvis (T) calculated dependence denoted by blue solid squares is shown in
Fig. 9. It can be seen that the value equal to Hvis = 390 ± 10 meV remained constant
within the indicated error above the entire temperature range under consideration.
As was stated earlier, a number of factors affect the position and half-width
of the bands during the optical absorption of QDs. To substantiate a quantitative
estimate of H, the low-energy wing of the exciton band was approximated by a
Gaussian function accounting for these contributions (see Fig. 8b). The centers of
the bands were set at energies E1 obtained using data of the second-order derivative
spectrophotometry [61, 63]. The inset in Fig. 8b displays the change in the optical
density DG corresponding to the maximum of the approximating Gaussian. As it
becomes clear from the insets in Fig. 8a, b, the temperature dependencies are quite
consistent with each other. The HG values derived from the described analysis of the
experimental curves are presented in Fig. 9 – they are denoted black solid circles.
For the QD-1 ensemble, the HG = 290 ± 20 meV half-width similar to Hvis (T) does
not change within the experimental error over the entire temperature range. In this
case HG < Hvis , which corresponds to the half-width of the luminescence band of the
quantum dots explored [63, 66, 67]. For the QD-2 exciton band, the estimation was
carried out using the second method. An example of an approximating Gaussian is
presented in the inset in Fig. 7b for 6.5 K. As in the case of QD-1, the half-width
does not change with temperature and amounts to HG = 370 ± 30 meV [63].

Fig. 9 The values of H for


various InP modifications
[65]. 1, Hvis (T) [61]; 2,
HG (T) [61]; 3, InP/ZnS [15];
4, InP/ZnS [4]; 5, 3.2 ÐÏ
InP/ZnS [6]; 6, 2.3 ÐÏ
InP/ZnS [28]; 7, InP/ZnS [5];
8, PL H(T) InP/ZnS [29].
Inset: 9, InP single crystal
[64]; 10, InP nanowires [56];
11, InP single crystal [69]
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 181

The temperature-independent half-width of the exciton band was previously


observed in the optical absorption spectra of CdSSe quantum dots with an average
size of 2.3 nm in a glassy matrix [68]. The values of H = 154 meV for the
first exciton band corresponding to the lowest electron–hole pair transitions almost
did not change upon cooling from 300 K to 20 K. As the authors indicate, this
is a consequence of the dominance of temperature-independent inhomogeneous
broadening mainly related to the f (a) size distribution of nanocrystals.
In Fig. 9, different symbols show the values of the visual half-width for the
first exciton absorption peak; we calculated using independent data: for InP/ZnS
quantum dots of various sizes [4–6, 15, 28], nanowires [56], and bulk InP single-
crystals [64, 69]. It can be seen that the Hvis value at room temperature (see
dependence 1 in Fig. 9) is consistent with the independent literature data on InP/ZnS
core/shell quantum dots (see symbols 3–7 in Fig. 9, the average nanocrystal size,
if known, indicated in the caption to the figure). In bulk InP single crystals, the
corresponding characteristic varies from 2 meV to 12 meV in the temperature range
of 5–300 K [69] (see the inset in Fig. 9). Thus, the half-width of the exciton band
at 6.5 K in InP/ZnS various-size quantum dots made by different manufacturers
is more than 100 times higher than the similar characteristic for bulk InP single
crystals.
The large value of the half-width of the exciton absorption band in InP/ZnS
quantum dots and its temperature independence evidence the inhomogeneous nature
of the broadening of the first exciton absorption band in the nanocrystals. In [16],
the authors note the dominant contribution of inhomogeneous broadening to the full
spectrum width for InP/ZnS QDs. They make this conclusion based on an analysis
of the stationary absorption and luminescence spectra within the Kennard–Stepanov
relation. It is also known that the composition and thickness of layers in multiple
quantum wells have an influence on the inhomogeneous width of the exciton absorp-
tion peak [70]. It should be emphasized that the temperature-independent behavior is
observed not only in low-dimensional systems but also typical of disordered massive
solids [71, 72]. In amorphous and vitreous states, the properties of optically active
point defects are significantly affected by inhomogeneous spectrum broadening
that is associated with the nonequivalence of centers’ local surrounding [73]. By
analogy with the disordered state, the nonequivalence of the excitons’ surrounding
being created due to the absorption of photons in each individual nanocrystal of QD
ensemble takes place.
Overall, inhomogeneous broadening in the absorption spectra is typical of
ensembles of nanocrystals and is a consequence of the size, shape, stoichiometry,
defect concentration, charge state, local environment distributions, etc. of nanocrys-
tals. All these factors cause essential differences in the potentials of each individual
nanocrystal [74]. Therefore, the energy structure of elementary excitations changes,
which ultimately manifests itself in the inhomogeneous broadening of the exciton
absorption bands.
To quantitatively analyze this effect, a model was proposed that describes well
the experimentally observed behavior of the first exciton absorption band and allows
one to estimate the influence of inhomogeneities in the ensembles of nanocrystals
by a magnitude of the broadening of optical spectra.
182 S. Savchenko et al.

6 Temperature Evolution of the First Exciton Absorption


Band of QD Ensemble

6.1 Static and Dynamic Disorder in Ensemble

The half-width h of an exciton optical absorption peak of a single nanocrystal at any


temperature can be represented as follows:

h(T ) = h0 + Δh (T ) (11)

Here, the first term reflects the natural line width at zero temperature. The second
summand accounts for the effects initiated by its temperature broadening. When
looking into a system of identical nanocrystals with the same energy e1 of the
first exciton transition, the corresponding absorption line homogeneously broadens
under the influence of temperature. Within the exciton–phonon interaction, the
corresponding contribution is given by [28, 29, 75]:
 −1
Δh (T ) = σ T + ALO exp (ωLO /kT ) − 1 (12)

where σ is the exciton–acoustic phonon coupling coefficient, eV/K, and the ALO
value reflects the force of interaction between excitons and longitudinal optical
LO-phonons with an energy -hωLO . In this case, the Δh value considered herein
quantitatively characterizes the dynamic disorder, which provides a temperature-
dependent contribution to the broadening of energy levels due to lattice vibrations
[61].
In real systems, nanocrystals in an ensemble possess different energies e1 due to
the scatter in the characteristics of QDs [74]. As a result, the H half-width of the
exciton absorption band for the ensemble turns out to be larger than h for individual
nanocrystals even at zero temperature. This occurs because of overlapping the
closely spaced peaks with different energies e1 . In this case, the optical absorption
band is inhomogeneously broadened, and the ΔI quantity quantitatively charac-
terizes the static disorder, which provides a temperature-independent contribution
to the broadening of energy levels [72, 76]. The static disorder is caused by the
f (X) distribution in the parameters of QDs in the ensemble tested. With increasing
temperature, the bands of individual nanocrystals broaden in accordance with Eq.
(12). This, in turn, affects the H width of the ensemble’s exciton band as a whole,
which leads to a homogeneous contribution of T (X, T) to the band broadening.
The temperature evolution of the H half-width of the optical absorption band of the
ensemble of nanocrystals with a certain distribution f in parameters can be written
as follows [65]:

H (X, T ) = h0 + ΔI (X) + ΔT (X, T ) . (13)


Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 183

In this case, the broadening is governed by the influence of both static and
dynamic types of atomic disorder. In what follows, we will call it temperature
broadening.

6.2 The Behavior of the Exciton Line of an Individual


Nanocrystal

To simulate the temperature behavior of an exciton band of an ensemble of


nanocrystals, it is necessary to know the change in the parameters of their individual
components. For this, the temperature dependencies should be preassigned for the
shift of the maximum of individual bands, their broadening, and area changes. The
individual spectral components for an ensemble of QDs are assumed to behave
identically, depending on temperature.
The shift of the e1 centers of individual Gaussian peaks corresponding to the
energy positions of the maxima of the absorption peaks of individual nanocrystals
of size a was modeled in accordance with Fan expression (4). As was claimed
above, this model well describes the shift of the first exciton absorption peak of
the QD ensemble for the following values of the parameters given in Table 4. The
indicated dependence is shown in Fig. 10, curve 1. Thus, the model adopted and

Table 4 Parameters of temperature evolution of the exciton peak in an InP/ZnS single nanocrystal
Shift [61] E1 (0) (eV) AF (eV) h̄ω (meV)
2.72 0.09 15
Broadening [29, 77] h0 (meV) σ (μeV/K) ALO (meV) h̄ωLO (meV)
6.3 172 60 40

Fig. 10 Model’s dependencies of the temperature behavior of the exciton absorption peak of
a single nanocrystal. (a) Shift (line 1 calculated according to Eq. (4)) and broadening (line 2
calculated according to Eq. (11)) of the exciton peak for an InP/ZnS individual nanocrystal. The
red triangle is the experimental data from [77]. (b) Change in area. Symbols are the experimental
estimates; solid line is the approximation according to Eq. (14)
184 S. Savchenko et al.

the experiment performed share the same shift magnitude for e1 in the band of an
individual nanocrystal and for E1 in the band of the ensemble as a whole.
The h(T) dependence was established through relations (11) and (12). Eq. (12)
includes the parameter values taken from [29], where they were determined by
approximating the H(T) dependence of the luminescence band of an ensemble of
InP/ZnS nanocrystals with an average size of 2.1 nm in the temperature range
of 2–510 K. The values are given in Table 4. The value of H = 73 meV for an
individual InP/ZnS core/shell nanocrystal at room temperature was found in [77]
using photon-correlation Fourier spectroscopy (see triangle in Fig. 10). The close
value of H(300) = 68 meV was previously obtained for InP nanocrystals when
analyzing the size dependence of the bandgap [78]. Figure 10 (curve 2) illustrates
the h(T) dependence calculated for a single nanocrystal in the ensemble. The value
of the fundamental half-width h0 in the model adopted amounts to 6.3 meV. The
line width at 6.5 K is h = 7.4 meV and exceeds the corresponding value for a single
crystal (2 meV). The values specified are in complete agreement with the data of
[68]. The latter emphasizes that the half-width of optical bands in quantum dots,
measured using the spectral hole burning technique, is three times larger than that
for bulk analogs.
The temperature-dependent area of the individual components corresponding
to individual nanocrystals is assumed to change in the same manner as the
experimentally observed area S of the integral peak of the QD ensemble. With
the half-width H remaining constant over a wide range of temperatures, the S(T)
dependence curve reflects the behavior of the optical density DG (T) at the maximum
of the exciton band. The S(T) experimental graph was built by evaluating the
area of the Gaussians used for approximating the first exciton band in the optical
absorption spectra measured (see Fig. 8). When simulating the temperature behavior
of the area of the optical component of an individual nanocrystal, the empirical
formula was utilized for the relative change in sr (T). This formula well describes
the experimentally observed dependence S(T):

S(T )  −1
sr (T ) = = 1 − B1 exp (B2 /T ) − 1 . (14)
S(0)

The values of the area beneath the Gaussian curves for the QD-1 ensemble,
normalized to S(0), and the corresponding model curve represented by expression
(14) that imitates the sr (T) dependence are shown in Fig. 10b. The values of
empirical parameters calculated during the approximation are S(0) = 0.22 eV,
B1 = 0.68, and B2 = 282.92 K. It can be seen that the area of the exciton peak
area decreases almost twice as the temperature increases in the range studied.
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 185

6.3 QD Size Distribution

A number of papers regard the dispersion of QDs in size a as the dominant factor
of inhomogeneous broadening [18, 22, 25, 26, 74]. Therefore, for definiteness, it is
the influence of this factor, X = f (a), that will be further associated with the static
disorder. The relationship between the size variation in the QD ensemble and the
width of the observe exciton absorption band is due, first and foremost, to quantum-
size effects. The estimate of the f (a) distribution function in the QD ensemble was
brought about using Eq. (1) and Gaussian curves intended for approximating the
exciton absorption bands at 6.5 K (see Fig. 8). In Eq. (1), the bandgap width for
a bulk InP single crystal was accepted as Eg (0) = 1.42 eV, which corresponds to
the value at 2 K [69]. The maximum position of the Gaussian band on the energy
axis corresponds to the energy e1 of the first exciton absorption peak of a particular
nanocrystal in the ensemble. Thus, the energy was recalculated into the radius of
the corresponding nanocrystal through the transformation e1 → a. In the process,
the optical density at certain energy is assumed to be proportional to the number of
nanocrystals of the corresponding size in the ensemble. Figure 11 compares the
normalized distributions f1 (a) and f2 (a) thus calculated for the QD-1 and QD-2
ensembles, respectively, and experimental data taken from independent works [79,
80]. The average values of the nanocrystal radii in the respective distributions are
ā1 = 2.1 nm and ā2 = 2.3 nm.
To characterize the quality of the size distributions, we computed the relative
width parameter using the following formula:

Δa
δ= 100% (15)
a

Fig. 11 Size distributions


and the δ parameter of the
studied InP/ZnS quantum
dots in comparison with the
data from independent works
[65]: 1, [79]; 2, [80]
186 S. Savchenko et al.

Table 5 Estimation of the Temperature (K) ā1 (nm) ā2 (nm)


size of nanocrystals according
to Eq. (1) at different 6.5 2.08 2.10
temperatures 296 2.26 2.32

where a is the width at half-height for the f (a) dependence. The δ-values obtained
are shown in Fig. 11 in brackets. It can be seen that in the ensembles concerned
a2 > a1 , f1 (a) and f2 (a) are quite typical of InP/ZnS nanocrystals.
It should be noted that such an approach in estimating the size distribution is
most appropriate for the absorption band since the latter’s position is not affected
by the Stokes shift, against the luminescence band. In addition, expression (1)
does not explicitly take into account the temperature factor. The latter is included
only through the Eg quantities for QDs and bulk semiconductors. However, the
nature of the influence of temperature on the behavior of the bandgap in the zero-
dimensional and three-dimensional cases may be different. This leads to an error
in estimating both the size itself and the distribution in size. We calculated the
sizes of nanocrystals both for room temperature and for 6.5 K in accordance with
the experimental data obtained. The results are listed in Table 5. For the QD-2
ensemble, where the dependence E1 (T) differs significantly from that of bulk InP,
the divergence in size estimates is larger in magnitude than for QD-1. Thus, to
reduce a temperature influence on the spectral position and half-width of the band,
the average size and size distribution need to be estimated at low temperatures. Then
the T (f, T) and E1 (T) contributions can be neglected.
Varying H of the absorption band underlies the simulation of the ensembles with
different size distributions f (a). In the case of a monodisperse QD ensemble, the
absorption band is formed by a single Gaussian-shaped component with the energy
e1 , half-width h, and the area s. The size distribution process leads to dramatically
raising the number of such components due to differences in the first exciton
transition energy to finally form a Gaussian-shaped band with energy E1 , half-width
H, and area S.

6.4 Contributions of Homogeneous and Inhomogeneous


Broadening

The influence of the degree of static disorder on broadening of the first exciton
absorption band can be analyzed premised on the proposed model by comparing
the homogeneous and inhomogeneous contributions for samples with different f (X)
distribution. The contribution of homogeneous broadening can be quantified as

ΔT (X, T )
CH = . (16)
H (X, T )

The contribution of inhomogeneous broadening, in turn, can be written as


Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 187

ΔI (X)
CI = . (17)
H (X, T )

Apart from the dimensional criterion δ, we also employed the optical criterion Q
that reflects the ratio of the homogeneous and inhomogeneous contributions to the
band broadening:

CH ΔT (X, T )
Q= = . (18)
CI ΔI (X)

In this case, it does not matter which factor causes the inhomogeneous broaden-
ing since we are dealing with their complex influence on the optical characteristics.
For the QD ensembles, this criterion is integral and monotonously increases
with decreasing static disorder in the system. Thus, the criterion can serve as a
universal tool of comparing the quality of low-dimensional systems with distributed
parameters.

6.5 Simulation of Experimental InP/ZnS Ensembles

Simulation of the temperature behavior of the exciton absorption band for a QD


ensemble with a given size distribution of f (a) included three stages. At first, the
OA band at zero temperature was formed by a set of Gaussian components. Each
of the latter had a half-width h0 and corresponded to a subset of nanocrystals of the
same size in the ensemble. The second stage predicted the behavior of an individual
Gaussian as the temperature increases. That is, the position of the e1 maximum,
the h half-width, and the s area were changed according to the assigned functional
dependencies (4), (12), and (14), respectively. At the end, at a fixed temperature,
the Gaussian components were summed up and, thus, a model optical exciton
absorption band was formed for the QD ensemble with parameters E1 , H, and S.
The model proposed was intended for describing the temperature behavior of the
first exciton absorption band of the experimentally studied QD-1 ensemble with
a size distribution of δ = 11.1%. The band with a half-width of 290 meV was
approximated by Gaussian components with an h = 7.4 meV half-width, which
corresponds to a temperature of 6.5 K. The minimum number of components to
describe the experimental data with high accuracy (adj. R2 > 0.999) amounted to
N = 180. Then, the exciton peaks of individual nanocrystals were simulated for
various experimental temperatures in the range 6.5–296 K. Figure 12 illustrates the
temperature evolution of such an individual spectral component. It can be seen that
the maximum shifted by e1 = 112 meV to the region of lower energies in the
temperature range of 6.5–296 K. Simultaneously, the area s decreased by 1.8 times,
and h increased by almost ten times from 7.4 meV to 73 meV.
After summing up the individual spectral components at different temperatures,
the temperature behavior of the model integrated band was found to describe well
188 S. Savchenko et al.

Fig. 12 Visualization of the


temperature change in the
absorption bands
corresponding to individual
nanocrystals in the ensemble.
For clarity, the optical density
is normalized to the
maximum value for a
component at 6.5 K [65]

Fig. 13 Experimental optical


absorption spectra (symbols)
and simulation of the first
exciton absorption band
(solid lines) for QD-1 at
specified temperatures [65]

the exciton absorption band experimentally observed for the QD-1 ensemble across
the entire temperature range (see Fig. 13). The resulting shifts in the position of
the model absorption band for the ensemble and a change in its area occur by the
same amount as in the experiment. It is important to underscore that the temperature
behavior of the indicated integral parameters coincides with the behavior of the
optical components for single nanocrystals. Thus, good agreement between the
experimental and simulation results confirms the validity of the assumptions made
earlier. The half-width of the total exciton absorption band for the InP/ZnS QD
ensemble remains unchanged within the experimental error in the temperature range
of 6.5–296 K (see the dependence for δ = 11.1% in Fig. 14 [hollow squares]). At the
same time, the h half-width of the optical peak for each of the nanocrystals in the
distribution varies by almost ten times. The proposed model correctly reproduces
the temperature behavior of the first exciton absorption band for the InP/ZnS QD
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 189

Fig. 14 Simulated H(T)


dependencies of the exciton
band of ensembles of
nanocrystals with different
size distributions (Table 6).
Hollow symbols indicate the
experimental dependence for
the QD-1 ensemble [65]

ensemble and can be applied for analyzing the influence of the size distribution of
nanocrystals on the broadening of optical spectra.

6.6 Disorder Effects in Temperature Band Broadening

Analyzing the influence of the degree of static disorder on the processes of


homogeneous and inhomogeneous broadening of the absorption band involved the
investigation of a number of QD ensembles with different half-widths H of the
first exciton band (see Table 6). The value of N indicates the number of Gaussian
components used for simulation of the corresponding integrated band. Moreover, the
latter’s maximum was at an energy of 2.72 eV, which conforms to the experimental
position for the ensemble of nanocrystals under study at 6.5 K. This means that the
computational experiments save the average particle size standing, and only the size
distribution changes, which is reflected in a change in H.
Figure 14 presents the simulation results of the temperature change in the half-
width of the exciton absorption band for ensembles of nanocrystals with different
degrees of static disorder, i.e., by different distribution of f (a), in the temperature
range of 6.5–296 K. It can be seen that as the width of f (a) for the nanocrystals
in the ensemble diminishes, the homogeneous broadening of the model band is
more pronounced. In the case of the band experimentally observed (the half-width
is 290 meV), the contribution of homogeneous broadening is CH = 3% and does
not exceed the experimental error of 7%. Reducing the degree of static disorder,
the contribution of homogeneous broadening rises. The findings of CT , CI , and Q
calculated using Eqs. (16), (17), and (18) at room temperature for the model range of
samples are shown in Table 6. As is clear from the table, as δ decreased, the behavior
of the H exciton absorption band of the ensemble tended to be the characteristic of
the only homogeneously broadened component. Otherwise speaking, H behaved as
190 S. Savchenko et al.

Table 6 The simulation results of the temperature behavior of the first exciton absorption band
of InP/ZnS quantum dot ensembles with different size distributions (data are shown in Fig. 14)
Curve H(6.5 K) (meV) N δ (%) CI (%) CH (%) Q
1 290 180 11.1 94.9 3.0 0.03
2 242 150 9.2 93.3 4.2 0.05
3 193 120 7.3 90.5 6.4 0.07
4 145 90 5.5 85.5 10.6 0.12
5 113 70 4.3 79.4 15.9 0.20
6 81 50 3.1 68.7 25.5 0.37
7 64 40 2.4 59.6 33.9 0.57
8 48 30 1.8 47.9 44.9 0.94
9 32 20 1.2 32.4 59.7 1.84
10 16 10 0.6 13.0 78.5 6.02
Single QD 7.4 1 0 0 89.9 –

matching to a monodisperse ensemble. On the other hand, the increase in the degree
of static disorder in the ensemble led to lesser and lesser impact of the homogenous
broadening of the peak of each individual nanocrystal on the width of the integrated
band with increasing temperature.
The simulated regularities also explain the temperature behavior of the H for
the QD-2 ensemble. For the latter, the half-width of the optical absorption band
amounts to 370 meV, which corresponds to a wider size distribution (δ = 17.3%)
than for QD-1. The contribution of CT to the temperature broadening in such an
ensemble is even less than in the case of QD-1.
It is worth emphasizing that the work in [29] reports on a change observed in the
half-width of the PL band of an ensemble of InP/ZnS nanocrystals. In particular,
for a sample with a half-width of ≈0.24 eV at 300 K, the temperature broadening
in the range of 2–300 K was ≈0.08 eV (see curve 8 in Fig. 9). This may be due to
the transfer of excitation energy over the ensemble of quantum dots from small to
larger [36]. As a consequence, a certain size subset of nanocrystals participates in
the radiative process, which leads to a decrease in the inhomogeneous broadening
contribution to the H value of the luminescence band of the ensemble. In this case,
the dynamic disorder contribution to the temperature broadening increases.
The analysis conducted indicates some limitations that should be considered
when analyzing the temperature evolution of the exciton absorption bands in the
ensembles of semiconductor nanocrystals. In describing the temperature curves for
the shift of the maxima, the values of the extracted fundamental parameters of the
exciton–phonon interaction are physically justified. As was shown in the framework
of the model proposed, the degree of static disorder in the system affects slightly
these dependencies. The shift of the maximum of the integrated band reflects the
displacement of the optical component of the peak for an individual nanocrystal.
However, when considering the temperature variations in the band half-width for
a QD ensemble, the static disorder plays a significant role. When forming, the
integrated absorption band losses the homogeneous broadening of the individual
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 191

nanocrystal components due to their overlap. The contribution of static disorder dis-
torts the corresponding parameter values of the exciton–phonon interaction, while
they characterize the ratio of the homogeneous and inhomogeneous broadening
processes. It is, therefore, mandatory to carry out estimates of them carefully. Thus,
for making an infallible estimate of the parameters of dynamic disorder, an analysis
of the temperature broadening processes in optical spectra needs to run accounting
for the contribution of inhomogeneous broadening factors.

7 Conclusion

In conclusion, we note that the chapter has explored the peculiarities of optical
absorption spectra in the temperature range of 6.5–296 K for ensembles of InP/ZnS
core/shell colloidal quantum dots with average particle radii of 2.1 (QD-1) and
2.3 (QD-2) nm. The approaches of derivative spectrophotometry are shown to
make it possible to successfully determine the spectral positions of hidden exciton
absorption bands of the nanocrystals. When formed by the precipitation method,
the films exhibit collective effects in the QD ensembles. This can be evidenced by
shifting the absorption bands for the shell, as compared to those for the solution.
For describing the temperature dependence of the energy gap in semiconductor
nanocrystals, well-known models were considered.
It is revealed that the shift in the position of the first exciton absorption band
E1 with temperature is due to the interaction with the effective vibrational modes
of 15- and 59-meV energies for QD-1 and QD-2, respectively. The H half-width
of the band for these ensembles is temperature-independent and amounts to 290
and 370 meV. A model is proposed that allows one to quantitatively describe the
temperature behavior of the first exciton absorption band of InP/ZnS QD ensembles,
as well as to analyze the contributions of homogeneous and inhomogeneous
broadening of this band. The numerical analysis conducted proved that the half-
width of the exciton absorption band for the ensembles remains unchanged in the
temperature range of 6.5–296 K, even accounting for the tenfold homogeneous
broadening of spectral components for individual nanocrystals.
The influence of static disorder associated with the δ quality factor of the size
distribution in the QD ensembles on the processes of broadening of model bands
is analyzed. It is shown that, when going over from a monodisperse ensemble of
nanocrystals to a δ = 11.1% distribution (for the QD-1 sample), the homogeneous
contribution to the integrated broadening drops from 90% to 3%. Thus, processes
of inhomogeneous broadening due to the high-degree static disorder in the QD
ensembles tested determine the observed temperature independence of the exciton
band half-width. The approach proposed allows examining various low-dimensional
systems whose optical characteristics are sensitive to the distribution of structural
parameters.
192 S. Savchenko et al.

Acknowledgments This research was supported by RFBR according to the research project no.
18-32-00664 and Act 211 Government of the Russian Federation, contract no. 02.A03.21.0006.

References

1. Alivisatos, A.P.: Semiconductor clusters, nanocrystals, and quantum dots. Science. 271, 933–
937 (1996). https://doi.org/10.1126/science.271.5251.933
2. Efros, A.L., Efros, A.L.: Interband absorption of light in a semiconductor sphere. Sov. Phys.
Semicond. 16, 772–775 (1982)
3. Brus, L.E.: Electron-electron and electron-hole interactions in small semiconductor crystal-
lites: the size dependence of the lowest excited electronic state. J. Chem. Phys. 80, 4403–4409
(1984). https://doi.org/10.1063/1.447218
4. Xie, R., Battaglia, D., Peng, X.: Colloidal InP nanocrystals as efficient emitters covering
blue to near-infrared. J. Am. Chem. Soc. 129, 15432–15433 (2007). https://doi.org/10.1021/
ja076363h
5. Lee, S.-H., Lee, K.-H., Jo, J.-H., Park, B., Kwon, Y., Jang, H.S., Yang, H.: Remote-type, high-
color gamut white light-emitting diode based on InP quantum dot color converters. Optic.
Mater. Express. 4, 1297–1302 (2014). https://doi.org/10.1364/OME.4.001297
6. Song, W.-S., Lee, S.-H., Yang, H.: Fabrication of warm, high CRI white LED using non-
cadmium quantum dots. Opt. Mater. Express. 3, 1468–1473 (2013). https://doi.org/10.1364/
OME.3.001468
7. Kamat, P.V.: Quantum dot solar cells. Semiconductor nanocrystals as light harvesters. J. Phys.
Chem. C. 112, 18737–18753 (2008). https://doi.org/10.1021/jp806791s
8. Scher, J.A., Bayne, M.G., Srihari, A., Nangia, S., Chakraborty, A.: Development of effective
stochastic potential method using random matrix theory for efficient conformational sampling
of semiconductor nanoparticles at non-zero temperatures. J. Chem. Phys. 149, 014103 (2018).
https://doi.org/10.1063/1.5026027
9. Rempel, A.A., Kozlova, E.A., Gorbunova, T.I., Cherepanova, S.V., Gerasimov, E.Y.,
Kozhevnikova, N.S., Valeeva, A.A., Korovin, E.Y., Kaichev, V.V., Shchipunov, Y.A.: Synthesis
and solar light catalytic properties of titania-cadmium sulfide hybrid nanostructures. Catal.
Commun. 68, 61–66 (2015). https://doi.org/10.1016/j.catcom.2015.04.034
10. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Luminescence parameters of
InP/ZnS@AAO nanostructures. AIP Conf. Proc. 1717(4943471), (2016). https://doi.org/
10.1063/1.4943471
11. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Optical properties of InP/ZnS quantum
dots deposited into nanoporous anodic alumina. J. Phys. Conf. Ser. 741, 012151 (2016). https:/
/doi.org/10.1088/1742-6596/741/1/012151
12. Klimov, V.I., Mikhailovsky, A.A., Xu, S., Malko, A., Hollingsworth, J.A., Leatherdale, C.A.,
Eisler, H.-J., Bawendi, M.G.: Optical gain and stimulated emission in nanocrystal quantum
dots. Science. 290, 314–317 (2000). https://doi.org/10.1126/science.290.5490.314
13. Anc, M.J., Pickett, N.L., Gresty, N.C., Harris, J.A., Mishra, K.C.: Progress in non-Cd quantum
dot development for lighting applications. ECS J. Solid State Sci. Technol. 2, R3071–R3082
(2013). https://doi.org/10.1149/2.016302jss
14. Resch-Genger, U., Grabolle, M., Cavaliere-Jaricot, S., Nitschke, R., Nann, T.: Quantum dots
versus organic dyes as fluorescent labels. Nat. Methods. 5, 763–775 (2008). https://doi.org/
10.1038/nmeth.1248
15. Hussain, S., Won, N., Nam, J., Bang, J., Chung, H., Kim, S.: One-pot fabrication of high-quality
InP/ZnS (core/shell) quantum dots and their application to cellular imaging. ChemPhysChem.
10, 1466–1470 (2009). https://doi.org/10.1002/cphc.200900159
16. Reiss, P., Protiere, M., Li, L.: Core/shell semiconductor nanocrystals. Small. 5, 154–168
(2009). https://doi.org/10.1002/smll.200800841
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 193

17. Brichkin, S.B.: Synthesis and properties of colloidal indium phosphide quantum dots. Colloid
J. 77, 393–403 (2015). https://doi.org/10.1134/S1061933X15040043
18. Song, W.-S., Lee, H.-S., Lee, J.C., Jang, D.S., Choi, Y., Choi, M., Yang, H.: Amine-derived
synthetic approach to color-tunable InP/ZnS quantum dots with high fluorescent qualities. J.
Nanopart. Res. 15, 1750 (2013). https://doi.org/10.1007/s11051-013-1750-y
19. Brichkin, S.B., Spirin, M.G., Tovstun, S.A., Gak, V.Y., Mart’yanova, E.G., Razumov,
V.F.: Colloidal quantum dots InP@ZnS: inhomogeneous broadening and distribution of
luminescence lifetimes. High Energy Chem. 50, 395–399 (2016). https://doi.org/10.1134/
S0018143916050064
20. Ayupova, D., Dobhal, G., Laufersky, G., Nann, T., Goreham, R.V.: An in vitro investigation
of cytotoxic effects of InP/ZnS quantum dots with different surface chemistries. Nano. 9, 135
(2019). https://doi.org/10.3390/nano9020135
21. Gaponenko, S.V.: Optical Properties of Semiconductor Nanocrystals. Cambridge University
Press, Cambridge (1998)
22. Weller, H.: Colloidal semiconductor Q-particles: chemistry in the transition region between
solid state and molecules. Angew. Chem. Int. Ed. Engl. 32, 41–53 (1993). https://doi.org/
10.1002/anie.199300411
23. Empedocles, S.A., Bawendi, M.G.: Quantum-confined stark effect in single CdSe quantum-
confined stark effect in single CdSe nanocrystallite quantum dots. Science. 278, 2114–2117
(1997). https://doi.org/10.1126/science.278.5346.2114
24. Ekimov, A.I.: Optical properties of semiconductor quantum dots in glass matrix. Phys. Scr. 39,
217–222 (1991). https://doi.org/10.1088/0031-8949/1991/T39/033
25. Murray, C.B., Norris, D.J., Bawendi, M.G.: Synthesis and characterization of nearly monodis-
perse CdE (E = S, Se, Te) semiconductor nanocrystallites. J. Am. Chem. Soc. 115, 8706–8715
(1993). https://doi.org/10.1021/ja00072a025
26. Efros, A.L., Rosen, M., Kuno, M., Nirmal, M., Norris, D., Bawendi, M.: Band-edge exciton
in quantum dots of semiconductors with a degenerate valence band: dark and bright exciton
states. Phys. Rev. B. 54, 4843–4856 (1996). https://doi.org/10.1103/PhysRevB.54.4843
27. Sugisaki, M., Ren, H.-W., Nishi, K., Masumoto, Y.: Optical properties of InP self-assembled
quantum dots studied by imaging and single dot spectroscopy. Jpn. J. Appl. Phys. Part. 1 Regul.
Pap. Short Notes Rev. Pap. 41, 958–966 (2002)
28. Narayanaswamy, A., Feiner, L.F., Van Der Zaag, P.J.: Temperature dependence of the
photoluminescence of InP/ZnS quantum dots. J. Phys. Chem. C. 112, 6775–6780 (2008).
https://doi.org/10.1021/jp800339m
29. Narayanaswamy, A., Feiner, L.F., Meijerink, A., Van Der Zaag, P.J.: The effect of temperature
and dot size on the spectral properties of colloidal InP/ZnS core-shell quantum dots. ACS
Nano. 3, 2539–2546 (2009). https://doi.org/10.1021/nn9004507
30. Biadala, L., Siebers, B., Beyasit, Y., Tessier, M.D., Dupont, D., Hens, Z., Yakovlev, D.R.,
Bayer, M.: Band-edge exciton fine structure and recombination dynamics in InP/ZnS colloidal
nanocrystals. ACS Nano. 10, 3356–3364 (2016). https://doi.org/10.1021/acsnano.5b07065
31. Norris, D.J., Bawendi, M.G.: Measurement and assignment of the size-dependent optical
spectrum in CdSe quantum dots. Phys. Rev. B. 53, 16338–16346 (1996). https://doi.org/
10.1103/PhysRevB.53.16338
32. Talsky, G.: Derivative Spectrophotometry: Low and Higher Order. VCH, Weinheim (1994)
33. O’Haver, T.C., Begley, T.: Signal-to-noise ratio in higher order derivative spectrometry. Anal.
Chem. 53, 1876–1878 (1981). https://doi.org/10.1021/ac00235a036
34. Savitzky, A., Golay, M.J.E.: Smoothing and differentiation of data by simplified least squares
procedure. Anal. Chem. 36, 1627–1639 (1964). https://doi.org/10.1021/ac60214a047
35. Kuś, S., Marczenko, Z., Obarski, N.: Derivative UV-VIS spectrophotometry in analytical
chemistry. Chem. Anal. 41, 899–929 (1996)
36. Mićić, O.I., Jones, K.M., Cahill, A., Nozik, A.J.: Optical, electronic, and structural properties
of uncoupled and close-packed arrays of InP quantum dots. J. Phys. Chem. B. 102, 9791–9796
(1998). https://doi.org/10.1021/jp981703u
194 S. Savchenko et al.

37. Fu, H., Zunger, A.: Excitons in InP quantum dots. Phys. Rev. B Condens. Matter Mater. Phys.
57, R15064–R15067 (1998). https://doi.org/10.1103/PhysRevB.57.R15064
38. Babichev, A.P.: Handbook of Physical Quantities. In: Grigor’ev, I.S., Meilikhov E.Z. (eds.)
Energoatomizdat, Moscow (1991). (in Russian)
39. Baskoutas, S., Terzis, A.F.: Size-dependent band gap of colloidal quantum dots. J. Appl. Phys.
99, 013708 (2006). https://doi.org/10.1063/1.2158502
40. Fu, H., Wang, L.W., Zunger, A.: Applicability of the k•p method to the electronic
structure of quantum dots. Phys. Rev. B. 57, 9971–9987 (1998). https://doi.org/10.1103/
PhysRevB.57.9971
41. Kayanuma, Y., Momiji, H.: Incomplete confinement of electrons and holes in microcrystals.
Phys. Rev. B. 41, 10261–10263 (1990). https://doi.org/10.1103/PhysRevB.41.10261
42. Pellegrini, G., Mattei, G., Mazzoldi, P.: Finite depth square well model: applicability and
limitations. J. Appl. Phys. 97, 073706 (2005). https://doi.org/10.1063/1.1868875
43. Baskoutas, S., Terzis, A.F.: Size dependent exciton energy of various technologically important
colloidal quantum dots. Mater. Sci. Eng. B-Solid State Mater. Adv. Technol. 147, 280–283
(2008). https://doi.org/10.1016/j.mseb.2007.09.041
44. Weller, H., Koch, U., Gutierrez, M., Henglein, A.: Photochemistry of colloidal metal sulfides.
7. Absorption and fluorescence of extremely small ZnS particles (the world of the neglected
dimensions). Berichte der Bunsengesellschaft/Physical Chemistry Chemical Physics. 88, 649–
656 (1984). https://doi.org/10.1002/bbpc.19840880715
45. Zhang, Y., Ma, M., Wang, X., Fu, D., Gu, N., Liu, J., Lu, Z., Ma, Y., Xu, L., Chen, K.:
First-order hyperpolarizability of ZnS nanocrystal quantum dots studied by hyper-Rayleigh
scattering. J. Phys. Chem. Solids. 63, 2115–2118 (2002). https://doi.org/10.1016/S0022-
3697(02)00259-7
46. Kho, R., Torres-Martınez, C.L., Mehra, R.K.: A simple colloidal synthesis for gram-quantity
production of water-soluble ZnS nanocrystal powders. J. Colloid Interface Sci. 227, 561–566
(2000). https://doi.org/10.1006/jcis.2000.6894
47. Norris, D.J., Sacra, A., Murray, C.B., Bawendi, M.G.: Measurement of the size dependent
hole spectrum in CdSe quantum dots. Phys. Rev. Lett. 72, 2612–2615 (1994). https://doi.org/
10.1103/PhysRevLett.72.2612
48. Woggon, U., Gaponenko, S.V.: Excitons in quantum dots. Phys. Status. Solidi B Basic Res.
189, 285–343 (1995). https://doi.org/10.1002/pssb.2221890202
49. Vainshtein, I.A., Zatsepin, A.F., Kortov, V.S.: Applicability of the empirical Varshni relation
for the temperature dependence of the width of the band gap. Phys. Solid State. 41, 905–908
(1999). https://doi.org/10.1134/1.1130901
50. Varshni, Y.P.: Temperature dependence of the energy gap in semiconductors. Physica. 34, 149–
154 (1967). https://doi.org/10.1016/0031-8914(67)90062-6
51. Fan, H.Y.: Temperature dependence of the energy gap in semiconductors. Phys. Rev. 82, 900–
905 (1951). https://doi.org/10.1103/PhysRev.82.900
52. Skettrup, T.: Urbach’s rule derived from thermal fluctuations in the band-gap energy. Phys.
Rev. B. 18, 2622–2631 (1978). https://doi.org/10.1103/PhysRevB.18.2622
53. Fan, H.Y.: Photon-Electron Interaction: Crystals Without Fields. Springer, Berlin (1967).
https://doi.org/10.1007/978-3-642-46074-6_3
54. Manoogian, A., Wooley, J.C.: Temperature dependence of the energy gap in semiconductors.
Can. J. Phys. 62, 285–287 (1984). https://doi.org/10.1139/p84-043
55. Viña, L., Logothetidis, S., Cardona, M.: Temperature dependence of the dielectric function of
germanium. Phys. Rev. B. 30, 1979–1991 (1984). https://doi.org/10.1103/PhysRevB.30.1979
56. Zilli, A., De Luca, M., Tedeschi, D., Fonseka, H.A., Miriametro, A., Tan, H.H., Jagadish, C.,
Capizzi, M., Polimeni, A.: Temperature dependence of interband transitions in wurtzite InP
nanowires. ACS Nano. 9, 4277–4287 (2015). https://doi.org/10.1021/acsnano.5b00699
57. O’Donnell, K.P., Chen, X.: Temperature dependence of semiconductor band gaps. Appl. Phys.
Lett. 58, 2924–2926 (1991). https://doi.org/10.1063/1.104723
58. Chen, L., Bao, H., Tan, T., Prezhdo, O.V., Ruan, X.: Shape and temperature dependence of hot
carrier relaxation dynamics in spherical and elongated CdSe quantum dots. J. Phys. Chem. C.
115, 11400–11406 (2011). https://doi.org/10.1021/jp201408m
Exciton–Phonon Interactions and Temperature Behavior of Optical Spectra. . . 195

59. Olkhovets, A., Hsu, R.-C., Lipovskii, A., Wise, F.W.: Size-dependent temperature variation
of the energy gap in lead-salt quantum dots. Phys. Rev. Lett. 81, 3539–3542 (1998). https://
doi.org/10.1103/PhysRevLett.81.3539
60. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Temperature dependence of the optical
absorption spectra of InP/ZnS quantum dots. Tech. Phys. Lett. 43, 297–300 (2017). https://
doi.org/10.1134/S1063785017030221
61. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Temperature-induced shift of the exciton
absorption band in InP/ZnS quantum dots. Optic. Mater. Express. 7, 354–359 (2017). https://
doi.org/10.1364/OME.7.000354
62. Alfrey, G.F., Borcherds, P.H.: Phonon frequencies from the Raman spectrum of indium
phosphide. J. Phys. C Solid State Phys. 5, L275–L278 (1972). https://doi.org/10.1088/0022-
3719/5/20/002
63. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Effect of temperature on the spectral
properties of InP/ZnS nanocrystals. J. Phys. Conf. Ser. 961, 012003 (2018). https://doi.org/
10.1088/1742-6596/961/1/012003
64. Turner, W.J., Reese, W.E., Pettit, G.D.: Exciton absorption and emission in InP. Phys. Rev. 136,
A1467–A1470 (1964). https://doi.org/10.1103/PhysRev.136.A1467
65. Savchenko, S.S., Weinstein, I.A.: Inhomogeneous broadening of the exciton band in optical
absorption spectra of InP/ZnS nanocrystals. Nano. 9, 716 (2019). https://doi.org/10.3390/
nano9050716
66. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Spectral features and lumines-
cence thermal quenching of InP/ZnS quantum dots within 7.5–295 K range. Optics
InfoBase Conference Papers Part F107-NOMA 2018. 140118 (2018). https://doi.org/10.1364/
NOMA.2018.NoW1J.4
67. Savchenko, S.S., Vokhmintsev, A.S., Weinstein, I.A.: Photoluminescence thermal quenching of
yellow-emitting InP/ZnS quantum dots. AIP Conf. Proc. 2015(020085), (2018). https://doi.org/
10.1063/1.5055158
68. Woggon, U., Gaponenko, S.V., Uhrig, A., Langbein, W., Klingshirn, C.: Homogeneous
linewidth and relaxation of excited hole states in II–VI quantum dots. Adv. Mater. Opt.
Electron. 3, 141–150 (1994). https://doi.org/10.1002/amo.860030121
69. Vaganov, S.A., Seisyan, R.P.: Temperature-dependent integral exciton absorption in semi-
conducting InP crystals. Tech. Phys. Lett. 38, 121–124 (2012). https://doi.org/10.1134/
S1063785012020174
70. Lee, D., Johnson, A.M., Zucker, J.E., Feldman, R.D., Austin, R.F.: Room temperature ex-
citonic absorption in CdZnTe/ZnTe quantum wells: contributions to exciton linewidth. J. Appl.
Phys. 69, 6722–6724 (1991). https://doi.org/10.1063/1.348895
71. Weinstein, I.A., Zatsepin, A.F., Shchapova, Y.V.: The phonon-assisted shift of the energy levels
of localized electron states in statically disordered solids. Physica B. 263–264, 167–169 (1999).
https://doi.org/10.1016/S0921-4526(98)01213-7
72. Weinstein, I.A., Zatsepin, A.F.: Modified Urbach’s rule and frozen phonons in glasses. Phys.
Status Solidi C. 1, 2916–2919 (2004). https://doi.org/10.1002/pssc.200405416
73. Skuja, L.: Defect studies in vitreous silica and related materials: optically active oxygen-
deficiency-related centers in amorphous silicon dioxide. J. Non-Cryst. Solids. 239, 16–48
(1998). https://doi.org/10.1016/0925-8388(96)02241-4
74. Salvador, M.R., Graham, M.W., Scholes, G.D.: Exciton-phonon coupling and disorder in the
excited states of CdSe colloidal quantum dots. J. Chem. Phys. 125, 184709 (2006). https://
doi.org/10.1063/1.2363190
75. Rudin, S., Reinecke, T.L., Segall, B.: Temperature-dependent exciton linewidths in semicon-
ductors. Phys. Rev. B. 42, 11218–11231 (1990). https://doi.org/10.1103/PhysRevB.42.11218
76. Weinstein, I.A., Zatsepin, A.F., Kortov, V.S.: Effects of structural disorder and Urbach’s rule in
binary lead silicate glasses. J. Non-Cryst. Solids. 279, 77–87 (2001). https://doi.org/10.1016/
S0022-3093(00)00396-3
77. Cui, J., Beyler, A.P., Marshall, L.F., Chen, O., Harris, D.K., Wanger, D.D., Brokmann,
X., Bawendi, M.G.: Direct probe of spectral inhomogeneity reveals synthetic tunability of
196 S. Savchenko et al.

single-nanocrystal spectral linewidths. Nat. Chem. 5, 602–606 (2013). https://doi.org/10.1038/


nchem.1654
78. Mićić, O.I., Curtis, C.J., Jones, K.M., Sprague, J.R., Nozik, A.J.: Synthesis and characteri-
zation of InP quantum dots. J. Phys. Chem. 98, 4966–4969 (1994). https://doi.org/10.1021/
j100070a004
79. Arias-Cerón, J.S., González-Araoz, M.P., Bautista-Hernández, A., Sánchez Ramírez, J.F.,
Herrera-Pérez, J.L., Mendoza-Álvarez, J.G.: Semiconductor nanocrystals of InP@ZnS: syn-
thesis and characterization. Superf. y Vacío. 25, 134–138 (2012)
80. Shen, W., Tang, H., Yang, X., Cao, Z., Cheng, T., Wang, X., Tan, Z., You, J., Deng, Z.:
Synthesis of highly fluorescent InP/ZnS small-core/thick-shell tetrahedral-shaped quantum
dots for blue light-emitting diodes. J. Mater. Chem. C. 5, 8243–8249 (2017). https://doi.org/
10.1039/c7tc02927f
Thick-Shell Core/Shell Quantum Dots

Lei Zhang, Wenbin Xiang, and Jiayu Zhang

Abstract Due to the Auger effect, traditional core/shell quantum dots exhibit
emission intermittency, which affects the application of quantum dots on lasers.
A thick shell could effectively inhibit the Auger nonradiation process, which
makes the quantum dots have high optical gain. In addition, the thick shell can
effectively eliminate the influence of the external environment on the excitons in the
nucleus, thus greatly improving the optical and chemical stability of the quantum
dots. Therefore, a number of research groups, including our research group, have
conducted extensive research on thick-shell core/shell quantum dots. This chapter
will review the important works of this kind of quantum dots according to the
following contents.

Keywords Core-thick shell · Quantum dots · Auger non-radiation · Optical


gain · Chemical stability

1 Thick-Shell Core/Shell Quantum Dots

1.1 History of Thick-Shell Core/Shell Quantum Dots

In recent decades, quantum dots (QDs), which are a kind of semiconductor


nanocrystals with particle size similar to the exciton Bohr radius, have attracted
much attention due to their unique physical, chemical, and luminescent properties.
However, the existence of a large number of active atoms on the QDs’ surface
can easily cause surface defects, which will seriously affect the photoluminescence

L. Zhang · W. Xiang · J. Zhang ()


Advanced Photonics Center, Southeast University, Nanjing, China
e-mail: leizhang.apc@seu.edu.cn; 220181289@seu.edu.cn; jyzhang@seu.edu.cn

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 197
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_6
198 L. Zhang et al.

quantum yields (PL QYs) and nonlinear optical properties of the QDs. In addition to
adjusting the nucleation and growth dynamics of quantum dots and controlling the
size and shape of quantum dots, the surfactants introduced in the preparation process
also play the role of surface ligands, which can not only provide enough space or
electrostatic repulsion force to make quantum dots disperse stably in solution and
prevent aggregation, but also passivates the surface trap states, so as to improve
the optical properties of quantum dots. Although the existence of surface organic
ligands can improve the surface properties and QYs, the optical stability of such
soft passivated QDs is relatively poor, and the surface ligand may also introduce
new defects on the QDs. In semiconductor QDs, the quantum confinement effect
leads to the discrete energy level structure, different energy levels correspond to
different angular momentums, and the exciton Auger recombination (AR) process
no longer needs to meet the dynamic quantity conservation [1–3]. Therefore, the
AR process of excitons in QDs is no longer constrained by the potential barrier and
even the carriers in the ground state can produce nonradiative AR. The nonradiative
AR is about much faster than the PL relaxation process. This ultrafast nonradiative
process will consume a large number of high-energy excitons and seriously hinder
the particle number inversion distribution of exciton state in QDs, thus limiting the
optical gain lifetime and gain bandwidth and reducing the optical gain performance.
For practical luminescent applications, inorganic shells are usually grown on the
surface of QDs to form core/shell heterostructure semiconductor QDs [4]. First, this
kind of inorganic shells can not only effectively passivate the surface trap states of
QDs but also isolate the surrounding environment to avoid the direct interference for
QDs. Therefore, the photophysical and photochemical stability of QDs are improved
significantly while obtaining high PL QYs. Second, the shell thickness has a great
influence on the exciton dynamics (AR and multiexciton process) of QDs, which
can adjust the optical properties of core/shell QDs by changing the shell thickness
according to the application requirements [5]. Therefore, the core/shell QDs have
important potential applications in the fields of light-emitting diode, fluorescence
labeling, solar cell, and microcavity lasers.
The optical properties of QDs can be further improved with the increase of the
shell thickness. For CdSe/CdS QDs, when the shell thickness is greater than 10 CdS
monolayers (MLs), the quantum dots are called “giant” QDs [5, 6]. In this kind
of thick-shell QDs, the nonradiative Auger process and the surface/interface states
can be effectively suppressed, so as to reduce the nonradiative loss of excitons
and enhance the gain performance of the QDs. At the same time, the PL blinking
would be reduced or even inhibited, and longtime laser radiation would not induce
a photobleaching phenomenon. Moreover, with the further increase of the shell
thickness, ultrathick-shell QDs would form the dot-in-bulk QD system, which could
show PL or stimulated emission from thick CdS shells due to the effect of the hole
potential barrier [7, 8].
Thick-Shell Core/Shell Quantum Dots 199

Fig. 1 Schematic illustration of the different types of core/shell heterostructures (top row) and
their energy band diagrams (bottom row) with possible electron (red curves) and hole (blue curves)
wave function

1.2 Basic Characteristics of Thick-Shell Core/Shell Quantum


Dots

Core/shell heterogeneous quantum dots are usually composed of two or more semi-
conductor materials. Generally, core/shell heterostructures are classified according
to the energy band relative position and arrangement of core/shell semiconductor
materials, mainly including type-I, type-II and quasi-type-II core/shell heterostruc-
ture, as shown in Fig. 1.
Type-I core/shell QDs refer to the shell material with a higher conduction band
energy level than the core material and the shell material with a lower valence
band energy level than the core material. That is to say, the highest valence band
energy level and the lowest conduction band energy level of the nuclear material are
all within the forbidden band of the shell material, such as CdSe/ZnS, CdSe/CdS,
CdS/ZnS, et al. The most commonly used shell materials are CdS, ZnS, and ZnSe
with a wide band gap. The emission of QDs is mainly determined by the band gap
of the core QDs. The biggest advantage of this structure is that it can inorganically
passivate the QD core surface. Most of the photogenerated carriers are confined in
the core. The excitons are blocked by the shell, which reduces the probability of
being trapped by the external surface defects. The QYs and PL stability of QDs are
improved obviously.
In the structural design of type-I core/shell QDs, in addition to the band gap
meeting the above requirements, the lattice mismatch between the core material and
200 L. Zhang et al.

the shell material should also be considered. For example, the lattice mismatch of
CdS and CdSe is about 4%. More than ten layers of CdS shell can be epitaxially
grown on the surface of a CdSe core [9]. When the lattice constant difference is
large, the epitaxial growth is difficult to be carried out. For example, the lattice
mismatch of ZnS and CdSe is ~12%, and only a few layers of epitaxial growth can be
completed. Thicker shell growth will lead to lattice dislocation and result in defects
by the stress release in the interface or shell, thus inducing the decrease of the PL
QYs [10, 11]. Although the lattice constants of CdS and ZnSe are smaller than ZnS,
they are easier to grow as thicker shells on the surface of the CdSe core, but their
band gap width is also smaller, which is not as good as the ZnS material in improving
the photochemical stability of the CdSe core. Whereas ZnS has a large bandwidth,
the lattice mismatch with the CdSe core is too large, which is difficult to meet the
high-quality epitaxial growth of thicker shells. In order to solve these problems, D.V.
Talapin et al. used CdS or ZnSe as the buffer layer of a ZnS shell epitaxy [10]. In
2005, R.G.Xie et al. further developed such design idea by introducing an alloyed
buffer layer with lattice constant between CdS and ZnS [11]. This kind of alloy
buffer structure with gradual change of component proportion realizes the gradual
change of lattice constant from core to shell, reduces the lattice mismatch between
core/shell materials, and effectively eliminates the defect states induced by lattice
stress release at the core/shell interface or shells. At the same time, due to the smooth
interface potential barrier, the excitons can rapidly relax from the shell to the core,
thus improving the exciton recombination efficiency in the cores [12]. The structure
design of the alloy buffer layer greatly improves the optical properties and promotes
the extensive research of the thick-shell quantum dots.
In type-II core/shell QDs, the energy levels of the QD core material and the
shell material are staggered, so the electrons are confined in the material with
lower conduction band, while the holes are confined in the material with higher
valence band. The emission of QDs is at the core/shell material interface, which is
determined by the energy levels of core and shell materials. So the exciton properties
would be controlled by both the energy levels of core and shell materials. The QDs
with type-II structure have photoelectric properties that are neither core nor shell
materials, so the type-II core/shell structure is called the “energy band engineering”
of QDs. In 2003, Bawendi group first reported CdTe/CdSe and CdSe/ZnTe type-II
core/shell QDs [13]. The confinement of electron/hole wave function in different
materials would reduce the radiation recombination probability between electrons
and holes. So this kind of heterostructure QDs usually has lower QYs and longer
lifetime. The type-II core/shell structure realizes the complete separation of carriers
in quantum dots, which is of great significance for the applications of QDs in fields
of solar cells, photocatalysis, and photodetectors.
In addition to the core/shell heterostructure with carrier common confinement
(type-I core/shell structure) and carrier separation (type-II core/shell structure), a
quasi-type-II core/shell structure is that one carrier confinement in the core material,
and the other carrier can be distributed in the core/shell materials. Such core/shell
heterostructures have conduction band or valence band with similar energy, such as
CdSe/CdS and CdSe/ZnSe. They have very high PL QYs. For example, the quantum
Thick-Shell Core/Shell Quantum Dots 201

yield of CdSe/CdS quantum dots can be up to 100% [14]. Due to the weakening of
the quantum confinement effect, electrons in the core of quasi-type-II core/shell
QDs can continuously diffuse to the shells, which is different from type-I core/shell
QDs, so the PL and absorption spectra gradually redshift with the growth of the
shells. It is worth noting that the degree of superposition of the electron hole wave
function in quasi-type-II core/shell QDs can be controlled by changing the thickness
of the shell. Among them, CdSe/CdS core/shell quantum dots are widely studied as
a classical model.
Generally, with the increase of the shell thickness of CdSe/CdS core/shell
quantum dots, the quantum dots transition from a structure similar to type-I to a
quasi-type-II structure and then further change to a structure similar to type II. In
this process, the electron wave function in the QDs gradually diffuses to the CdS
shells, while the hole is still mainly confined in the CdSe core, so the degree of
overlap of the electrons and hole wave function gradually decreases. In the thick-
shell QDs, in addition to the reduced superposition of the exciton wave functions,
the Coulomb repulsion interaction among the multiexcitons is also conducive to
the suppression of the nonradiative Auger process and the improvement of the
optical gain performance of the QDs. At the same time, the Coulomb interaction
between the multiexcitons could be controlled by changing the size of the core and
shell thickness, thus causing the relative blue-, red-, and none-shift of the amplified
spontaneous emission (ASE) relative to the peak of spontaneous emission [15]. For
the past 10 years, the optical properties of thick-shell QDs have attracted more
attention from researchers, whether it is the design concept of alloyed buffer derived
from type-I core/shell QDs or the adjustment of the distribution of exciton wave
function through the regulation of core size and shell thickness in quasi-type-II
core/shell QDs.
In addition to the influence of the band gap between different core/shell materials
on the band structure, the thick shells can also greatly increase the absorption cross
section of the QDs. For example, the single-photon (∼1.6 × 10−13 cm2 ) and two-
photon (~1.1 × 105 GM) absorption cross section of the thick-shell CdSe/CdS QDs
is at least one order of magnitude higher than that of the CdSe core QDs [16, 17].
High absorption cross section contributes to the generation of more photogenerated
carriers in the shells, which could relax to the CdSe cores rapidly and construct
the population inversion, thus reducing the optical gain thresholds. Besides, the
thick shells could effectively avoid the influence of the external environment on
the luminescent center of the QD core, thus enhancing the photochemistry and
photophysical stability of the QDs [6]. However, in the ultrathick-shell QDs, the
stress-release-induced defect states would lead to the degradation of the optical
properties, such as the decrease of QYs and more serious PL blinking with the
increase of the shell [18]. Therefore, the most excellent optical properties of thick-
shell QDs can be obtained with appropriate shell thickness.
202 L. Zhang et al.

2 Synthesis of Thick-Shell Core/Shell Quantum Dots

2.1 Successive Ionic Layer Adsorption and Reaction Method

With the development of high-temperature organometallic synthesis, the size and


shape of core/shell heterostructures can be controlled, and the types of material
combinations are far more than that synthesized by the low-temperature approach.
In 2003, X. G. Peng group applied the “successive ionic layer adsorption and
reaction” (SILAR) method for the epitaxial growth of shell materials on the surface
of QDs and obtained high-quality CdSe/CdS core/shell QDs with different shell
thicknesses [19]. Soon, Peng group measured the extinction coefficients of some
binary quantum dots in different sizes, which paves a foundation for the accurate
calculation of the shell thickness [20]. The thickness of the shell can be controlled
more accurately by calculating the amount of precursor needed to grow each shell
monolayer on the surface of the QD core and then injecting the anion and cation
precursor alternately. The injection amount of the shell precursor is just completely
adsorbed on the surface of the QDs, which successfully avoids the problem of self-
nucleation of the shell precursor. In the synthesis process, the growth defects can
be reduced by the annealing process for each single molecular shell layer, so as to
improve the optical properties of quantum dots. The SILAR method is simple and
easy to operate, which has become the most commonly used method to synthesize
thick-shell quantum dots. In 2008, Hollingsworth et al. synthesized ultrathick-shell
CdSe/CdS QDs with 19 CdS MLs by the SILAR method [6]. Compared with the
traditional quantum dots, the thick-shell quantum dots exhibit the high optical and
chemical stability. The QYs of the QDs after purification for several times are
nearly unchanged (±5%). At the same time, there is no photobleaching under laser
irradiation for several hours.
However, the volume of the shell precursor increases with the increase of the
shell thickness, indicating that the chemical potential of the monomer in the solution
is too high, which makes the quantum dots tend to anisotropically grow, and the
shape of the QDs is not easy to control. Then, in 2007, the researchers report the
improved SILAR and proposed the thermal cycling alternate ionic layer adsorption
and reaction (TC-SILAR) method [21]. This method uses the principle of crystal
growth kinetics. For the same precursor concentration, low precursor activity in
a low-temperature environment helps in the uniform adsorption of monomers on
the surface of quantum dots; a high-temperature environment helps to improve the
crystallinity of quantum dots and reduce lattice defects. The TC-SILAR method
is not only used for the preparation of various complex core/shell QDs, but also
for the synthesis of metal nanocrystals and rare-earth-doped core/shell crystals
[22]. With the development of synthesis methods, the growth of the core/shell
heterostructure with controllable size and shape has been realized, and the types of
material combinations have become diverse, such as CdSe/CdS/ZnS core/shell/shell
structure, CdSe/CdZnS core/alloy-shell, and CdSe/Cdx Zn1-x S core/shell/gradient
shell structure, while the optical properties have also been greatly improved.
Thick-Shell Core/Shell Quantum Dots 203

2.2 Syringe Pump Injection

The synthesis method of thick-shell QDs is chosen according to the precursor


activity and crystal structure growth requirements. In the preparation of phase-
pure CdSe/CdS core/shell QDs, the higher growth quality and excellent optical
properties of QDs are related to the slower precursor injection and higher reaction
temperature. Moreover, the simultaneous and continuous injection of anionic and
cationic precursors by syringe can further save manpower and time cost.
In 2013, Bawendi group reported the synthesis of such high-quality CdSe/CdS
core/shell QDs in an optimized process that maintains a slow growth rate of the
shell through the use of low-activity octanethiol and cadmium oleate as precursors.
The cation and anion precursor were injected into the core/shell QDs of phase-
pure wurtzite CdSe/CdS by syringe at 310 ◦ C high temperature [23]. The alloyed
interface layer was gradually formed during the growth of CdS shells at high temper-
ature, which helped minimize defects within the shell and on the surface/interface,
thus obtaining the high optical quality QDs. In 2014, Christodoulou and Moreels et
al. synthesized phase-pure wurtzite CdSe/CdS giant-shell nanocrystals with a CdS
shell thickness of up to 7–8 nm (equivalent to about 20 MLs of CdS) [24]. They
take sulfur powder and cadmium oleate as precursors and injected the precursors
continuously through a syringe at a high temperature of 300 ◦ C. The average
diameter of QDs could be up to 18.1 nm over the growth course of 4 h. Both the
core and shell have a wurtzite crystal structure, yielding epitaxial growth of the
shell and nearly defect-free crystals. As a result, the PL QYs is as high as 90%. The
results show that the slow injection of precursor is beneficial to obtain high optical
performance ultrathick-shell QDs.

2.3 “Flash” Hot Injection

With the trend of industrial production and application of QDs, thick-shell core/shell
QDs have become a main choice because of its excellent photostability. In order to,
as far as possible, save time on the preparation of thick-shell QDs, the researchers
put forward some experimental scheme for the rapid synthesis of thick-shell QDs.
In 2014, Cirillo group developed a “flash” method to rapidly synthesize thick-shell
CdSe/CdS core/shell QDs [25]. It only takes 3 min and the shell thickness can reach
~20 MLs. However, the rapid shell growth results in the occurrence of excessive
defects, with the PL QYs at only 15%.
A thick-shell heterostructure also plays an important role in Mn-doped core/shell
QDs. In order to avoid lattice exclusion for Mn2+ dopants, the reaction temperature
should not be too high, and the growth time of shells should not be too long. In
2015, Xu et al. studied the flash preparation of Mn-doped CdS/ZnS QDs with thick
shells [26]. When the temperature is over 300 ◦ C, the crystallization of the shell
is improved, while the defect states are decreased. In addition, the high precursor
204 L. Zhang et al.

concentration and few QD core seeds could promote the growth of thick shells.
In this synthesis system, high concentration of oleic acid can be used to avoid
the nucleation of ZnS shell materials. The thick ZnS shell was overcoated at the
injection and growth temperature of 340 and 315 ◦ C, respectively, and ZnS shell
with thickness up to 18 MLs could be grown within only 9 min. The PL QYs of
CdS/ZnS QDs with a thick shell of a 14 ZnS molecular layer doped with manganese
can reach 36%. The experimental results of optical stability show that the thick-shell
QDs have high optical stability and narrow size distribution.
The narrower size distribution and higher crystallinity of Mn-doped QDs can
be prepared by nucleation doping at high reaction temperature [27]. Especially, for
the Mn2+ -doped QD crystal nucleus prepared by the high-temperature hot injection
method, the distribution of the Mn2+ ion is close to the inner layer of the crystal
nucleus. The nucleus diffusion is relatively matched with the high temperature of
shell growth; it can promote the diffusion of the Mn2+ ion and ensure a higher
rate of exciton energy transfer to the Mn2+ ion, which is conducive to obtaining
a higher PL QY [28]. The perfect combination strategy of hot-injection nucleation
doping and optimized “flash” synthesis goes beyond the combination strategy of
one-pot growth doping and typical “flash” synthesis, which led to an increase in PL
QYs of giant Mn-doped CdS/ZnS QDs (~18 MLs) from ~20% to 40%. The PLQY
was enhanced to 45% by light annealing. Using traditional LED as the reference,
these simply-encapsulated QDs exhibited the higher photostability. Therefore, in
order to find an efficient way to prepare QD materials that can meet the practical
application requirements, the synthesis approaches for thick-shell core/shell QDs
are also always in development.

3 Optical Properties of Thick-Shell Core/Shell Quantum


Dots

3.1 Fundamental Optical Properties

In the thick-shell core/shell QDs, the inorganic wide band gap shells can passivate
the surface defects of QDs and reduce the probability of exciton trapping on the
surface/interface, thus improving the optical stability of the QDs.
First of all, the thick shells can passivate the surface defect states and reduce the
nonradiative loss from the surface trapping of exciton. As shown in Fig. 2, with the
increase of the shell thickness, the single exciton relaxation process in the thick-
shell CdSe/CdS QDs gradually tends to the single exponential dynamic behavior,
which can be well fitted by the single exponential function. χ2 is smaller than 1.2,
suggesting the acceptability of the PL decay fitting, thus showing that the thick
CdS shells effectively reduce the surface states related to the exciton nonradiative
recombination [16]. Therefore, the QYs of single exciton in thick-shell CdSe/CdS
Thick-Shell Core/Shell Quantum Dots 205

Fig. 2 PL decay trace of the


CdSe cores and their
corresponding core/shell QDs
with different shell
thicknesses. The solid line
stands for the
single-exponential fittings
using the function
I(t) = Aexp(−t/τ) + B [16]

Fig. 3 Integrated PL
intensities of CdZnSeS cores
and CdZnSeS/ZnS (3, 11, and
17 ML) core/shell QDs as a
function of exposure time
[30]

QDs could be as high as 100% due to the effective suppression of the exciton
nonradiative process [14].
Secondly, the ultraviolet (UV) light stability of CdSe/CdS core/shell QDs is
greatly enhanced with increasing shell thickness [29]. It is considered that the
thicker shell can effectively prevent the surface-related recombination and improve
the photostability of the QDs. Similarly, the shell thickness is also important for the
optical stability of CdZnSeS/ZnS alloy core/shell QDs [30]. As shown in Fig. 3,
the integrated PL intensities were normalized to the initial values of each sample
(CdZnSeS cores and with 3, 11, and 17 MLs). The core emission experienced
a large drop (by ∼70%) after 3 h of UV illumination and decreased to 17% of
the initial intensity after illumination for 136 h. In contrast, the photostability of
206 L. Zhang et al.

CdZnSeS/ZnS QDs with 3 MLs of ZnS was significantly improved, and only a
small drop (by ∼20%) could be found after 136 h of illumination. However, as the
illumination time was continuously increased, the PL intensity constantly decreased
and retained only 47% of its initial intensity after 660 h of UV exposure. Contrary
to the QDs with 3 MLs, the PL intensity of the sample with 11 MLs of ZnS slightly
increased during light irradiation, and the enhanced PL was maintained after 660 h
of illumination, which was about 1.3 times stronger than the initial intensity. The
enhanced emission may be due to photo-oxidation-induced passivation of defect
sites on QDs. The sample with 17 MLs of ZnS underwent less pronounced decrease
of the emission intensity (25% decrease). These data demonstrated that a thick ZnS
shell could effectively improve the photostability of QDs. This is mainly because
the coating of the shell makes the exciton wave function of the QDs far away from
the surface defects. The smaller overlap between the exciton and the surface trap
wave function in thicker shells leads to lesser sensitivity of QDs to the external
environment. The thick-shell core/shell green QDs with 11 ML ZnS exhibit the best
optical stability. However, in ultrathick-shell CdZnSeS/ZnS QDs coated with 17 ML
ZnS, with the further increase of the ZnS shell, the lattice dislocations are generated
at the core/shell interface and release part of the stress, which leads to the generation
of interface defects and the reduction of QYs.

3.2 Exciton Dynamics Process

In 2009, Klimov group reported a kind of “giant” QDs to suppress AR. The
“giant” QD consists of a small-size CdSe core (3–4 nm) and a thick CdS shell
(more than 10 ML CdS) with a wide band gap growing on the surface of the
core, as shown in Fig. 4a [31]. Thick-shell CdSe/CdS QDs exhibit a significant

Fig. 4 (a) A schematic of “giant” CdSe/CdS quantum dots; (b) spatial probability distribution of
the hole and electron. The inset shows a contour plot of the electron-hole overlap integral [31]
Thick-Shell Core/Shell Quantum Dots 207

(orders of magnitude) suppression of Auger decay rates, thus greatly improving the
radiative recombination efficiency of multiexcitons in QDs. The biexciton QYs of
the ultrathick-shell CdSe/CdS QDs with 19 ML CdS and the thick-shell CdSe/CdS
QDs with gradient shell can be as high as 100% [32, 33]. Klimov et al. analyzed the
physical mechanism behind this huge inhibition and believed that it was the result
of at least three factors:
1. Giant volume
In general, the Auger lifetime is directly proportional to the quantum dot volume
[34]. In giant QDs, the electron wave function expands in a huge space, which will
lead to the increase of the lifetime of the biexciton.
2. Partial separation of electron and hole wave function
In giant QDs, the electron delocalization to the shell region and the hole is
mainly confined in the core (Fig. 4b) with the increase of shell thickness. Therefore,
the spatial separation of electron and hole wave function can reduce the Coulomb
interaction between electrons and holes and thus help to reduce the Auger rates
[35].
3. Alloy interface
The alloy interface properties of the giant QDs are very similar to those of the
epitaxial grown quantum dots. Although the synthesis temperature is not high, the
long reaction time (about 40 h for the synthesis of QDs with 10 ML CdS in shell
thickness) makes the anions diffuse significantly at the core/shell interface and form
a transition alloy layer with a smoothed interface potential, which would contribute
to the diffusion of electron wave function, thus inhibiting the Auger process of
excitons [36].
Generally, the biexciton dynamics in the thick-shell CdSe/CdS QDs could be
extracted by subtracting the low-fluence (average QD occupancy N   1) PL
traces from the “tail-normalized” traces measured for N  slightly above unity
[34]. The biexciton lifetimes (τ2 ) were derived from the obtained decay traces
by fitting function. The biexciton lifetime can reach 29 ns with increasing shell
thickness from 0 to 19 MLs [37]. The biexciton Auger lifetime (τ2A ) can be further
estimated using these results. τ2A can be obtained from τ2A −1 = τ2 −1 −τ2r −1 . The
measured biexciton lifetime, τ2 , is limited by both biexciton radiative lifetime (τ2r )
and biexciton Auger lifetime (τ2A ) (here nonradiative decay channels other than
AR are neglected). The τ2r in QDs is shorter than τ1r by a factor β (range from 3 to
4), which relates to the relative number of recombination routes [36]. The obtained
values of τ2A indicate more than a thousandfold increase in the Auger lifetime with
increasing shell thickness. It changes from 67 ps for the core-only QDs to 102–
593 ns (depending on β) for the QDs with 19 ML CdS shell, which indicates that
the nonradiative AR of excitons is effectively suppressed.
In addition, the PL intermittence of quantum dots usually comes from the
photoionization of excitons and the trapping of excitons by defect states. Therefore,
208 L. Zhang et al.

Fig. 5 Average number of excitons per particle (N ) as a function of pump intensity measured for
CdSe/CdS QDs (5, 11, and 15 MLs). The solid lines are on behalf of their theoretical relationships.
The dashed lines denote the variation trend of the number of higher-order excitons at higher pump
intensities [17]

the PL blinking can be obviously suppressed in the thick-shell QDs, because


the thick-shell structure can effectively reduce the surface defect states and AR.
Particularly, in wurtzite CdSe/CdS QDs, there are almost no defects in the core/shell
interface due to the alloy interface layer formed at high temperature for a long
reaction time. Thereby, the phase-pure CdSe/CdS QDs have a very low non-blinking
volume threshold (~390 nm3 ), which is significantly lower than the traditional thick-
shell QDs (~750 nm3 ) [5, 23].
Furthermore, the complicated fast multiexciton dynamics in phase-pure wurtzite
CdSe/CdS QDs can be analyzed from the view of the superradiation mechanism
[17]. The relationships between the average number of excitons per QD (N ) and
pump intensity can be obtained by fitting PL decay traces using a superradiance
decay function: I(t) = Im csch2 ( t+ττr ) + Bexp(−αt) + C, as shown in Fig. 5. For all
d

of the QD samples (with 5, 11, and 15 ML CdS), there was a common characteristic
that the N value was proportional to pump intensity within a certain range of
pump intensity. At the same time, the data points were almost in perfect conformity
with their own theoretical linear relationships (solid lines). However, when the pump
intensity continued to increase, the average number of excitons (>6) in QDs deviated
from the theoretical values and started to exhibit different saturation trends (dashed
lines), because the nonradiation losses of excitons increase with the pump intensity.
At the same time, the ASE measurements of CdSe/CdS QDs (11 MLs) prove that
the fitting results of N  (2, 7.6, and 15.8) at the three ASE thresholds are basically
conformed to the theory (2, 6, and 14), [38] further indicating the correctness of the
fitting scheme. These results give us more insights into the multiexciton states in
the fast PL decay process and provide a new way from the view of superradiance
mechanism to study the multiexciton dynamics of CdSe/CdS QDs.
Thick-Shell Core/Shell Quantum Dots 209

3.3 Quantum Confined Stark Effect

In 1984, the quantum confinement stark effect was first observed in quantum well
structures [39]. In 1997, Bawendi group first reported the quantum confinement
stark effect of a single CdSe QD [40]. The variation of PL spectra of CdSe QD with
the applied electric field was studied at a temperature of 10 K. Under the applied
electric field, the excited electron-hole pairs were polarized. In the experiment, a
spectral shift of two orders of magnitude wider than the line could be observed.
At the same time, it was also observed that the local field around the quantum
dot changes with the time, resulting in the spectra shift with time. The random
local electric field mainly originates from the charged state generated by the surface
trapping states of excitons, AR process, and the ionization of exciton under a strong
electric field. In 2012, Park et al. studied the quantum confinement stark effect of
a single quantum dot with different heterostructures at room temperature [41]. The
research results show that the initial separation distance between electrons and holes
in a quantum dot determines the strength of the quantum confinement stark effect
under an external electric field.
Semiconductor QDs are always suggested as promising candidates for electric
field (voltage) sensing via the quantum-confined Stark effect (QCSE) [40–43].
However, the charged carriers on or near the QD-surface-induced randomly oriented
local electric field would lead to spectral broadening in the ensemble, which
is difficult for QDs to exhibit an obvious ensemble Stark effect for potential
applications. If the local field is eliminated, QCSE in QD ensembles can be readily
observable at room temperature, and the response of the absorbing state can be
possibly characterized in the ensemble measurement. The phase-pure wurtzite
CdSe/CdS core/shell QDs have been proven to efficiently inhibit AR and the
consequent PL blinking [23, 37]. Hence, the thick-shell wurtzite CdSe/CdS QDs
can inhibit the random local field induced by nonradiative AR or defect states in
favor of the applications based on QCSE, such as optical switch and optoelectronic
sensors.
Figure 6a shows the PL spectra of thick-shell CdSe/CdS QDs (11 MLs) with
increasing field strength [44]. The PL spectra showed an obvious Stark shift in
the peak position with a redshift of ∼2.3 nm. The spectral shifts could be well
fitted by a purely quadratic function of the electric field (Fig. 6b). At the same
time, the PL intensity showed remarkable quenching of ∼17% and the PL spectra
slightly broadened by ∼6%, as expected features of QCSE in semiconductor
nanoparticles. Figure 6c shows the absorption spectra of thick-shell CdSe/CdS
QDs under different electric fields. On increasing the electric field strength, the
redshift (∼2.1 nm) in the first exciton absorption peak also exhibited quadratic
dependence on the external electric field (Fig. 6d), concomitant with ∼9% decrease
in the absorption intensity and ∼10% increase in the full width at half maximum.
The results provide a possibility to efficiently tune the absorption properties by
an external field, which is meaningful for the applications of QD-based QCSE in
electroabsorption modulators. However, the field-dependent optical properties in the
210 L. Zhang et al.

Fig. 6 (a) PL spectra of thick-shell CdSe/CdS QDs (11 MLs) in the QCSE device. (b) The Stark
shift versus electric field for the PL spectra in (a). (c) Absorption spectra of thick-shell CdSe/CdS
QDs under the electric field. The spectra are vertically shifted for clarity. (d) The Stark shift in the
first excitonic absorption peak for the spectra in (c) [44]

moderate-shell (5 ML) CdSe/CdS QDs showed nearly no spectral shifts, while the
shifts for ultrathick-shell (15 ML) CdSe/CdS QDs were not as obvious as that for the
thick-shell QDs. The time-dependent PL measurements at the single-particle level
revealed that moderate-shell and ultrathick-shell QDs were likely to suffer from the
randomly local electric field, stemmed from a large number of deep surface traps and
strain-induced defect states, respectively. Therefore, the study suggested that thick-
shell CdSe/CdS QDs could minimize the adverse impact of the random local electric
field induced by the charged exciton states, which is crucial for the applications of
QCSE in QD ensembles.

3.4 Optical Gain Performance

The suppression of AR can greatly improve the QDs’ optical gain performance.
At the same time, the large absorption cross section of the thick-shell QDs is also
Thick-Shell Core/Shell Quantum Dots 211

Fig. 7 Band structure of a


CdSe/CdS core/shell g-QD
features an energetic barrier
in the valence band associated
with an interfacial ZB CdS
layer, which separates a thick
WZ CdS outer layer from the
ZB CdSe core [8]

helpful to reduce the threshold of ASE. Therefore, the ultralow gain threshold
(~26 μJ cm−2 ) and larger optical gain bandwidth (>500 meV) is realized in the
“giant” CdSe/CdS QDs with 11 ML CdS [31].
In order to suppress the Auger process and improve the photostability of QDs,
further studies show that the quantum dots need to be coated with thicker inorganic
shells [6, 33]. However, there is a conventional problem in the synthesis of ultrathick
CdSe/CdS core/shell QDs. In the schematic diagram shown in Fig. 7, the CdSe
core has a zinc-blende phase crystal structure at the usual synthesis method, the
first layers of CdS shells will grow into zinc-blende phase due to their epitaxial
growth on the surface of CdSe core, and then the outer CdS shells will change into
the wurtzite phase due to the higher growth temperature and long reaction time.
The CdS shell with zinc-blende phase plays a role of hole potential barrier in the
valence band, which will block the relaxation rate of the hole from the shells to
the core. Therefore, the radiative recombination of excitons among the high energy
levels of shells becomes competitive. Under the excitation of strong subpicosecond
pulses, these ultrathick-shell QDs exhibit an ASE from CdS shells similar to the
bulk CdS material, but ASE cannot be realized in the quantum-confined CdSe cores
[7, 8]. For quantum dot laser applications, the ASE from the quantum confinement
core is better than that from the shells. In 2005, Liao et al. reported that the
interfacial potential barrier in a thick-shell CdSe/CdS QDs could be removed by
fs laser annealing and consequently transferred ASE from a bulk-like CdS shell to a
quantum-confined CdSe core [8].
To obtain the high-performance ASE from the quantum-confined CdSe core,
the interface hole barrier can be completely eliminated in phase-pure wurtzite
CdSe/CdS core/shell QDs [37]. Emission from the phase-pure wurtzite CdSe/CdS (4
and 11 ML) QD film edge with increasing pump intensity exhibits a clear transition
from spontaneous emission to ASE through an abrupt increase in output intensity
and spectral narrowing (Fig. 8a–e). In the QDs with a 4 ML CdS shell, the ASE
peak (654 nm) is redshifted (∼11 nm) with respect to the spontaneous emission
peak, because of attractive exciton-exciton interaction. In the QDs with an 11 ML
CdS shell, the longer-wavelength ASE peak (635 nm) corresponding to the first (1S)
electron quantization shell of the CdSe core is blueshifted (∼14 nm), because of the
212 L. Zhang et al.

Fig. 8 (a–c) Per-pulse pump intensity-dependent emission spectra of a close-packed film of


CdSe/CdS core/shell QDs with 4, 11, and 19 ML CdS shell, respectively. Emission intensity versus
pump intensity at the positions of ASE and spontaneous peaks observed for CdSe/CdS QDs with
(d) 4 and 11 ML and (e) 19 ML CdS shell. (f) ASE thresholds for biexcitonic gain as a function of
CdS shell thickness [37]

formation of quasi-type-II band alignment and subsequent repulsive exciton-exciton


interaction. As the pump intensity further increases, two shorter-wavelength ASE
peaks (at 580 and 535 nm) corresponding to optical transitions involving the second
(1P) and the third (1D) quantization shells also appear. Within a simple particle-in-
a-box model [38], the third electron-quantized level becomes a population inversion
only if the QD contains at least 14 excitons. High-order multiexcitons exhibit high
emission efficiencies and contribute to optical gain, indicating that AR is effectively
suppressed.
The ASE threshold for biexcitonic gain (Fig. 8f) decreased dramatically with the
increase of the CdS shell toward 11 MLs, which results from increased absorption
cross section and passivated surface, in addition to suppressed Auger process. The
ASE thresholds are increased with further shell growth because strain-induced
defects will weaken the suppression of AR. In addition, the ultrathick shells would
more reduce wave function overlap between shell-based electron states and core-
localized hole states, therefore slowing the relaxation of holes into the core. At the
same time, both nonradiative recombination involving surface and interface trap
states and radiative recombination involving shell-based states become competitive.
In the QDs with a 19 ML CdS shell, the ultrathick shell slows the relaxation to a
degree which is sufficient for obtaining population inversion of shell-based states
and then observed an ASE from the CdS shell instead of the CdSe core.
Thick-Shell Core/Shell Quantum Dots 213

The transient spectra obtained with a streak camera exhibit clearly the evolution
of ASE emission of QDs with the pump intensity [37]. At low pump intensity, only
a broad long-decayed spontaneous peak was observed. As the intensity increases, a
narrower ASE peak appears. As expected, the emission lifetime is shortened to near
the camera time resolution of ~10 ps, which is more than 3 orders of magnitude
faster than the Auger lifetime and implies highly efficient ASE. Noticeably, at
the position of the gain emission band, the optical gain lifetime exceeds 1000 ps,
which is primarily limited by the multiexciton AR [45]. The extremely long gain
lifetime (>1000 ps) of the phase-pure QDs is about an order of magnitude longer
than that of thin-shell QDs (>100 ps) [46] and more than 2 times greater than that
of conventional giant QDs [47] and similar dot/rods (>400 ps) [48]. As the pump
intensity further increases, the bandwidth of optical gain of the QDs exceeds 170 nm
because of the participation of higher-order multiexcitons. These results imply that
using single-size QDs one can tune the lasing color almost over the entire range of
visible wavelengths by simply adjusting the laser cavity.
In 2018, Klimov group reported population inversion and optical gain in contin-
uously graded colloidal CdSe/Cdx Zn1 − x Se/ZnSe0.5 S0.5 core/shell QDs under the
direct-current electrical pumping [49]. The considerable suppression of Auger decay
can be outpaced by electrical injection. The special device architecture allows them
to produce high current densities up to ~18 A cm−2 without damaging either the
QDs or the injection layers. The quantitative analysis of electroluminescence and
current-modulated transmission spectra indicates that the population inversion of
the band-edge states was achieved with current densities of 3–4 A cm−2 .

3.5 Stimulated Radiation from Thick-Shell Core/Shell


Quantum Dots

Particularly, for the CdSe/CdS core/shell structure, both thick CdS shells and
alloyed interfacial layers with a smooth confinement potential have been proven
to be efficient ways to suppress AR. Thick-shell CdSe/CdS QDs with quasi-type-
II band alignment, concurring strong electron delocalization, efficiently decrease
the rate of AR and simultaneously reduce or even suppress blinking [5, 6, 31].
Therefore, the ASE regime will be realized in thick-shell QDs with a lowered gain
threshold. Besides, CdSe-based QDs show almost temperature-independent optical
gain performance because of the well-separated electronic states [50]. All of these
properties clearly state that thick-shell CdSe/CdS QDs have many incomparable
merits that make them a superior laser gain medium.
Several optical microstructures have been applied as resonant cavities in laser
devices, such as microring [37, 51], microsphere resonators [52, 53], and distributed
feedback (DFB) gratings [46, 54], which have been used as optical feedback
structures in QD lasers. With the assistance of these resonant cavities, the laser per-
formance can be improved markedly. Among the available options, DFB structures
214 L. Zhang et al.

Fig. 9 The dashed frame shows the fabrication process of DFB structures by LIA. Atomic force
microscopy (AFM) images of the DFB structures. Lasing emission spectra of the DFB devices [16]

have the advantages of single-mode emission, low threshold, high quality factor,
and tunable lasing wavelength by changing the grating period and thickness of the
gain layer. Nanoimprint lithography and soft lithography are common approaches
to construct DFB gratings. Furthermore, DFB lasers can be fabricated by laser
interference ablation (LIA) based on thick-shell CdSe/CdS core/shell QDs; the
schematic of the experimental setup for fabricating DFB QDs gratings is shown
in Fig. 9 [16]. The prepared films were exposed to a laser interference pattern
for about 15 min, and their surfaces were structured to acquire periodic relief
patterns with uniform structure distribution and good repeatability due to the laser-
induced ablation effect. The single-mode laser emission with high performance of
the QDs’ distributed feedback structure is realized under single-photon and two-
photon excitation. The laser emission peaks are all at 645 nm, the linewidth is about
~1 nm, and the laser threshold is 0.028 and 1.03 mJ cm−2 , respectively.
In 2015, Park et al. reported random lasers of thick-shell CdSe/CdS core/shell
QDs with alloyed interface layer that realized single-mode lasing emission with low
threshold of ~18 μJ cm−2 [55]. In 2015, Moreels et al. fabricated ultralow threshold
(~10 μJ cm−2 ) nanorod microcavity lasers based on a coffee-ring microstructure by
using CdSe/CdS quantum dot-in-rods as gain medium. The laser linewidth was as
narrow as ~0.79 nm, and the quality factor was estimated to be as high as ~1000
according to Q = λ/Δλ, where λ and Δλ are the wavelength and line width of
the lasing peak, respectively [56]. The fabrication process of the self-assembled
microstructure is very simple, which can achieve high-performance laser emission.
Liao et al. also reported self-assembled phase-pure thick-shell CdSe/CdS QDs
with low-threshold gain performance into a coffee-ring microcavity (Fig. 10a) and
studied its laser emission properties [37]. The evaporation dynamics of the droplets
Thick-Shell Core/Shell Quantum Dots 215

Fig. 10 (a) Optical microscopy image of a CdSe/CdS QD (11 ML CdS shell) coffee ring. The
scale bar represents 200 μm. (b) AFM image of the bottom left part of the coffee ring shown in
panel a. (c) Per-pulse pump intensity-dependent emission spectra of the QD coffee ring shown in
panel a. The inset shows the emission intensity versus pump intensity at the position of the lasing
peak [37]

are governed by the “coffee-ring effect,” which leads to the formation of well-
defined micrometer-size rings. The AFM image of the microlaser (Fig. 10b) shows
that the ring has a full width at half maximum of 7.7 μm and a height of 350 nm. As
a result, the microlaser displays single-mode operation (at 636 nm, with line width
of ∼0.3 nm) and an ultralow threshold of ∼2 μJ cm−2 (Fig. 10c). The quality factor
of the microlaser is as high as ∼2000, indicating the high quality of the coffee-ring
microlaser.
With the development of synthesis technology and exciton dynamics theory of
QDs, the relevant research tends to the single exciton low-threshold lasing. In 2019,
Klimov group reported single-mode lasing with low thresholds using thick-shell,
multilayered CdSe/Cdx Zn1-x Se QDs combined with second-order DFB resonators
[57]. An extremely low and sub-single-exciton lasing threshold <Nlas > ≈0.3 was
achieved by charging the sample with about three extra electrons per dot on average.
216 L. Zhang et al.

The result is more than fourfold reduction compared with the neutral QDs. Further-
more, they also show a reversible tuning of the lasing threshold by controlling the
degree of QD charging. Their approaches facilitate the development of solution-
processable lasing devices, thereby extending the reach of lasing technologies into
areas not accessible with traditional or epitaxially grown semiconductor materials.

4 Conclusion

In summary, the optical properties of quantum dots are improved greatly in thick-
shell heterostructures compared to traditional core/shell quantum dots. Thick shells
not only efficiently decrease the rate of AR and simultaneously reduce or even
suppress blinking, but also decrease the surface nonradiative channel and increase
PL QYs and absorption cross sections. Therefore, thick-shell QDs exhibit superior
optical gain performance and photostability, thus contributing to the application
of quantum dots in the fields of microlaser, LED, optoelectronics sensors, and
fluorescence labeling.

References

1. Norris, D.J., Sacra, A., Murray, C.B., et al.: Measurement of the size dependent hole spectrum
in CdSe quantum dots. Phys. Rev. Lett. 72(16), 2612–2615 (1994)
2. Wang, L.W., Califano, M., Zunger, A., et al.: Pseudopotential theory of Auger processes in
CdSe quantum dots. Phys. Rev. Lett. 91(5), 056404 (2003)
3. Kharchenko, V.A., Rosen, M.: Auger relaxation processes in semiconductor nanocrystals and
quantum wells. J. Fluoresc. 70, 158–169 (1996)
4. Reiss, P., Protière, M., Li, L.: Core/Shell semiconductor Nanocrystals. Small. 5(2), 154–168
(2009)
5. Ghosh, Y., Mangum, B.D., Casson, J.L., et al.: New insights into the complexities of
shell growth and the strong influence of particle volume in nonblinking “giant” core/shell
nanocrystal quantum dots. J. Am. Chem. Soc. 134(23), 9634–9643 (2012)
6. Chen, Y., Vela, J., Htoon, H., et al.: “Giant” multishell CdSe nanocrystal quantum dots with
suppressed blinking. J. Am. Chem. Soc. 130(15), 5026–5027 (2008)
7. Galland, C., Brovelli, S., Bae, W.K., et al.: Dynamic hole blockade yields two-color quantum
and classical light from dot-in-bulk nanocrystals. Nano Lett. 13(1), 321–328 (2013)
8. Liao, C., Fan, K., Xu, R., et al.: Laser-annealing-made amplified spontaneous emission of
“giant” CdSe/CdS core/shell nanocrystals transferred from bulk-like shell to quantum-confined
core. Photonics Res. 3(5), 200–205 (2015)
9. Mahler, B., Spinicelli, P., Buil, S., et al.: Towards non-blinking colloidal quantum dots. Nat.
Mater. 7(8), 659–664 (2008)
10. Talapin, D.V., Mekis, I., Götzinger, S., et al.: CdSe/CdS/ZnS and CdSe/ZnSe/ZnS core-shell-
shell nanocrystals. J. Phys. Chem. B. 108(49), 18826–18831 (2004)
11. Xie, R., Kolb, U., Li, J., et al.: Synthesis and characterization of highly luminescent CdSe-Core
CdS/Zn0.5 Cd0.5 S/ZnS multishell nanocrystals. J. Am. Chem. Soc. 127(20), 7480–7488 (2005)
12. Pinchetti, V., Meinardi, F., Camellini, A., et al.: Effect of core/shell interface on carrier
dynamics and optical gain properties of dual-color emitting CdSe/CdS nanocrystals. ACS
Nano. 10(7), 6877–6887 (2016)
Thick-Shell Core/Shell Quantum Dots 217

13. Kim, S., Fisher, B., Eisler, H., et al.: Type-II quantum dots: CdTe/CdSe (core/shell) and
CdSe/ZnTe (core/shell) heterostructures. J. Am. Chem. Soc. 125(38), 11466–11467 (2003)
14. Javaux, C., Mahler, B., Dubertret, B., et al.: Thermal activation of non-radiative Auger
recombination in charged colloidal nanocrystals. Nat. Nanotechnol. 8(3), 206–212 (2013)
15. Cihan, A.F., Kelestemur, Y., Guzelturk, B., et al.: Attractive versus repulsive excitonic
interactions of colloidal quantum dots control blue- to red-shifting (and non-shifting) amplified
spontaneous emission. J. Phys. Chem. Lett. 4(23), 4146–4152 (2013)
16. Zhang, L., Liao, C., Lv, B., et al.: Single-mode lasing from “Giant” CdSe/CdS core-shell
quantum dots in distributed feedback structures. ACS Appl. Mater. Interfaces. 9(15), 13293–
13303 (2017)
17. Zhang, L., Li, H., Liao, C., et al.: New insights into the multiexciton dynamics in phase-pure
thick-Shell CdSe/CdS quantum dots. J. Phys. Chem. C. 122(43), 25059–25066 (2018)
18. Omogo, B., Gao, F., Bajwa, P., et al.: Reducing blinking in small core−multishell quantum
dots by carefully balancing confinement potential and induced lattice strain: the “goldilocks”
effect. ACS Nano. 10(4), 4072–4082 (2016)
19. Li, J.J., Wang, Y.A., Guo, W., et al.: Large-scale synthesis of nearly monodisperse CdSe/CdS
core/shell nanocrystals using air-stable reagents via successive ion layer adsorption and
reaction. J. Am. Chem. Soc. 125(41), 12567–12575 (2003)
20. Yu, W.W., Qu, L., Guo, W., et al.: Experimental determination of the extinction coefficient of
CdTe, CdSe, and CdS nanocrystals. Chem. Mater. 15(14), 2854–2860 (2003)
21. Blackman, B., Battaglia, D.M., Mishima, T.D., et al.: Control of the morphology of complex
semiconductor nanocrystals with a type II heterojunction, dots vs peanuts, by thermal cycling.
Chem. Mater. 19(15), 3815–3821 (2007)
22. Li, X., Shen, D., Yang, J., et al.: Successive layer-by-layer strategy for multi-Shell epitaxial
growth: Shell thickness and doping position dependence in upconverting optical properties.
Chem. Mater. 25(1), 106–112 (2013)
23. Chen, O., Zhao, J., Chauhan, V.P., et al.: Compact high-quality CdSe-CdS core-shell nanocrys-
tals with narrow emission linewidths and suppressed blinking. Nat. Mater. 12(5), 445–451
(2013)
24. Christodoulou, S., Vaccaro, G., Pinchetti, V., et al.: Synthesis of highly luminescent wurtzite
CdSe/CdS giant-shell nanocrystals using a fast continuous injection route. J. Mater. Chem. C.
2(17), 3439–3447 (2014)
25. Marco, C., Tangi, A., Raquel, G., et al.: “Flash” synthesis of CdSe/CdS core–shell quantum
dots. Chem. Mater. 26(2), 1154–1160 (2014)
26. Xu, R., Liao, C., Zhang, H., et al.: “Flash” synthesis of “giant” Mn-doped CdS/ZnS nanocrys-
tals for high photostability. RSC Adv. 5(108), 88921–88927 (2015)
27. Xu, R., Huang, B., Wang, T., et al.: Bright and high-photostable inner-Mn-doped core/giant-
shell quantum dots. Superlattice. Microst. 111, 665–670 (2017)
28. Chen, H.Y., Maiti, S., Son, D.H., et al.: Doping location-dependent energy transfer dynamics
in Mn-doped CdS/ZnS nanocrystals. ACS Nano. 6(1), 583–591 (2012)
29. Huang, B., Xu, R., Zhuo, N., et al.: “Giant” red and green core/shell quantum dots with high
color purity and photostability. Superlattice. Microst. 91, 201–207 (2016)
30. Huang, B., Yang, H., Zhang, L., et al.: Effect of surface/interfacial defects on photostability of
thick-shell CdZnSeS/ZnS quantum dots. Nanoscale. 10(38), 18331–18340 (2018)
31. Garcia-Santamaria, F., Chen, Y.F., Vela, J., et al.: Suppressed Auger recombination in “giant”
nanocrystals boosts optical gain performance. Nano Lett. 9(10), 3482–3488 (2009)
32. Park, Y.S., Malko, A.V., Vela, J., et al.: Near-unity quantum yields of biexciton emission from
CdSe/CdS nanocrystals measured using single-particle spectroscopy. Phys. Rev. Lett. 106(18),
187401 (2011)
33. Nasilowski, M., Spinicelli, P., Patriarche, G., et al.: Gradient CdSe/CdS quantum dots with
room temperature biexciton unity quantum yield. Nano Lett. 15(6), 3953–3958 (2015)
34. Klimov, V.I., Mikhailovsky, A.A., McBranch, D.W., et al.: Quantization of multiparticle Auger
rates in semiconductor quantum dots. Science. 287(5455), 1011–1013 (2000)
218 L. Zhang et al.

35. Jain, A., Voznyy, O., Hoogland, S., et al.: Atomistic design of CdSe/CdS core-shell quantum
dots with suppressed Auger recombination. Nano Lett. 16(10), 6491–6496 (2016)
36. Garcia-Santamaria, F., Brovelli, S., Viswanatha, R., et al.: Breakdown of volume scaling in
Auger recombination in CdSe/CdS heteronanocrystals: the role of the core-shell interface.
Nano Lett. 11(2), 687–693 (2011)
37. Liao, C., Xu, R., Xu, Y., et al.: Ultralow-threshold single-mode lasing from phase-pure
CdSe/CdS core/shell quantum dots. J. Phys. Chem. Lett. 7(24), 4968–4976 (2016)
38. Efros, A.L., Efros, A.L.: Interband absorption of light in a semiconductor sphere. Semiconduc-
tors. 16(7), 772–775 (1982)
39. Miller, D.A.B., Chemla, D.S., Damen, T.C., et al.: Band-edge electroabsorption in quantum
well structures: the quantum-confined stark effect. Phys. Rev. Lett. 53(22), 2173–2176 (1984)
40. Empedocles, S.A., Bawendi, M.G.: Quantum-confined stark effect in single CdSe nanocrystal-
lite quantum dots. Science. 278(5346), 2114–2117 (1997)
41. Park, K.W., Deutsch, Z., Li, J.J., et al.: Single molecule quantum-confined stark effect
measurements of semiconductor nanoparticles at room temperature. ACS Nano. 6(11), 10013–
10023 (2012)
42. Kuo, Y., Li, J., Michalet, X., et al.: Characterizing the quantum-confined stark effect in semi-
conductor quantum dots and nanorods for single-molecule electrophysiology. ACS Photonics.
5(12), 4788–4800 (2018)
43. Achtstein, A.W., Prudnikau, A.V., Ermolenko, M.V., et al.: Electroabsorption by 0D, 1D,
and 2D nanocrystals: a comparative study of CdSe colloidal quantum dots, nanorods, and
nanoplatelets. ACS Nano. 8(8), 7678–7686 (2014)
44. Zhang, L., Lv, B., Yang, H., Xu, R., et al.: Quantum confined stark effect in ensemble of phase-
pure CdSe/CdS quantum dots. Nanoscale. 11(26), 12619–12625 (2019)
45. Xu, Y., Chen, Q., Zhang, C., et al.: Two-photon-pumped perovskite semiconductor nanocrystal
lasers. J. Am. Chem. Soc. 138(11), 3761–3768 (2016)
46. Todescato, F., Fortunati, I., Gardin, S., et al.: Soft-lithographed up-converted distributed
feedback visible lasers based on CdSe-CdZnS-ZnS quantum dots. Adv. Funct. Mater. 22(2),
337–344 (2012)
47. Marceddu, M., Saba, M., Quochi, F., et al.: Auger recombination and optical gain in CdSe/CdS
nanocrystals. Nanotechnology. 23(1), 015201 (2012)
48. Zavelani-Rossi, M., Lupo, M.G., Tassone, F., et al.: Suppression of biexciton Auger recombi-
nation in CdSe/CdS Dot/Rods: role of the electronic structure in the carrier dynamics. Nano
Lett. 10(8), 3142–3150 (2010)
49. Lim, J., Park, Y.S., Klimov, V.I.: Optical gain in colloidal quantum dots achieved with direct-
current electrical pumping. Nat. Mater. 17(1), 42–49 (2018)
50. Moreels, I., Raino, G., Gomes, R., et al.: Nearly temperature-independent threshold for
amplified spontaneous emission in colloidal CdSe/CdS quantum dot-in-rods. Adv. Mater.
24(35), OP231–OP235 (2012)
51. Kiraz, A., Chen, Q., Fan, X.: Optofluidic lasers with aqueous quantum dots. ACS Photonics.
2(6), 707–713 (2015)
52. Chan, Y., Steckel, J.S., Snee, P.T., et al.: Blue semiconductor nanocrystal laser. Appl. Phys.
Lett. 86(7), 073102 (2005)
53. Snee, P.T., Chan, Y., Nocera, D.G., et al.: Whispering-gallery-mode lasing from a semiconduc-
tor nanocrystal/microsphere resonator composite. Adv. Mater. 17(9), 1131–1136 (2005)
54. Saliba, M., Wood, S.M., Patel, J.B., et al.: Structured organic−inorganic perovskite toward a
distributed feedback laser. Adv. Mater. 28(5), 923–929 (2016)
55. Park, Y.S., Bae, W.K., Baker, T., et al.: Effect of Auger recombination on lasing in heterostruc-
tured quantum dots with engineered core/shell interfaces. Nano Lett. 15(11), 7319–7328
(2015)
56. Stasio, F.D., Grim, J.Q., Lesnyak, V., et al.: Single-mode lasing from colloidal water-soluble
CdSe/CdS quantum dot-in-rods. Small. 11(11), 1328–1334 (2015)
57. Kozlov, O.V., Park, Y., Roh, J., et al.: Sub–single-exciton lasing using charged quantum dots
coupled to a distributed feedback cavity. Science. 365(6454), 672–675 (2019)
Core/Shell Quantum-Dot-Sensitized
Solar Cells

Gurpreet Singh Selopal

Abstract Colloidal quantum dots (QDs) exhibit size-/shape-/composition-


dependent optoelectronic properties due to quantum confinement and considered
as building blocks for several optoelectronic devices. However, QDs possess a
high density of surface trap states, which act as non-radiative carrier recombination
centers, thereby reducing the overall performance of the optoelectronic device.
Surface passivation of QDs by the epitaxial growth of an outer shell of different
materials or composition, so-called core/shell QDs, has proven to be an effective
approach to reduce the surface trap states as well as to tune their optoelectronic
properties. Resulting core/shell QDs offer broader absorption spectrum, improved
quantum yield (QY), and prolonged photoluminescence (PL) lifetime with better
thermal, chemical, and photophysical stability compared to core QDs. In addition,
these optoelectronic properties can be controlled by tuning the size and shape
of core QDs, thickness and compositions of the shell layer, and electronic band
edge alignment between core and shell. This chapter provides a comprehensive
overview of the recent development in QDs-sensitized solar cells (QDSCs) based
on different types of colloidal core/shell QDs including type-I, type-II, and quasi-
type-II core/shell QDs as light-harvesting materials.

Keywords Core/shell quantum dots · Band alignment · Carrier dynamic ·


Quantum dot sensitized solar cells

G. S. Selopal ()
Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology
of China, Chengdu, People’s Republic of China
Institut National de la Recherché Scientifique, Centre Énergie, Matériaux et Télécommunications,
Montreal, QC, Canada
e-mail: gurpreet.selopal@emt.inrs.ca

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 219
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_7
220 G. S. Selopal

1 Introduction

The world’s major dependency on fossil fuels for energy is causing its rapid
depletion and environmental pollution and global warming [1, 2]. Therefore, the
scientific community is putting their best efforts to substitute fossil fuels with
renewable, clean energy sources like wind, hydro-energy, biomass, and solar energy.
Among all, solar energy is considered as one of the most promising, reliable,
clean, and economical renewable energy resources due to several reasons. For
instance, the amount of solar energy that reaches the surface of the earth in
1 h is larger than the energy consumed by the whole world in 1 year [3]. This
highlights that the efficient utilization of solar energy is of great importance to
fulfil future energy demands. There are several approaches such as photovoltaic
(PV), photoelectrochemical, and photothermal that are communally available to
utilize this promising energy source. Among these renewable technologies, PV is
considered as the most appealing approach that converts solar radiation directly into
electricity based on the photoelectric effect. In addition, PV devices are lightweight
and can be used everywhere without the need of any transmission line and also
the large-scale application of PV devices can partially address the growing energy
demand and environment-related issues [4]. The PV devices are mainly categorized
into three generations based on the nature of the materials, cost per watt, and
obtainable maximum photoconversion efficiency (PCE). In brief, first-generation
PV devices are based on single-crystal silicon features with high PCE of 24.4%
(close to the theoretical limit) [5] due to the broad absorption spectrum range and
high carrier mobility. To date, silicon solar cell technology covers more than 90%
of the PV market [6]. However, the high production cost of single-crystal silicon
for first-generation PV increases the overall cost of the electricity compared to the
electricity produced from fossil fuels. Thus, with the aim to reduce the production
cost of the first-generation PV, the second-generation solar cells were developed
based on thin-film technology. Amorphous silicon (a-Si), copper indium gallium
selenide (CIGS), copper indium selenide (CIS), and cadmium telluride (CdTe)
are the most commonly reported second-generation solar cells. Although the PCE
of the second-generation solar cells is comparable to the silicon solar cells, still
they possess less than 10% share of the PV market due to the problem of the
module scale implementation and their limited stability. In the last few decades,
third-generation PV technology attained great attention due to the possibility of
achieving high PCE while maintaining low production cost with the application of
novel nanomaterials and device architecture. In addition, utilization of inexpensive
and environment-friendly materials and large-area fabrication are considered as
motivation assets for third-generation PV technology. Due to these appealing
features, third-generation PV devices can represent perhaps the future of solar cell
technology, and even these devices are at the pre-commercial stage. Currently, dye-
sensitized solar cells (DSCs), organic (or polymer) solar cells (OSCs), perovskite
solar cells (PSCs), and quantum dot (QD)-based solar cells belong to the third-
generation PV technology. Among all, QD-based solar cells are of particular interest
Core/Shell Quantum-Dot-Sensitized Solar Cells 221

Fig. 1 Schematic illustration of QD-based solar cell configurations and energy band diagram:
(a) QDSC, (b) Schottky junction solar cell, (c) QD polymer hybrid solar cell, and p-n junction
solar cells: (d) heterojunction and (e) homojunction. (Reproduced with permission from Ref. [9].
Copyright 2015, American Chemical Society)

due to unique and promising optoelectronic properties of QDs as a light-harvesting


material such as size-, shape-, or composition-dependent absorption edge, high
absorption coefficient, high photoluminescence quantum yield (PL QY), and large
intrinsic dipole moment. Also the possibility of multiple exciton generation with
single-photon absorption and hot electron extraction before thermalization, which
could boost the PCE of QD-based solar cells beyond the Shockley–Queisser limit of
32.9% for single-absorber-based solar cells [7, 8]. Based on the device architecture,
QD-based solar cells can be categorized into four types such as p-n junction solar
cells (homo- or heterojunction), hybrid QD-polymer solar cells, Schottky junction
solar cells, and QD-sensitized solar cells (QDSCs) [9]. Figure 1 displays the device
architecture of different types of QD-based solar cells and electronic band alignment
with carrier separation dynamic at interfaces of corresponding devices. Among
these, QDSCs have attracted increasing attention due to simple, low-cost fabrication
and great potential to obtain high PCE, which is confirmed from the increasing
trend of the number of publications in the last few years (Fig. 2a). In addition,
QDSCs borrowed their architecture from their analogous counterpart DSCs and
can possibly be utilized for large area scalability. In the last few years, significant
progress in the PCE has occurred and recorded value reached more than 13% due
to the development of novel materials and device architecture (Fig. 2b) [10].

2 Brief History of QDSCs

The present form of QDSC was developed by the concept of sensitization of


the semiconductor. For the first time, in 1986, Gerischer et al. [11] introduced
the idea of sensitization of a wide-bandgap semiconductor (e.g., TiO2 ) with
222 G. S. Selopal

Fig. 2 (a) Number of published articles per year based on the keywords “quantum dot sensitized
solar cell” data is collected from ISI web of knowledge. (b) Trend of record PCE of liquid junction
QDSCs based on a standard two-electrode configuration and tested under one sun illumination
(AM 1.5, 100 mW/cm−2 ) from 2003 to 2019 [6]

narrow-bandgap materials (e.g., CdS) and demonstrated the enhancement in the


photocurrent of the electrode due to the broader light absorption range. Later on,
Vogel et al. [12] fabricated a three-electrode-based photoelectrochemical cell based
on TiO2 mesoporous-film-sensitized quantum-sized CdS particles and reported the
open-circuit voltage (Voc ) of 395 mV, short-circuit current density (Jsc ) of 175
mAm−2 , and fill factor (FF) of 0.75 under the monochromatic light illumination
(λ = 460 nm). In 1998, Zaban et al. [13] reported for the first time a complete QDSC
by using a nanoporous TiO2 film sensitized with pre-synthesized InP colloidal QDs,
with iodine/iodide redox couple as an electrolyte and Pt as a counter electrode
(CE). Yu et al. [14] reported a PCE of 0.3% for QDSCs based on InAs QDs.
These reports highlight that more attention should be paid toward other device
components, such as redox couple electrolyte and CE, to fabricate efficient QDSCs.
For instance, the metal chalcogenide (CdS, CdSe, PbS, PbSe, etc.) QDs applied
as light harvesters are not stable in an iodine/iodide redox couple electrolyte due
to their chemical incompatibility. In 2007, a significant breakthrough occurs in
PCE of QDSCs by replacing the iodine /iodide redox couple electrolyte with a
polysulfide electrolyte. Diguna et al. [15] reported a PCE of 2.7% for QDSCs based
on CdSe QD-sensitized TiO2 inverse opal electrode with polysulfide electrolyte.
In addition, they highlight that the thin overcoating of the ZnS layer on the CdSe
QD-sensitized TiO2 photoanode enhances the device performance due to reduced
carrier recombination. Thereafter, Lee et al. [16] demonstrated the concept of TiO2
sensitization by using a bifunctional linker molecule, followed by a chemical bath
deposition (CBD) process for CdSe layer growth and Au-coated conducting glass
Core/Shell Quantum-Dot-Sensitized Solar Cells 223

as CE and reported a PCE of 2.9%. Later on, the same group used a CdS/CdSe co-
sensitized photoanode via successive ion layer adsorption and reaction (SILAR) and
boost the PCE of QDSCs over 4% [17]. During this period, the scientists realized
that the sulfur species of the polysulfide electrolyte have a poisoning effect over
noble metal-based CEs due to the adsorption of S2− onto the surface of the CE,
which reduces its catalytic performances. In addition, the high cost of noble metals
increases the overall production cost of the device that is an issue for its large-scale
commercialization. In this context, Gimenez et al. [18] explored Cu2 S CE prepared
by treating brass with HCl and demonstrated that the PCE of QDSCs is significantly
improved from 0.65% with Pt CE to 1.83% with Cu2 S CE. In 2011, Zhang et al.
[19] reported the PCE of 4.92% by using Cu2 S/brass CE with optimized TiO2
film electrode sensitized with CdS/CdSe QDs. Fan et al. [20] prepared efficient CE
based on mesoporous carbon nanofibers and reported a PCE of 4.81%. Afterward,
Cu2 S CEs prepared by different approaches have achieved considerable attention
to enhance the performance of QDSCs [21–26]. These studies demonstrated that
performance of QDSCs can be improved by using a suitable combination of
polysulfide-based redox couple electrolytes (which guarantee for improved QD
stability) [27] with Cu2 S CE (better catalytic activity toward polysulfide electrolyte)
[28]. In the last few years, a significant improvement in PCE from 5% to more
than 13% has been reported due to advanced development in the material design,
device architecture through interface engineering, and in-depth understanding of
carrier mechanism of QDSCs. Among all, it is worth mentioning, Kamat et al. [29]
reported a PCE of 5.4% for QDSCs fabricated by using Mn-doped CdS/CdSe QD-
sensitized TiO2 mesoscopic film as photoanode and Cu2 S/graphene oxide (GRO)
as CEs with polysulfide as the electrolyte. Zhong et al. [30] synthesized core–
shell colloidal CdTe/CdSe QD and fabricated QDSCs with PCE of 6.76%. The
design and development of ternary- or quaternary-alloyed and core/shell QDs leads
to a significant breakthrough in the PCE of QDSCs. For example, in 2014, a
new record of PCE as high as 7.04% has been reported for QDSCs based on
type-I CuInS2 /ZnS core/shell QDs [31]. Then Zhao et al. [32] fabricated QDSCs
using ternary a CdSex Te1−x QD-sensitized TiO2 mesoporous film with additional
double layer capping of ZnS and SiO2 to reduce the carrier recombination at the
TiO2 /QDs/electrolyte interface. The resulting QDSCs exhibit a certified PCE of
8.21% under one sun illumination. In 2016, Du et al. [33] synthesized a mesoporous
carbon (MC) deposited on titanium (Ti) mesh and used as CEs. QDSCs based on
MC-Ti CEs yield a PCE of 11.51% by using a CdSex Te1−x QD-sensitized TiO2
mesoporous film as photoanode and polysulfide as the electrolyte. This PCE value
was further enhanced to 12.45% by exploring Zn-Cu-In-Se QDs as light harvester
and nitrogen-doped MC-Ti as CE [34].
Recently, Pan et al. [35] fabricated QDSCs by employing a co-sensitization
approach based on Zn-Cu-In-Se and Zn-Cu-In-Se colloidal QDs and reported an
average PCE of 13.18%. This significant improvement in QDSC performance
makes this promising technology one step toward commercialization and compet-
itive with the other emerging solar cell technology. Figure 2b displays the overall
progress in PCE of QDSCs based on sandwich-type two-electrode configuration
224 G. S. Selopal

under one sun illumination (AM 1.5G). All calculated photovoltaic parameters of
the corresponding QDSCs are reported in Table 1.

3 QDSC Architecture

The device architecture of QDSCs is analogous to DSCs, in which the dye molecule
is replaced by semiconductor QDs as a light-harvesting material. Typically, QDSCs
consist of a wide-bandgap semiconductor-based anode as the electron transport
layer; QDs as a light harvester; a hole transport medium, so-called electrolyte; and
CE (Fig. 1a).

3.1 Anode

Anode is mainly composed of a wide-bandgap semiconductor film deposited on a


transparent conducting oxide glass [e.g., fluorine-doped tin oxide (FTO) or indium-
doped tin oxide (ITO)], which acts as a scaffold to support QDs, extract the electron
from the photoexcited QD, and then transport them toward the conducting oxide
glass substrate. For the ideal anode, a semiconductor film should have a high
specific surface area, high electron mobility, appropriate band alignment with the
QDs, and compatibility with the other device components. Usually, the thickness
of the electron transport layer is slightly lower compared to the thickness used in
DSCs due to the high extinction coefficient of semiconductor QDs. However, the
optimization of the thickness of the electron transport layer can be varied according
to the nature, structure of the materials, and design of the device architecture. There
are different types of wide-bandgap semiconductors that can be used as an electron
transport layer: (i) The TiO2 semiconductor is most communally used as an electron
transport layer due to several promising features such as low cost, non-toxic,
and chemically stable materials [42, 43]. Usually, a standard TiO2 mesoporous
electron transport layer is composed of two layers: The first layer is the 20-nm-
sized TiO2 nanoparticles with high specific surface area to load the QDs, so-called
active layer, and second layer is the 200–400-nm-sized nanoparticles, so-called
scattering layer, to increase the time spent within the photoanode and enhance the
possibility of harvesting more sunlight photon by QDs. To date, QDSCs based on
the TiO2 nanoparticle electron transport layer yield the highest PCE. However, the
photoinjected electron undergoes transport within nanoparticles and the mesoporous
film faces grain boundaries of nanoparticles, therefore increasing the possibility of
electron recombination. Thus, with the aim to facilitate faster electron transport,
one-dimensional (1D) nanostructures of TiO2 such as nanorods, nanowires, and
nanotubes have been synthesized and applied as the electron transport layer [44–
46]. These 1D nanostructures improve the electron transport within the photoanode
and reduce the carrier recombination. However, the lower amount of QD loading
Table 1 Summary of the development in photovoltaic performance of QDSCs based on two-electrode sandwich configuration, tested under one sun
illumination (AM 1.5G)
Year QDs Electrolyte CE Voc (V) Jsc (mAcm−2 ) FF PCE (%) Ref.
1998 InP I− /I3 − Pt – – 0.685 – [13]
2002 PbS Spiro-OMeTAD – 0.24 – – 0.49 [36]
2006 InAs Co2+ /Co3+ Pt 0.35 1.75 0.48 0.30 [14]
2007 CdSe Polysulfide Pt 0.71 7.51 0.50 2.7 [15]
2008 CdS/CdSe Polysulfide Au 0.503 11.66 0.49 2.9 [16]
2009 CdS/CdSe Polysulfide Au 0.514 16.80 0.49 4.22 [17]
2011 CdS/CdSe Polysulfide Cu2 S/brass 0.575 13.68 0.63 4.92 [19]
2012 Mn-CdSe/CdSe Polysulfide 0.558 20.70 0.47 5.42 [29]
Core/Shell Quantum-Dot-Sensitized Solar Cells

Cu2 S-RGO
2013 CdSex Te1−x Polysulfide Cu2 S/brass 0.571 19.35 0.58 6.36 [37]
2013 CdTe/CdSe Polysulfide Cu2 S/brass 0.606 19.59 0.57 6.76 [30]
2014 CuInS2 /ZnS Polysulfide Cu2 S/brass 0.586 20.65 0.58 7.04 [31]
2015 CdSex Te1−x Polysulfide Cu2 S/brass 0.653 20.78 0.61 8.21 [32]
2015 CdSex Te1−x Polysulfide Cu2 S/FTO 0.702 20.78 0.64 9.28 [38]
2016 CdSex Te1−x Polysulfide Cu2 S/brass 0.720 21.04 0.64 9.73 [39]
2016 CdSex Te1−x Polysulfide MC-Ti 0.807 20.69 0.69 11.51 [33]
2016 ZCISe Polysulfide MC-Ti 0.745 25.49 0.63 11.91 [40]
2017 ZCISe Polysulfide N-MC-Ti 0.759 25.67 0.64 12.45 [34]
2018 ZCISe-CdSe Polysulfide MC-Ti 0.752 27.39 0.62 12.75 [41]
2019 ZCISe-ZCIS Polysulfide MC-Ti 0.762 25.93 0.66 12.98 [35]
225
226 G. S. Selopal

is due to a limited specific area of 1D nanostructures that reduces the overall PCE
of QDSCs. (ii) ZnO is also widely studied as an electron transport layer due to its
higher electron mobility (1–5 cm2 V−1 s−1 ) compared to TiO2 (0.1–4 cm2 V−1 s−1 )
and position of the conduction band (CB) edge, and it is relatively easy to grow
at different morphologies (nanowires, nanotubes, and neotheropods) via the wet-
chemical approach [47, 48]. However, the overall PCE of QDSCs based on ZnO is
still lower than that of QDSCs based on TiO2 due to several issues. In addition, the
chemical instability of ZnO as the electron transport layer in QDSCs reduces the
long-term stability. Furthermore, the dissolution of ZnO in the acidic medium limits
the possibility to attach the bi-linker groups to load pre-synthesized colloidal QDs.
Despite these issues, it is reported that the highest obtained Voc of QDSCs based on
ZnO is 0.836 V, which is 0.2 V higher than the Voc of QDSCs based on TiO2 due
to its CB edge position [49]. This highlights that ZnO possesses a great potential
to be applied as an electron transport layer to fabricate high-performing QDSCs
by optimizing the QD structure and interfacial engineering by surface passivation.
(iii) SnO2 is also considered as promising electron transport materials in QDSCs
due to its excellent physical, chemical, optical, and electrical properties such as
less sensitive toward UV light and better chemical stability and electron mobility
compared to both TiO2 and ZnO [50, 51]. However, the PCE of the QDSCs based
on SnO2 electron transport layer is still lower than QDSCs based on TiO2 .

3.2 Light Harvester

QDs are considered as the heart of QDSCs and applied as light harvester materials.
Ideal QDs should have broader light absorption range, high absorption coefficient,
and suitable band alignment with the electron transport layer and hole transport
materials that allows the efficient separation of exciton generated with the absorp-
tion of a photon. In addition, the stability of QDs in the electrolyte and under light
illumination and heat is also considered as a crucial requirement to fabricate efficient
and stable QDSCs.
Generally, there are several approaches available to deposit QDs (in situ and
ex situ) on the mesoporous electron transport layer depending on the nature
of QDs. The most commonly used QDs are classified into four categories: (i)
binary QDs (InP, InAs, CdS, CdSe, CdTe, PbS, and Ag2 S) [12–15, 52, 53]; (ii)
alloyed QDs (II–VI, CdSx Se1−x , CdSex Te1−x , etc.; I, III, VI, CuInS2 , AgInS2 ,
CuInSe2 , etc.) [32, 37–40]; (iii) doped QDs (CIS/Mn:CdS; Mn-CdSex Te1-x ; Co2+ -
ion-doped CdS/CdSe, etc.) [29, 54], and (iv) core/shell QDs (type-I, CuInS2 /ZnS,
CdSeTe/CdS, etc.; reverse type-I, CdS/CdSe; and type-II, CdSex Te1−x /CdS; more
details about the core/shell QDs are reported in Sect. 6) [30, 31, 55].
Core/Shell Quantum-Dot-Sensitized Solar Cells 227

3.3 Electrolyte

The space between the two electrodes of a sandwich cell structure is filled with
the redox couple electrolyte, which regenerates the oxidized QDs to its original
form and acts as hole transport media during cell operation. For highly efficient
devices, the redox couple electrolyte possesses the following properties: (i) high
solubility and diffusion coefficient of the solvent used in electrolyte preparation
for better charge carrier transport, (ii) oxidized and reduced forms that are stable
and highly reversible, and (iii) chemically inert toward the other cell components
and spectrally toward the visible light spectrum (absence of light absorption
from the visible light spectrum). The most commonly used liquid electrolyte is
polysulfide redox couple (S2− /Sn2 ), which is composed of Na2 S, S, and NaOH
or KCL [15–17, 29–32]. Also, Co2+ /Co3+ , (CH3 )4 N2 S/(CH3 )4 N2 Sn , Fe2+ /Fe3+ ,
and Fe(CN)6 3− / Fe(CN)6 4− redox couple electrolytes are used to improve the
performance of QDSCs [14, 56–58]. However, leakage and volatile nature of the
liquid electrolytes reduce the stability of QDSCs. To solve these issues, quasi-
solid-state [gel polysulfide electrolyte using gelators: Dextran, sodium polyacrylate
(PAAS), and sodium carboxymethylcellulose (CMC-Na)] and solid-state (Spiro-
OMeTAD, poly (3-hexylthiophene), CuSCN, etc.) electrolytes are also used in
QDSCs [36, 59–61].

3.4 Counter Electrode

CE is a thin layer of catalytic active materials deposited on the conducting substrate,


which mainly speeds up the reduction of the oxidized electrolyte. Several parame-
ters of the CE that effect the overall device performance such as: (i) electro-catalytic
activity, (ii) chemical stability in redox couple electrolyte, and (iii) morphology and
surface roughness of exposed facet. The FF of the device is significantly influenced
by the properties of CE, so by improving the CE properties, we can improve
the performance of QDSCs. For efficient CE, there should be less charge transfer
resistance (Rct ) from the CE to the electrolyte, which reduces the series resistance
leading to improvement in FF, which results in higher PCE. There are several types
of materials used as CEs in QDSCs: (i) noble metals (Pt and Au) [15, 17], (ii) metal
chalcogenides (Cu2 S/brass, Cu2 S/FTO or ITO, Cu2 Se, CoS, FeS, NiS, etc.) [19, 38,
62–64], (iii) carbon materials (active carbon, conductive carbon black, mesoporous
carbon nanofibers, mesoporous carbon supported on Ti mesh, etc.) [33, 34], and (iv)
composites (Cu2 S-RGO, CuInS2 /carbon composite, PbS/carbon black, etc.) [29, 65,
66].
228 G. S. Selopal

Fig. 3 Schematic illustration: (a) device structure and working principle of QDSC composed of
a photoanode as electron transport layer, QDs as light harvester, redox couple electrolyte as hole
transport medium, and counter electrode; (b) current density–voltage (J-V) curves of QDSCs under
one sun illumination (AM 1.5G, 100 mW/cm−2 )

4 Working Principle

The working principle of QDSCs is quite different from conventional p-n junction
solar cells due to the difference in their fundamental physical phenomena behind
their operation and device construction. In the conventional p-n junction solar cells,
light absorption and charge transport occur within the same region of the material,
whereas in the QDSC, light is absorbed by the QDs and transport occurs through
the metal oxide electrode and electrolyte. In addition, carrier separation in QDSC
mainly occurs due to energetic and entropic forces at the anode/QD/electrolyte
interface, whereas in the conventional p-n junction solar cell, it is controlled by
the electric field across the p-n junction. A complete one cycle of QDSC operation
is shown in Fig. 3a and consists of the following process:

4.1 Light Absorption

The absorption of a light photon (hν ≥ Eg of QD) by QDs via an electronic excita-
tion process (1). In brief, an electron–hole pair is generated with the absorption of
photon and electron jumps from the higher energy levels of valence band (VB) to
the lowest energy levels of CB. During this process, QD is get transferred from the
ground state (QD) to the excited state (QD∗ ) as shown below:

QD + hϑ → QD∗
Core/Shell Quantum-Dot-Sensitized Solar Cells 229

4.2 Carrier Separation

Now the photogenerated electron–hole pair in QD with the absorption of photon is


separated through the injection of the electron from the CB energy level of excited
QD to CB of metal oxide (2) and a hole transport from the VB energy level of
excited QD to the electrolyte (3), as shown below:

QD∗ → QD+ + e− CB

2QD+ + S2− → 2QD + S

The electron injection process occurs on the femtosecond scale, whereas the hole
transport process occurs on the microsecond range.

4.3 Carrier Transport

After the electron–hole separation, the injected electrons travel through a meso-
porous electron transport layer via diffusion (4) to the transparent conducting oxide
substrate. Several theoretical and experimental studies have been carried out to
understand the exact electron transport mechanism through the mesoporous electron
transport layer. In brief, the concept of electron transport by a built-in electric field
has been discarded due to the small size of the nanoparticles of the mesoporous
electron transport layer and the screening effect of the surrounding electrolyte. Thus,
electron transport is assumed to occur via the diffusion process and is considered
as the most realistic [67, 68]. The electron diffusion process involves trapping
and de-trapping of the electron from one particle to another [69]. It has been
demonstrated that an electron undergoes 103 –106 trapping/de-trapping events to
reach the transparent conducting oxide [70]. These events (electron trapping/de-
trapping) occur through the localized energy trap states having different depths and
lies just below the CB of the electron transport layer. These energy states (traps/de-
traps) play a significant role during electron transport. At low light intensity, deep
traps participate in the electron transport causing low diffusion coefficient, whereas
at higher light intensity (≈ solar illumination intensity), deep traps are filled and
shallow traps contribute to electron transport resulting in a high diffusion coefficient
[71]. Finally, these electrons are collected at the conducting substrate and reach the
CE through the external load.
On the other hand, the hole in the VB energy level of excited QD is scavenged by
the polysulfide redox couple electrolyte (3) and regenerates the QD. The oxidized
electrolyte reaching the CE is immediately reduced by gaining the electron that
reaches at CE through an external circuit (5), hence completing one cycle. This
electrolyte reduction is catalyzed by catalytic CE materials as shown below:
230 G. S. Selopal

Sx 2− + 2e− (CE) → Sx−1 2− + S2−

The time scale at which the electron and hole transport occurs is in 100
microseconds and less than 10 microseconds, respectively.

4.4 Recombination

Apart from these favorable carrier-transport processes in QDSCs, there are series
of non-radiative carrier recombination paths during their transport (Fig. 3a). These
recombination processes depend on various factors such as the morphology of
the mesoporous electron transport network, type and nature of QDs, electrolyte
composition, and CE and are responsible for the loss in overall PV performance
of QDSCs. The most predominant recombination processes occur at the electron
transport layer/QD/electrolyte interface and within the QDs, for example, (i)
recombination of the photoinjected electron in the CB of the electron transport layer
with the oxidized species in the electrolyte at the electron transport layer/electrolyte
interface (6), (ii) recombination of the electron in the CB of the electron transport
layer with hole in the VB of QDs at the electron transport layer/QD interface (7), (iii)
recombination inside the QDs through surface trap states (8), and (iv) recombination
of the excited electrons in the CB of QDs with the oxidized species in the redox
couple electrolyte at the QD/electrolyte interface (9) [72–74].

5 Photovoltaic Characterizations

The key parameters of the PV performance of QDSC are obtained from J-V
measurements taken under standard one sun illumination conditions (AM 1.5 G).
The intensity of the solar simulator spectrum (AM 1.5 G) should be optimized using
a standard silicon solar cell (100 mWcm−2 , one sun) that allows the systematic
comparisons of the obtained results from different research labs. The J-V curve is
obtained by applying an external potential and measuring the current, then plotted
as the current density (current/area of the device) vs voltage (Fig. 3b). From the J-V
curves, Jsc and Voc values can be obtained from the intercepts of the J-V curves at
the y-axis (when V = 0) and x-axis (when J = 0), respectively. There is a point in a
J-V curve which intercepts the maxima of the photocurrent (Imp ) and photovoltage
(Vmp ) drawn and their product gives maximum obtainable power (Pmax ). The FF is
defined as the ratio of the Pmax to the product of the Voc and Jsc (theoretical value),
which is in the range of 0–1. The overall PCE of the QDSCs can be calculated from
the ratio of the Pmax to the total power of the incident light (Pin ) as given below:
Core/Shell Quantum-Dot-Sensitized Solar Cells 231

Pmax VMP × JMP F F × Voc × Jsc


PCE = = =
Pin Pin Pin

Apart from these PV parameters, incident photon to current conversion efficiency


(IPCE) is also an important measurement to know the conversion efficiency of the
solar cells at particular wavelength and is the combination of the light-harvesting
efficiency (LHE), electron injection efficiency (ηinj ) from the QDs to the CB of
electron transport layer, and photogenerated charge collection efficiency (ηcol ) [75]:

IPCE = LHE (λ) .ηinj (λ) .ηcol (λ)

The Isc measured under short-circuit conditions is the integrated sum of IPCE (λ)
measured over the entire solar spectrum:
 ∝
Isc (λ) = IPCE (λ) Isun (λ) dλ
0

where Isun (λ) is the incident irradiance as a function of the wavelength λ (nm). This
integrated current Isc (λ) is used to cross-check with the result obtained with I-V
measurements. The IPCE is also called external quantum efficiency (EQE).
In addition, the absorbed photon to current conversion efficiency (APCE) is the
ratio of the number of generated electrons (Ne ) and the number of absorbed photons
(Na ). The APCE is also called the internal quantum efficiency (IQE) and reflects the
efficiency of absorbed photons by a QD being converted into photocurrent excluding
the effect of the light-harvesting efficiency of the photoanode. The APCE can be
calculated by the following equation:

Ne IPCE (λ)
APCE (λ) = = = ϕinj ϕcc
Na LHE (λ)

Carrier dynamics measurements of QDSCs can be performed using charge


extraction (CE), transient photocurrent decay (TCD) and photovoltage decay
(TVD), electrochemical impedance spectroscopy (EIS), open-circuit voltage decay
(OCVD), intensity-modulated photocurrent spectroscopy (IMPS), and photovoltage
spectroscopy (IMVS).

6 Core/Shell QDs

QDSCs have the possibility to boost the PCE beyond the Shockley–Queisser limit
of 32.9% due to promising optoelectronic properties of QDs as light-harvesting
materials. However, the record value of PCE (≥ 13%) of QDSCs is still lower
than that of the calculated theoretical limit of PCE [76]. This substantially lower
value of PCE of QDSCs is mainly attributed to undesirable non-radiative carrier
232 G. S. Selopal

recombination during the operation under illumination (Sect. 4.4). It can be seen
in Fig. 2 that the major non-radiative recombination processes [(7), (8), and (9)] in
QDSCs are mainly dependent on the nature of the QDs, in particular, the ability
of exciton generation with the absorption of a light photon, efficient dissociation
of photogenerated exciton at the interface of QD/carrier scavengers (TiO2 or elec-
trolyte), and regeneration of QDs for the next cycle [72, 77]. Usually, QDs process
a high density of surface traps due to high surface-to-volume ratio, which promotes
these non-radiative recombination processes [(7), (8), and (9)] through surface trap
states that act as recombination center and reduces the overall performance of
QDSCs. Also, the bandgap tuning of QDs by tailoring their size has several issues.
In particular, in solar cell application, QDs with a narrow bandgap offer a broader
light-harvesting range, but the efficiency of the photogenerated electrons from the
CB of QDs to the CB of the electron transport layer is due to unfavorable band
alignment between QDs and carrier scavengers [78]. Thus, there is an urgent need
to design and synthesize high-quality QDs for the fabrication of high-efficiency and
stable QDSCs. Till date, different approaches have been reported to minimizing
these non-radiative recombinations. Among all, the core/shell architecture of QDs
is considered the most effective strategy [79–82]. In this core/shell architecture,
the surface of the core QDs is passivated by the growth of the shell of different
materials and thickness. The core/shell QDs offer a reduced density of surface traps,
broader light absorption, better thermal and photochemical/physical stability to the
surrounding environment, and longer PL lifetime compared to core QDs [80, 82,
83]. In addition, these optoelectronic properties of the core/shell QD system can
be tuned by the appropriate design and selection of the core and shell materials,
thickness, and composition. Thus, the band alignment between the core and shell
materials and understanding of carrier dynamics of the core/shell QD system are
crucial to fabricate high-performing QDSCs.

6.1 Classification of Core/Shell QD Systems

The core/shell QD systems are typically classified into three systems, such as type-I,
reverse type-I, and type-II, based on the relative band alignment between the CB
and VB edges of the core and shell materials (Fig. 4) [84]. In type-I core/shell
QDs, the bandgap of the shell materials is larger than that of the core materials
and both the CB and VB edges of the core are located in the bandgap of the shell
materials. Therefore, both electrons and holes are confined in the core region (Fig.
4a). The shell is used not only to passivate the active surface of the core QD but also
improves its optoelectronic properties such as PLQY and photophysical/chemical
stability toward its surrounding environment [85–89]. However, the PL emission of
the type-I core/shell QD system depends on the nature of the core materials as both
electron and hole are localized in the core region. The reverse type-I core/shell QD
system is opposite to the type-I core/shell QDs, in which the bandgap of the shell
materials is lower than that of core materials (Fig. 4b). Both electrons and holes
Core/Shell Quantum-Dot-Sensitized Solar Cells 233

Fig. 4 Schematic illustration of the electronic band alignment for different types of core/shell QD
systems and below electron and hole wave functions correspond to the position of the CB and
VB of the core (center) and shell materials (outer), respectively: (a) type-I, (b) reverse type-I, (c)
type-II, (d) quasi-type-II

are partially or entirely delocalized in the shell region that depends on the thickness
of the shell layer [90, 91]. The absorption spectrum of the reverse type-I core/shell
QD system can be tuned by changing the shell thickness that allows the efficient
extraction of the photogenerated electrons and improves the electron injection rate
[78, 92]. In type-II core/shell QD system, either the CB edge or VB edge of the
shell is located in the bandgap of the core (Fig. 4c) [93, 94]. The electrons and
holes are spatially separated into different regions of the core and shell [95, 96]
This spatial carrier separation in type-II core/shell QDs leads to a substantial shift
of the emission wavelength and prolonged PL lifetime decay with shell growth.
The emission of the type-II core/shell QD system can be controlled by the effective
band offset of core and shell materials. In addition, these core/shell QDs, there is
a particular case of core/shell structured QDs called quasi-type-II in which one
type of carrier is confined in the core, whereas the other is delocalized over the
entire core/shell structure (Fig. 4d). In the case of “giant” core/shell QDs (shell
thickness ≥ 1.5 nm to several tens of nm), type-I core/shell QDs can be transferred
to quasi-type-II core/shell QDs, by tuning the thickness and composition of the shell
layer.

6.2 Synthesis of Core/Shell QDs

The size-/shape-/composition-dependent optoelectronic properties of QDs allow to


design and synthesize novel colloidal QDs as per requirements for the fabrication of
234 G. S. Selopal

Fig. 5 Schematic representation of the synthesis of colloidal core/shell QDs. First, the core QDs
can be synthesized via the hot-injection approach (left side), then core/thin shell, core/thick shell,
and core/alloyed shell/shell structure (right side) synthesized by the growth of shell over the core
QDs either via cation exchange/SILAR or a combination of both the cation exchange and SILAR
approach

efficient and stable optoelectronic devices. In the last few decades, considerable
efforts have been devoted to developing novel wet chemical approaches to syn-
thesize colloidal QDs with uniform size distribution to replace the expensive and
complicated physical methods. Core/shell QDs are commonly synthesized in a two-
step approach as shown in Fig. 5: (i) first, the synthesis of high-quality colloidal core
QDs via the well-established hot-injection approach. In brief, one of the reaction
precursor is heated to a high temperature that depends on the nature of the precursor
and then subsequent injection of another precursor at room temperature. These
mixed precursors decomposed to form nucleation seeds and the slight lowering in
reaction temperature allows the growth of seeds into QDs. The QDs synthesized by
the hot-injection approach are monodisperse with a narrow size distribution, high
QY, and better chemical/photostability. The size of the colloidal QDs can be tuned
by varying the reaction temperature, duration, and precursor concentration [97, 98].
Then the as-synthesized core QDs are purified and re-dispersed in the organic
solvent for the shell growth. (ii) Several approaches have been reported for shell
growth over the pre-synthesized core QDs. The cation exchange and SILAR
approaches are the most preferred approaches for shell growth. In the cation
exchange approach, the mixture of the cationic and anionic precursor is injected in
the core QD solution at the growth temperature and only the new cationic precursor
of the shell material is gradually replaced by the cation of the core QDs during the
shell growth. The overall size of the as-synthesized core/shell QDs via the cation
exchange method does not change significantly compared to core QDs (Fig. 5).
While in the SILAR method, the shell is grown over the pre-synthesized core QDs
Core/Shell Quantum-Dot-Sensitized Solar Cells 235

by alternative injection of cationic and anionic precursors. The thickness of the


shell layer can be tuned by changing the growth temperature, duration, and amount
of shell precursor. The overall size of core/shell QDs synthesized via the SILAR
approach is increased and depends on the number of grown shell layers. In addition,
the combination of cation exchange and SILAR is used to engineer the interface
between the core and shell QDs via the synthesis of the core/alloyed shell/shell QDs,
which offer better optoelectronic properties compared to simple core/shell QDs (Fig.
5). In these approaches, the shell growth temperature (T2 ) is generally lower than
the temperature used for the core QD synthesis to avoid uncontrolled ripening of
the core QDs and homogeneous nucleation of the shell material. In addition, the
concentration of core QD dispersion is required to inject shell precursor amount to
obtain the desired shell thickness.

7 Photovoltaic Performance of Core/Shell QDSCs

The major undesirable non-radiative recombination processes of QDSC solar cells


are mainly associated with QDs. Therefore, high-quality QDs with reduced surface
trap states and efficient carrier dynamics are required to fabricate efficient and stable
QDSCs. In the last few years, core/shell QDs achieved considerable attention to be
applied as light-harvesting materials due to their versatile optoelectronic properties
(discussed in Sect. 6). All three types of core/shell QDs (type-I, reverse type-I, and
type-II) have been widely applied as a light harvester in QDSCs and are discussed
below.

7.1 Type-I Core/Shell QDs

In type-I core/shell QDs, the CB and VB edges of shell materials are higher than that
of core materials as discussed in Sect. 6.1 and both electron and hole are confined
in the core region (Fig. 4a). The shell layer passivates the surface of the core QDs
that reduces the surface traps and improves the photophysical/chemical stability and
optoelectronic properties. For example, the carrier dynamics measurements of type-
I CdSe/ZnS core/shell QDs as a function of ZnS shell thickness highlight that the
ZnS shell layer acts as a tunnelling barrier for the electron and hole transfer can
lower their transfer rates due to reduced electronic coupling between the type-I
core/shell QDs and carrier scavengers [83]. Thus, optimization of the shell layer
thickness with efficient carrier transfer rates should be required for the fabrication
of high-performance QDSCs. In this context, Pan et al. [31] synthesized heavy
metal-free CuInS2 core QDs via the hot-injection method at 180 ◦ C and then a thin
shell layer of ZnS was overcoated via the partial cation exchange approach. Figure
6a–b displays the TEM images of CuInS2 core QDs and CuInS2 /ZnS core/shell
QDs, which confirm the comparable size of both core and core/shell QDs with an
236 G. S. Selopal

Fig. 6 TEM images of as-synthesized (a) CuInS2 QDs and (b) CuInS2 /ZnS core/shell QDs.
Inset displays the cation exchange process for the synthesis of CuInS2 /ZnS core/shell QDs.
Optical characterizations: (c) UV−vis absorption and (d) PL emission spectra with emission
wavelength = 400 nm. (e) J-V curves of QDSCs based on CuInS2 and CuInS2 /ZnS core/shell
QDs; (f) IPCE spectra of corresponding QDSCs. Electrochemical impedance spectroscopy char-
acterization of the QDSCs based on CuInS2 and CuInS2 /ZnS core/shell QDs: (g) recombination
resistance (Rrec ) on applied voltage (Vappl ). (h) Nyquist plots of both cells at −0.55 V forward bias.
Reproduced with permission from Ref [31]. Copyright 2014, American Chemical Society

average size of 5.1 ± 0.4 nm. Then bi-linker molecule mercaptopropionic (MPA)-
capped water-soluble CuInS2 /ZnS core/shell QDs were prepared by phase transfer
via the ligand exchange approach. The absorption spectra of resulting CuInS2 /ZnS
core/shell QDs extend to the near-infrared region of 850 nm (Fig. 6c). As can be
seen, the PL QY of the CuInS2 /ZnS core/shell QDs is increased more than tenfold
with the addition of a thin ZnS shell layer over the CuInS2 QDs due to reduced
surface traps (Fig. 6d). The QDSC-based MPA-capped water-soluble CuInS2 /ZnS
core/shell QDs yield a PCE of 7.04% under one sun illumination (AM 1.5G), which
is 40% higher than the PCE of QDSCs with CuInS2 core QDs (Fig. 6e). This
enhanced photovoltaic performance of QDSCs is mainly attributed to improved
optoelectronic properties and reduced surface trap states with the surface passivation
of CuInS2 QDs with a thin ZnS shell layer. The IPCE response of QDSCs based
on CuInS2 /ZnS core/shell QDs (75%) is significantly higher compared to QDSCs
based on CuInS2 QDs (60%) between the 350- and 700-nm range (Fig. 6f). The
effect of the ZnS shell layer on the carrier dynamics at the QD/TiO2 /electrolyte
interface of QDSCs is evaluated using EIS at different applied voltages. It can be
seen that the calculated Rrec value of QDSCs based on CuInS2 /ZnS core/shell QDs
is significantly (threefold) higher than that of the QDSCs based on CuInS2 QDs,
which confirms the reduced carrier recombination (Fig. 6g). At particular forward
bias voltage (−0.55 V, close to the Voc of the device), comparison of Nyquist plots
between QDSCs based on CuInS2 /ZnS core/shell QDs and CuInS2 QDs shows the
clear difference in the carrier dynamics of the respective devices (Fig. 6h).
Core/Shell Quantum-Dot-Sensitized Solar Cells 237

Yang et al. [55] also reported the synthesis of type-I CdSeTe/CdS core/shell QDs
via a combination of two approaches: hot injection for the synthesis of CdSeTe core
QDs and CdS shell layer grown by the SILAR method at moderate temperature
condition. Then as-synthesized oil-soluble QDs were transferred to thioglycolic
acid (TGA)-capped water-soluble QDs through the ligand exchange process and
anchored on TiO2 mesoporous film via the self-assembly deposition technique.
QDSCs based on type-I CdSeTe/CdS core/shell QDs yield a PCE of 8.02%, which
is 13% higher than the PCE of QDSCs based on CdSeTe QDs. This significant
improvement in the PCE of QDSCs is ascribed to reduced carrier recombination
within QDs and QD/TiO2 /electrolyte interface with the grown CdS shell layer over
CdSeTe QDs. In addition, the capping of a QD/TiO2 photoanode with SiO2 and
amorphous TiO2 layers was applied to further improve the interface between the
QDs/TiO2 and electrolyte. Resulting QDSCs achieved a PCE of 9.48%, which was
a new record of the PCE for liquid-junction QDSCs.
Overall, these studies highlight that in the case of type-I core/shell QDs,
overcoating of the shell layer on core QDs improves the optoelectronic properties
and photochemical/physical stability due to reduced surface trap states. Although
the performance of the QDSCs is enhanced due to these appealing features of type-I
core/shell QDs, the shell layer acts as a tunnelling barrier for the carrier separation
process and reduces their transfer rates.

7.2 Reverse Type-I Core/Shell QDs

In type-I core/shell QDs, both electrons and holes are confined in the core region
and the shell layer acts as a tunnelling barrier that reduces the carrier transfer rates
(Fig. 4b). Thus, to solve these issues, another type of core/shell QDs in which the
CB and VB gap of the core materials is higher than that of shell materials is the
so-called “reverse type-I” core/shell QDs (also called “inverted type-I” core/shell
QDs). Among the various reverse type-I core/shell QDs, CdS/CdSe core/shell QDs
have been widely studied and applied as a light harvester in QDSCs due to several
appealing features of the constituents: CdS offers higher Voc and CdSe offers
broader light absorption near 720 nm. In addition, charge carriers (both electron
and hole) are delocalized in the CdSe shell region that allows efficient extraction
of photogenerated electrons and holes by carrier scavengers and enhances their
injection rates [99].
In the last few years, reverse type-I CdS/CdSe core/shell QD-based solar cells
achieved great attention due to the above-mentioned promising optoelectronic
properties [100–103]. In these studies, reverse type-I CdS/CdSe QDs were directly
grown over the mesoporous wide-bandgap semiconductors (e.g., TiO2 or SnO2 )
via SILAR or chemical bath deposition. However, it is hard to control the size
and shape of the directly grown CdS/CdSe QDs due to the uncontrollability of
nucleation and growth processes of the QDs on the mesoporous film surfaces. In
addition, these QDs are not offered the exact inverted type-I core/shell configuration.
238 G. S. Selopal

Thus, the overall PCE of QDSCs based on directly grown CdS/CdSe QDs is in the
range of 2–4%. To solve these issues, high-quality pre-synthesized reverse type-I
colloidal QDs were developed and their carrier dynamics were studied [99, 104].
Both fluorescence lifetime decay and femtosecond transient absorption (TA) spec-
troscopy measurements of reverse type-I CdS/CdSe core/shell QDs demonstrated
that both electron and hole are localized in the CdSe shell region and suitable for
the fabrication of high-efficiency QDSCs. Pan et al. [78] synthesized high-quality
reverse type-I CdS/CdSe core/shell QDs by the hot-injection approach for the CdS
core, followed by the SILAR method for the CdSe shell of tunable thickness.
It can be seen that both CdS core and CdS/CdSe core/shell QDs with tunable
CdSe shell thickness show a nearly spherical shape with narrow size distribution
(Fig. 7a–e). To improve the QD loading, as-synthesized oil-soluble high-quality
CdS/CdSe core/shell QDs were transferred to water-soluble QDs via the ligand
exchange approach by using a bifunctional mercaptopropionic acid (MPA) linker
molecule. As a proof of concept, water-soluble MPA-capped CdS/CdSe core/shell
QDs were loaded to TiO2 mesoporous film. It is clear that absorption spectra profiles
of CdS/CdSe core/shell QDs with different CdSe shell thicknesses anchored to TiO2
is consistent with MPA-capped QD aqueous solution (Fig. 7g). This highlights that
the high-temperature synthesis method has better control over the size and structure
of QDs, which is absent in the directly grown reverse type-I CdS/CdSe QDs. QDSCs
based on MPA-capped CdS/CdSe core/shell QDs with optimized absorption onset
reported a new record PCE of 5.32% under one sun illumination (Fig. 7h). The
trend of the IPCE values of the corresponding QDSCs between the 400- and 650-nm
range is consistent with J-V measurements (Fig. 7i). This significant improvement in
the QDSC performance is mainly ascribed to the extended light-harvesting range,
efficient charge injection, and suppressed charge recombination in reverse type-I
CdS/CdSe core/shell QDs compared to core QDs.

7.3 Type-II Core/Shell QDSCs

Type-II core/shell QDs are composed of a core QD passivated by the shell layer
with the CB or VB edge located in the bandgap of the core material (Fig. 4c).
As discussed previously, the spatial separation of electron and hole wave function
enables the fast electron injection rate from the QDs to the electron transport layer
and reduces the carrier recombination processes as the shell layer acts as a tunnelling
barrier for the hole localized in the core region. In addition, the prolonged lifetime of
the charge carrier in type-II core/shell QDs can create a dipole moment through the
accumulation of a negative charge in the electron transport layer (e.g., TiO2 or ZnO)
and a positive charge in the core of the QDs. This dipole moment leads to the upshift
in the CB of the electron transport layer and resulting solar cells yield higher open-
circuit photovoltage compared to core QD-based solar cells [30, 105]. A significant
redshift of the absorption edge of type-II core/shell QDs due to the exciplex state
offers a new route to broaden the light absorption range by reducing the effective
Core/Shell Quantum-Dot-Sensitized Solar Cells 239

Fig. 7 TEM images of as-synthesized (a) CdS core QDs with size of 2.9 nm; (b)–(e) CdS/CdSe
core/shell QDs with sizes of 4.6, 5.7, 6.1, and 6.7 nm. (f) Schematic illustration of band
edge alignment of reverse type-I CdS/CdSe core/shell QD-sensitized TiO2 -based solar cell. (g)
Absorption spectra of MPA-capped water-soluble CdS/CdSe QD-sensitized TiO2 films of 4-μm
thickness. Inset displays photographs of corresponding photoanode films. (h) J-V curves of QDSCs
based on reverse type-I CdS/CdSe core/shell QDs; (i) IPCE spectra corresponding to QDSCs.
(Reproduced with permission from Ref [78]. Copyright 2012, American Chemical Society)

bandgap [106, 107]. All these appealing features of the type-II core/shell QDs make
them as promising light-harvesting materials for QDSCs and other optoelectronic
devices. For the first time, Ning et al. [108] synthesized type-II ZnSe/CdS core/shell
QDs and fabricated QDSCs with prominent absorbed photon to current conversion
efficiency and PCE of 0.27%. Luo et al. [109] designed and developed a microwave-
assisted aqueous approach to synthesize type-II CdSex Te1-x /CdS core/shell QDs.
Resulting QDSC-based CdSex Te1-x /CdS core/shell QDs show a PCE of 5.04%
under one sun illumination (AM 1.5 G).
240 G. S. Selopal

Later, Bang et al. [110] also synthesized environmentally friendly type-II


ZnTe/ZnSe core/shell QDs that are explored as light-harvesting materials in solar
cell application. However, the PCE of the QDSCs based on these type-II core/shell
QDs is still lower than that of the predicted theoretical limit of PCE for type-
II core/shell QDs due to limited loading of colloidal QDs in the mesoporous
electron transport layer. Then, Wang et al. [30] synthesized type-II CdTe/CdSe
core/shell QDs synthesized via the combination of the hot-injection and SILAR
methods. Briefly, first, CdTe core QDs of the average size of 2.7 ± 0.2 nm were
synthesized via the hot-injection method in which CdO-tetradecylphosphonic acid
(TDPA) reacts with trioctylphosphine (TOP)-Te in the 1-octadecene medium at high
temperature (Fig. 8a). As-synthesized CdTe core QDs were purified and dispersed
in a chloroform solution. Then CdSe shell layers were grown over the CdTe core
QDs via the SILAR approach. After the three cycles of CdSe shell layers, an
average size of 4.9 ± 0.3 nm of CdTe/CdSe core/shell QDs was obtained (Fig.
8b). Finally, the oil-soluble type-II CdTe/CdSe core/shell QDs were transferred to
MPA-capped water-soluble QDs. The absorption spectrum of CdTe core QDs shows
the first excitonic absorption peak at 540 nm and PL emission peak at 558 nm,
whereas with the growth of three CdSe shell layers, CdTe/CdSe core/shell QDs
show broadening of the absorption spectrum toward a longer wavelength with first
excitonic peak at 785 nm and PL emission peak redshift from 558 to 795 nm (Fig.
8c). QDSCs based on MPA-capped type-II CdTe/CdSe core/shell QDs reported
impressive PCE of 6.75%, a new record for type-II core/shell QD-based QDSCs.
However, the relatively low value of Voc still needs to be improved to further
boost the PCE of the QDSCs based on type-II core/shell QDs. Thus, Jiao et al.
[111] explored the novel ZnTe/CdSe type-II core/shell QDs as a light-harvesting
material in the fabrication of QDSCs. As-synthesized ZnTe/CdSe type-II core/shell
QDs show an average size of 5.3 nm with ZnTe core of 3.2 nm and CdSe shell of
2.1 nm (Fig. 8d). Figure 8e displays the TEM image of QD-sensitized mesoporous
TiO2 film. It can be seen that ZnTe/CdSe type-II core/shell QDs display better
optoelectronic and relative band alignment compared to the CdTe/CdSe core/shell
QDs. In particular, ZnTe/CdSe core/shell QDs show much higher CB offset potential
(1.22 eV) compared to the (0.27 eV) CdTe/CdSe core/shell QDs (Fig. 8f). The
higher band offset of ZnTe/CdSe than CdTe/CdSe QDs produces an increase of
charge accumulation at QD/TiO2 interfaces under illumination, which results in
creating a stronger dipole moment and leads to a greater upward shift in the CB edge
of QD-sensitized TiO2 photoanodes and enhances the Voc of QDSCs. In addition,
the reduced effective bandgap of type-II ZnTe/CdSe core/shell QDs broadens the
light absorption spectrum from the visible to the NIR region (Fig. 8g). Resulting
QDSCs based on the MPA-capped aqueous type-II ZnTe/CdSe core/shell QDs yield
a PCE of 7.17%, which is 8% higher than that of previously designed type-II
CdTe/CdSe core/shell QDs (Fig. 8h). This significant improvement in the PCE of
QDSCs based on type-II ZnTe/CdSe core/shell QDs is mainly derived from the
enhanced Voc of 0.646 V compared to 0.597 V for CdTe/CdSe QD-based QDSCs
through the band engineering approach. The IPCE of the respective QDSCs shows
an ~ 75% in the range of 360–680 nm for both CdSe, CdTe/CdSe and ZnTe/CdSe,
Core/Shell Quantum-Dot-Sensitized Solar Cells 241

Fig. 8 Structural and optical characterization: TEM images of as-synthesized QDs: (a) CdTe
core QDs; (b) CdTe/CdSe core/shell QDs; and (c) normalized absorption (solid lines) and PL
spectra (dashed line) with emission wavelength of 360 nm of CdTe (black lines) and type-II
CdTe/CdSe core/shell QDs (red lines). Reproduced with permission from Ref [30]. Copyright
2013, American Chemical Society. TEM images: (d) as-synthesized ZnTe/CdSe core/shell QDs
with 3 monolayers of CdS shell; (e) ZnTe/CdSe core/shell QD-sensitized TiO2 mesoporous film.
(f) Schematic diagram of the bandgap and band offsets (in eV) for the interfaces between bulk
ZnTe/CdSe and CdTe/CdSe core/shell QDs. (g) Absorption spectra of ZnTe/CdSe, CdTe/CdSe, and
CdSe QD-sensitized TiO2 films with a 6.0-μm transparent layer. Inset display of the photographs
of corresponding QD-sensitized TiO2 film electrodes. Photovoltaic performance of QDSCs based
on ZnTe/CdSe and reference CdTe/CdSe and CdSe QDs: (h) J-V curves of champion cells; (i)
IPCE curves of corresponding devices. Reproduced with permission from Ref [111]. Copyright
2015, American Chemical Society

but the QDSCs based on CdTe/CdSe and ZnTe/CdSe core/shell QDs also show
IPCE response in the broader range of 350–900 nm (Fig. 8i).These results highlight
that the band engineering of QDs through the core/shell (type-I, reverse type-I, and
type-II) approach is the most effective to improve the optoelectronic properties of
the QDs and hence to enhance the performance of QDSCs based on the respective
core/shell QDs.
242 G. S. Selopal

7.4 Quasi-Type-II Core/Thick-Shell QDs (“Giant” Core/Shell


QDs)

The performance of solar energy conversion devices depends on the efficiency of


photogenerated electrons injected from the photoexcited QDs to the wide-bandgap
mesoporous electron transport layer. In particular, to fabricate high-performing
QDSCs, understanding of separation and injection of photogenerated carriers
(electron and hole) at the QD/scavenger interface is a key point. In core/shell QDs,
the band alignment between core and shell materials depends on the nature of
the constitute materials (type-I, reverse type-I, and type-II, as discussed above)
and can be tuned by tailoring the core size and shell thickness. The tunable
shell thickness of core/shell QDs allows us to precisely tune the carrier transfer
rates from the photoexcited QDs to the carrier scavengers. In addition, there is a
possibility of formation of quasi-type-II core/shell QD system with the increase
of the shell thickness (1.5 nm up to tens of nm) in which one type of carrier is
confined in the core, whereas the other is delocalized over the entire core/shell
structure (Fig. 4d). This spatial separation reduces the carrier recombination and
hence boosts the performance of QDSCs. In this context, Selopal et al. [112]
designed and synthesized “giant” CdSe/CdS core/shell QDs with a CdSe core
of 1.65 nm in size and a CdS shell of tunable shell thickness from 0.66 to
4.51 nm to investigate the carrier dynamics as a function of CdS shell thickness
and photovoltaic performance of QDSCs based on respective “giant” CdSe/CdS
core/shell QDs (Fig. 9a). Typically, the first CdSe core QDs were synthesized by the
hot-injection approach. Then CdSe/CdS core/shell QDs were prepared by growing
a thick CdS shell layer over CdSe core QDs by the SILAR approach at 240 ◦ C
under continuous N2 flow. The diameter of the as-synthesized CdSe core QDs is
3.30 ± 0.29 nm (Fig. 9b). The final diameter of CdSe/CdS core/shell QDs increases
with the growth of 2, 6, and 13 monolayers of CdS at 4.6 ± 0.5 nm, 7.2 ± 0.5 nm
and 12.3 ± 1.1 nm, respectively (Fig. 9c–d).
The electron (ψ e (r)) and hole (ψ h (r)) wave function of CdSe/CdS core/shell
QDs with tunable CdSe shell thickness is calculated by solving the stationary
Schrödinger equation in spherical geometry. Results demonstrated with the increase
of CdS shell thickness, the ψ e (r) shows enhanced leakage of the electrons into
the shell region, while the ψ h (r) displayed that the hole is confined in the CdSe
core (Fig. 9e). This highlights that with the increase of CdS shell thickness, type-
I CdSe/CdS core/shell QDs displays the quasi-type II band alignment and spatial
overlap area between electron and hole wave functions decreasing from 94% for
CdSe core QDs to 57% for “giant” CdSe/CdS core/shell QDs with CdS shell
thickness of 4.51 nm. This significantly enhances the electron lifetime of “giant”
CdSe/CdS core/shell QDs compared to CdSe core QDs, which is further confirmed
by the trend of electron lifetime calculated from the transient PL decay [112].
To investigate the effect of the CdS shell layer on the carrier transfer of
CdSe/CdS core/shell QDs, as-synthesized CdSe/CdS core/shell QDs with different
shell thicknesses were grafted on mesoporous semiconductor metal oxide film,
Core/Shell Quantum-Dot-Sensitized Solar Cells 243

Fig. 9 (a) Schematic illustration of CdSe/CdS core/shell QDs with tunable CdS shell thickness
(yellow color), with a constant CdSe core (red color) radius of 1.65 nm. TEM images of as-
synthesized: (b) CdSe core, (c) CdSe/CdS core/shell with 6-shell layers, and (d) CdSe/CdS
core/shell with 13-shell layers. (e) Spatial probability distribution [ρ (r)] of electron and hole wave
functions of the corresponding QDs as a function of the QD radius (R + H, nm). (f) PL intensity
decay for the core/shell QDs of different shell thicknesses. Calculated carrier dynamics parameters
as a function of CdS shell thickness, from the electron lifetime measurement of core/shell QDs: (g)
electron lifetime and (h) electron transfer rate. (i) Current density–voltage curves of QDSCs based
on CdSe/CdS core/shell QDs as a function of CdS shell thickness; (j) open-circuit voltage decay
as the function of time; (k) electron lifetime as a function of Voc of the corresponding QDSCs.
(Reproduced with permission from Ref [112]. Copyright 2017, John Wiley and Sons Ltd)
244 G. S. Selopal

measuring the transient PL decay rates (Fig. 9f). Results demonstrated that with the
increase of CdS shell thickness from 0 to 4.51 nm, the electron lifetime increases
from 20 ± 2 to 36 ± 1 ns, respectively (Fig. 9g). This enhanced the electron lifetime
value with the growth of the CdS shell that is mainly ascribed to the reduced spatial
electron–hole overlap due to the leakage of the electron in the CdS shell region and
the hole confined in the CdSe core region. On the other hand, the electron transfer
rate decreases from 2.2 ± 0.1 × 107 s−1 to 0.6 ± 0.1 × 107 s−1 with the CdS shell
thickness increasing from 0.0 to 4.51 nm, respectively (Fig. 9h). This is mainly
attributed to large barrier potential of the thick CdS shell and undesirable carrier
confinement at the CdSe core and CdS shell interface due to large lattice mismatch
of 4.4%. QDSCs were fabricated based on these CdSe/CdS core/shell QDs with
tunable shell thickness from 0.66 to 4.51 nm. The photovoltaic performance of
QDSCs was initially increasing with the increase in CdS shell thickness from 0.0
to 1.96 nm and then decreases with the further increase in CdS shell thickness
from 1.96 to 4.51 nm (Fig. 9i). Briefly, QDSCs based on CdSe/CdS core/shell
QDs with a shell thickness of 1.96 nm yield a PCE of 3.01% under one sun
illumination (AM 1.5 G, 100 mWcm−2 ), which is significantly higher than bare
CdSe (PCE = 1.22%). This significant improvement in the PCE is mainly attributed
to reduced carrier recombination at the QD/TiO2 /electrolyte interface and enhanced
electron lifetime (Fig. 9j–k) and broadening of the light absorption spectrum
compared to CdSe core QDs [112]. Recently, Ghosh et al. [113] also synthesized
CdSe/CdS core/shell QDs with tunable CdS shell thickness and investigated the
modulation of carrier dynamics with shell thickness by using ultrafast transient
absorption (TA) and correlated the PV performance of QDSCs. TA measurement
demonstrated that with the increase in CdS shell thickness, carrier recombination
reduces due to the decoupling of the electron–hole increases (electron delocalized
in the shell region, while the hole confined in the core region) and carrier cooling
time enhances from 500 fs (100%) for bare CdSe QDs to biexponential 1 ps
(72%) and 6 ps (28%) for CdSe/CdS core/shell QDs (6ML). However, the electron
transfer decreases from 750 fs (−67%) to 1.75 ps (−69%), 2.5 ps (−65%), and
3.5 ps (−63%) for CdSe, CdSe/CdS (2 ML), CdSe/CdS (4 ML), and CdSe/CdS
(6 ML), respectively, as the CdS shell layer acts as a potential barrier for the
electron transfer. Furthermore, carrier dynamics measurement of “giant” CdSe/CdS
core/shell QDs (CdSe core = 1.65 nm and shell thickness = 4.3 nm) coupled with
TiO2 demonstrated that the electron transfer rate is 1.5- to 1.9-fold lower than that
of CdSe core QDs, whereas the hole transfer rate is four- to fivefold lower than that
of CdSe core QDs [114]. This difference in the decrease in the value of electron
transfer rate compared with the decrease in hole transfer rate for “giant” CdSe/CdS
core/shell QDs leads to the formation of quasi-type-II band alignment [115, 116].
In addition to thick shell core/shell with constant core size, PbS/CdS core/shell
QDs with different PbS core size and CdS shell thicknesses were synthesized [117].
Carrier dynamics of as-synthesized PbS/CdS core/shell QDs were investigated by
coupling QDs with different types of wide-bandgap semiconductor metal oxide
films such as SiO2 , TiO2 , and SnO2 and evaluated the trend of electron and hole
transfer rates by using transient PL spectroscopy. Results demonstrated that the fast
Core/Shell Quantum-Dot-Sensitized Solar Cells 245

charge injection rate for QD-sensitized TiO2 mesoporous film is in the range of 110–
250 ns and for sensitized SnO2 mesoporous film is in the range of 100–170 ns for
PbS core with diameters in the 3–4.2-nm range and CdS shell thickness of 0.3 nm
of PbS/CdS core/shell QDs.
Therefore, the optimization of the shell thickness and core size of core/shell QDs
is a crucial factor to tune the carrier dynamics and broaden the absorption spectrum
toward a longer wavelength with a suitable electron transfer rate and hole tunnelling
through a shell barrier.

7.5 Core and Shell Interface Optimization

A core/shell heterostructured system offers promising optoelectronic properties and


superior photophysical/chemical stability due to reduced surface trap/defect states
by the surface passivation of core QDs with the shell of different materials and
thickness [114, 118–119]. However, core/shell QDs suffer from the formation of
interfacial defects at the core and shell sharp interface during the shell growth due
to lattice mismatch between the core and shell materials (e.g., 4.4% for CdSe and
CdS) [120]. These interfacial defects cause the undesirable carrier recombination
during the device operation and reduce the overall performance [121]. In addition,
the growth of the shell layer over the core QDs reduces the electron and hole
transfer rate as the shell layer acts as an energy barrier potential for the electron
and hole injection process [122]. Thus, understanding and optimization of these
interfacial strains between the core and shell are required to fabricate efficient
core/shell QD-based solar cells. In this context, interfacial engineering of core/shell
QDs by addition of an interfacial layer between the core and shell has attracted great
attention to reduce the lattice mismatch and modulate the band alignment to improve
the electron and hole transfer rate and reduce the undesirable carrier recombination.
Recently, different types of core/alloyed layer/shell QD systems were synthesized
and their structural and optoelectronic properties were studied through several
spectroscopic and theoretical measurements. Among all, “giant” core/shell QDs
with interfacial layers such as PbS/CdS/(CdS)n [123], CdSe/Pbx Cd1-x S/CdS [124],
and CdSe/CdSex S1-x /CdS [125–128] are the most commonly studied due to their
special feature and suitability for a broad range of applications in optoelectronic
devices. These core/shell QDs with alloyed interfacial layers can be synthesized
either by the combination of the cation exchange and SILAR approach or only by
the SILAR approach depending on the final structure and size of QDs (Fig. 4).
For example, “giant” PbS/CdS/(CdS)n core/shell/shell QDs were synthesized by
the combination of the cation exchange and SILAR approach [123]. For the dual-
color-emission PbS/CdS/(CdS)n core/shell/shell QDs, the first PbS core QDs were
synthesized by hot injection using OLA as ligands and purified using ethanol and
re-dispersed in toluene. Then PbS/CdS core/thin-shell QDs were synthesized by
two-step cation exchange from PbS QDs. After the growth of a thin CdS shell,
a thick CdS shell layers were grown over the PbS/CdS core/thin-shell QDs via
246 G. S. Selopal

the SILAR approach by adding equal molar ratio Cd and S source precursors.
On the other hand, single-color-emission core/shell/shell QDs were synthesized
using the Cd and S precursors with a molar ratio of 1:0.8. Then PbS/CdS/CdS
core/shell/shell QDs were washed with ethanol and re-dispersed in toluene for
characterizations and device applications. Selopal et al. [126] also design and
synthesized CdSe/(CdS)6 core/shell QDs and CdSe/CdSex S1−x /CdS core/shell QDs
with different numbers of monolayers of interfacial layer CdSex S1−x (x = 0.5
or x = 0.9 ~ 0.1). Figure 10a–c displays the schematic of band alignment of
CdSe/(CdS)6 core/shell QDs (denoted by CS) and CdSe/CdSex S1−x /CdS core/shell
QDs with x = 0.5 or x = 0.9 ~ 0.1. Briefly, the first CdSe core QDs of diameter
3.3 nm were synthesized by hot injection. Then four cycles of CdSe1−x Sx -alloyed
interfacial layer (x = 0.5) were grown over the CdSe core QDs via the SILAR
approach by adding the mixture of 0.2 M Se and S in ODE instead of 0.2 M S.
Finally, two monolayers of CdS shell were grown over the CdSe/(CdSe0.5 S0.5 )4
core/alloyed shell QDs via the SILAR approach at 240 ◦ C under continuous
N2 flow to synthesize CdSe/(CdSe0.5 S0.5 )4 /(CdS)2 (denoted as CSA1). Similarly,
CdSe/(CdSex S1−x )5 /(CdS)1 (denoted as CSA2) were synthesized via the SILAR
approach by growing five interfacial layers of CdSex S1−x (x = 0.9 ~ 0.1) and one
monolayer of CdS. It can be seen that the final diameter of as-synthesized CS
core/shell QDs is 7.2 ± 0.5 nm, whereas for CSA1 core/alloyed shell/shell QDs
is 7.4 ± 0.7 nm and CSA2 core/graded alloyed shell/shell QDs is 7.6 ± 1.3 nm
(Fig. 10d–f). The clearly visible lattice fringes shown in the high-resolution TEM
(HRTEM) image of corresponding QDs [inset of each Fig. 10d–f] confirm the high
crystallinity of each QDs.
The effect of the interface engineering on the carrier dynamics was carried
out by transient PL spectroscopic decay measurements. Results demonstrated that
addition of CdSex S1−x interfacial layers between the CdSe core and CdS shell
leads not only to improve the electron and hole transfer rates (Fig. 10h–i), but also
to broaden the absorption spectrum compared to the reference “giant” CdSe/CdS
core/shell QDs. However, the electron lifetime of core/shell QDs decreases with
the incorporation of CdSex S1−x interfacial layers from 29 ± 0.4 ns for CS QDs to
24 ± 0.4 ns and 14 ± 0.4 ns for CSA1 and CSA2 QDs (Fig. 10g). This decrease in
the electron lifetime values is mainly attributed to the favorable band alignment of
CSA1 and CSA2 QDs that allows the leakage of both electron and holes in the shell
region, whereas in CS QDs only electrons leak into the shell region. This enhanced
possibility of the holes leaking into the shell region is mainly attributed to reduced
overall interfacial barrier potential with the incorporation of CdSex S1−x interfacial
layers that could slow down or even block hole transfer from the CdSe core into the
shell region. This leads to the fast electron–hole recombination, thus decreasing
the overall electron lifetime values for CSA1 and CSA2 QDs compared to CS
QDs. This increased electron and hole wave function overlapping was confirmed by
the theoretical simulation. QDSCs were fabricated using these specially designed
core/shell QDs as a light harvester. Resulting QDSCs yield a PCE of 3.08% for
CS QDs, 5.52% for CSA1, and 7.14% for CSA2 QDs under one sun simulated
sunlight (AM 1.5 G) (Fig. 10j), which is consistent with optoelectronic properties
Core/Shell Quantum-Dot-Sensitized Solar Cells 247

Fig. 10 Schematic illustration of internal interfacial structures and carrier confinement potentials
of QDs: (a) CdSe/(CdS)6 (R = 1.65 nm, H = 1.96 nm), (b) CdSe/(CdSex S1−x )4 /(CdS)2
(x = 0.5 for all monolayers, R = 1.65 nm, H1 = 1.39 nm, H2 = 0.66 nm), and (c)
CdSe/(CdSex S1−x )5 /(CdS)1 (x = 0.9–0.1, R = 1.65 nm, H3 = 0.82 nm, H4 = 0.33 nm). R is radius
of CdSe QDs and H is the shell thickness. TEM images of as-synthesized QDs: (d) CdSe/(CdS)6 ,
(e) CdSe/(CdSex S1−x )4 /(CdS)2 , and (f) CdSe/(CdSex S1−x )5 /(CdS)1 . Inset of each figure displays
the HR-TEM images of corresponding QDs. Comparison of the calculated parameters from the
transient PL curves of each types of QDs: (g) electron lifetime, (h) electron transfer rate, and (i)
hole transfer rate. (j) Comparison of current density versus voltage curves of QDSCs based on these
QDs under one sun irradiation (AM 1.5G, 100 mW/cm−2 ); (k) electron lifetime (τ ) as a function
of Voc , calculated from transient photovoltage decay measurements. (l) Variation of Jsc (mA/cm2 )
(black color, left) and PCE (%) (red color, right) of QDSCs with QD structure. (Reproduced with
permission from Ref [126]. Copyright 2019 Elsevier)

of the respective QDs. The calculated electron lifetime of QDSCs from the transient
photovoltage decay measurements confirms the reduced recombination in QDSCs
by tailoring the structure of core/shell QDs via adding the alloyed and graded
alloyed interfacial layers at the core/shell interface (Fig. 10k). The trend of Jsc
and PCE as a function of the structure of different types of QDs is shown in Fig.
10l. Zhou et al. [129] reported the synthesis of CdS/CdSe core/shell QDs with a
CdSx Se1−x interfacial layer between the CdS core and CdSe shell. Resulting graded
alloyed CdS/CdSx Se1−x /CdSe core/shell QD-based QDSCs yield a PCE of 5.06%,
which is 14% higher than the conventional CdS/CdSe core/shell QD-based QDSCs
248 G. S. Selopal

(4.41%). This enhanced photovoltaic performance of QDSCs is mainly attributed to


improved charge transfer rate and reduced recombination with the passivation of the
interfacial defects and interphase strains.
Therefore, the results of these studies demonstrated that the interface engineering
of core/shell QDs is a promising approach to tune the carrier dynamics of QDs
without alerting their size and shape.

8 Conclusions

In conclusion, this chapter presents an overview of the recent progress in the


development of different types of colloidal core/shell QDs including type-I, type-
II, and quasi-type-II core/shell QDs as promising light-harvesting materials to boost
the performance of QDSCs. In particular, we described how the band alignment
between core and shell materials plays a major role to tune the optoelectronic
properties of QDs and to reduce the carrier dynamics within the QDs and at the
QD/carrier scavenger (metal oxide and electrolyte) interfaces in QDSCs. Although,
in the last few years, due to considerable development in materials design, device
architecture and technology advance, the performance of QDSCs has reached above
13%. This record value of PCE is still lower than that of expected theoretical value
due to several unrevealed recombination processes in the different components
(mesoporous anode, electrolyte, and CEs) and interfaces of QDSCs (QDs/TiO2
or ZnO, QDs/electrolyte, electrolyte/TiO2 or ZnO, and CE/electrolyte). Thus, the
optimization of each component and interfaces of QDSCs along with engineering
of QDs is required to make further breakthroughs in the performance of QDSCs.

Acknowledgments The author acknowledges funding support from the UESTC, China and
INRS-EMT, Canada. The Author is greatful to Dr. Gopal Singh for his valuable discussion. The
Author also thanks to Mrs. Ravneet Kaur for her support.

References

1. Bickerstaff, K., Walker, G.: Public understandings of air pollution: the ‘localisation’ of
environmental risk. Glob. Environ. Chang. 11, 133–145 (2001)
2. Nayak, P.K., Garcia-Belmonte, G., Kahn, A., Bisquert, J., Cahen, D.: Photovoltaic efficiency
limits and material disorder. Energy Environ. Sci. 5, 6022–6039 (2012)
3. Morton, O.: A new day dawning?: Silicon Valley sunrise. Nature. 443, 19–22 (2006)
4. Holdren, J.P.: Science and technology for sustainable well-being. Science. 319, 424–434
(2008)
5. Zhao, J., Wang, A., Greeen, M.A., Ferrazza, F.: 19.8% efficient “honeycomb” textured
multicrystalline and 24.4% monocrystalline silicon solar cells. App. Phys. Lett. 73, 1991–
1993 (1998)
6. Pan, Z., Rao, H., Mora-Sero, I., Bisquert, J., Zhong, X.: Quantum dot-sensitized solar cells.
Chem. Soc. Rev. 47, 7659–7702 (2018)
Core/Shell Quantum-Dot-Sensitized Solar Cells 249

7. Nozik, A.J., Beard, M.C., Luther, J.M., Law, M., Ellingson, R.J., Johnson, J.C.: Semiconduc-
tor quantum dots and quantum dot arrays and applications of multiple Exciton generation to
third-generation photovoltaic solar cells. Chem. Rev. 110, 6873–6890 (2010)
8. Semonin, O.E., Luther, J.M., Choi, S., Chen, H.-Y., Gao, J., Nozik, A.J., Beard, M.C.: Peak
external photocurrent quantum efficiency exceeding 100% via MEG in a quantum dot solar
cell. Science. 334, 1530–1533 (2011)
9. Kim, M.R., Ma, D.: Quantum-dot-based solar cells: recent advances, strategies, and chal-
lenges. J. Phys. Chem. Lett. 6, 85–99 (2015)
10. Wang, D., Yin, F., Du, Z., Han, D., Tang, J.: Recent progress in quantum dot-sensitized solar
cells employing metal chalcogenides. J. Mater. Chem. A. 7, 26205–26226 (2019)
11. Gerischer, H., Lubke, M.: A particle size effect in the sensitization of TiO2 electrodes by a
CdS deposit. J. Electroanal. Chem. 204, 225–227 (1986)
12. Vogel, R., Pohl, K., Weller, H.: Sensitization of highly porous, polycrystalline TiO2 electrodes
by quantum sized CdS. Chem. Phys. Lett. 174, 241–246 (1990)
13. Zaban, A., Micic, O.I., Gregg, B.A., Nozik, A.J.: Photosensitization of Nanoporous TiO2
electrodes with InP quantum dots. Langmuir. 14, 3153–3156 (1998)
14. Yu, P., Zhu, K., Norman, A.G., Ferrere, S., Frank, A.J., Nozik, A.J.: Nanocrystalline TiO2
solar cells sensitized with InAs quantum dots. J. Phys. Chem. B. 110, 25451–25454 (2006)
15. Diguna, L.J., Shen, Q., Kobayashi, J., Toyoda, T.: High efficiency of CdSe quantum-dot-
sensitized TiO2 inverse opal solar cells. Appl. Phys. Lett. 91, 023116 (2007)
16. Lee, Y.-L., Huang, B.-M., Chien, H.-T.: Highly efficient CdSe-sensitized TiO2 photoelectrode
for quantum-dot-sensitized solar cell applications. Chem. Mater. 20, 6903–6905 (2008)
17. Lee, Y.-L., Lo, Y.-S.: Highly efficient quantum-dot-sensitized solar cell based on co-
sensitization of CdS/CdSe. Adv. Funct. Mater. 19, 604–609 (2009)
18. Gimenez, S., Mora-Sero, I., Macor, L., Guijarro, N., Lana-Villarreal, T., Gomez, R., Diguna,
L.J., Shen, Q., Toyoda, T., Bisquert, J.: Improving the performance of colloidal quantum-dot-
sensitized solar cells. Nanotechnology. 20, 295204 (2009)
19. Zhang, Q., Guo, X., Huang, X., Huang, S., Li, D., Luo, Y., Shen, Q., Toyoda, T., Meng,
Q.: Highly efficient CdS/CdSe-sensitized solar cells controlled by the structural properties of
compact porous TiO2 photoelectrodes. Phys. Chem. Chem. Phys. 13, 4659–4667 (2011)
20. Fang, B., Kim, M., Fan, S.-Q., Kim, J.H., Wilkinson, D.P., Ko, J., Yu, J.-S.: Facile synthesis of
open mesoporous carbon nanofibers with tailored nanostructure as a highly efficient counter
electrode in CdSe quantum-dot-sensitized solar cells. J. Mater. Chem. 21, 8742–8748 (2011)
21. Selopal, G.S., Concina, I., Milan, R., Natile, M.M., Sberveglieri, G., Vomiero, A.: Hierarchi-
cal self-assembled Cu2 S nanostructures: fast and reproducible spray deposition of effective
counter electrodes for high efficiency quantum dot solar cells. Nano Energy. 6, 200–210
(2014)
22. Jiang, Y., Yu, B.B., Liu, J., Li, Z.H., Sun, J.K., Zhong, X.H., Hu, J.S., Song, W.G., Wan, L.J.:
Boosting the open circuit voltage and fill factor of QDSSCs using hierarchically assembled
ITO@Cu2 S nanowire array counter electrodes. Nano Lett. 15, 3088–3095 (2015)
23. Zhang, H., Wang, C., Peng, W.X., Yang, C., Zhong, X.H.: Quantum dot sensitized solar
cells with efficiency up to 8.7% based on heavily copper-deficient copper selenide counter
electrode. Nano Energy. 23, 60–69 (2016)
24. Wang, S.X., Shen, T., Bai, H.W., Li, B., Tian, J.J.: Cu3 Se2 nanostructure as a counter electrode
for high efficiency quantum dot-sensitized solar cells. J. Mater. Chem. C. 4, 8020–8026
(2016)
25. Ghosh, D., Halder, G., Sahasrabudhe, A., Bhattacharyya, S.: A microwave synthesized Cux S
and graphene oxide nanoribbon composite as a highly efficient counter electrode for quantum
dot sensitized solar cells. Nanoscale. 8, 10632–10641 (2016)
26. Selopal, G.S., Chahine, R., Mohammadnezhad, M., Navarro-Pardo, F., Benetti, D., Zhao, H.,
Wang, Z.M., Rosei, F.: Highly efficient and stable spray assisted nanostructured Cu2 S/Carbon
paper counter electrode for quantum dots sensitized solar cells. J. Power Sources. 436, 226849
(2019)
250 G. S. Selopal

27. Chakrapani, V., Baker, D., Kamat, P.V.: Understanding the role of the sulfide redox couple
(S2–/Sn 2–) in quantum dot-sensitized solar cells. J. Am. Chem. Soc. 133, 9607–9615 (2011)
28. Radich, J.G., Dwyer, R., Kamat, P.V.: Cu2 S reduced graphene oxide composite for high-
efficiency quantum dot solar cells. Overcoming the redox limitations of S2 –/Sn 2– at the
counter electrode. J. Phys. Chem. Lett. 2, 2453–2460 (2011)
29. Santra, P.K., Kamat, P.V.: Mn-doped quantum dot sensitized solar cells: a strategy to boost
efficiency over 5% J. Am. Chem. Soc. 134, 2508–2511 (2012)
30. Wang, J., Mora-Seró, I., Pan, Z., Zhao, K., Zhang, H., Feng, Y., Yang, G., Zhong, X., Bisquert,
J.: Core/shell colloidal quantum dot exciplex states for the development of highly efficient
quantum-dot-sensitized solar cells. J. Am. Chem. Soc. 135, 15913–15922 (2013)
31. Pan, Z., Mora-Sero, I., Shen, Q., Zhang, H., Li, Y., Zhao, K., Wang, J., Zhong, X., Bisquert, J.:
High-efficiency “green” quantum dot solar cells. J. Am. Chem. Soc. 136, 9203–9210 (2014)
32. Zhao, K., Pan, Z., Mora-Sero, I., Canovas, E., Wang, H., Song, Y., Gong, X., Wang, J., Bonn,
M., Bisquert, J., Zhong, X.: Boosting power conversion efficiencies of quantum-dot-sensitized
solar cells beyond 8% by recombination control. J. Am. Chem. Soc. 137, 5602–5609 (2015)
33. Du, Z., Pan, Z., Fabregat-Santiago, F., Zhao, K., Long, D., Zhang, H., Zhao, Y., Zhong, X.,
Yu, J., Bisquert, J.: Carbon counter-electrode-based quantum-dot-sensitized solar cells with
certified efficiency exceeding 11%. J. Phys. Chem. Lett. 7, 3103–3111 (2016)
34. Jiao, S., Du, J., Du, Z., Long, D., Jiang, W., Pan, Z., Li, Y., Zhong, X.: Nitrogen-doped
mesoporous carbons as counter electrodes in quantum dot sensitized solar cells with a
conversion efficiency exceeding 12%. J. Phys. Chem. Lett. 8, 559–564 (2017)
35. Pan, Z., Yue, L., Rao, H., Zhang, J., Zhong, X., Zhu, Z., Jen, A.K.-Y.: Boosting the perfor-
mance of environmentally friendly quantum dot-sensitized solar cells over 13% efficiency by
dual sensitizers with cascade energy structure. Adv. Mater. 31, 1903696 (2019)
36. Plass, R., Pelet, S., Krueger, J., Grätzel, M., Bach, U.: Quantum dot sensitization of
organic−inorganic hybrid solar cells. J. Phys. Chem. B. 106, 7578–7580 (2002)
37. Pan, Z., Zhao, K., Wang, J., Zhang, H., Feng, Y., Zhong, X.: Near infrared absorption of
CdSex Te1–x alloyed quantum dot sensitized solar cells with more than 6% efficiency and high
stability. ACS Nano. 7, 5215–5222 (2013)
38. Ren, Z., Wang, J., Pan, Z., Zhao, K., Zhang, H., Li, Y., Zhao, Y., Mora-Sero, I., Bisquert, J.,
Zhong, X.: Amorphous TiO2 buffer layer boosts efficiency of quantum dot sensitized solar
cells to over 9%. Chem. Mater. 27, 8398–8405 (2015)
39. Ren, Z., Wang, Z., Wang, R., Pan, Z., Gong, X., Zhong, X.: Effects of metal oxyhydroxide
coatings on photoanode in quantum dot sensitized solar cells. Chem. Mater. 28, 2323–2330
(2016)
40. Du, J., Du, Z., Hu, J.S., Pan, Z., Shen, Q., Sun, J., Long, D., Dong, H., Sun, L., Zhong, X.,
Wan, L.J.: Zn-Cu-In-Se quantum dot solar cells with a certified power conversion efficiency
of 11.6%. J. Am. Chem. Soc. 138, 4201–4209 (2016)
41. Wang, W., Feng, W., Du, J., Xue, W., Zhang, L., Zhao, L., Li, Y., Zhong, X.: Cosensitized
quantum dot solar cells with conversion efficiency over 12%. Adv. Mater. 30, 1705746 (2018)
42. O’Regan, B., Grätzel, M.: A low-cost, high-efficiency solar cell based on dye-sensitized
colloidal TiO2 films. Nature. 353, 737–740 (1991)
43. Bai, Y., Mora-Seró, I., De Angelis, F., Bisquert, J., Wang, P.: Titanium dioxide nanomaterials
for photovoltaic applications. Chem. Rev. 114, 10095–10130 (2014)
44. Chen, C., Xie, Y., Ali, G., Yoo, S.H., Cho, S.O.: Improved conversion efficiency of CdS quan-
tum dots-sensitized TiO2 nanotube array using ZnO energy barrier layer. Nanotechnology. 22,
015202 (2011)
45. Liu, B., Sun, Y., Wang, X., Zhang, L., Wang, D., Fu, Z., Lin, Y., Xie, T.: Branched hierarchical
photoanode of anatase TiO2 nanotubes on rutile TiO2 nanorod arrays for efficient quantum
dot-sensitized solar cells. J. Mater. Chem. A. 3, 4445–4452 (2015)
46. Qiu, Q., Li, S., Jiang, J., Wang, D., Lin, Y., Xie, T.: Improved Electron transfer between
TiO2 and FTO interface by N-doped anatase TiO2 nanowires and its applications in quantum
dot-sensitized solar cells. J. Phys. Chem. C. 121, 21560–21570 (2017)
Core/Shell Quantum-Dot-Sensitized Solar Cells 251

47. Zhang, Q., Dandeneau, C.S., Zhou, X., Cao, G.: Zno nanostructures for dye-sensitized solar
cells. Adv. Mater. 21, 4087–4108 (2009)
48. Djurisic, A.B., Chen, X., Leung, Y.H., Ng, A.M.C.: ZnO nanostructures: growth, properties
and applications. J. Mater. Chem. 22, 6526–6535 (2012)
49. Xu, J., Yang, X., Yang, Q.-D., Wong, T.-L., Lee, S.-T., Zhang, W.-J., Lee, C.-S.: Arrays of
CdSe sensitized ZnO/ZnSe nanocables for efficient solar cells with high open-circuit voltage.
J. Mater. Chem. 22, 13374–13379 (2012)
50. Park, N.G., Kang, M.G., Ryu, K.S., Kim, K.M., Chang, S.H.: Photovoltaic characteristics of
dye-sensitized surface-modified nanocrystalline SnO2 solar cells. J. Photochem. Photobiol.
A. 161, 105–110 (2004)
51. Xiong, L., Guo, Y., Wen, J., Liu, H., Yang, G., Qin, P., Fang, G.: Review on the application of
SnO2 in perovskite solar cells. Adv. Funct. Mater. 28, 1802757 (2018)
52. Chava, R.K., Kang, M.: Ag2 S quantum dot sensitized zinc oxide photoanodes for environment
friendly photovoltaic devices. Mater. Lett. 199, 188–191 (2017)
53. Yang, J., Zhong, X.: CdTe based quantum dot sensitized solar cells with efficiency exceeding
7% fabricated from quantum dots prepared in aqueous media. J. Mater. Chem. A. 4, 16553–
16561 (2016)
54. Wang, J., Li, Y., Shen, Q., Izuishi, T., Pan, Z., Zhao, K., Zhong, X.: Mn doped quantum dot
sensitized solar cells with power conversion efficiency exceeding 9%. J. Mater. Chem. A. 4,
877–886 (2016)
55. Yang, J.W., Wang, J., Zhao, K., Izuishi, T., Li, Y., Shen, Q., Zhong, X.H.: CdSeTe/CdS type-I
Core/Shell quantum dot sensitized solar cells with efficiency over 9%. J. Phys. Chem. C. 119,
28800–28808 (2015)
56. Li, L., Yang, X.C., Gao, J.J., Tianand, H.N., Zhao, J.Z.: Highly efficient CdS quantum dot-
sensitized solar cells based on a modified polysulfide electrolyte. J. Am. Chem. Soc. 133,
8458–8460 (2011)
57. Evangelista, R.M., Makuta, S., Yonezu, S., Andrews, J., Tachibana, Y.: Semiconductor
quantum dot sensitized solar cells based on ferricyanide/ferrocyanide redox electrolyte
reaching an open circuit photovoltage of 0.8 V. ACS Appl. Mater. Interfaces. 8, 13957–13965
(2016)
58. Choi, H., Nicolaescu, R., Paek, S., Ko, J., Kamat, P.V.: Supersensitization of CdS quantum
dots with a near-infrared organic dye: toward the design of panchromatic hybrid-sensitized
solar cells. ACS Nano. 5, 9238–9245 (2011)
59. Levy-Clement, C., Tena-Zaera, R., Ryan, M.A., Katty, A., Hodes, G.: CdSe-sensitized p-
CuSCN/nanowire n-ZnO heterojunctions. Adv. Mater. 17, 1512–1515 (2005)
60. Im, S.H., Lim, C.S., Chang, J.A., Lee, Y.H., Maiti, N., Kim, H.J., Nazeeruddin, M.K., Grätzel,
M., Seok, S.I.: Toward interaction of sensitizer and functional moieties in hole-transporting
materials for efficient semiconductor-sensitized solar cells. Nano Lett. 11, 4789–4793 (2011)
61. Park, J., Heo, J., Im, S., Kim, S.: Highly efficient solid-state mesoscopic PbS with embedded
CuS quantum dot-sensitized solar cells. J. Mater. Chem. A. 4, 785–790 (2016)
62. Quan, L., Li, W., Zhu, L., Geng, H., Changand, X., Liu, H.: A new in-situ preparation method
to FeS counter electrode for quantum dots-sensitized solar cells. J. Power Sources. 272, 546–
553 (2014)
63. Faber, M., Park, K., Caban-Acevedo, M., Santra, P., Jin, S.: Earth-abundant cobalt pyrite
(CoS2 ) thin film on glass as a robust, high-performance counter electrode for quantum dot-
sensitized solar cells. J. Phys. Chem. Lett. 4, 1843–1849 (2013)
64. Chen, X., Li, Z., Bai, Y., Sun, Q., Wangand, L., Dou, S.: Room-temperature synthesis of
Cu2−x E (E=S, se) nanotubes with hierarchical architecture as high-performance counter
electrodes of quantum-dot-sensitized solar cells. Chemistry. 21, 1055–1063 (2015)
65. Yang, Y., Zhu, L., Sun, H., Huang, X., Luo, Y., Li, M., Meng, Q.: Composite counter electrode
based on nanoparticulate PbS and carbon black: towards quantum dot-sensitized solar cells
with both high efficiency and stability. ACS Appl. Mater. Interfaces. 4, 6162–6168 (2012)
252 G. S. Selopal

66. Zhang, X., Huang, X., Yang, Y., Wang, S., Gong, Y., Luo, Y., Li, D., Meng, Q.: Investigation
on new CuInS2 /carbon composite counter electrodes for CdS/CdSe cosensitized solar cells.
ACS Appl. Mater. Interfaces. 5, 5954–5960 (2013)
67. Luque, A., Hegedus, S.: Handbook of photovoltaic science and engineering. Wiley (2011)
68. Katoh, R., Furube, A.: Electron injection efficiency in dye-sensitized solar cells. J Photochem
Photobiol C: Photochem Rev. 20, 1–16 (2014)
69. Hagfeldt, A., Gratzel, M.: Light-induced redox reactions in nanocrystalline systems. Chem.
Rev. 95, 49–68 (1995)
70. Benkstein, K.D., Kopidakis, N., Van de Lagemaat, J., Frank, A.J.: Influence of the percolation
network geometry on electron transport in dye-sensitized titanium dioxide solar cells. J. Phys.
Chem. B. 107, 7759–7767 (2003)
71. Hagfeldt, A., Grätzel, M.: Molecular photovoltaics. Acc. Chem. Res. 33, 269–277 (2000)
72. Mora-Sero, I., Gimenez, S., Fabregat-Santiago, F., Gomez, R., Shen, Q., Toyoda, T., Bisquert,
J.: Recombination in quantum dot sensitized solar cells. Acc. Chem. Res. 42, 1848–1857
(2009)
73. Hines, D.A., Forrest, R.P., Corcelli, S.A., Kamat, P.V.: Predicting the rate constant of electron
tunneling reactions at the CdSe–TiO2 interface. J. Phys. Chem. B. 119, 7439–7446 (2015)
74. Zhao, K., Pan, Z.X., Zhong, X.H.: Carbon counter-electrode-based quantum-dot-sensitized
solar cells with certified efficiency exceeding 11%. J. Phys. Chem. Lett. 7, 406–417 (2016)
75. Gentilini, D., D’Ercole, D., Gagliardi, A., Brunetti, A., Reale, A., Brown, T., Di Carlo, A.:
Analysis and simulation of incident photon to current efficiency in dye sensitized solar cells.
Superlattice. Microst. 47, 192–196 (2010)
76. Sanehira, E.M., Marshall, A.R., Christians, J.A., Harvey, S.P., Ciesielski, P.N., Wheeler, L.M.,
Schulz, P., Lin, L.Y., Beard, M.C., Luther, J.M.: Enhanced mobility CsPbI3 quantum dot
arrays for record-efficiency, high-voltage photovoltaic cells. Sci Adv. 3, eaao4204 (2017)
77. Hetsch, F., Xu, X.Q., Wang, H.K., Kershaw, S.V., Rogach, A.L.: Semiconductor nanocrystal
quantum dots as solar cell components and photosensitizers: material, charge transfer, and
separation aspects of some device topologies. J. Phys. Chem. Lett. 2, 1879–1887 (2011)
78. Pan, Z.X., Zhang, H., Cheng, K., Hou, Y.M., Hua, J.L., Zhong, X.H.: Highly efficient inverted
type-I CdS/CdSe Core/Shell structure QD-sensitized solar cells. ACS Nano. 6, 3982–3991
(2012)
79. Peng, X.G., Schlamp, M.C., Kadavanich, A.V., Alivisatos, A.P.: Epitaxial growth of highly
luminescent CdSe/CdS core/shell nanocrystals with photostability and electronic accessibil-
ity. J. Am. Chem. Soc. 119, 7019–7029 (1997)
80. Reiss, P., Protiere, M., Li, L.: Core/Shell semiconductor nanocrystals. Small. 5, 154–168
(2009)
81. Smith, A.M., Nie, S.M.: Semiconductor nanocrystals: structure, properties, and band gap
engineering. Acc. Chem. Res. 43, 190–200 (2010)
82. Chaudhuri, R.G., Paria, S.: Core/shell nanoparticles: classes, properties, synthesis mecha-
nisms, characterization, and applications. Chem. Rev. 112, 2373–2433 (2012)
83. Zhu, H.M., Song, N.H., Lian, T.Q.: Controlling charge separation and recombination rates
in CdSe/ZnS Type I core−shell quantum dots by shell thicknesses. J. Am. Chem. Soc. 133,
8762–8771 (2011)
84. De Geyter, B., Justo, Y., Moreels, I., Lambert, K., Smet, P.F., Van Thourhout, D., Houtepen,
A.J., Grodzinska, D., Donega, C.D., Meijerink, A., Vanmaekelbergh, D., Hens, Z.: The
different nature of band edge absorption and emission in colloidal PbSe/CdSe core/shell
quantum dots. ACS Nano. 5, 58–66 (2011)
85. Dabbousi, B.O., RodriguezViejo, J., Mikulec, F.V., Heine, J.R., Mattoussi, H., Ober, R.,
Jensen, K.F., Bawendi, M.G.: (CdSe)ZnS Core−Shell quantum dots: synthesis and char-
acterization of a size series of highly luminescent nanocrystallites. J. Phys. Chem. B. 101,
9463–9475 (1997)
86. Lambert, K., De Geyter, B., Moreels, I., Hens, Z.: PbTe|CdTe core|shell particles by cation
exchange, a HR-TEM study. Chem. Mater. 21, 778–780 (2009)
Core/Shell Quantum-Dot-Sensitized Solar Cells 253

87. Huang, K., Demadrille, R., Silly, M.G., Sirotti, F., Reiss, P., Renault, O.: Internal structure
of InP/ZnS nanocrystals unraveled by high-resolution soft X-ray photoelectron spectroscopy.
ACS Nano. 4, 4799–4805 (2010)
88. Ivanov, S.A., Piryatinski, A., Nanda, J., Tretiak, S., Zavadil, K.R., Wallace, W.O., Werder, D.,
Klimov, V.I.: Type-II core/shell CdS/ZnSe nanocrystals: synthesis, electronic structures, and
spectroscopic properties. J. Am. Chem. Soc. 129, 11708–11719 (2007)
89. Chen, Y., Vela, J., Htoon, H., Casson, J.L., Werder, D.J., Bussian, D.A., Klimov, V.I.,
Hollingsworth, J.A.: “Giant” multishell CdSe nanocrystal quantum dots with suppressed
blinking. J. Am. Chem. Soc. 130, 5026–5027 (2008)
90. Shariati, M.R., Samadi-Maybodi, A., Colagar, A.H.: Dual co-catalyst loaded reverse type-I
core/shell quantum dots for photocatalytic antibacterial applications. J. Mater. Chem. A. 6,
20433–20443 (2018)
91. Balet, L.P., Ivanov, S.A., Piryatinski, A., Achermann, M., Klimov, V.I.: Inverted core/shell
nanocrystals continuously tunable between type-I and type-II localization regimes. Nano Lett.
4, 1485–1488 (2004)
92. Kim, S., Park, J., Kim, T., Jang, E., Jun, S., Jang, H., Kim, B., Kim, S.W.: Reverse
Type-I ZnSe/InP/ZnS core/shell/shell nanocrystals: cadmium-free quantum dots for visible
luminescence. Small. 7, 70–73 (2010)
93. Jones, M., Kumar, S., Lo, S.S., Scholes, G.D.: Exciton trapping and recombination in Type II
CdSe/CdTe nanorod heterostructures. J. Phys. Chem. C. 112, 5423–5431 (2008)
94. Dorfs, D., Franzl, T., Osovsky, R., Brumer, M., Lifshitz, E., Klar, T.A., Eychmuller, A.: Type-
I and type-II nanoscale heterostructures based on CdTe nanocrystals: a comparative study.
Small. 4, 1148–1152 (2008)
95. Piryatinski, A., Ivanov, S.A., Tretiak, S., Klimov, V.I.: Effect of quantum and dielectric
confinement on the exciton−exciton interaction energy in Type II core/shell semiconductor
nanocrystals. Nano Lett. 7, 108–115 (2007)
96. Tong, X., Zhou, Y.F., Jin, L., Basu, K., Adhikari, R., Selopal, G.S., Tong, X., Zhao, H.G.,
Sun, S.H., Vomiero, A., Wang, Z.M., Rosei, F.: Heavy metal-free, near-infrared colloidal
quantum dots for efficient photoelectrochemical hydrogen generation. Nano Energy. 31, 441–
449 (2017)
97. Peng, Z.A., Peng, X.G.: Formation of high-quality CdTe, CdSe, and CdS nanocrystals using
CdO as precursor. J. Am. Chem. Soc. 123, 183–184 (2001)
98. Hines, M.A., Scholes, G.D.: Colloidal PbS nanocrystals with size-tunable near-infrared
emission: observation of post-synthesis self-narrowing of the particle size distribution. Adv.
Mater. 15, 1844–1849 (2003)
99. Ning, Z., Tian, H., Qin, H., Zhang, Q., Ågren, H., Sun, L., Fu, Y.: Wave-function engineering
of CdSe/CdS core/shell quantum dots for enhanced electron transfer to a TiO2 substrate. J.
Phys. Chem. C. 114, 15184–15189 (2010)
100. Lee, H.J., Bang, J., Park, J., Kim, S., Park, S.-M.: Multilayered semiconductor
(CdS/CdSe/ZnS)-sensitized TiO2 mesoporous solar cells: all prepared by successive ionic
layer adsorption and reaction processes. Chem. Mater. 22, 5636–5643 (2010)
101. Yang, Z., Chen, C.-Y., Liu, C.-W., Chang, H.-T.: Electrocatalytic sulfur electrodes for
CdS/CdSe quantum dot-sensitized solar cells. Chem. Commun. 46, 5485–5487 (2010)
102. Hossain, M.A., Jennings, J.R., Koh, Z.Y., Wang, Q.: Carrier generation and collection in
CdS/CdSe-sensitized SnO2 solar cells exhibiting unprecedented photocurrent densities. ACS
Nano. 5, 3172–3181 (2011)
103. Zhu, G., Pan, L., Xu, T., Sun, Z.: CdS/CdSe-cosensitized TiO2 photoanode for quantum-dot-
sensitized solar cells by a microwave-assisted chemical bath deposition method. ACS Appl.
Mater. Interfaces. 3, 3146–3151 (2011)
104. Maity, P., Debnath, T., Ghosh, H.N.: Ultrafast charge carrier delocalization in CdSe/CdS
Quasi-Type II and CdS/CdSe inverted Type I core–shell: a structural analysis through carrier-
quenching study. J. Phys. Chem. C. 119, 26202–26211 (2015)
105. Kim, S., Fisher, B., Eisler, H.J., Bawendi, M.: Type-II quantum dots: CdTe/CdSe (core/shell)
and CdSe/ZnTe (core/shell) heterostructures. J. Am. Chem. Soc. 125, 11466–11467 (2003)
254 G. S. Selopal

106. Lo, S.S., Mirkovic, T., Chuang, C.H., Burda, C., Scholes, G.D.: Emergent properties resulting
from type-II band alignment in semiconductor nanoheterostructures. Adv. Mater. 23, 180–197
(2011)
107. Scholes, G.D., Jones, M., Kumar, S.: Energetics of photoinduced electron-transfer reactions
decided by quantum confinement. J. Phys. Chem. C. 111, 13777–13785 (2007)
108. Ning, Z., Tian, H., Yuan, C., Fu, Y., Qin, H., Sun, L., Agren, H.: Solar cells sensitized with
type-II ZnSe-CdS core/shell colloidal quantum dots. Chem. Commun. 47, 1536–1538 (2011)
109. Luo, J., Wei, H., Li, F., Huang, Q., Li, D., Luo, Y., Meng, Q.: Microwave assisted aqueous
synthesis of core-shell CdSe(x)Te(1-x)-CdS quantum dots for high performance sensitized
solar cells. Chem. Commun. 50, 3464–3466 (2014)
110. Bang, J., Park, J., Lee, J.H., Won, N., Nam, J., Lim, J., Chang, B.Y., Lee, H.J., Chon, B.,
Shin, J., Park, J.B., Choi, J.H., Cho, K., Park, S.M., Joo, T., Kim, S.: ZnTe/ZnSe (core/shell)
Type-II quantum dots: their optical and photovoltaic properties. Chem. Mater. 22, 233–240
(2010)
111. Jiao, S., Shen, Q., Mora-Seró, I., Wang, J., Pan, Z., Zhao, K., Kuga, Y., Zhong, X., Bisquert,
J.: Band engineering in core/shell ZnTe/CdSe for photovoltage and efficiency enhancement
in exciplex quantum dot sensitized solar cells. ACS Nano. 9, 908–915 (2015)
112. Selopal, G.S., Zhao, H.G., Tong, X., Benetti, D., Navarro-Pardo, F., Zhou, Y.F., Barba, D.,
Vidal, F., Wang, Z.M., Rosei, F.: Highly stable colloidal “giant” quantum dots sensitized solar
cells. Adv. Funct. Mater. 27, 1701468 (2017)
113. Dana, J., Maiti, S., Tripathi, V.S., Ghosh, H.N.: Direct correlation of excitonics with efficiency
in a core–shell quantum dot solar cell. Chemistry. 24, 2418–2425 (2018)
114. Adhikari, R., Jin, L., Navarro-Pardo, F., Benetti, D., AlOtaibi, B., Vanka, S., Zhao, H.G.,
Mi, Z.T., Vomiero, A., Rosei, F.: High efficiency, Pt-free photoelectrochemical cells for solar
hydrogen generation based on “giant” quantum dots. Nano Energy. 27, 265–274 (2016)
115. Brovelli, S., Schaller, R.D., Crooker, S.A., Garcia-Santamaria, F., Chen, Y., Viswanatha, R.,
Hollingsworth, J.A., Htoon, H., Klimov, V.I.: Nano-engineered electron–hole exchange inter-
action controls exciton dynamics in core–shell semiconductor nanocrystals. Nat. Commun. 2,
280 (2011)
116. Abdi, F.F., Van de Krol, R.: Nature and light dependence of bulk recombination in Co-Pi-
Catalyzed BiVO4 photoanodes. J. Phys. Chem. C. 116, 9398–9404 (2012)
117. Zhao, H., Fan, Z., Liang, H., Selopal, G.S., Gonfa, B.A., Jin, L., Soudi, A., Cui, D., Enrichi,
F., Natile, M.M., Concina, I., Ma, D., Govorov, A.O., Rosei, F., Vomiero, A.: Controlling
photoinduced electron transfer from PbS@CdS core@shell quantum dots to metal oxide
nanostructured thin films. Nanoscale. 6, 7004–7011 (2014)
118. Pal, B.N., Ghosh, Y., Brovelli, S., Laocharoensuk, R., Klimov, V.I., Hollingsworth, J.A.,
Htoon, H.: ‘Giant’ CdSe/CdS core/shell nanocrystal quantum dots as efficient electrolumi-
nescent materials: strong influence of shell thickness on light-emitting diode performance.
Nano Lett. 12, 331 (2012)
119. Navarro-Pardo, F., Zhao, H., Wang, Z.M., Rosei, F.: Structure/property relations in “giant”
semiconductor nanocrystals: opportunities in photonics and electronics. Acc. Chem. Res. 51,
609–618 (2017)
120. Smith, A.M., Mohs, A.M., Nie, S.: Tuning the optical and electronic properties of colloidal
nanocrystals by lattice strain. Nat. Nanotechnol. 4, 56–63 (2009)
121. Pinchetti, V., Meinardi, F., Camellini, A., Sirigu, G., Christodoulou, S., Bae, W.K., De Donato,
F., Manna, L., Zavelani-Rossi, M., Moreels, I., Klimov, V.I., Brovelli, S.: Effect of core/shell
interface on carrier dynamics and optical gain properties of dual-color emitting CdSe/CdS
nanocrystals. ACS Nano. 10, 6877–6887 (2016)
122. Jiao, S., Wang, J., Shen, Q., Li, Y., Zhong, Z.H.: Surface engineering of PbS quantum dot
sensitized solar cells with a conversion efficiency exceeding 7%. J. Mater. Chem. A. 4, 7214–
7221 (2016)
Core/Shell Quantum-Dot-Sensitized Solar Cells 255

123. Jin, L., Sirigu, G., Tong, X., Camellini, A., Parisini, A., Nicotra, G., Spinella, C., Zhao, H.G.,
Sun, S.H., Morandi, V., Zavelani-Rossi, M., Rosei, F., Vomiero, A.: Engineering interfacial
structure in “giant” PbS/CdS quantum dots for photoelectrochemical solar energy conversion.
Nano Energy. 30, 531–541 (2016)
124. Adhikari, R., Basu, K., Zhou, Y.F., Vetrone, F., Ma, D.L., Sun, S.H., Vidal, F., Zhao, H.G.,
Rosei, F.: Heterostructured quantum dot architectures for efficient and stable photoelectro-
chemical hydrogen production. J. Mater. Chem. A. 6, 6822–6829 (2018)
125. Bae, W.K., Padilha, L.A., Park, Y.S., McDaniel, H., Robel, I., Pietryga, J.M., Klimov, V.I.:
Controlled alloying of the core–shell interface in CdSe/CdS quantum dots for suppression of
Auger recombination. ACS Nano. 7, 3411–3419 (2013)
126. Selopal, G.S., Zhao, H.G., Liu, G.J., Zhang, H., Tong, X., Wang, K.H., Tang, J., Sun, X.H.,
Sun, S.H., Vidal, F., Wang, Y.Q., Wang, Z.M., Rosei, F.: Interfacial engineering in colloidal
“giant” quantum dots for high-performance photovoltaics. Nano Energy. 55, 377–388 (2019)
127. Zhao, H.G., Liu, G.J., Vidal, F., Wang, Y.Q., Vomiero, A.: Colloidal thick-shell pyramidal
quantum dots for efficient hydrogen production. Nano Energy. 53, 116–124 (2018)
128. Zhao, H.G., Liu, J.B., Vidal, F., Vomiero, A., Rosei, F.: Tailoring the interfacial structure of
colloidal “giant” quantum dots for optoelectronic applications. Nanoscale. 10, 17189–17197
(2018)
129. Zhou, R., Wan, L., Niu, H.H., Yang, L., Mao, X.L., Zhang, Q.F., Miao, S.D., Xu, J.Z.,
Cao, G.Z.: Tailoring band structure of ternary CdSx Se1-x quantum dots for highly efficient
sensitized solar cells. Sol. Energy Mater. Sol. Cells. 155, 20–29 (2016)
Core/Shell Quantum-Dot-Based
Solar-Driven Photoelectrochemical Cells

Ali Imran Channa, Xin Li, Xin Tong, and Zhiming M. Wang

Abstract Colloidal quantum dots (QDs) are promising for a variety of optoelec-
tronic applications due to their size/shape/composition optical properties. Among
various QDs, the core/shell QDs have demonstrated improved optical properties
and enhanced photo-/chemical stability with respect to the bare QDs. Specifically,
the band structure of the core/shell QDs can be tailored by choosing the appropriate
core and shell material, changing the chemical composition, or varying the shell
thickness. In this chapter, we review the recent advances of the synthetic approaches,
optical properties engineering, and charge dynamics of core/shell QDs with type
I, type II, and quasi-type II band structure and highlight their application in
solar-driven photoelectrochemical (PEC) cells. The device performance of these
core/shell QD-based PEC cells is discussed in detail and their prospective devel-
opments toward future device optimizations and commercialization are proposed as
well.

Keywords Colloidal quantum dots · Core/shell architecture · Band structure


engineering · Photoelectrochemical cell · Hydrogen generation

1 Introduction

The energy consumption of the modern world has been increasing steadily over
the past few decades [1, 2]. The use of fossil fuels may not be able to meet
the required energy needs in an environmentally sustainable way for decades to

Author Contribution Authors Ali Imran Channa and Xin Li have been equally contributed to this
chapter.

A. I. Channa · X. Li · X. Tong () · Z. M. Wang ()


Institute of Fundamental and Frontier Sciences, University of Electronic Science and Technology
of China, Chengdu, People’s Republic of China
e-mail: xin.tong@uestc.edu.cn; zhmwang@uestc.edu.cn

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 257
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_8
258 A. I. Channa et al.

come. The solar energy incident on earth in 1 h is more than the total energy
consumed by the world in 1 year, suggesting that the solar energy conversion is
a promising approach to address the future energy needs [2–4]. Hydrogen is a
promising high-density fuel with no toxic effects on the environment [5, 6]. During
the combustion of hydrogen, water is the only by-product that is generated. Solar-
driven hydrogen generation via water splitting has become one of the hot topics after
the pioneering work by Fujishima and Honda [7]. Fujishima et al. have reported that
a semiconductor material (i.e., TiO2 ) could be used for the production of hydrogen
via photoelectrochemical (PEC) water splitting [7]. TiO2 possesses numerous merits
such as considerable photo-/chemical stability, inertness toward biological/chemical
corrosion, and nontoxicity and cost-effectiveness [8]. However, due to the wide
bandgap of TiO2 (~3.2 eV), the solar energy conversion was limited to the ultraviolet
(UV) region of the solar spectrum [7]. To establish a sustainable and efficient solar
energy conversion system, semiconductor materials with a narrow bandgap and
broad light absorption spectrum are required [9, 10]. Moreover, efficient charge
carrier separation/transfer and long-term stability are vital parameters to achieve
high-efficiency solar energy conversion [11, 12].
Semiconductor colloidal quantum dots (QDs) are nanocrystals with size typically
smaller than ~20 nm with numerous outstanding features such as size-dependent
optical bandgap due to quantum confinement effect [13], high absorption coef-
ficients [14, 15], cost-effective synthesis [16], facile and large-scale production,
etc. [17, 18]. To date, colloidal QDs have been applied for various applications
such as QD-sensitized solar cells (QDSCs) [19], light-emitting diodes (LEDs) [20],
photodetectors [21], luminescent solar concentrators (LSCs) [22], and PEC cells
for hydrogen generation [23]. Among the various types of QDs, the core/shell QDs
have been demonstrated as promising building blocks in applications due to their
favorable physical/chemical and optical properties [24]. In general, the surface of
the bare QD is very sensitive to the ambient environment (light, moisture, oxygen,
temperature, etc.); such sensitivity can induce the dangling/broken bonds, leading
to the degradation of QDs and affecting its chemical stability as well as the optical
properties [25, 26]. The organic ligands on the surface of bare QDs can offer a good
degree of surface passivation. However, the stability of organic ligand passivated
QDs is also dependent on the ligand stability itself, which cannot guarantee long-
term stability [27]. During various QD-based device fabrication processes, the
organic ligands can be damaged; hence, the ligands offer inefficient passivation of
both anionic and cationic surface sites. Moreover, the surface-related defects/traps
can result in non-radiative charge recombination for poor charge separation/transfer
capability. This effect is detrimental for device applications which require better
charge separation and transfer characteristics as well as for the long-term stability
of QDs [28].
An elegant approach to solve this problem is to overcoat the bare QD’s surface
with a second inorganic semiconductor material, to form core/shell structured QDs,
wherein the bare QDs are effectively passivated through an inorganic shell. Several
types of colloidal core/shell QDs have been developed with different morphologies
including spherical core/shell, dot-in-rod, and rod/rod core/shell QDs [29]. In
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 259

Fig. 1 Schematic diagrams of type I, type II, and quasi-type II band structures. (Adapted from
Ref. [31])

contrast to the bare QDs, core/shell QDs allow further tuning of the physical
properties, such as the photoluminescence quantum yield (PLQY) and excitonic
lifetime, which are considered as important parameters for the application of QDs
in optoelectronic devices. The band alignment of the two materials (i.e., core and
shell) can be tuned and different carrier-localization regions can be obtained which
affects the QD’s optoelectronic properties [30]. Hence, different kinds of band-
engineered core/shell QDs can be targeted for diverse device applications. The
typical band structures of core/shell QDs with type I, type II, and quasi-type II band
alignment (as shown in Fig. 1) can be obtained via tuning the core size, chemical
composition, or thickness of the shell. The electronic band structure of bare QDs
is simple and the excitations or de-excitations of the charge carriers occur within
the bandgap. However, the growth of the shell on the bare QDs alters the band
structure and creates new pathways for the de-excitation process of charge carriers.
For instance, in type I core/shell QDs, a wide bandgap semiconductor material is
coated on the core of the narrow-bandgap material, in which the conduction band
(CB) and valence band (VB) of the core reside within the bandgap of the shell
material. Therefore, the excitation and de-excitation of charge carriers occur within
the bandgap of the core [31]. In the type II band structure, the CB and the VB of the
shell stagger with that of the core, allowing the spatial separation of electrons and
holes, with holes residing in the core and electrons into the shell [32]. This type of
core/shell structure often exhibits long-lived lifetime of excitons [33]. Quasi-type II
band alignment is a special case in which either the CB or the VB of the shell and
core materials possess a small band offset. This band structure allows the existence
of one type of charge carrier in the entire core/shell structure and another type of
carrier confined in the core [34, 35]. Core/shell QDs with both type II and quasi-type
II band structures have demonstrated efficient spatial separation of photoinduced
electrons and holes, leading to efficient charge separation and transfer in QD-based
PEC systems, which are beneficial to boost the efficiency of QD-based solar-driven
PEC cells for hydrogen generation [36, 37].
260 A. I. Channa et al.

1.1 Conversion of Solar Energy into Hydrogen

The overall world’s energy resources include fossil fuel, nuclear fuel, and the
renewable ones. Among these sources, the renewable energy resources are available
throughout the years. However, nonrenewable energy resources are predicted to
be eventually depleted. In addition, the immense use of nonrenewable sources
has polluted the environment by producing huge quantity of toxic gases in the
atmosphere (e.g., COx , NOx , SOx , Cx Hy ) and radioactive pollution [38]. Therefore,
the hazardous effects of nonrenewable energy sources as well as their risk of being
depleted have led to the exploitation of various renewable energy sources including
wind, water, sun, and biomass. From these renewable energy sources, chemical
fuel such as hydrogen gas can be produced by thermal, electrical, photonic, and
biochemical energy.
Hydrogen can be generated via thermal energy, playing a main role in steam
reforming of fossil fuels, which is not suitable due to the utilization of fossil fuels.
Hydrogen generation via water splitting is not a spontaneous process and requires an
external energy source to establish a sustainable water splitting reaction. Hydrogen
can also be obtained by using different methods with solar energy as an external
driving force to establish a sustainable water splitting reaction, as demonstrated
in Fig. 2. Thermolysis is the direct splitting of water at high temperatures for
hydrogen production, though it offers a drawback due to the possibility of a rapid
back reaction of H2 and O2 at higher temperatures, thus preventing the thermolysis
method from being a viable approach. Actually, this method was initially developed
to utilize the waste heat from nuclear reactors [39]. Biomass conversion into
hydrogen via solar energy is not considered as a reasonable approach due to its low
conversion efficiency. It implies that in order to achieve higher efficiency to generate
substantial amounts of hydrogen, the system must be designed in large inappropriate
dimensions. However, if the biomass from the waste by-product is used for this
purpose, then this could perhaps be the least expensive technique for hydrogen
generation. Wind energy and photovoltaic (PV) systems connected to electrolyzers
have shown to be adaptable and reliable approaches for hydrogen production in the
near future. An electrolyzer is a type of device that splits water into H2 and O2 by
electricity. Currently, these systems are being used in industries to produce high-
purity hydrogen and are very expensive. Researches are being devoted to develop
small and inexpensive electrolyzer systems for individual use. In contrast to the wind
and PV systems coupled to the electrolyzers, a photolysis system consists of a single
unit combining the two separate steps of electrical generation and electrolysis into a
single step. The photolysis systems include photoelectrolysis and photobiologically
based systems and depend purely on solar energy to split water [40].
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 261

Fig. 2 Methods for sustainable hydrogen generation

1.2 Working Principle of Solar-Driven PEC Cell for Hydrogen


Generation

PEC hydrogen generation via photoelectrolysis of water is a promising technique to


achieve considerable hydrogen production in an environmentally sustainable man-
ner [41, 42]. This technique exhibits promising features such as simple architecture,
cost-effectiveness, and easy large-scale implementation. Figure 3 shows a schematic
diagram of a typical PEC cell, which exhibits the key components including a
light-sensitive semiconductor as a working electrode, a counter electrode, and
an electrolyte. When the PEC cell is exposed to light, electron–hole pairs are
generated in the semiconductors upon the absorption of light. The photoexcited
electrons from the CB of the semiconductor migrate to the counter electrode
under an external applied electric field. The electrons then reduce water at the
counter electrode/electrolyte interface to produce hydrogen. The holes however
oxidize water at the (photoanode) working electrode/electrolyte interface for oxygen
evolution [43]. In case if the working electrode is a photocathode, then the process
will be reversed, i.e., hydrogen will be generated at the photocathode side, whereas
the oxygen evolution reaction will occur at the counter electrode side [44]. The
262 A. I. Channa et al.

Fig. 3 Scheme of a typical


PEC cell for solar-driven
hydrogen evolution

following reactions explain the complete chemical processes [45].

2γ → 2e− + 2h+ Photon − induced electron − hole pair generation

1
Working electrode/electrolyte interface : H2 O + 2h+ → O2 (gas) + 2H+
2
(1)

Counter electrode/electrolyte interface : 2H+ + 2e− → H2 (gas) (2)

1
Overall reaction : H2 O + 2γ → O2 (gas) + H2 (gas) (3)
2

G = +237.18 kJ mol−1 Standard Gibbs free energy (4)

G
Vrev = = 1.23 V Standard reversible potential (5)
nF

Vop = Vrev + ηa + ηc + η" + ηsys Operating voltage with overpotential losses


(6)

Here γ represents the photon energy, e− is electron, h+ is a hole, and G is


the standard Gibbs free energy. Gibbs free energy is the minimum energy required
for the water decomposition reaction to occur under a standard temperature of
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 263

Fig. 4 Schematic diagram of


a representative modern
three-electrode PEC cell

25 ◦ C and a pressure of 1 bar. The standard Gibbs free energy change required
for water splitting reaction is +237.18 kJ mol−1 . The positive sign of Gibbs free
energy (G) deduces that the energy must be provided to the system for the
electrolysis of water. Vrev in Eq. 5 is the standard reversible potential, i.e., the
minimum electrical potential required to carry out the reversible photoelectrolysis.
The minimum reversible potential for initiating a reversible photoelectrolysis is
given as 1.23 eV. However, this potential does not include the overpotential losses;
therefore, the water splitting reaction cannot occur at this potential. Vop represented
in Eq. 6 is the operational voltage and ηa , ηc , η" , and ηsys are the overpotentials
related to the anode, cathode, and ionic conductivity of the electrolyte and other
system losses, respectively [45]. “n” denotes the total number of electrons involved
in the reaction for water photoelectrolysis. “F” is the Faraday constant with a value
of 96484.34 C/mol.
During the operation of a PEC cell, it is necessary to maintain a fixed or
controlled potential on the electrodes. However, it is difficult for electrodes to
maintain a constant potential while passing current simultaneously. To resolve this
issue, the passage of current and maintaining a fixed potential are separated by the
introduction of an additional electrode called reference electrode [46]. The reference
electrode with a known reduction potential serves as a half cell and provides a
reference for measuring and controlling the potential of the working electrode with
no current passing through it. However, the counter electrode passes all the current
needed to balance the current observed at the working electrode [46]. Hence, the
modern voltammetry system consists of three electrodes, the working electrode,
reference electrode, and counter electrode, as shown in Fig. 4.
264 A. I. Channa et al.

2 Synthesis and Optical Properties of Core/Shell QDs

2.1 Synthesis of Core/Shell QDs

Generally, the bare QDs have large surface-to-volume ratio due to their very small
size, typically around or less than ~20 nm [47], rendering the surface a very
important factor for determining the optical properties of the QDs. The organic
ligands attached on the surface of the QDs can stabilize them by passivating the
dangling bonds [47]. However, there exists a weak bonding between the organic
ligands and the QDs’ surface, which can be affected by various parameters such
as light exposure, temperature, and humidity, thus consequently affecting the
properties which are highly dependent on the surface conditions including the
optical properties. The surface deformations can generate surface-related defects
that act as the non-radiative recombination centers, henceforth reducing both the
PLQY and photostability of the bare QDs [16]. The bare QDs, therefore, face
many challenges for their practical applications. One of the solutions to address this
problem is the formation of core/shell structured QDs. In such type of QDs, the bare
QDs are coated by another inorganic shell, and such efficient surface passivation
by the inorganic shell can lead to improved PLQY and photostability compared
to the bare QDs [47–49]. Moreover, the formation of the shell on bare QDs not
only improves the optical properties but also creates a new band structure, which
can be tuned by the choice of composition and thickness of the shell materials. In
this case, reduced non-radiative recombination and efficient spatial separation of the
electron–hole pairs can be achieved, which is the important parameter for achieving
high-efficiency PV devices [50]. Core/shell QDs are mainly categorized as type I,
type II, and quasi-type II depending upon the relative band alignment of the CB and
VB of the shell material with respect to the core QD, as discussed in the Introduction
[49].
The growth of the core/shell QDs can be achieved by injecting the cationic and
anionic solutions of shell precursors together or separately in parts into the core
QD solution [16, 51–53]. Jo et al. have reported the growth of CuGaS2 (CGS)/ZnS
core/shell QDs by using the injection of preprepared cationic–anionic solutions
for the ZnS shell [54]. The CGS core QDs were first synthesized at 180 ◦ C for
5 min, and then two different kinds of ZnS shell solutions prepared beforehand
were injected. The first ZnS shell solution was prepared by dissolving 4 mmol of Zn
(OAc)2 in 4 mL of oleic acid (OA), 2 mL of 1-dodecanethiol (DDT), and 2 mL of
1-octadecene (ODE) at 190 ◦ C. The second ZnS shell solution was synthesized by
dissolving 8 mmol of Zn (St)2 in 4 mL of DDT and 8 mL of ODE. The first stock
solution was injected into the pre-synthesized CGS core QDs at 220 ◦ C; after an
interval of 30 min, the second stock solution was then injected at 250 ◦ C; and the
reaction was held for 60 min [54]. The obtained CGS/ZnS core/shell QDs exhibited
white luminescence with PLQY of about ~75%. A similar methodology was also
employed for the growth of CGS/ZnS by Jalalah et al. with slight modifications in
the chemical composition and reaction parameters [55].
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 265

Successive ion layer adsorption and reaction (SILAR) is a well-known synthetic


technique to prepare the core/shell QDs [56]. In this technique, the cationic and
anionic solutions of shell precursors are alternatively injected into the core QD
solution with a suitable time interval in between [56]. For example, Xie et al. first
synthesized the Cu-doped InP core QDs at 210 ◦ C [57]. In order to grow the ZnSe
shell, they used 0.1 M Zn stearate and 0.1 M trioctylphosphine (TOP)-Se solutions.
Both these solutions were injected into the InP core solution at 150 ◦ C at an interval
of 10 min followed by the increase of the temperature to 220 ◦ C for the growth
of the ZnSe shell [57]. Similarly, they injected the second round of cationic and
anionic precursor solutions to increase the ZnSe shell thickness at 150 ◦ C and
then increased the temperature to 220 ◦ C for the growth of the second shell layer
[57]. Tessier et al. also synthesized InP/ZnE (E = S, Se) core/shell QDs; they first
synthesized InP core QDs at 180 ◦ C [58] and then used Zn stearate dissolved in ODE
and TOP-(Se, S) as the shell precursor solutions for the reaction. The TOP-(Se, S)
solution was injected dropwise at the same temperature and left for 60 min under
continuous stirring. After 60 min, the temperature was increased to 200 ◦ C and
maintained for 60 min, followed by injection of the Zn stearate solution at 200 ◦ C
and increasing the temperature to 220 ◦ C and maintained for 30 min [58]. In this
way, they consequently obtained a thick layer of ZnSe(S) shell on the InP core QDs.
The shell on the core can be grown by a two-step method as well. In this method,
the core QDs are grown first and purified to get rid of the unreacted precursor and
then redispersed in an organic solvent such as ODE and degassed to evacuate the
humidity and air. The reaction mixture is then heated to the optimum shell growth
temperature whereupon the shell precursor solution is injected dropwise. Such
method is reported by Dabbousi et al.; they first synthesized the CdSe core QDs
through hot-injection method [51]. The purified CdSe QDs were then redispersed in
a mixed solvent of ODE and octadecylamine (ODA) and heated to the shell growth
temperature (240 ◦ C) after the conventional degassing process and N2 purging. The
Zn and S precursor mixture solutions with identical molarity were slowly injected
into the core QD solution [51]. The main advantage of this overgrowth technique is
that the shell thickness can be precisely controlled. The shell thickness of 1.5 nm
was obtained in the above-reported CdSe/ZnS core/shell QDs [51]. Using this
technique, numerous core/shell QDs with well-controlled shell thickness have been
grown such as CdTe/CdSe, CdSe/ZnTe, CdSe/Zn1-x Cdx S, PbSe/PbS, and InP/ZnS
[52, 53, 59, 60]. Specifically, the as-synthesized core/shell QDs with very thick
shell, typically known as the “giant” core/shell QDs, possess outstanding properties
such as higher chemical and photostability as compared to both bare QDs and thin-
shell QDs [56, 61, 62]. The band structure in such “giant” core/shell QDs can be
suitably tailored by controlling the shell thickness and composition to obtain long-
lived lifetime of photoexcited charge carriers as compared to the pure and thin-shell
QDs [56, 61, 62]. Moreover, the existence of type II or quasi-type II band alignment
in these “giant” core/shell QDs makes them excellent candidates for PV applications
[63, 64].
Cation exchange method is another method to grow a very thin shell on the
core to passivate the surface defects [65]. In this method, the overall size of the
266 A. I. Channa et al.

Fig. 5 HRTEM images of PbTe/CdTe core/shell QDs along (a) (111), (b) (100), and (c) (211)
directions. (Adapted from Ref. [67])

core/shell QDs remains almost the same as the initial size of the core QDs [66–
68]. The reason behind this is the gradual replacement of cationic constituents of
the core material by cations of the shell material. Pietryga et al. demonstrated the
formation of PbSe/CdSe core–shell QDs via the cation exchange method [69]. Bare
PbX (X = S, Se, Te) QDs were synthesized and purified to get rid of the unreacted
precursor and then redispersed in toluene. The surface treatment of the PbX QDs
was performed at 100 ◦ C by injecting Cd oleate solution which was separately
prepared beforehand at 155 ◦ C under a N2 environment. Then the Cd oleate was
injected into PbX QDs at 100 ◦ C [69]. Lambert et al. have also reported the synthesis
of PbTe/CdTe core/shell QDs via the cation exchange method; the representative
high-resolution transmission electron microscope (HRTEM) images of PbTe/CdTe
core/shell QDs are displayed in Fig. 5, wherein the clear different lattice fringes
from the core and the shell, which were observed along (100) and (211) orientations
of PbTe/CdTe core/shell QDs, indicate the formation of the core/shell structure [67].
Growth of a very thin shell of ZnS (~0.1 nm) on CuInSex S2 − x via cation exchange
has also been reported by Tong et al. [66]. After the synthesis, CuInSex S2 − x QDs
were purified and redispersed in ODE. A 0.25 M Zn oleate solution was used for the
surface treatment of CuInSex S2 − x QDs in ODE at 100 ◦ C for 10 min to obtain the
CuInSex S2 − x /ZnS core/shell QDs [66].

2.2 Optical Properties of Core/Shell QDs


2.2.1 Type I Core/Shell QDs

Numerous type I core/shell QDs, including CdSe/ZnS [51, 70], CdS/ZnS [71], and
InP/ZnS [72] core/shell/shell QDs [73, 74], have been reported. In type I core/shell
QDs, both the electrons and the holes are well confined within the core QDs due
to the wider bandgap of the shell material. Therefore, the optical properties of
such core/shell structures are not very sensitive to the surrounding environments
and are maintained even after multiple purifications, film deposition, or ligand
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 267

Fig. 6 (a) Absorption and (b) emission spectrum of type I CdSe/ZnS core/shell QDs. (Adapted
from Ref. [76])

exchange [72, 73, 75]. Figure 6a, b presents the absorption and emission spectra
of type I CdSe/ZnS core/shell QDs with variable ZnS shell thickness [76]. With
the increasing ZnS shell thickness, a slight redshift in both the absorption (Fig. 6a)
and photoluminescence (PL) spectrum (Fig. 6b) of CdSe/ZnS core/shell QDs can
be observed, which was attributed to the extension of the electron and hole wave
functions into the shell material [51, 76].
The ZnS coating on CdSe suppressed the defects in the CdSe core and enhanced
the PLQY from 10% to over 40% in CdSe/ZnS (1.7 MLs) (inset in Fig. 6b). Another
study based on CdSe/ZnS core/shell QDs with type I band structure revealed that
the formation of type I band structure maintains the original optical properties of
the bare QDs, including the absorption onset and the PL peak position, while the
core/shell structure exhibited enhanced PLQY compared to the bare QDs [70].
Figure 7 shows the absorption and emission spectrum of bare CdSe and CdSe/ZnS
core/shell QDs, in which the CdSe/ZnS core/shell QD showed the absorption and PL
features at the same peak positions as the bare CdSe QDs, but with much enhanced
peak intensities.

2.2.2 Type II Core/Shell QDs

In type II core/shell QDs, the lowest-energy CB electrons and VB holes are localized
in different regions within the QD (i.e., one type of charge carrier resides in the
core, whereas another in the shell). A new band for the interfacial charge transfer
(CT) between the two materials is formed in addition to the electronic transitions
of individual materials [77]. This CT band is situated at the lower band energy
than the individual bandgaps of the core and shell materials, extending both the
absorption onset and the emission spectrum of the type II core/shell QDs to longer
wavelength as compared to core QDs. Figure 8a shows the UV–vis absorption and
emission spectra of CdTe core QDs and CdTe/CdSe core/shell QDs in heptane.
CdTe core QDs exhibited characteristic first (1 s) exciton peak at around ∼560 nm
of wavelength. However, CdTe/CdSe core/shell QDs exhibited a new emission at
∼650 nm with diminished previous 1 s exciton peak. As compared to bare CdTe
268 A. I. Channa et al.

Fig. 7 (a) Absorption spectrum of the CdSe QDs (dotted line) and the CdSe/ZnS core/shell QDs
(solid line), PL spectrum of the CdSe/ZnS core/shell (solid line). (b) PL spectrum of CdSe QDs
(dotted line) and CdSe/ZnS core/shell QDs (solid line). (Adapted from Ref. [70])

core QDs, the CdTe/CdSe core/shell QDs also exhibited broad absorption spectrum
extended to near infrared (NIR). In addition, the emission spectrum of CdTe/CdSe
core/shell QDs was observed to be considerably red-shifted (∼800 nm) as compared
to the CdTe core QDs (∼580 nm) [77]. It is also demonstrated that tunable optical
properties can be obtained by varying the core size and shell thickness [52, 78].
Typically, in type II core/shell structures, the absorption spectrum exhibits the
transition of charge carriers between both materials (core and shell). As shown in
Fig. 8a, four absorption bands of B1, B2, B3, and C can be observed at 770, 650,
500, and 560 nm, respectively, for CdTe/CdSe core/shell QDs. These bands are more
obvious in the TA spectrum (blue spectrum in Fig. 8a) at the initial delay times (such
as 1 ps) exhibiting bleaches at these bands. These bleaches can be assigned to the
filling of CB electron levels [77]. For instance, the absorption band B1 centered at
∼770 nm (near the emission peak position) can be assigned to the CT transition
between the CdTe (core) VB edge (1sh ) and CdSe (shell) CB edge (1se ) [77].
The TA spectra speculate B1, B2, and B3 transitions to follow exactly the same
bleach formation strategy and decay kinetics, which suggest that these transitions
are sharing identical electron level. The C band corresponds to transition between a
core-localized VB 1 s hole level and a delocalized CB electron level.
Type II band alignment can control the excited state dynamics of photoexcited
electrons and holes. After the excitation, the electron–hole pairs undergo charge
separation and localize in different regions such as electrons in the shell and holes
in the core [36]. Figure 8c, d shows the TA spectra of CdTe/CdSe core/shell QDs,
exhibiting bleaches at B1, B2, and B3 with a time constant of 0.7 ps. The formation
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 269

Fig. 8 (a) UV–vis absorption and emission spectra of CdTe core and CdTe/CdSe core/shell QDs.
(b) Band structure diagram of CdTe/CdSe core/shell QDs illustrating the transitions from different
energy levels. (c) Transient absorption spectra of CdTe/CdSe core/shell QDs at indicated delay
time windows after 400-nm excitation. (d) Transient decay kinetics at indicated transitions in (c).
(Adapted from Ref. [77])

of these bleaches refer to the separation of excited electrons at the shell, which
is consistent with other similar core/shell structures with type II band alignment
[79–82]. The PL lifetime of QDs with a type II core/shell structure has also been
observed to be lengthened with respect to the bare core QDs. The extended lifetimes
of hundreds of ps, nanoseconds, or even in the scale of microseconds have been
reported [36, 83, 84]. A recent study on CdTe/CdSe core/shell QDs has revealed that
the exciton lifetime depends on the volume and electron–hole overlap integral [84].
Therefore, in type II core/shell structures, due to the reduction in the electron–hole
overlap region and increased volume, the lifetime can be significantly prolonged
[84].

2.2.3 Quasi-Type II Core/Shell QDs

In a typical quasi-type II band structure, a small CB and large VB offset exists


between the core and the shell materials. The holes with the lowest energy are
localized in the core, whereas the electrons in both core and shell [63, 83, 85].
270 A. I. Channa et al.

Fig. 9 (a) Schematic energy level diagram and lowest energy electron and hole wave functions in
CdSe/CdS (core/shell) quasi-type II QDs. (b) Static absorption and emission spectra with transient
absorption spectrum (at 1 ns, green line) spectra of CdSe/CdS core/shell QDs. (c) Transient
absorption spectra and (d) bleach formation kinetics of T0 (575 nm) and T1 (475 nm) bands of
CdSe/CdS QDs at indicated delay times (0–3 ps) after 400 nm excitation. (Adapted from Ref.
[90])

Examples of quasi-type II core/shell QDs include CdSe/CdS [86], CdSe/ZnSe [87],


PbSe/CdSe [88], Zn- CuInSe2 /CuInS2 [89], CuInSex S2 − x /CdSeS/CdS [37], etc. It
has been observed that due to the existence of a small CB offset, no noticeable
interfacial CT band is formed in case of quasi-type II core/shell structures. The
theoretical calculation studies for CdSe/CdS core/shell QDs by Zhu et al. revealed
the existence of quasi-type II band alignment [90]. Figure 9a displays the schematic
energy level diagram of quasi-type II band alignment for CdSe/CdS core/shell QDs,
in which 1 s electrons are delocalized throughout the whole particle (core and shell),
whereas 1 s holes are strongly confined within the core [90].
Figure 9b shows the absorption spectrum obtained for CdSe/CdS core/shell
QDs with weak absorption band at 575 nm (denoted as T0 ) and strong absorption
band with onset around ∼475 nm (denoted as T1 ). These bands are further clearly
observed in the TA spectrum (Fig. 9c). The T0 band being closer to the emission
spectrum can be ascribed to the transition between the lowest energy CB electron
(1se ) and VB hole (1sh ) levels in the core/shell structure (1se -1sh ). In Fig. 9b
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 271

(TA spectrum at 1 ns), the bleaches at both the T1 and T0 bands suggest these
transitions to involve the 1se level. The T1 band therefore can be assigned to the
transition between the delocalized 1se level and the lowest energy hole level above
the VB band edge of the CdS shell [90]. This can be further supported by identical
relaxation kinetics of T0 and T1 bleaches, as illustrated in Fig. 9d [90]. In summary,
similar to type II core/shell QDs, the quasi-type II core/shell QDs can also exhibit
prolonged lifetime due to the reduction of the electron–hole overlap integral [91,
92].

2.3 Charge Dynamics of Core/Shell QDs


2.3.1 Type I Core/Shell QDs

The charge transfer from different core/shell QDs coupled with semiconductor films
or molecular acceptors has been widely studied [93–96]. An investigation of type I
CdSe/ZnS core/shell QDs and CdSe-sensitized QDSCs has revealed almost similar
power conversion efficiencies (PCE) from both CdSe core and CdSe/ZnS core/shell
QDs [97], while some studies have also demonstrated the enhanced PCE of the
QDSCs via coating ZnS overlayers on the QD-sensitized photoanode surface [98–
100]. This suggests that the ZnS surface coating plays a crucial role in achieving
efficient charge separation/transfer in QDSCs. The CdSe/ZnS core/shell QD–
anthraquinone (AQ) complex was studied as a model type I QD–electron acceptor
system by Lian et al. [76]. In this work, the charge separation (kCS ) and charge
recombination (kCR ) kinetics as a function of ZnS shell thickness were measured by
TA spectroscopy (Fig. 10a). Both kCS and kCR rates exhibited exponential decay
with variable shell thickness (Fig. 10b), mathematically represented as follows:
k(d) = k0 exp(−βd), where k0 denotes charge separation or recombination rate of
bare CdSe QDs and β represents decay constant. The findings of this study suggested
that the ZnS shell in CdSe/ZnS type I QDs slows down the electron and hole transfer
rates and the carrier extraction is controlled by the surface carrier density [101–
103]. It is worth noting that the decrease in the charge recombination rate is faster
compared to the charge separation rate with increasing shell thickness, which can
be ascribed to heavier hole effective mass in the ZnS shell [76]. As the rates of
charge separation are considerably faster than excited state relaxation, a high charge
separation yield can be achieved via increasing the shell thickness.

2.3.2 Type II and Quasi-Type II Core/Shell QDs

In type II and quasi-type II core/shell QDs, such as CdTe/CdSe (type II) and
CdSe/CdS (quasi type II) core/shell QDs, the electron is either localized in the shell
(in case of type II) or delocalized over both the core and shell (in case of quasi-
type II). Therefore, the electron holds greater chance to overlap with the adsorbed
272 A. I. Channa et al.

Fig. 10 (a) The kinetics of the excited QD population (NQD∗ (t)) for the QD−AQ complexes with
different shell thicknesses (x-axis is logarithmic in the right panel and linear in the left panel). (b)
Plot of the logarithm of charge separation (red circles) and recombination (blue triangles) rates as
a function of the ZnS shell thicknesses. (Adapted from Ref. [76])

electron acceptors coupled with the shell surface. The hole is confined in the core for
both cases (type II and quasi-type II), due to which electronic coupling of the hole at
the surface is reduced. As compared to bare QDs, the type II and quasi-type II QDs
can exhibit enhanced or retained charge separation rate (kCS ) but reduced charge
recombination rates (kCR ). A study based on CdTe/CdSe core/shell QDs (type II)
has revealed that type II carrier distribution can facilitate charge separation and
significantly prolonged carrier lifetimes compared to bare CdSe QDs [77]. Likewise,
quasi-type II CdSe/CdS core/shell QDs exhibited almost similar ultrafast charge
separation rate as the CdSe core, while the charge recombination rate lowered to
about three orders of magnitude [90]. This is different from the case of type I QDs
(such as CdSe/ZnS) in which both charge separation and recombination rates are
reduced [76]. The type II QDs with the internal electron–hole separation are similar
to the molecular donor–acceptor complexes which were employed to enhance the
efficiency of dye-sensitized solar cells [104, 105]. Thus, the type II or quasi-type II
core/shell QDs, with their built-in charge separation properties, are very promising
for light-harvesting applications in PV and photocatalytic devices.

2.3.3 Charge Dynamics from QDs to Semiconductor Films

In QD-based PV devices such as a QDSC or PEC photoanode, QDs are normally


attached with a wide-bandgap metal oxide semiconductor such as TiO2 , which acts
as an electron transport layer as well as a vessel for QDs [19, 106]. As the QDs
are surrounded by the organic ligands on the surface, the contact of QD-TiO2 is
indirect and linked together via the organic ligands. In this case, the photoexcited
electrons have to tunnel through an insulating barrier (i.e., linker organic molecule)
even though the linker molecules are important in attaching and increasing the
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 273

loading capacity of QDs onto the substrate and to achieve a homogenous distribution
[107]. The direct contact of QD–TiO2 can alleviate the electron tunnel barrier [108].
Watson et al. studied the effect of linker molecule length on the electron transfer
from QDs to TiO2 and demonstrated much decreased electron transfer rates for
long-chain linker molecule in QD–TiO2 assemblies [109]. In addition, Bisquert et
al. have demonstrated that the enhanced concentration of QDs on TiO2 can lead to
the increased incident photon-to-current efficiency (IPCE) for QDSCs [110]. From
these studies, it can be deduced that the charge recombination can possibly occur
at the QD–TiO2 interface due to poor charge transfer characteristics of these linker
molecules. The long lifetime of charge carriers and high charge carrier mobility can
minimize the charge recombination losses. In this regard, the band alignment of the
core/shell QDs can be tuned to exhibit spatial charge separation for extended PL
lifetime [111]. The energy band alignment can be tuned by changing the core QD
size/composition and the shell thickness [111]. For example, CdSe/CdS core/shell
QDs (with shell thickness <1 nm) exhibits type I band alignment [63]. However, via
increasing the CdS shell thickness, the electrons can be made to delocalize into the
shell which leads to an efficient electron/hole separation [64]. Further increasing the
shell thickness, however, can create a physical barrier to slow down the hole transfer
rate. In this situation, the hole transport can be enhanced via attaching different
molecules with superior transporting capability [112]. Alivisatos et al. demonstrated
the hole transfer from the CdSe/CdS core/shell QDs through a covalently linked
thiolate binding group attached on the QD surface [112]. Therefore, a promising
approach to address the charge transport from QDs to the attached metal oxide
semiconductor is to form a core/shell QD system with optimized band alignment.
The surface of the core/shell QDs can also be modified by attaching hole acceptor
molecules to obtain the improved hole transport capability. In addition, the charge
transport of both kinds of carriers can also be improved without the additional
hole acceptor molecules, but via the interfacial engineering of the core/shell QDs
to form the core/interfacial layer/shell type of structure [37]. Tong et al. and
Zhao et al. have demonstrated the interfacial engineered CuInSeS/CdSeS/CdS
and CdSe/CdSeS/CdS core/interfacial-layer/shell QDs without any hole acceptor
molecule to be efficient for transporting both kinds of carriers [37, 113].

3 Solar-Driven PEC Cells Based on Core/Shell QDs

Numerous core/shell structured QDs have been applied for the fabrication of
photoanodes for PEC hydrogen generation because of their outstanding photo-
/chemical stability and tunable optoelectronic properties as compared to the bare
QDs [23]. By appropriate tuning of the thickness and composition of the shell,
core/shell QDs can be made to exhibit enhanced PLQY and prolonged PL lifetime,
leading to efficient charge separation and reduced non-radiative recombination rates,
making them potential candidate for solar-driven PEC cells [47, 48].
274 A. I. Channa et al.

Fig. 11 Photocurrent density as a function of V versus reversible hydrogen electrode (RHE)


of (a) PbS/CdS core/shell QD-sensitized photoelectrode and (b) PbS/CdS core/shell QD-based
photoelectrode with a TiO2 blocking layer between FTO and mesoporous TiO2 film. (c) Mea-
sured photocurrent density as a function of time for the samples TiO2 /(PbS/CdS core/shell
QDs)/CdS/ZnS, TiO2 /(PbS/CdS core/shell QDs)/CdS, TiO2 /(PbS/CdS core/shell QDs) at 0.2 V
versus RHE (AM1.5G, 100 mW cm−2 ). (Adapted from Ref. [117])

3.1 Core/Thin-Shell QDs

The core/thin-shell QDs basically refer to the core/shell QDs with core QDs coated
by an ultrathin shell (such as by a cation exchange treatment). PbS/CdS core/thin-
shell QDs have attracted increasing attention in solar technologies due to their
size-dependent broad NIR light absorption (typically from 300 to 2000 nm) and
good photo-/chemical stability [114], which can be applied in PEC cells for efficient
solar-driven H2 generation [115, 116]. Jin et al. synthesized PbS/CdS core/thin-
shell QDs via a cation exchange method [117]. The PbS core QDs with an average
diameter of ~3 nm were first obtained by using hot-injection method. After the
cation exchange process for CdS shell growth, the blueshift in the absorption/PL
spectrum was observed, which was attributed to the decrease in PbS core QD
size during the replacement of Pb2+ by Cd2+ [118]. The CdS shell thickness was
estimated to be ~0.1 nm. PbS/CdS core/thin-shell QDs were then deposited in
mesoporous TiO2 film by electrophoretic deposition (EPD) approach. This process
was followed by the deposition of CdS and ZnS layer by SILAR method. The PEC
performance of the PbS/CdS core/shell QD-based photoanode was tested. Initially,
in the PbS/CdS core/shell QD-sensitized TiO2 photoanode, a saturated photocurrent
density of 2 mA cm−2 was observed. The PEC performance was further enhanced
to 7.3 mA cm−2 after coating the surface of PbS/CdS-sensitized photoanode by CdS
and ZnS layers, as shown in Fig. 11a.
The PEC performance of the PbS/CdS core/shell QD-based photoanode was
further boosted to 11.2 mA cm−2 after the introduction of a blocking layer of
TiO2 between fluorinated tin oxide (FTO) and mesoporous TiO2 film, as shown
in Fig. 11b. The durability of the PbS/CdS core/shell QD-based photoanode with
and without CdS and CdS/ZnS coatings was tested, showing higher stability for the
TiO2 /(PbS/CdS core/shell QDs)/CdS/ZnS-based photoanode (Fig. 11c). Herein, the
photocurrent density obtained from the PbS/CdS QD-based photoanode was limited
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 275

Fig. 12 (a) PEC performance of CGS/CdS core/shell QD-based photoanode and (b) photocur-
rent density as a function of time for CGS/CdS core/shell QD-based photoanode (AM 1.5G
100 mW cm−2 ). (Adapted from Ref. [119])

to 2 mA/cm2 , which was boosted to 7.3 mA cm−2 by further coating the photoanode
surface with CdS. It indicates that the major contribution for the photocurrent is
from the CdS layer rather than the PbS/CdS core/shell QDs. It is therefore more
significant to develop QDs with higher solar energy conversion efficiency for the
fabrication of QD-based PEC cells.
Channa and coworkers have reported the synthesis of a new type of CuGaS2
(CGS)/CdS core/shell QDs for application in solar-driven PEC cells [119]. As-
synthesized CGS/CdS core/shell QDs (having CdS shell thickness ~1 nm) showed
improved optical properties as compared to the CGS core QDs, for instance, the
long-lived PL lifetime (~4.8 μs), which infers the effective passivation of the
surface defects/traps on CGS core QDs by CdS shell overcoating. The long-lived
exciton lifetime can give rise to the efficient charge separation in the core/shell
QDs, which are important to boost the efficiency of QD-based PV devices [36,
37]. As a proof of concept, CGS/CdS core/shell QDs were used for the fabrication
of the PEC cells and corresponding device performance is displayed in Fig. 12a,
exhibiting a saturated photocurrent density as high as ~6.5 mA cm−2 , which is
much higher than the bare CGS QD-based PEC cells (~0.83 mA cm−2 ). Moreover,
the CGS/CdS core/shell QD-based photoanode showed higher device stability by
maintaining ~60% of the initial photocurrent density after 2 h of operation with
respect to the CGS core QD-based photoanode (~50%) as shown in Fig. 12b. This
work demonstrates that building a core/thin-shell structure can optimize the optical
properties and enhance the photo-/chemical stability of QDs, thus leading to the
improved device performance of QD-based PEC cells.
276 A. I. Channa et al.

3.2 “Giant” Core/Shell QDs

“Giant” core/shell QDs (g-QDs) are core/shell structured QDs with a very thick shell
typically ranging from 1.5 nm to tens of nm. By appropriately tailoring the chemical
composition and shell thickness, the band structure of g-QDs can be tuned to exhibit
prolonged exciton lifetime and reduced electron–hole recombination, for instance,
type II or quasi-type II g-QDs (e.g., CdSe/CdS and ZnSe/CdS) [120], which are
predicted as favorable for optoelectronic technologies including solar-driven PEC
H2 generation [121, 122].
Adhikari et al. designed and synthesized CdSe/CdS g-QDs with variable core
and shell thickness, which are labeled as Giant #1 (core radius of 1.85 nm, shell
thickness of 3.2 nm) and Giant #2 (core radius of 1.65 nm, shell thickness of 4.3 nm)
QDs [64]. These g-QDs were then used to sensitize mesoporous TiO2 films as
photoanodes for the fabrication of solar-driven PEC cells. The PEC performance
of g-QD-sensitized photoanodes is displayed in Fig. 13. Photocurrent density as a
function of applied voltage (V versus RHE) for bare CdSe, Giant #1, and Giant
#2 QD-based PEC photoanode without and with ZnS coating is shown in Fig.
13a, b, respectively. The Giant #2 QD-based PEC cells exhibit a superior saturated
photocurrent density as high as ~10 mA cm−2 , which is much higher than the
photocurrent density of bare CdSe QD-based photoanode. Figure 13c shows the
stability of these QD-based PEC photoanodes, wherein much improved stability was
obtained from Giant #2 QDs with two layers of ZnS coating on the QD-sensitized
photoanode. The results indicate that the band structure of g-QDs can be engineered
by varying the shell thickness during the synthesis of core/shell QDs, resulting in an
efficient charge transfer from the QDs to TiO2 due to efficient charge separation in
core/shell QDs. However, the applicability of such QDs still faces a challenge with
respect to device stability, for instance, the Giant #2 QD-based photoanode with two
ZnS layers only retained about 52% of its initial photocurrent density after 2 hours
of operation. These results demonstrate that the ZnS layer can significantly enhance
the stability of the g-QDs against the corrosive electrolyte in the presence of light.
However, an overly thick shell can inhibit the efficient transport of holes and lead
to the self-oxidation of g-QDs, thus resulting in lower performance of g-QD-based
PEC cells [64].

3.3 Core/Alloyed-Shell/Shell QDs

The g-QDs have been demonstrated as promising building blocks for achieving
high-performance solar-driven PEC cells. However, an overly thick shell can
hinder the hole transport and result in the self-oxidation of g-QDs for reduced
device stability. Adhikari et al. developed a core/multishell QD system by the
introduction of a hole transport layer at the interface of the core and shell to
facilitate efficient hole transfer from the core QDs to the electrolyte. In this
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 277

Fig. 13 Photocurrent density as a function of applied voltage (versus RHE) for bare CdSe, Giant
#1, and Giant #2 QD-based PEC cells (a) without and (b) with ZnS coating in the dark and under
one sun illumination (AM1.5G, 100 mW/cm2 ). (c) Stability measurements of these QD-based PEC
devices. (Adapted from Ref. [64])

work, CdSe/Pbx Cd1 − x S/CdS core/alloyed-shell/shell g-QDs were synthesized with


variable shell thicknesses, which were labeled as pure shell-6, alloyed shell-6, pure
shell-13, and alloyed shell-13 [123]. The hole transfer capability of the interfacial
layer Pbx Cd1-x S was investigated via transient fluorescence spectroscopy, which
revealed a threefold enhancement in the hole transfer rate as compared to CdS with
similar shell thickness.
They also performed theoretical calculations to further study the role of interfa-
cial layer Pbx Cd1-x S. For pure shell-6, alloyed shell-6, pure shell-13, and alloyed
shell-13, the degree of hole leakage from the core to the shell was found to be 12%,
22%, 11.5%, and 12%, respectively. In alloyed shell-6, the holes were found to leak
efficiently into the shell as compared to other samples, as a result of the introduced
Pbx Cd1-x S alloyed shell. However, in the thick shell sample (i.e., alloyed shell-13),
the reduction in hole leakage was attributed to the existence of thick CdS shell [123].
278 A. I. Channa et al.

Table 1 Summary of photocurrent density and device stability


Retention of Retention of
Photocurrent photocurrent (after photocurrent after
Samples density (mA cm−2 ) first 50 s) (%) 2 h (%)
Pure shell-6 7.2 75 54.1
Alloyed shell-6 6.0 70 66.3
Pure shell-13 8.9 98 87.0
Alloyed shell-13 8.0 98 94.9
Pure shell-13 9.9 – –
(graphene-TiO2
photoanode)
Alloyed shell-13 9.5 – –
(graphene-TiO2
photoanode)

The CdSe/Pbx Cd1 − x S/CdS core/alloyed-shell/shell g-QD-based photoanode


was fabricated by the sensitization of g-QDs in a mesoporous TiO2 structure
followed by the deposition of two ZnS layers to protect the g-QDs from pho-
todegradation. Figure 14a, b presents the PEC performance of the pure shell-6,
alloyed shell-6, pure shell-13, and alloyed shell-13 g-QD-based photoanodes,
among which the alloyed shell-13 g-QD-based photoanode exhibited the highest
PEC performance. Moreover, Fig. 14c, d displays the device stability of pure shell-
6, alloyed shell-6, pure shell-13, and alloyed shell-13 g-QD-based photoanodes
after 800 s and 7200 s. The alloyed shell-13 g-QD-based photoanode exhibited the
highest retention of the initial photocurrent density. The results obtained from the
PEC performance and the stability measurements are tabulated above (Table 1).
Herein, they also modified the traditional TiO2 porous film by 0.01 wt% graphene
addition; as a result, the photocurrent density shows a significant enhancement
(about ∼16%) for the alloyed shell-13 g-QDs. These findings provide insights that
could be useful in developing the core/shell QDs with high-performance and stable
QD-based PEC cells. It can be also concluded that along with the rational design of
the core/shell QDs, the modifications in the mesoporous TiO2 can lead to improved
device performance.
Wang et al. have reported the synthesis of CdSe/CdSeS/CdS core/alloyed-
shell/shell g-QDs to achieve high-performance PEC cells [113]. Two samples with
variable CdSeS and CdS shell thickness and composition labelled as CdSe/Alloy#1
(CdSe/(CdSe0.5 S0.5 )4 /(CdS)2 ) and CdSe/Alloy#2 (CdSe/(CdSex S1 − x )5 /CdS) were
synthesized, in which the purpose of using alloyed gradient shell was to provide
accelerated exciton dissociation via the stepwise band energy level engineering, as
depicted in Fig. 15a. Moreover, the gradient CdSeS alloyed g-QDs exhibited light
absorption extended to the longer wavelengths compared to the control sample of
CdSe/CdS core/shell g-QDs.
Figure 15b displays the PEC performance of the CdSe/Alloy#2 g-QD-based
photoanode. The saturated photocurrent density of 17.5 mA cm−2 was obtained
under standard one sun illumination (AM 1.5 G, 100 mW cm−2 ). Several synergistic
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 279

Fig. 14 PEC performance of (a, b) various photoanodes fabricated by loading different g-QDs
with variable shell thickness. (c, d) Stability measurements of these QD-based PEC devices.
(Adapted from Ref. [123])

effects can be responsible for such an unprecedented value of photocurrent density


such as (1) the extension of absorption spectrum due to incorporation of Se in the
alloyed shell, which in turn harvests more photons of light, (2) the stepwise gradient
280 A. I. Channa et al.

Fig. 15 (a) Schematic diagram of the band structure and carrier transition of the CdSe/Alloy#2
(CdSe/CdSeS/CdS core/alloyed-shell/shell) QD-based photoanode. (b) PEC performance of
CdSe/Alloy#2 QD-based photoanode. (c) Stability measurements of CdSe/Alloy#2 QD-based PEC
devices under different surface treatments. (Adapted from Ref. [113])

band alignment leading to more electrons and holes extracted from QDs and
injected into TiO2 , and (3) the efficient spatial separation of electrons and holes for
suppressed charge recombination. The stability of the CdSe/Alloy#2 g-QD-based
photoanode with two layers of ZnS at the surface was measured and a retention of
70% of the initial photocurrent density was obtained. In order to further optimize the
device stability, Wang et al. modified the surface of the CdSe/Alloy#2 g-QD-based
photoanode by various treatments such as vacuum annealing and thicker ZnS and
silica coating (Fig. 15c). By annealing the CdSe/Alloy#2 photoanode with 3ZnS
monolayers in vacuum, they found the improvement in the photocurrent density
retention after 2 h. The same processes were repeated for QD-based photoanodes
with increased number of ZnS layers. Although the retention in photocurrent density
was improved with the increased ZnS thickness, the initial photocurrent density
showed reduction as well. The photoanodes with 3 and 4 layers of ZnS coatings
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 281

exhibited 86% and 82% retention in the initial photocurrent density, respectively. A
layer of silica was further deposited on the photoanode to promote hole transport and
block the electrons. This strategy was found to be useful and the CdSe/Alloy#2 g-
QD-based photoanode retained 50% of the initial photocurrent density even after
39 h of continuous operation. These results indicate that core/alloyed-shell/shell g-
QDs are promising for long-term stable PEC operation for hydrogen generation.

4 Heavy Metal-Free Core/Shell QD-Based PEC Cells

The toxic heavy metals of Pb and Cd in the above-discussed core/shell and


core/alloyed-shell/shell QD-based PEC systems impart harmful effects to the
environment as well as to human health. Environmentally friendly core/shell QDs
(QDs without highly toxic and heavy metal elements) are thus more favorable
and should be developed for solar-driven QD-based PEC cell applications in
terms of their future commercialization. Multinary I–III–VI colloidal QDs have
gained much attention due to their less toxic nature, size/shape/composition tunable
optical properties, and excellent performance in PV/optoelectronic devices [124–
126]. Specifically, the colloidal QDs of I–III–VI possess great potential to replace
the highly toxic, heavy metal-based QDs such as CdS, PbS, etc. due to their
high absorption coefficients and narrow bandgaps [124–126]. In this regard, Tong
et al. reported the synthesis of heavy metal-free CuInSeS/ZnS core/shell QDs
[66]. The thickness of the ZnS shell was kept around ~0.1 nm to suppress the
surface traps/defects and enhance the stability of CuInSeS QDs. The CuInSeS/ZnS
core/shell QDs exhibited improved optical properties including enhanced emission
efficiency and prolonged lifetime as compared to bare CuInSeS. These QDs were
utilized as a photosensitizer in mesoporous TiO2 films to fabricate QD-based
PEC cells for solar-driven H2 production. The PEC performance of QD-sensitized
photoanodes was tested and the results are shown in Fig. 16. Figure 16a displays
the schematic diagram and estimated band alignment of QDs/TiO2 . The PEC
performance of bare CuInSeS QD-based photoanode was tested, demonstrating
a saturated photocurrent density of ~2.57 mA cm−2 (Fig. 16b). In contrast, the
CuInSeS/ZnS core/shell QD-based photoanode showed a significantly enhanced
photocurrent density of ~4.3 mA cm−2 compared to their bare CuInSeS QD
counterpart (Fig. 16c). The enhanced photocurrent density in the core/thin-shell QD
system was attributed to the efficient surface passivation of CuInSeS core QDs by
ZnS for reduced charge recombination. Furthermore, Fig. 16d presents the device
stability performance of CuInSeS bare QDs and CuInSeS/ZnS core/shell QD-based
photoanodes. The CuInSeS/ZnS core/shell QD-based photoanode maintained about
~77% of the initial photocurrent density, which is remarkably higher than the
CuInSeS QD-based photoanodes (~38% retention of the initial photocurrent density
after 2 h), indicating the efficient surface passivation of CuInSeS QDs via the ZnS
shell for enhanced stability of QDs. In addition, the optical photographs of the
CuInSeS/ZnS core/shell QD-based photoanodes before and after the stability test
282 A. I. Channa et al.

Fig. 16 (a) Schematic diagram and approximated band alignment of a CuInSeS/ZnS core/shell
QD-based photoanode. Photocurrent density versus applied voltage (V vs RHE) for (b) CISeS
QDs and (c) CuInSeS/ZnS core/thin-shell QD-based photoanode (AM 1.5 G, 100 mW/cm2 ). (d)
Stability measurements of bare CuInSeS and core/thin-shell CISeS/ZnS QD-modified photoanodes
(inset shows photographs of CuInSeS/ZnS core/shell QD-based photoanodes before (left) and after
(right) stability measurement (9 h, AM 1.5 G, 100 mW/cm2 ). (Adapted from Ref. [66])

were displayed in the inset of Fig. 16d, showing no noticeable color change of the
active area, which demonstrates no significant structural or chemical changes of
CuInSeS/ZnS core/shell QD-based cells after long-term device operation.
Moreover, the NIR heavy metal-free CuInSe2 /CuInS2 (CISe/CIS) core/shell g-
QDs were developed as another member of I–III–VI QDs with great potential in
PV devices due to broad absorption extended to the NIR region [89]. The optical
properties of such g-QDs can be readily tailored via tuning CIS shell thickness,
as demonstrated by Tong et al. [89]. In this work, a sequential cation exchange
procedure was used to synthesize CISe/CIS core/shell QDs using CdSe/CdS
core/shell QDs as initial templates [127]. CIS shell with variable thickness was
grown on CISe core QDs (i.e., CISe/6CIS and CISe/13CIS) via tuning the shell
thickness of CdSe/CdS core/shell QDs. The resulting CISe/CIS core/shell QDs
exhibited very broad UV–visible–NIR absorption up to 1100 nm. The PL spectra
of the CISe core and CISe/6CIS and CISe/13CIS core/shell QDs exhibited emission
peaks at the wavelengths of 765, 1075, and 1100 nm, respectively [89]. The CISe
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 283

core, CISe/6CIS, and CISe/13CIS core/shell QDs also exhibited prolonged average
lifetime with increasing shell thickness, inferring the existence of a quasi-type II
band alignment for efficient photocarrier separation in the CISe/CIS core/shell QDs
[89].
These quasi-type II g-QDs were then applied for the fabrication of photoanodes
for hydrogen generation. Although the CISe/CIS/ZnS core/shell/shell g-QD-based
photoanode only exhibited a saturated photocurrent density of ~3 mA cm−2 , the
CISe/CIS/ZnS core/shell/shell g-QDs with thicker shell maintained 80% of the
photocurrent density after 2 h. This indicates that these g-QDs are promising due
to their superb stability during device operation and the lower photocurrent density
suggests that there is still room left for the improvement of the device performance.

5 Conclusions and Perspectives

Hydrogen fuel is a promising high-energy-density fuel with zero hazardous effects


on the environment. Hydrogen combustion process can yield enormous amounts
of energy with water as the only by-product. Among various methods to produce
hydrogen, solar-driven PEC hydrogen generation has become one of the most
important research topics. Pioneering work in this field was carried out by Fujishima
and Honda who reported that semiconductor materials such as TiO2 have the
ability to drive the PEC reaction for hydrogen evolution. However, the solar
conversion efficiency of a TiO2 -based PEC photoanode was limited due to the
wide bandgap of TiO2 (typically around ~3.2 eV) that results in the inadequate
light absorption only in the UV region. The sensitization of TiO2 with narrow-
bandgap semiconductor materials can lead to enhanced photoconversion efficiency.
In this perspective, colloidal QDs were found to be very useful due to their
relatively narrow bandgaps, high absorption coefficients, and proper band alignment
with TiO2 . Among various types of QDs, the core/shell QDs have demonstrated
promising potential in applications due to their enhanced stability as compared to
bare QDs. Moreover, the optical properties of the core/shell QDs can be easily
tailored by tuning the core size and varying the shell thickness. Different kinds
of core/shell structures such as type I, type II and quasi-type II can be obtained,
wherein the type II and quasi-type II core/shell QDs with efficient charge separation
capability are more promising for PV applications.
Until now, various core/shell QDs including PbS/CdS, CdSe/CdS, CdSe/CdSeS/
CdS, etc. are utilized to achieve high-performance solar-driven PEC cells to produce
hydrogen. However, the heavy metals such as Pb and Cd are hazardous and can
cause dangerous chemical pollution in case of material disposal. Therefore, heavy
metal-free core/shell QDs such as CuInSeS/ZnS and CuInSe2 /CuInS2 core/shell
QDs were recently developed for the fabrication of PEC cells and show, however,
relatively low solar energy conversion efficiency. Yet the stability of the core/shell
QD-based PEC systems still cannot meet the requirement for industry proposes.
Future work may focus on exploring new heavy metal-free and ultrastable core/shell
284 A. I. Channa et al.

QDs to achieve breakthrough in the performance of QD-based PEC cells and their
consequent commercialization as well as real-life usage.

Acknowledgments This work was supported by the National Key Research and Development
Program of China (2019YFB2203400), the “111 Project” (B20030), and the UESTC Shared
Research Facilities of Electromagnetic Wave and Matter Interaction (Y0301901290100201).

References

1. Hoffert, M.I., Caldeira, K., Benford, G., et al.: Science. 298(5595), 981 (2002)
2. Lewis, N.S.: Science. 351(6271), aad1920 (2016)
3. Lewis, N.S., Nocera, D.G.: Proc Natl Acad Sci U S A. 103(43), 15729 (2006)
4. Blankenship, R.E., Tiede, D.M., Barber, J., et al.: Science. 332(6031), 805 (2011)
5. Singh, S., Jain, S., Ps, V., et al.: Renew Sustain Energy Rev. 51, 623–633 (2015)
6. Abdalla, A.M., Hossain, S., Nisfindy, O.B., et al.: Energ Conver Manage. 165, 602–627
(2018)
7. Fujishima, A., Honda, K.: Nature. 238(5358), 37–38 (1972)
8. Seery, M.K., George, R., Floris, P., et al.: J Photochem Photobiol A. 189(2), 258–263 (2007)
9. Kong, D., Zheng, Y., Kobielusz, M., et al.: Mater Today. 21(8), 897–924 (2018)
10. Alqahtani, M., Sathasivam, S., Chen, L., et al.: Sustain Energ Fuels. 3(7), 1720–1729 (2019)
11. Maeda, K., Domen, K.J.: Phys Chem Lett. 1(18), 2655–2661 (2010)
12. Walter, M.G., Warren, E.L., McKone, J.R., et al.: Chem Rev. 110(11), 6446–6473 (2010)
13. Moreels, I., Lambert, K., Smeets, D., et al.: ACS Nano. 3(10), 3023–3030 (2009)
14. Ahn, S., Chung, H., Chen, W., et al.: J Chem Phys. 151(23), 234705 (2019)
15. Cademartiri, L., Montanari, E., Calestani, G., et al.: J Am Chem Soc. 128(31), 10337–10346
(2006)
16. Alivisatos, A.P.: Science. 271(5251), 933–937 (1996)
17. Pu, Y., Cai, F., Wang, D., et al.: Ind Eng Chem Res. 57(6), 1790–1802 (2018)
18. Zhou, S., Liu, Z., Wang, Y., et al.: J Mater Chem C. 7(6), 1575–1583 (2019)
19. Nozik, A.J.: Phys E. 14(1), 115–120 (2002)
20. Song, J., Wang, O., Shen, H., et al.: Adv Funct Mater. 29(33), 1970226 (2019)
21. Livache, C., Martinez, B., Goubet, N., et al.: Nat Commun. 10(1), 2125 (2019)
22. You, Y., Tong, X., Wang, W., et al.: Adv Sci. 6(9), 1801967 (2019)
23. Zhao, H., Rosei, F.: Chem. 3(2), 229–258 (2017)
24. Vasudevan, D., Gaddam, R.R., Trinchi, A., et al.: J Alloys Compd. 636, 395–404 (2015)
25. Inerbaev, T.M., Masunov, A.E., Khondaker, S.I., et al.: J Chem Phys. 131(4), 044106 (2009)
26. Giansante, C., Infante, I.J.: Phys Chem Lett. 8(20), 5209–5215 (2017)
27. Aldana, J., Wang, Y.A., Peng, X.J.: Am Chem Soc. 123(36), 8844–8850 (2001)
28. Moon, H., Lee, C., Lee, W., et al.: Adv Mater. 31(34), 1804294 (2019)
29. Cozzoli, P.D., Pellegrino, T., Manna, L.: Chem Soc Rev. 35(11), 1195–1208 (2006)
30. Zhu, H., Lian, T.: Energ Environ Sci. 5(11), 9406–9418 (2012)
31. Mastria, R., Rizzo, A.J.: Mater Chem C. 4(27), 6430–6446 (2016)
32. Piryatinski, A., Ivanov, S.A., Tretiak, S., et al.: Nano Lett. 7(1), 108–115 (2007)
33. Leontiadou, M.A., Tyrrell, E.J., Smith, C.T., et al.: Sol Energy Mater Sol Cells. 159, 657–663
(2017)
34. Neo, D.C.J., Cheng, C., Stranks, S.D., et al.: Chem Mater. 26(13), 4004–4013 (2014)
35. Jia, Y., Chen, J., Wu, K., et al.: Chem Sci. 7(7), 4125–4133 (2016)
36. Tong, X., Li, X., Channa, A.I., et al.: Sol RRL. 3(10), 1900186 (2019)
37. Tong, X., Kong, X.-T., Wang, C., et al.: Adv Sci. 5(8), 1800656 (2018)
Core/Shell Quantum-Dot-Based Solar-Driven Photoelectrochemical Cells 285

38. Azwar, M.Y., Hussain, M.A., Abdul-Wahab, A.K.: Renew Sustain Energy Rev. 31, 158–173
(2014)
39. Turner, J.A.: Science. 285(5428), 687–689 (1999)
40. Khaselev, O., Turner, J.A.: Science. 280(5362), 425 (1998)
41. Koumi Ngoh, S., Njomo, D.: Renew Sustain Energy Rev. 16(9), 6782–6792 (2012)
42. Osterloh, F.: E Chem Soc Rev. 42(6), 2294–2320 (2013)
43. Lianos, P.: J Hazard Mater. 185(2–3), 575–590 (2011)
44. Huang, Q., Ye, Z., Xiao, X.J.: Mater Chem A. 3(31), 15824–15837 (2015)
45. Grätzel, M.: Nature. 414(6861), 338–344 (2001)
46. Elgrishi, N., Rountree, K.J., McCarthy, B.D., et al.: J Chem Educ. 95(2), 197–206 (2018)
47. Reiss, P., Protière, M., Li, L.: Small. 5(2), 154–168 (2009)
48. Zhao, H., Chaker, M., Ma, D.J.: Mater Chem. 21(43), 17483–17491 (2011)
49. Ghosh Chaudhuri, R., Paria, S.: Chem Rev. 112(4), 2373–2433 (2012)
50. Etgar, L., Yanover, D., Čapek, R.K., et al.: Adv Funct Mater. 23(21), 2736–2741 (2013)
51. Dabbousi, B.O., Rodriguez-Viejo, J., Mikulec, F.V., et al.: J Phys Chem B. 101(46), 9463–
9475 (1997)
52. Kim, S., Fisher, B., Eisler, H.-J., et al.: J Am Chem Soc. 125(38), 11466–11467 (2003)
53. Lifshitz, E., Brumer, M., Kigel, A., et al.: J Phys Chem B. 110(50), 25356–25365 (2006)
54. Jo, D.-Y., Yang, H.: Chem Commun. 52(4), 709–712 (2016)
55. Jalalah, M., Al-Assiri, M.S., Park, J.-G.: Adv Energy Mater. 8, 20 (2018)
56. Zhao, H., Sirigu, G., Parisini, A., et al.: Nanoscale. 8(7), 4217–4226 (2016)
57. Xie, R., Peng, X.J.: Am Chem Soc. 131(30), 10645–10651 (2009)
58. Tessier, M.D., Dupont, D., De Nolf, K., et al.: Chem Mater. 27(13), 4893–4898 (2015)
59. Lim, J., Jeong, B.G., Park, M., et al.: Adv Mater. 26(47), 8034–8040 (2014)
60. Xu, S., Ziegler, J., Nann, T.J.: Mater Chem. 18(23), 2653–2656 (2008)
61. Ghosh, Y., Mangum, B.D., Casson, J.L., et al.: J Am Chem Soc. 134(23), 9634–9643 (2012)
62. Chen, Y., Vela, J., Htoon, H., et al.: J Am Chem Soc. 130(15), 5026–5027 (2008)
63. Brovelli, S., Schaller, R.D., Crooker, S.A., et al.: Nat Commun. 2(1), 280 (2011)
64. Adhikari, R., Jin, L., Navarro-Pardo, F., et al.: Nano Energy. 27, 265–274 (2016)
65. McDaniel, H., Koposov, A.Y., Draguta, S., et al.: J Phys Chem C. 118(30), 16987–16994
(2014)
66. Tong, X., Zhou, Y., Jin, L., et al.: Nano Energy. 31, 441–449 (2017)
67. Lambert, K., Geyter, B.D., Moreels, I., et al.: Chem Mater. 21(5), 778–780 (2009)
68. McDaniel, H., Fuke, N., Makarov, N.S., et al.: Nat Commun. 4, 2887 (2013)
69. Pietryga, J.M., Werder, D.J., Williams, D.J., et al.: J Am Chem Soc. 130(14), 4879–4885
(2008)
70. Hines, M.A., Guyot-Sionnest, P.J.: Phys Chem. 100(2), 468–471 (1996)
71. Steckel, J.S., Zimmer, J.P., Coe-Sullivan, S., et al.: Angew Chem Int Ed. 43(16), 2154–2158
(2004)
72. Li, L., Reiss, P.J.: Am Chem Soc. 130(35), 11588–11589 (2008)
73. Xie, R., Kolb, U., Li, J., et al.: J Am Chem Soc. 127(20), 7480–7488 (2005)
74. Kim, S., Kim, T., Kang, M., et al.: J Am Chem Soc. 134(8), 3804–3809 (2012)
75. Lim, S.J., Chon, B., Joo, T., et al.: J Phys Chem C. 112(6), 1744–1747 (2008)
76. Zhu, H., Song, N., Lian, T.J.: Am Chem Soc. 132(42), 15038–15045 (2010)
77. Zhu, H., Song, N., Lian, T.J.: Am Chem Soc. 133(22), 8762–8771 (2011)
78. Kim, S., Lim, Y.T., Soltesz, E.G., et al.: Nat Biotechnol. 22(1), 93–97 (2004)
79. Chen, C.-Y., Cheng, C.-T., Yu, J.-K., et al.: J Phys Chem B. 108(30), 10687–10691 (2004)
80. Dooley, C.J., Dimitrov, S.D., Fiebig, T.J.: Phys Chem C. 112(32), 12074–12076 (2008)
81. Hewa-Kasakarage, N.N., Kirsanova, M., Nemchinov, A., et al.: J Am Chem Soc. 131(3),
1328–1334 (2009)
82. Chuang, C.-H., Doane, T.L., Lo, S.S., et al.: ACS Nano. 5(7), 6016–6024 (2011)
83. García-Santamaría, F., Chen, Y., Vela, J., et al.: Nano Lett. 9(10), 3482–3488 (2009)
84. Oron, D., Kazes, M., Banin, U.: Phys Rev B. 75(3), 035330 (2007)
85. Rainò, G., Stöferle, T., Moreels, I., et al.: ACS Nano. 5(5), 4031–4036 (2011)
286 A. I. Channa et al.

86. Maity, P., Debnath, T., Ghosh, H.N.J.: Phys Chem C. 119(46), 26202–26211 (2015)
87. Pandey, A., Guyot-Sionnest, P.: Science. 322, 5903–5929 (2008)
88. De Geyter, B., Justo, Y., Moreels, I., et al.: ACS Nano. 5(1), 58–66 (2011)
89. Tong, X., Kong, X.-T., Zhou, Y., et al.: Adv Energy Mater. 8(2), 1701432 (2018)
90. Zhu, H., Song, N., Rodríguez-Córdoba, W., et al.: J Am Chem Soc. 134(9), 4250–4257 (2012)
91. García-Santamaría, F., Brovelli, S., Viswanatha, R., et al.: Nano Lett. 11(2), 687–693 (2011)
92. Htoon, H., Malko, A.V., Bussian, D., et al.: Nano Lett. 10(7), 2401–2407 (2010)
93. Jin, S., Lian, T.: Nano Lett. 9(6), 2448–2454 (2009)
94. Song, N., Zhu, H., Jin, S., et al.: ACS Nano. 5(11), 8750–8759 (2011)
95. Jin, S., Hsiang, J.-C., Zhu, H., et al.: Chem Sci. 1(4), 519–526 (2010)
96. Song, N., Zhu, H., Jin, S., et al.: ACS Nano. 5(1), 613–621 (2011)
97. Sambur, J.B., Parkinson, B.A.J.: Am Chem Soc. 132(7), 2130–2131 (2010)
98. Li, T.-L., Lee, Y.-L., Teng, H.: Energ Environ Sci. 5(1), 5315–5324 (2012)
99. Shen, Q., Kobayashi, J., Diguna, L.J., et al.: J Appl Phys. 103(8), 084304 (2008)
100. Diguna, L.J., Shen, Q., Kobayashi, J., et al.: Appl Phys Lett. 91(2), 023116 (2007)
101. Dworak, L., Matylitsky, V.V., Breus, V.V., et al.: J Phys Chem C. 115(10), 3949–3955 (2011)
102. Jiang, Z.-J., Kelley, D.F.J.: Phys Chem C. 115(11), 4594–4602 (2011)
103. Xu, Z., Hine, C.R., Maye, M.M., et al.: ACS Nano. 6(6), 4984–4992 (2012)
104. Yella, A., Lee, H.-W., Tsao, H.N., et al.: Science. 334(6056), 629 (2011)
105. Bessho, T., Zakeeruddin, S.M., Yeh, C.-Y., et al.: Angew Chem Int Ed. 49(37), 6646–6649
(2010)
106. Kamat, P.V., Phys, J.: Chem Lett. 4(6), 908–918 (2013)
107. Pernik, D.R., Tvrdy, K., Radich, J.G., et al.: J Phys Chem C. 115(27), 13511–13519 (2011)
108. Hines, D.A., Forrest, R.P., Corcelli, S.A., et al.: J Phys Chem B. 119(24), 7439–7446 (2015)
109. Dibbell, R.S., Watson, D.F.J.: Phys Chem C. 113(8), 3139–3149 (2009)
110. Guijarro, N., Lana-Villarreal, T., Mora-Seró, I., et al.: J Phys Chem C. 113(10), 4208–4214
(2009)
111. Zhao, H., Fan, Z., Liang, H., et al.: Nanoscale. 6(12), 7004–7011 (2014)
112. Ding, T.X., Olshansky, J.H., Leone, S.R., et al.: J Am Chem Soc. 137(5), 2021–2029 (2015)
113. Wang, K., Tong, X., Zhou, Y., et al.: J Mater Chem A. 7(23), 14079–14088 (2019)
114. Concina, I., Manzoni, C., Grancini, G., et al.: J Phys Chem Lett. 6(13), 2489–2495 (2015)
115. Zhao, H., Jin, L., Zhou, Y., et al.: Nanotechnology. 27(49), 495405 (2016)
116. Zhang, H., Selopal, G.S., Zhou, Y., et al.: Nanoscale. 9(43), 16843–16851 (2017)
117. Jin, L., AlOtaibi, B., Benetti, D., et al.: Adv Sci. 3(3), 1500345 (2016)
118. Zhao, H., Wang, D., Zhang, T., et al.: Chem Commun. 46(29), 5301–5303 (2010)
119. Channa, A.I., Tong, X., Xu, J.-Y., et al.: J Mater Chem A. 7(17), 10225–10230 (2019)
120. Navarro-Pardo, F., Zhao, H., Wang, Z.M., et al.: Acc Chem Res. (2017)
121. Zhu, H.M., Song, N.H., Lv, H.J., et al.: J Am Chem Soc. 134(28), 11701–11708 (2012)
122. Perera, D., Lorek, R., Khnayzer, R.S., et al.: J Phys Chem C. 116(43), 22786–22793 (2012)
123. Adhikari, R., Basu, K., Zhou, Y., et al.: J Mater Chem A. 6(16), 6822–6829 (2018)
124. Allen, P.M., Bawendi, M.G.J.: Am Chem Soc. 130(29), 9240–9241 (2008)
125. Berends, A.C., Mangnus, M.J.J., Xia, C., et al.: J Phys Chem Lett. 10(7), 1600–1616 (2019)
126. Li, F., Guo, C., Pan, R., et al.: Opt Mater Express. 8(2), 314–323 (2018)
127. van der Stam, W., Bladt, E., Rabouw, F.T., et al.: ACS Nano. 9(11), 11430–11438 (2015)
Core/Shell Quantum-Dot-Based
Luminescent Solar Concentrators

Guiju Liu, Xiaohan Wang, Guangting Han, and Haiguang Zhao

Abstract Luminescent solar concentrators (LSCs) can serve as large-area sun-


light collectors, suitable for applications in high-efficiency and low-cost energy
conversion devices. LSCs can also provide adaptability to the needs of architects
for building-integrated photovoltaics, which makes them an attractive option for
structural buildings as transparent or nontransparent electricity generators. The
optical efficiency of large-area LSCs significantly depends on the optical properties
of the fluorophores. Among various types of fluorophores used in LSCs, core/shell
quantum dots (QDs) are promising candidates as a new type of absorber/emitter
in LSCs, due to their size-tunable wide absorption spectrum, narrow emission
spectrum, high quantum yield, and structure-engineered large Stokes shift compared
to organic dyes and polymers. In this chapter, we first introduce the working
principle of an LSC, and then we discuss the design and synthesis of core/shell QDs
with high Stokes shift and high fluorescence quantum yield, and core/shell structure-
dependent band energy alignment. We further discuss in details the relationship
between structure and optical properties, which is a key requirement for their
applications in LSCs. We conclude with a detailed account of the latest research
progress in structure, materials, and performance of LSCs based on colloidal
core/shell QDs and a further perspective on the remaining key issues and open
opportunities in the field.

Keywords Core/shell quantum dots · Luminescent solar concentrators · Optical


property · Stokes shift · Quantum yield · Stability

G. Liu · H. Zhao ()


College of Physics & State Key Laboratory of Bio-Fibers and Eco-Textiles, Qingdao University,
Qingdao, People’s Republic of China
e-mail: hgzhao@qdu.edu.cn
X. Wang · G. Han
State Key Laboratory of Bio-Fibers and Eco-Textiles & College of Textiles & Clothing, Qingdao
University, Qingdao, People’s Republic of China
e-mail: kychgt@qdu.edu.cn

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 287
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4_9
288 G. Liu et al.

1 Introduction of Luminescent Solar Concentration

With the rapid development of global economy and the rapid growth of population,
the demand for energy is increasing day by day and the problems of fossil
energy exhaustion, greenhouse effect, and environmental pollution are increasingly
serious. Changing the energy structure, developing new energy, and preventing
environmental pollution are the great challenges for today’s social sustainable
development. Solar energy is a clean and renewable energy with huge reserves. Solar
energy technologies represent an opportunity to solve the energy crisis and alleviate
the ecological environment problems caused by the consumption of fossil energy
[1–7]. Among the various types of solar energy technologies, solar-to-electricity
conversion is the most dynamic and fastest developing research field in recent
years. The power conversion efficiency (PCE) of silicon solar cells has reached over
25%, but the cost of the electricity generated by commercial silicon photovoltaic
(PV) devices is still high compared to traditional power plants (like thermal power
generation) [8–9]. There are still several challenges that limit the development of
the solar energy, such as fabrication, cost, efficiency, and stability. Concentrated PV
technology has become an emerging technology to improve the PCE of PV devices
[10–11]. However, the traditional concentrated PV devices must be equipped with
the solar tracking system to follow the sun, which increase the cost of a module.
Luminescent solar concentrators (LSCs) are attracting great interest as a platform
for solar energy harvesting [10, 12–20]. LSCs can serve as large-area sunlight
collectors for PV cells to reduce the power generation cost by decreasing the use
of expensive PV cell materials [21–22]. A typical LSC consists of waveguide
transparent/semitransparent materials, such as glass or polymer, doped with fluo-
rescent materials, e.g., organic dyes, polymer, and quantum dots (QDs) [14, 23–30].
The working principle of an LSC is that the fluorophores, which are embedded
in a waveguide or spin coated on the surface of the waveguide, absorb sunlight
and reemit emissions with a shorter or longer wavelength and propagated to the
edge of the waveguide through total internal reflection (TIR). The concentrated
emissions on the edges of the LSC are collected by PV cells placed at the edges
of the waveguide, and they can be eventually converted to electricity, as shown
in Fig. 1 [31]. Compared with the traditional concentrated PV systems, the LSC
technology can concentrate both direct and diffused solar radiations, which is not
limited by the angle of sunlight incidence. There is no need to add a solar tracking
system, which reduces the complexity of the concentrated light system and the
cost of power generation. The LSCs can be widely used as power generation units
in semitransparent solar windows, modernized agricultural greenhouses, and glass
curtain walls to develop building-integrated PVs (BIPVs) to realize the building
energy self-sufficiency by tuning the type and concentration of fluorophores to
adjust transparent and emitting color [20, 22, 32–41].
The efficiency of an LSC mainly depends on the photon conversion and transport
over a macroscopic distance; thus, it suffers from several loss mechanisms, as shown
in Fig. 1. For example, sunlight is lost in the colloidal QD-based LSCs in the
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 289

Fig. 1 Schematic representation of a QD-based LSC. (Reproduced with permission from Ref.
[31]. Copyright 2016, Wiley-VCH)

following ways: (1) only a fraction of incident light can be absorbed by the QDs
due to the limited overlap between the QDs’ absorption and solar spectrum; (2)
part of the sunlight is reflected at the front surface of the devices; (3) because the
fluorescence quantum yield (QY) of the QDs is below 100%, the absorbed sunlight
cannot be completely converted into fluorescent photons; (4) the fraction of emitted
photons falling into the escape cone as defined by Snell’s law is lost [42]; and (5) the
fraction of photons trapped in the waveguide can still be lost by reabsorption, which
is caused by the overlap between the absorption and emission spectra of the QDs.
Usually, reabsorption is a crucial loss factor because it occurs repetitively. It happens
when the photons traveling within the waveguide can be absorbed by other QDs and
reemitted with similar energy band edges multiple times before reaching an edge.
The reabsorption limits the implementation of the large-scale LSCs. Other minor
loss could be due to the absorption of emitted light by the polymer matrix or the
emissions are quenched by the defects in the waveguide. At present the preparation
of LSCs with high efficiency is still a huge challenge.
In the LSCs, optical efficiency (ηopt ) is used to quantify the fraction of photons
reaching the edges. It is defined as the ratio between the optical power of the emitted
photons transmitting to the edges and that of the incident photons. By accounting
for the various loss mechanisms present in an LSC, the ηopt can be calculated by the
following equation [17, 43–45]:

ηopt = ηAbs · ηinter = ηAbs · ηQY · ηtrap · ηself (1)

where ηAbs is the fraction of absorbed sunlight, and ηinter is the internal quantum
efficiency of an LSC. The ηinter can be presented as ηinter = ηQY ηtrap ηself , where
ηQY is the fluorescence QY of emitters; ηtrap is the efficiency of light trapping into
the LSC, which is determined by comparing the emission angle to the critical angle
290 G. Liu et al.

of the reflection; and ηself is related to self-absorption due to the spectral overlap of
absorption and emission.
A common method to characterize the optical performance of an LSC is to couple
a PV cell at the edge of the LSC. The ηopt can be calculated as [35, 45–46]:

ILSC • APV ILSC


ηopt = = (2)
ISC • ALSC ISC • G

where ILSC is the short-circuit current generated by the PV cell coupled with the
LSC, ISC is the short-circuit current of the same PV cell under direct illumination,
and G is the geometric factor, defined as the ratio between the area of the top of the
LSC (ALSC ) and the area of the edge of the LSC (APV ). So far, the research direction
on LSCs focuses on the improvement of their optical efficiency by improving the
optical properties of the fluorophores and configuration of the LSCs.
During the last thirty years, various types of fluorescent materials have been
used as light converters in LSCs, including dyes, polymers, and QDs [13–14,
23, 29, 47–52]. In early studies, organic dyes with high fluorescence QY are the
most commonly used fluorophores, mainly contains rhodamines, coumarins, and
perylene derivatives [13–14, 53–55]. However, the performance of these LSCs is
limited by the poor stability of the dyes and the large reabsorption loss due to the
significant overlap between absorption and emission spectra of these dyes [56–57].
Quite recently, inorganic fluorophores, especially fluorescent QDs, have been used
as emitters in LSCs due to their size/compositional tunable absorption/emission
spectra, high QY, wide absorption spectrum, good optical/chemical stability, and
solution processability [31, 36, 58]. More importantly, the Stokes shift (energy gap
between the first absorption peak and emission peak) of QDs can be engineered to
be very large, which is a key parameter to realize large-area high-efficiency LSCs
as the spectral overlap often leads to reabsorption energy loss [31, 59–60]. Usually,
compared with the bare QDs, the core/shell QDs have higher QY, better spectral
separation between emission and absorption spectra, and improved stability. In this
chapter, we will discuss different types of core/shell QDs based on their bandgap
alignment and optical properties to highlight their potential benefits as emitters in
LSCs. Various types of core/shell QD-based LSCs will be introduced, including
PbS/CdS, CdSe/CdS, InP/ZnO, and CuInS2 /ZnS core/shell QDs.

2 Band Alignment and Optical Properties of Core/Shell


Quantum Dots

2.1 Exciton Dynamic in Core/Shell Quantum Dots

In high-performance LSCs, an extremely promising type of fluorescent QDs are


core/shell QDs, which consist of core QDs covered by the shell of a different semi-
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 291

Fig. 2 (a) Schematic representation of the energy-level alignment in different core/shell systems.
The lower and upper edges of the rectangles correspond to the positions of the valence band and
conduction band edges of the core and shell materials, respectively. (b) Electron and hole wave
function of core/shell QDs. (Ref. [65]. Copyright 2010, American Chemical Society)

conductor material. Depending on the relative band alignment of the conduction-


band edge (CB) and valence-band edge (VB) of core and shell semiconductors,
the core/shell QDs can be classified into three types: type I, type II and reverse
type I, as shown in Fig. 2 [61–66]. In type I structure, the bandgap of the shell is
larger than that of the core and both the electrons and holes are localized in the
core, resulting in chemically stable and well-passivated QDs with high QY. The
reverse type I is exactly the opposite of type I, where the electrons and holes are
partially or completely confined in the shell depending on the thickness of the
shell. In the type II structure, either the CB or the VB of the shell materials is
located in the bandgap of the core, inducing the spatial separation of the electrons
and holes. Due to the lower overlap of the electron and hole wave functions, the
PL decay times are strongly prolonged in the type II structure [64–65]. Usually,
the type I core/shell QDs have higher QY and better stability compared to the
type II core/shell structure. The most extensively studied type I core/shell QDs
are CdSe/ZnS, PbS/CdS, and CdSe/CdS QDs [67–69]. For the type II system,
due to the well separation of excited electrons and holes, they are more beneficial
in photoelectrochemical hydrogen generations. The type II structured QDs (e.g.,
CdTe/CdSe, CdTe/CdS, and CdS/ZnSe QDs) have been extensively studied [62,
70–72]. The band structure of the core/shell QDs can be controlled by tuning the
core size, shell thickness, and chemical composition of both core and shell [73].
292 G. Liu et al.

A special case is called quasi-type II structure. In this case, one type of carriers is
delocalized over the entire core/shell structure, while the other type of carriers is
confined in the core (Fig. 2b) [65]. For example, in CdSe/CdS core/thick-shell QDs
(shell thickness, H > 1.5 nm), by increasing the shell thickness, the electrons exhibit
increased leakage from the core to the shell region, while the holes remain confined
in the core, and the band alignment of the core/shell structure changes from type I
to quasi-type II structure [74]. Similarly in CuInSe2 /CuInS2 core/thick-shell QDs,
with the increase of CuInS2 shell thickness, the QDs also exhibit a quasi-type II
band alignment structure [75]. The quasi-type II band alignment is advantageous for
optoelectronic devices, as this structure leads to an efficient electron-hole separation,
an increased PL lifetime, and a strong PL red shift, leading to enlarged Stokes shift,
which is a dominant factor for the fabrication of high-quality LSCs [60, 73–74, 76].

2.2 Stokes Shift in Core/Shell Quantum Dots

The Stokes shift is the spectral difference between the first excitonic absorption peak
and maximum emission, as shown in Fig. 3 [58]. The reabsorption energy loss can
be largely reduced by the increase of Stokes shift, especially for large-scale LSCs.

Fig. 3 (a) Normalized absorption and PL spectra of PbS/CdS QDs with core radius of 1.5 nm and
2.3 nm with tunable CdS shell thickness (H, 0–0.7 nm). (b) Approximate electronic band structure
of a core/shell PbS/CdS QD. (c) Stokes shift of QDs with various shell thicknesses as a function
of the PL peak position (related to the core size diameter from 3 to 6 nm). (Ref. [58]. Copyright
2016, Wiley-VCH)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 293

Several strategies have been proposed to increase the Stokes shift of QDs, containing
the design and synthesis of ternary I-III-VI2 semiconducting nanocrystals (e.g.,
CuInSex S2–x QDs) [77–78], doped core/shell nanocrystals (e.g., Mn2+ -doped
ZnSe/ZnS QDs) [79–80], and core/shell QDs (e.g., CdSe/CdS, PbS/CdS QDs) [58,
60, 81]. As mentioned above, coating QDs with a wide-bandgap semiconductor
to form a quasi-type II structure can increase the Stokes shift due to the spatial
separation of e-h wave function [58]. As shown in Fig. 3, the PbS/CdS core/shell
QDs exhibit a larger Stokes shift compared to the PbS core QDs. Meanwhile, the
Stokes shift could be tunable by varying core size and shell thickness. For example,
in PbS/CdS QDs with a larger core size (R = 2.3 nm), there is a very slight variation
of the Stokes shift by tuning the shell thickness. In contrast, for the QDs with a core
radius of 1.5 nm, the Stokes shift exhibits a significant increase with the increase of
shell thickness. The PbS/CdS core/shell QDs with small core size and thicker shell
layer have larger Stokes shift (200–250 meV) than that of core QDs or core/thin-
shell QDs. With the increase of the shell thickness, more electrons are delocalized in
the shell region and the absorption becomes dominated by the thick shell, leading to
an increase of the Stokes shift [58]. Similar to the core/thick-shell quasi-type II QDs,
core/thick shell type I QDs could also have large Stokes shift due to the leakage
of the electron. In the thick-shell QDs, the absorption of the QDs is dominant in
the shell materials, while the emission is governed by both the shell and core. For
example, the thick-shell CdSe/Cd1-x Znx S QDs have a PL peak at 628 nm and the
onset of absorption at 460 nm, which corresponds to a Stokes shift of 720 meV [47].
It is sufficiently large to reduce the reabsorption energy loss in large-scale LSCs.

2.3 Quantum Yield in Core/Shell Quantum Dots

Core/shell QDs are usually more suitable for optoelectronic applications due to
their greater environmental tolerance and higher QY compared to bare QDs, which
often suffer from the surface oxidation. The QY is defined as the ratio between the
emitted photons and absorbed photons by QDs. It is also a very important factor
to determine the optical efficiency in LSCs [68, 82–84]. The QY of the QDs are
strongly affected by the structure of the QDs. The presence of shell materials usually
can give a very well surface passivation of the core, thus enhancing the optical
properties of QDs. For example, for the PbS/CdS QDs, as the CdS shell thickness
increases, the QY increases first, due to the surface passivation, and then decreases
quickly because of the lattice mismatch, which leads to defect-related non-radiative
decay. The maximum QY of 33% was obtained for the QDs in water with optimized
shell thickness of 0.7 nm [68]. When the QDs were dispersed in toluene, the QY is
increased to 67% after a 0.7-nm CdS-shell coating [85]. For the thicker-shell QDs
(H > 1 nm), the introduction of defects causes a significant drop of QY (Fig. 4) [68].
For the synthesis of CdSe/CdS QDs via a successive ionic layer adsorption reaction
(SILAR) approach, long annealing times and asymmetric distribution of annealing
times between cadmium and sulfur precursors significantly affect QY in both thin-
294 G. Liu et al.

Fig. 4 QY of PbS/CdS QDs


before and after water
transfer as a function of CdS
shell thickness. (Ref. [68].
Copyright 2011, The Royal
Society of Chemistry)

and thick-shell regimes [82]. Hanifi et al. [84] synthesized CdSe/CdS core/shell
QDs by a SILAR approach using oleylamine, which helps to maintain a high ligand
surface coverage, reduce the surface traps, and preserve a high radiative efficiency.
With the increase of CdS shell thickness, the QY increases first and then tends to be
stable. For the QDs with a 2-nm-thick shell, the QY is 90 ± 2%, and when the shell
thickness increases to 4 nm, the QY consistently exceeds 97 ± 3% [84].

2.4 Stability of Core/Shell Quantum Dots

The stability of the QDs plays an important role in the potential applications
of QD-based optoelectronic devices. The photostability of the QDs is usually
characterized by the variation of PL intensity under continuous illumination. In
general, the core/shell QDs show much better stability as compared to the shell-
free QDs due to the improved surface passivation of the core QDs [68, 74–75].
For example, the PbS/CdS core/shell QDs exhibit markedly better photostability
under continuous ultraviolet (UV) illumination and better thermal stability under
heat treatment than the shell-free PbS QDs dispersed in organic solvent (Fig. 5a, b)
[85]. Compared to the CdZnSeS alloy cores, the CdZnSeS/ZnS core/shell QDs with
3, 11, and 15 monolayers of the ZnS shell show better photostability [86]. When
the shell thickness further increased to 17 MLs, the PL intensity decreases due to
the increased interfacial defects caused by stress release, but the decrease rate is
much slower than that of the QDs coated with 3 MLs of ZnS. Hence, the thick
shell could effectively improve the photostability of QDs. A similar phenomenon
was also found in the system of InP/ZnS QDs. For the InP/ZnS QDs with thin
shell, the PL intensity shows a rapid decrease within 10 h of UV illumination,
while for the thick-shell QDs, even after 25 h of continuous irradiation, little
photodegradation can be seen [87]. The core/thick-shell QDs can be engineered to
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 295

Fig. 5 (a) The variation in the PL intensity of PbS and PbS/CdS QDs in ODE after 4 h of
continuous illumination. (b) The PL spectra of PbS and PbS/CdS QDs before and after heat
treatment in ODE (PbS, 100 ◦ C for 5 min; PbS/CdS, 150 ◦ C for 30 min). (Ref. [85]. Copyright
2011, The Royal Society of Chemistry)

exhibit outstanding chemical/photostability for the fabrication of high-performance


optoelectronic devices [69, 74, 88–89].

3 Luminescent Solar Concentrator Based on Core/Shell


Quantum Dots

3.1 Near-Infrared PbS/CdS Core/Shell QD-Based LSCs

As mentioned above, QDs especially core/shell QDs usually have high QY and good
stability due to the appropriate surface passivation. The Stokes shift can be tuned by
controlling the absorption and emission spectra which are determined by the band
alignment of core/shell QDs. Usually, the quasi-type II core/shell QDs are excellent
emitters for near-infrared (NIR) wavelength, which is well suited for harvesting
solar radiation. It is able to realize the fabrication of high-transparency colorless
LSCs. In addition, the NIR wavelength emitted by QDs perfectly match with a
typical Si PV cell which exhibits high PCE in the 400–1000-nm range. Zhou et al.
[58] synthesized the NIR PbS/CdS QDs by hot injection for PbS QDs and flowing a
cation exchange approach for core/shell QDs. The absorption and emission spectra
can be well controlled by tuning the size of the core and the thickness of the shell
(Fig. 3). The PbS/CdS QDs with small core size and large shell thickness have
larger Stokes shift due to the efficient leakage of electrons [58, 68]. The PbS/CdS
QDs with high QY and large Stokes shift can alleviate the reabsorption, making
them promising candidates as emitters for high-efficiency large-area LSCs.
296 G. Liu et al.

Zhou et al. [58] fabricated the PbS QDs and PbS/CdS QD-based LSCs
(5 × 1.5 × 0.2 cm3 ) by embedding the QDs in a polymer matrix. Typically,
the monomer precursors of lauryl methacrylate and ethylene glycol dimethacrylate
were first mixed at a mass loading of ~20%. Then the UV initiator [diphenyl(2,4,6-
trimethylbenzoyl)phosphine oxide] was added in the above solution to form a clear
solution. Subsequently, the solution was mixed with the dried QDs powder and
sonicated until a clear solution was obtained. The mixture was further injected
into a mold consisting of two glass slides separated by a flexible rubber spacer
and illuminated by a UV lamp to form an LSC. With the increase of the optical
path, the normalized PL peak position of PbS QD-based LSC shows a redshift due
to the strong overlap between the absorption and emission spectra, which causes
strong reabsorption (Fig. 6a). However, for the PbS/CdS QD-based LSC, the PL
spectra remain unchanged along the optical path as long as 5 cm due to the large
Stocks shift, which suppresses the reabsorption (Fig. 6b). Eventually, the large-

Fig. 6 Normalized absorption and PL spectra measured at different optical paths for the samples
of (a) the large-sized PbS QDs and (b) the small-sized PbS/CdS core/shell QDs. Optical efficiency
of PbS/CdS QD-based LSC coupled with Si diode as a function of (c) side length or (d) G factor.
(e) Photograph of a QD-polymer-based LSC (dimension: 10 cm × 1.5 cm × 0.2 cm) comprising
PbS/CdS QDs in ambient light. Scale bar is 1 cm. (Ref. [58]. Copyright 2016, Wiley-VCH)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 297

Fig. 7 (a) Absorption and PL spectra of PbS/CdS in solution and a polymer matrix. (b) Opti-
cal/quantum efficiency of PbS/CdS QD-embedded LSCs with different G factors. (c) Photograph
of an LSC device (scale bar: 1 cm). (d) Scheme of the experimental setup for exciting and taking
the photograph of an illuminating LSC device (region of q angle is the escape cone of emitted
light). (e) Photograph of an illuminating LSC, captured by a NIR camera with a 780-nm long-pass
filter. (Ref. [90]. Copyright 2017, The Royal Society of Chemistry)

area semitransparent PbS/CdS QD-based LSC (10 × 1.5 × 0.2 cm3 ) was prepared
(Fig. 6e). The ηopt shows an exponential decrease with increasing side length or
geometric factor (G) of the LSC (Fig. 6c, d). The semitransparent PbS/CdS QD-
based LSC with G of 10 shows the ηopt of 6.1%, which is over 15 times higher than
that of PbS QD-based LSC (G = 4.2) with similar concentration of QDs (~20 μM).
Similarly, Tan et al. [90] fabricated NIR LSCs using ultrasmall PbS/CdS core/shell
QDs with a large Stokes shift (0.36 eV) and good stability (no absorption and PL
spectra change before and after QDs transfer into the polymer matrix) (Fig. 7). The
LSC can absorb the sunlight and reemit photons in the 700–1100-nm wavelength
range, matching well with the absorption range of the Si PV cell (Fig. 7a). The ηopt
298 G. Liu et al.

of the core/shell PbS/CdS QD-based LSC with QD concentration of 24 μM is ~4%


at G factor of 10 and decreases to ~1.2% at G factor of 50 (10 cm in length). The
optical efficiency decreases with the increasing length of the LSC mainly due to
the small overlap between the absorption and emission spectra, which leads to a
reabsorption phenomenon (Fig. 7b).

3.2 Visible CdSe/CdS Core/Shell QD-Based LSCs

Benefiting from the structured engineered optical properties in visible core/thick-


shell CdSe/CdS QDs, they have been widely used as high-quality emitters for
large-scale LSCs [91–94]. Meinardi et al. [60] used CdSe/CdS QDs to prepare
LSCs. Similar to PbS/CdS QDs, the Stokes shift of CdSe/CdS QDs can be
also enhanced by increasing the shell thickness. The thicker the shell layer, the
smaller the spectral overlap and the larger the Stokes shift. The intensities of
both PL and 835-nm scattered light exhibit the same distance (d)-dependence,
confirming that emission losses due to reabsorption are virtually nonexistent. Monte
Carlo simulation shows a 100-fold increase of emitted photons propagated to the
edges of a core/shell giant QD-based LSC, due to the reduced reabsorption and
increased QY, compared to the LSC comprising core-only QDs. The prepared LSC
(21.5 × 1.3 × 0.5 cm3 ) shows a remarkable quantum efficiency (QE, defined as
the ratio between the number of photons collected by the photodiode and the total
number of photons absorbed by the LSC) of 10.2% and ηopt of ~1% with negligible
reabsorption losses, which achieved an essential requirement for applications as
PV windows. Coropceanu and Bawendi fabricated the LSCs using core/thick-shell
CdSe/CdS QDs with very high QY (86%) [59], and the QDs were embedded in a
polymer matrix to achieve highly transparent LSCs (Fig. 8a). The EQE spectrum
exhibits an excellent agreement with the absorbance of the LSC, suggesting the
minimal energy dependence of the QY (Fig. 8b). The LSC exhibits optical efficiency
as high as 48% at 400 nm. With the wavelength increases, the optical efficiency of
the LSCs decreases, mainly due to the reduction of absorption at longer wavelengths
(Fig. 8b, c). Compared to Meinardi’s device, the significantly higher efficiency
of CdSe/CdS QD-based LSCs by Coropceanu and Bawendi is mainly due to the
smaller device area (2 × 2 × 0.2 cm2 vs. 21.5 × 1.3 × 0.5 cm3 ) and the higher QY
(86% vs. 40%), which leads to less reabsorption energy losses.
In the simplest planner LSC architecture, the emitted photons propagated in
random directions which implies that a number of photons strike the interfaces
within an incident angle smaller than the critical cone, leading to the escape and
scattering losses. Several adjustments have been exploited to reduce the escape loss
in order to achieve higher efficiency, such as the use of diffuse or specular mirror
[92–95]. Bronstein et al. [16] developed a strategy to couple a photonic mirror
with CdSe/CdS QD-based LSC to form an optical cavity where emitted photons
were trapped (Fig. 9). In the design, a tuned wavelength-selective mirror transmits
blue light and reflects red emissions that trap the PL inside the cavity, increasing
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 299

Fig. 8 (a) CdSe/CdS QD-based LSCs in ambient light (left) and under UV illumination (right)
with the edges clear (top) and blocked by carbon paint (bottom). (b) External quantum efficiency
(EQE, defined as the ratio of the number of emitted photons and the total number of incident
photons) spectra of the LSC prototype. (c) Optical efficiency of the LSC. (d) Distribution of
outcomes for photons incident on the LSC from the Monte Carlo simulation. (Ref. [59]. Copyright
2014, American Chemical Society)

the intensity of emitted light inside the cavity (Fig. 9a). Subsequently, they used
the core/thick-shell CdSe/CdS QDs with QY of 68% to prepare a microcell-LSC
integrated with a photonic mirror and a trench-shaped diffuse trench reflector, which
increases the collection efficiency of red photons, mitigates the escape and scattering
losses, and improves the power output of red photons (Fig. 9b, c). Connell et al.
[94] also used core/thick-shell CdSe/CdS QDs as fluorophore to prepare LSCs with
reflectors. They explored the design of the spectrally selective top mirror on the
performance of the LSCs and found that the mirrors designed to trap emitted light
result in higher performance than mirrors designed for sunlight transmission. When
the QY is higher, the concentration of the fluorophores is lower, the lateral size is
larger, and the overlap between the absorption and emission spectra is lower. The
integration of mirrors and LSCs can trap both the incident and emitted light into the
LSCs and enhance the optical efficiency of the LSCs due to the increasing photons
collected by the PV cells. However, the application of LSCs in BIPVs is limited
because of the lack of transparency.
300 G. Liu et al.

Fig. 9 (a) Graphic showing a typical transmission electron micrograph and schematic of giant
CdSe/CdS QDs, incorporated into a traditional LSC (open top) and the luminescent concentrator
cavity (with mirror). The black rectangle is a PV cell. (b) J-V characteristics of the LSC device
with and without the photonic mirror. (c) Experimental and simulated photon concentration ratios
at different optical densities of QD, with a G of 61. (Ref. [16]. Copyright 2015, American Chemical
Society)

Another LSC architecture is the multiple LSCs, including tandem structure and
sandwich structure [21, 23–24, 33, 45, 96–99]. In the tandem configuration, each
layer absorbs the solar spectrum selectively to improve the optical efficiency of
LSCs attributed to the increased absorption efficiency of the devices [21, 97].
Recently, Liu et al. [98] fabricated the tandem LSC based on carbon dots (C-dots)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 301

and core/thick-shell CdSe/CdS QDs to improve the optical efficiency of the LSCs
(Fig. 10a). First, they synthesized C-dots and CdSe/CdS QDs by a hydrothermal
method and SILAR approach, respectively, with high QY (40% for C-dots and
50% for CdSe/CdS QDs). To experimentally investigate the optical efficiency of
the LSCs, they fabricated tandem LSC (10 × 10 cm2 ) comprising C-dots (top
layer) and CdSe/CdS QDs (bottom layer). The C-dot-based LSC was prepared
by spin-coating a C-dot/polyvinylpyrrolidone (PVP) polymer on a glass substrate,
and the CdSe/CdS QD-based LSC was prepared by embedding the giant QDs in
a poly(lauryl methacrylate-co-ethylene glycol dimethacrylate) (PLMA-co-EGDM)
polymer matrix, respectively. No significant changes in the absorption and PL peak
position were observed in the investigated QDs in solution and embedded in the
polymer (Fig. 10b). The tandem device shows a ηopt of 1.4% under simulated one
sun illumination (100 mW/cm2 ), which is improved 16% of the optical efficiency
compared to the single-layer CdSe/CdS QD-based LSCs. The enhancement of the
optical efficiency for the tandem structure is mainly due to the increased absorption
of sunlight by both C-dots and QDs and the decreased escape loss, as the escaped
emission from the top layer can be partially reabsorbed by the bottom layer. In
addition, the tandem structure largely enhances the photostability of the CdSe/CdS
QD-based LSC due to the protection of the C-dot layer. After 70 h of continuous
UV illumination, the PL intensity of the CdSe/CdS QD-based LSC with the C-dot
protective layer maintains 75% of its initial value compared to 43% of the LSC
without the C-dot layer (Fig. 10c).
Although the tandem LSCs mentioned above can improve the optical efficiency
and photostability of the device, the long-term stability of the LSC under surround-
ing environments (light, moisture, etc.) still needs to increase due to the sensitivity
of the QDs. Liu et al. [45] further fabricated the sandwich structured (or laminated)
LSC. CdSe/CdS QD/polymer film was laminated between two sheets of glass (Fig.
10d). The structure can isolate the QDs from the ambient environments so as to
improve the long-term stability of the device. After 3 months, the efficiency of
the sandwich structure LSC maintains 85% of its initial efficiency value, while
for the single-layer device, it drops more than 25%. The PL intensity of the
sandwich structure LSC increases compared to that of the single-film LSC, which
demonstrates the enhancement of optical efficiency of sandwich structure LSC (Fig.
10e). Compared to the single-layer LSCs, the improvement of optical efficiency for
sandwich structure LSCs is mainly attributed to the lower G factor, less photon
escape, and reduced reabsorption loss due to the propagation of emitted light in
the glass (Fig. 10f). The optimized optical efficiency of the large-scale sandwich
structure LSCs (10 × 10 cm2 ) based on CdSe/CdS QDs is 2.95%, increasing
75% compared to the single-layer LSCs. The sandwich structure LSCs provide a
promising pathway to achieve power generation in different application places, such
as BIPVs and plastic greenhouses.
302 G. Liu et al.

Fig. 10 (a) Photograph (left) and schematic diagram (right) of the tandem LSC comprising C-
dots and CdSe/CdS QDs. (b) Absorption and PL spectra of the C-dots (top) and CdSe/CdS QDs
(bottom) in solution and LSC, respectively. (c) PL spectra of the CdSe/CdS QD-based LSCs before
and after 70 h of UV illumination with or without the C-dot protective layer (top) and the absorption
and PL spectra of C-dot-based LSC before and after 70 h of illumination under ambient conditions
(bottom) [98]. Copyright 2018, The Royal Society of Chemistry. (d) Photograph (left and right)
and schematic diagram (middle) of the sandwich structure LSC based on CdSe/CdS QDs. (e)
Schematic diagram of PL spectra of single and sandwich structure LSCs. (f) Experimental and
calculated optical efficiency of different LSCs based on CdSe/CdS QDs. “G” and “F” represent the
glass and QD film, respectively. (Ref. [45]. Copyright 2019, Elsevier Ltd.)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 303

3.3 Eco-Friendly InP/ZnO Core/Shell QD-Based LSCs

The above mentioned core/shell QDs in LSCs usually contain toxic heavy metal
elements, such as Pb and Cd, which is not eco-friendly and limits their practical use.
Heavy metal free QDs including InP/ZnO, CuInS2 /ZnS, AgInS2 /ZnS, etc. are more
suitable for optoelectronic devices due to their eco-friendly and superior optical
properties [25, 33, 48]. InP QDs are one of the typical binary eco-friendly QDs
with a bulk bandgap of 1.35 eV, which could tune the emission wavelength from the
visible to the NIR region by changing their size [100]. Formation of a core/shell
structure can largely improve the QY and photostability of InP QDs. The QY
exceeds 70% after coating the InP core with appropriate shell materials, such as
InP/ZnS, InP/ZnSeS, InP/GaP/ZnS, etc. [101–103] Thus, they are very promising
materials in optoelectronic applications.
Karatum et al. [104] synthesized type II InP/ZnO QDs by adjusting the shell
coverage for optimum in-film QE. They found that even though InP/2ZnO (i.e., InP
core surrounded by 2 monolayers of ZnO shell) achieves lower efficiency loss due
to large Stokes shift, InP/1ZnO (i.e., InP core surrounded by 1 monolayer of ZnO
shell) shows higher in-film efficiency because of higher QY. This study presents an
opportunity for the ability of engineering nontoxic QDs for minimizing solid-state
efficiency loss by forming shells owing to type II band alignment. Sadeghi et al. used
binary nontoxic InP/ZnO core/shell QDs to fabricate LSCs [25]. They investigated
the shell thickness effect in order to optimize the Stokes shift of the core/shell QDs.
The emission peak of the QDs redshifts with increasing the ZnO shell thickness
due to the type II band alignment, in which the electron tends to delocalize toward
the ZnO shell, while the hole was confined in the core (Fig. 11a). As a result, with
the shell thickness increased, the spectral overlap between the absorption and PL
emission decreased, indicating the reduction of reabsorption loss. Figure 11b shows
the schematic and photographs of the InP/ZnO QD-based LSC. The LSC under
ambient light exhibits a good transparency which could appropriately reduce the
incident sunlight and transmit it to the inside of the rooms. Under UV light, the PL
emitted by the QDs was especially visible in the edges. Moreover, the LSC shows
great structural flexibility which indicates that the QD-polymer slabs can be useful
to be integrated on curved surfaces and to capture the sunlight all day.
Figure 11c shows the dependence of the output optical intensity of the InP/2ZnO
and InP/5ZnO (i.e., InP core surrounded by 5 monolayers of ZnO shell) QD-
based LSCs versus the illuminated area, which could illustrate the effect of the
Stokes shift on the LSC performance. The InP/2ZnO-based LSC shows a near-
exponential behavior for the output optical intensity as the coverage area increased
due to the reabsorption loss of InP/2ZnO, while the InP/5ZnO-based LSC exhibits
a linear response of optical intensity with the increasing illuminated area due to the
negligible reabsorption loss (Fig. 11c). Furthermore, thanks to the low reabsorption
loss in the LSC based on InP/5ZnO QDs, the optical output decay was mainly due
to the scattering loss (Fig. 11d). Figure 11e shows the simulated optical efficiency
of the LSCs with a different QE and G factor. The simulated optical efficiency of
304 G. Liu et al.

Fig. 11 (a) Absorption and PL spectra of the InP/ZnO QDs with increasing shell layers (0, 2, 5
monolayers) in hexane. (b) Schematic and photographs of InP/ZnO QD-LSC under ambient light
(middle) and UV irradiation (right). (c) Dependence of the optical output intensity by varying the
illuminated area. (d) Optical output intensity of polydimethylsiloxane (PDMS) and InP/5ZnO QD-
LSC. (e) Simulations of optical efficiency for the QE levels of QDs as 16.2, 30, 60, and 100%.
(Ref. [25]. Copyright 2018, American Chemical Society)

LSCs is improved with higher QE of QDs and lower G factor. The optical efficiency
of the InP/5ZnO QD-based LSC with QD concentration of 0.05 wt.% reached 1.4%
at G factor of 5 and decreased to 0.225% at G factor of 30.

3.4 Cu-Based Ternary or Quaternary Eco-Friendly Core/Shell


QD-Based LSCs

Ternary eco-friendly core/shell QDs are emerging alternatives for heavy metal-
based toxic QDs due to their environmental-friendly, broad light absorption and
tunable emission by the size effect, composition ratio, and the surface ion situations
[33, 48, 77, 105]. Li et al. [106] fabricated the heavy metal free CuInS2 /ZnS QD-
based LSC with an optical efficiency of 26.5% (LSC size: 2.2 × 2.2 × 0.3 cm3 ).
The one-pot method was used to synthesize the core/shell QDs with a QY of 81%
and large Stokes shift of 150 nm. The as-prepared CuInS2 /ZnS QDs are able to
convert short-wavelength light (<450 nm) into long-wavelength light (~550 nm),
which could be absorbed by the crystalline-silicon (c-Si) PV cells more efficiently
(Fig. 12a). By integrating the c-Si PV cell with the pure polymethyl methacrylate
(PMMA) and CuInS2 /ZnS QDs-LSC, the PCE was measured (Fig. 12b). The PCE
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 305

of the CuInS2 /ZnS QD-LSC-PV device increases more than threefold from 2.73%
for the pure PMMA-PV device to 8.71% for the QD-LSC-PV device. For the same-
size c-Si PV cell, the device can harvest 4.91-fold solar energy with the assistance of
the CuInS2 /ZnS QD-based LSC, which is able to reduce the cost of PV electricity.
Bergren et al. [33] used high QY (>90%) and NIR-emitting CuInS2 /ZnS QDs
as emitters to fabricate the laminated glass LSCs. The absorption and emission
spectra of the CuInS2 /ZnS QDs show a very large Stokes shift (>550 meV) with
band edge at ~2.0 eV and emission peak at ~1.44 eV (862 nm) (Fig. 12c). The PL
emission spectrum is well matched with the peak EQE of c-Si solar cells (850–
950 nm) to achieve high PCE. The large Stokes shift has practical implications
for the high-efficiency LSCs as the QDs have little reabsorption and thus minimal
intrinsic optical losses. By incorporating the high QY NIR QDs into the polymer
interlayer between two sheets of glass, a laminated LSC with high optical efficiency
was achieved (Fig. 12d). The I-V curves and respective QE curves of the LSC with
a mirrored background and a black background show that the LSC with a mirror
substrate exhibits better performance than that with a black background (Fig. 12e, f).
The as-fabricated laminated LSC based on NIR CuInS2 /ZnS QDs reaches an optical
efficiency of 8.1% and PCE of 2.94% and 2.18% (with and without a reflective
substrate) for a 10 × 10 cm2 device. Due to its high efficiency and high capability
of integration into buildings, the laminated LSCs have the potential to realize net-
zero power consumption for BIPVs.
Beside the binary and ternary semiconductors, the ternary alloyed heavy metal
free QDs (CuInSex S2-x , Zn and Al co-doped CuInS2 ) have also been used as
emitters to fabricate LSCs [46, 77]. Meinardi et al. [77] used NIR-emitting
CuInSex S2-x /ZnS (CISeS/ZnS) QDs as high-quality fluorophores to fabricate LSCs
(12 × 12 × 0.3 cm3 ). Both the absorption and PL emission spectra of the
QDs in the polymer are identical to those in the solution, which reveals that the
spectral and dynamical properties of CISeS/ZnS QDs are unaffected by the radical
polymerization procedure. The core/shell QDs exhibit a large Stokes shift (up to
530 nm), but the overlap between the absorption and PL emission spectra is still
clear, suggesting a partial energy loss due to reabsorption. Due to the scattering at
optical imperfections, the PL intensity decreased with increasing distance from the
edge. The normalized PL spectra slightly changed with varying distance, indicating
that the reabsorption loss is very small. The PL intensity drops with increasing
distance, due to the combined effect of light escaping from the waveguide and
reabsorption by the QDs and the polymer matrix [77].
The QD-based LSC usually only comprises pure QDs without other impurities.
Liu et al. [107] fabricated CuInS2 /ZnS QD-based LSC with SiO2 particles as the
scattering center to realize highly efficient scattering-enhanced LSC devices. The
schematic and photographs for scattering-enhanced LSC are shown in Fig. 13e.
SiO2 particles could change the direction of incident sunlight, which would increase
the probability of light captured and absorbed by QDs. Moreover, the emission
light reemitted from QDs will also be redirected by the SiO2 particles, which will
be more likely guided to the PV cells installed at the edge of LSCs. Since SiO2
particles induced scattering effect, they achieved a PCE of 4.2%, which showed
306 G. Liu et al.

Fig. 12 (a) The relationships of AM 1.5 G spectrum, c-Si PV cell-responsive spectrum, and the
absorption as well as emission spectra of CuInS2 /ZnS QDs. (b) Photographs of the LSC prototype
(22 mm × 22 mm × 3 mm) non-containing and containing CuInS2 /ZnS QDs under UV light
(365 nm) and different schematic of LSC-PV devices [106]. Copyright 2015, Springer Nature.
(c) Absorption and PL spectra for NIR-emitting CuInS2 /ZnS QDs in toluene. (d) Diagram of the
laminated glass LSC design. (e) National Renewable Energy Laboratory (NREL)-certified I–V
curves measured under AM1.5 illumination in a solar simulator for LSC with an absorbing black
background (dashed line) and a mirrored background behind the window (solid line), simulating
a low-e coating. (f) Normalized QE measured and averaged from four locations on the device for
both a black and mirrored background. The inset shows a photo of the LSC device. (Ref. [33].
Copyright 2018, American Chemical Society)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 307

Fig. 13 (a) Schematic for scattering-enhanced LSC. (b) The photographs of pure PMMA light
guide bulk and scattering-enhanced LSC under daylight (left) and scattering-enhanced LSC
contained SiO2 (5 μm) under UV (365 nm) light (right). (Ref. [107])

an improvement of 60.3% compared with the pure QD-based LSC without SiO2
particles, while the presence of a SiO2 nanoparticle will affect the transparency of
the LSCs.

4 Conclusions and Perspectives

In this chapter, we introduce various types of core/shell QDs and their application in
LSCs. In the LSCs, the energy loss is mainly due to the reabsorption induced by the
overlap between the absorption and emission of the fluorescent materials and their
QY which is typically lower than 100%. In the core/shell QDs, by tuning the core
size and shell thickness, one can tune the band alignment, further tuning the optical
308 G. Liu et al.

properties of QDs. Although high-quality fluorophores (e.g., CdSe/CdS, PbS/CdS


QDs) have been used to fabricate high-efficiency LSCs with excellent stability, these
QDs contain toxic Cd and Pb. Compared to heavy metal QDs, heavy-metal-free QDs
become more promising candidates for potential applications in LSCs due to their
low toxicity. The binary and ternary core/shell eco-friendly QDs (e.g., InP/ZnO,
CuInS2 /ZnS QDs) have been used for LSCs benefitting from their size-dependent
optical properties.
Beside the tremendous effort in exploring the structure, properties, and appli-
cations of core/shell QDs in LSCs, there are still several critical limitations and
challenges that need to be resolved: (i) The optical properties of the core/shell QDs
still need to be optimized. And the effective synthetic approaches for core/shell
QDs with large Stokes shift, high QY, and long-term stability should be developed.
(ii) Although the eco-friendly core/shell QDs are nontoxic materials, the organic
precursors are still hazardous. Eco-friendly core/shell QDs with green synthesis
process need to be developed. (iii) The optical efficiency of the large-scale LSCs
is still low, which is difficult to realize further commercialization. (iv) The size
of the QD-based LSCs is still small and should be enlarged while simultaneously
maintaining their efficiency.
The LSCs can be applied in buildings as part of the roofs, facades, or windows,
to form net-zero energy buildings. The LSCs can be also used in solar noise
barriers which serve as electricity-generating device and noise barriers. Besides
the application in BIPVs, the LSCs can be used as solar concentrated chemical
photoreactors as well [18, 108]. Specially, the LSCs can convert the sunlight into
selected emission light to match the active spectrum range of a photocatalyst or
photosensitizer. To date, the challenges related to the applications of LSCs are the
cost, durability and efficiency. In addition, the replacement of toxic materials with
green and earth-abundant materials for fabrication of LSCs is important to address
environmental sustainability. With further optimizations of both the fluorescent
materials and devices’ structure, highly efficient, long-term stable and large-area
LSCs based on core/shell QDs need to be achieved in order to realize a net-zero
power consumption of future buildings.

Acknowledgments H. Zhao acknowledges the start funding support from Qingdao University and
the funding from the Natural Science Foundation of Shandong Province (ZR2018MB001).

References

1. Zhang, S., Wu, S., Chen, W., Zhu, H., Xiong, Z., Yang, Z., Chen, C., Chen, R., Han, L.,
Chen, W.: Solvent engineering for efficient inverted perovskite solar cells based on inorganic
CsPbI2 Br light absorber. Mater Today Energy. 8, 125–133 (2018)
2. Zhao, H., Rosei, F.: Colloidal quantum dots for solar technologies. Chem. 3(2), 229–258
(2017)
3. Giustino, F., Snaith, H.J.: Toward lead-free perovskite solar cells. ACS Energy Lett. 1(6),
1233–1240 (2016)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 309

4. Song, S., Kang, G., Pyeon, L., Lim, C., Lee, G.-Y., Park, T., Choi, J.: Systematically optimized
bilayered electron transport layer for highly efficient planar perovskite solar cells (η =
21.1%). ACS Energy Lett. 2, 2667–2673 (2017)
5. Wu, L., Chen, S.Y., Fan, F.J., Zhuang, T.T., Dai, C.M., Yu, S.H.: Polytypic nanocrystals of
Cu-based ternary chalcogenides: colloidal synthesis and photoelectrochemical properties. J
Am Chem Soc. 138(17), 5576–5584 (2016)
6. Feng, H.P., Tang, L., Zeng, G.M., Zhou, Y., Deng, Y.C., Ren, X., Song, B., Liang, C., Wei,
M.Y., Yu, J.F.: Core-shell nanomaterials: applications in energy storage and conversion. Adv
Colloid Interface Sci. 267, 26–46 (2019)
7. Chen, D., Wang, A., Buntine, M.A., Jia, G.: Recent advances in zinc-containing colloidal
semiconductor nanocrystals for optoelectronic and energy conversion applications. Chem-
ElectroChem. 6(18), 4709–4724 (2019)
8. NRE, Best research-cell efficiency chart, plotted from 1976 to the present. 20191106. https://
www.nrel.gov/pv/cell-efficiency.html
9. Green, M.A.: Third generation photovoltaics: advanced solar energy conversion. Springer,
Berlin (2006)
10. Wehrspohn, R.B., Gombert, A., Gombert, A., Heile, I., Wüllner, J., Gerstmaier, T., van Riesen,
S., Gerster, E., Röttger, M., Lerchenmüller, H.: Recent progress in concentrator photovoltaics.
Photonic Solar Energ Syst III. 7725, 772508 (2010)
11. Kirkpatrick, D., Eisenstadt, E., Haspert, A.: DARPA pushes for 50% efficient photovoltaics to
power soldiers’ small tools. SPIE-The International Society for Optical Engineering (2006)
12. Weber, W.H., Lambe, J.: Luminescent greenhouse collector for solar radiation. Appl Optics.
15(10), 2299–2300 (1976)
13. Currie, M.J., Mapel, J.K., Heidel, T.D., Goffri, S., Baldo, M.A.: High-efficiency organic solar
concentrators for photovoltaics. Science. 321(5886), 226–228 (2008)
14. Desmet, L., Ras, A.J.M., de Boer, D.K.G., Debije, M.G.: Monocrystalline silicon photovoltaic
luminescent solar concentrator with 4.2% power conversion efficiency. Opt Lett. 37, 3087–
3089 (2012)
15. Corrado, C., Leow, S.W., Osborn, M., Chan, E., Balaban, B., Carter, S.A.: Optimization of
gain and energy conversion efficiency using front-facing photovoltaic cell luminescent solar
concentrator design. Sol Energ Mat Sol C. 111, 74–81 (2013)
16. Bronstein, N.D., Yao, Y., Xu, L., O’Brien, E., Powers, A.S., Ferry, V.E., Alivisatos, A.P.,
Nuzzo, R.G.: Quantum dot luminescent concentrator cavity exhibiting 30-fold concentration.
ACS Photonics. 2(11), 1576–1583 (2015)
17. Klimov, V.I., Baker, T.A., Lim, J., Velizhanin, K.A., McDaniel, H.: Quality factor of lumi-
nescent solar concentrators and practical concentration limits attainable with semiconductor
quantum dots. ACS Photonics. 3(6), 1138–1148 (2016)
18. Cambie, D., Zhao, F., Hessel, V., Debije, M.G., Noel, T.: A leaf-inspired luminescent solar
concentrator for energy-efficient continuous-flow photochemistry. Angew Chem Int Ed Engl.
56(4), 1050–1054 (2017)
19. Brennan, L.J., Purcell-Milton, F., McKenna, B., Watson Trystan, M., Gun’ko, Y.K., Evans,
R.C.: Large area quantum dot luminescent solar concentrators for use with dye-sensitised
solar cells. J Mater Chem A. 6(6), 2671–2680 (2018)
20. Talite, M.J., Huang, H.Y., Wu, Y.H., Sena, P.G., Cai, K.B., Lin, T.N., Shen, J.L., Chou,
W.C., Yuan, C.T.: Greener luminescent solar concentrators with high loading contents based
on in situ cross-linked carbon nanodots for enhancing solar energy harvesting and resisting
concentration-induced quenching. ACS Appl Mater Interfaces. 10(40), 34184–34192 (2018)
21. Wu, K., Li, H., Klimov, V.I.: Tandem luminescent solar concentrators based on engineered
quantum dots. Nat Photonics. 12(2), 105–110 (2018)
22. AbouElhamd, A.R., Al-Sallal, K.A., Hassan, A.: Review of core/shell quantum dots technol-
ogy integrated into building’s glazing. Energies. 12(6), 1058 (2019)
23. Sadeghi, S., Melikov, R., Bahmani Jalali, H., Karatum, O., Srivastava, S.B., Conkar, D., Firat-
Karalar, E.N., Nizamoglu, S.: Ecofriendly and efficient luminescent solar concentrators based
on fluorescent proteins. ACS Appl Mater Interfaces. 11(9), 8710–8716 (2019)
310 G. Liu et al.

24. Sol, J., Dehm, V., Hecht, R., Wurthner, F., Schenning, A., Debije, M.G.: Temperature-
responsive luminescent solar concentrators: tuning energy transfer in a liquid crystalline
matrix. Angew Chem Int Ed Engl. 57(4), 1030–1033 (2018)
25. Sadeghi, S., Bahmani Jalali, H., Melikov, R., Ganesh Kumar, B., Mohammadi Aria, M.,
Ow-Yang, C.W., Nizamoglu, S.: Stokes-shift-engineered indium phosphide quantum dots for
efficient luminescent solar concentrators. ACS Appl Mater Interfaces. 10(15), 12975–12982
(2018)
26. Luo, X., Ding, T., Liu, X., Liu, Y., Wu, K.: Quantum cutting luminescent solar concentrators
using ytterbium doped perovskite nanocrystals. Nano Lett. 19, 338–341 (2018)
27. Sharma, M., Gungor, K., Yeltik, A., Olutas, M., Guzelturk, B., Kelestemur, Y., Erdem,
T., Delikanli, S., McBride, J.R., Demir, H.V.: Near-unity emitting copper-doped colloidal
semiconductor quantum wells for luminescent solar concentrators. Adv Mater. 29(30),
1700821 (2017)
28. Mateen, F., Oh, H., Jung, W., Lee, S.Y., Kikuchi, H., Hong, S.-K.: Polymer dispersed liquid
crystal device with integrated luminescent solar concentrator. Liq Cryst. 45(4), 498–506
(2017)
29. Gutierrez, G.D., Coropceanu, I., Bawendi, M.G., Swager, T.M.: A low reabsorbing lumi-
nescent solar concentrator employing pi-conjugated polymers. Adv Mater. 28(3), 497–501
(2016)
30. Zhang, J., Wang, M., Zhang, Y., He, H., Xie, W., Yang, M., Ding, J., Bao, J., Sun, S., Gao, C.:
Optimization of large-size glass laminated luminescent solar concentrators. Sol Energy. 117,
260–267 (2015)
31. Zhao, H., Benetti, D., Jin, L., Zhou, Y., Rosei, F., Vomiero, A.: Absorption enhancement in
“giant” core/alloyed-shell quantum dots for luminescent solar concentrator. Small. 12(38),
5354–5365 (2016)
32. Meinardi, F., Bruni, F., Brovelli, S.: Luminescent solar concentrators for building-integrated
photovoltaics. Nat Rev Mater. 2(12), 17072 (2017)
33. Bergren, M.R., Makarov, N.S., Ramasamy, K., Jackson, A., Guglielmetti, R., McDaniel, H.:
High-performance CuInS2 quantum dot laminated glass luminescent solar concentrators for
windows. ACS Energy Lett. 3(3), 520–525 (2018)
34. You, Y., Tong, X., Wang, W., Sun, J., Yu, P., Ji, H., Niu, X., Wang, Z.M.: Eco-friendly colloidal
quantum dot-based luminescent solar concentrators. Adv Sci. 6(9), 1801967 (2019)
35. Mazzaro, R., Gradone, A., Angeloni, S., Morselli, G., Cozzi, P.G., Romano, F., Vomiero, A.,
Ceroni, P.: Hybrid silicon nanocrystals for color-neutral and transparent luminescent solar
concentrators. ACS Photonics. 6(9), 2303–2311 (2019)
36. Zhou, Y., Zhao, H., Ma, D., Rosei, F.: Harnessing the properties of colloidal quantum dots in
luminescent solar concentrators. Chem Soc Rev. 47(15), 5866–5890 (2018)
37. Moraitis, P., Schropp, R.E.I., van Sark, W.G.J.H.M.: Nanoparticles for luminescent solar
concentrators – a review. Opt Mater. 84, 636–645 (2018)
38. Reinders, A.H.M.E., de la Grée, G. D., Papadopoulos, A., Rosemann, A., Debije, M. G., Cox,
M., Krumer, Z.: Leaf roof – designing luminescent solar concentrating PV roof tiles. 2016
IEEE 43rd Photovoltaic Specialists Conference (PVSC), 3447–3451 (2016)
39. van Sark, W., Moraitis, P., Aalberts, C., Drent, M., Grasso, T., L’Ortije, Y., Visschers, M.,
Westra, M., Plas, R., Planje, W.: The “electric mondrian” as a luminescent solar concentrator
demonstrator case study. Sol RRL. 1, 1600015 (2017)
40. Debije, M.G., Tzikas, C., Rajkumar, V.A., de Jong, M.M.: The solar noise barrier project:
2. The effect of street art on performance of a large scale luminescent solar concentrator
prototype. Renew Energy. 113, 1288–1292 (2017)
41. Kanellis, M., de Jong, M.M., Slooff, L., Debije, M.G.: The solar noise barrier project: 1.
Effect of incident light orientation on the performance of a large-scale luminescent solar
concentrator noise barrier. Renew Energy. 103, 647–652 (2017)
42. Batchelder, J.S., Zewail, A.H., Cole, T.: Luminescent solar concentrators 1- theory of
operation and techniques for performance. Appl Optics. 18(18), 3090–3110 (1979)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 311

43. Wilton, S.R., Fetterman, M.R., Low, J.J., You, G., Jiang, Z., Xu, J.: Monte Carlo study of PbSe
quantum dots as the fluorescent material in luminescent solar concentrators. Opt Express.
22(1), A35–A43 (2014)
44. Zhou, Y., Benetti, D., Tong, X., Jin, L., Wang, Z.M., Ma, D., Zhao, H., Rosei, F.: Colloidal
carbon dots based highly stable luminescent solar concentrators. Nano Energy. 44, 378–387
(2018)
45. Liu, G., Mazzaro, R., Wang, Y., Zhao, H., Vomiero, A.: High efficiency sandwich structure
luminescent solar concentrators based on colloidal quantum dots. Nano Energy. 60, 119–126
(2019)
46. Zhu, M., Li, Y., Tian, S., Xie, Y., Zhao, X., Gong, X.: Deep-red emitting zinc and aluminium
co-doped copper indium sulfide quantum dots for luminescent solar concentrators. J Colloid
Interface Sci. 534, 509–517 (2019)
47. Li, H., Wu, K., Lim, J., Song, H.-J., Klimov, V.I.: Doctor-blade deposition of quantum dots
onto standard window glass for low-loss large-area luminescent solar concentrators. Nat
Energy. 1, 16157 (2016)
48. Chen, W., Li, J., Liu, P., Liu, H., Xia, J., Li, S., Wang, D., Wu, D., Lu, W., Sun, X.W., Wang,
K.: Heavy metal free nanocrystals with near infrared emission applying in luminescent solar
concentrator. Solar RRL. 1(6), 1700041 (2017)
49. Zhao, H., Sun, R., Wang, Z., Fu, K., Hu, X., Zhang, Y.: Zero-dimensional perovskite
nanocrystals for efficient luminescent solar concentrators. Adv Funct Mater. 29(30), 1902262
(2019)
50. Mateen, F., Ali, M., Oh, H., Hong, S.-K.: Nitrogen-doped carbon quantum dot based
luminescent solar concentrator coupled with polymer dispersed liquid crystal device for smart
management of solar spectrum. Sol Energy. 178, 48–55 (2019)
51. Khan, A.H., Pinchetti, V., Tanghe, I., Dang, Z., Martín-García, B., Hens, Z., Van Thourhout,
D., Geiregat, P., Brovelli, S., Moreels, I.: Tunable and efficient red to near-infrared photolumi-
nescence by synergistic exploitation of core and surface silver doping of CdSe nanoplatelets.
Chem Mater. 31(4), 1450–1459 (2019)
52. El-Bashir, S.M.: Coumarin-doped PC/CdSSe/ZnS nanocomposite films: a reduced self-
absorption effect for luminescent solar concentrators. J Lumin. 206, 426–431 (2019)
53. Rowan, B.C., Wilson, L.R., Richards, B.S.: Advanced material concepts for luminescent solar
concentrators. IEEE J Sel Top Quant Electronics. 14(5), 1312–1322 (2008)
54. Liang, H., Zeng, Z., Li, Z., Xu, J., Chen, B., Zhao, H., Zhang, Q., Ming, H.: Fabrication
and amplification of rhodamine B-doped step-index polymer optical fiber. J Appl Polym Sci.
93(2), 681–685 (2004)
55. Dienel, T., Bauer, C., Dolamic, I., Brühwiler, D.: Spectral-based analysis of thin film
luminescent solar concentrators. Sol Energy. 84(8), 1366–1369 (2010)
56. Slooff, L.H., Bende, E.E., Burgers, A.R., Budel, T., Pravettoni, M., Kenny, R.P., Dunlop, E.D.,
Büchtemann, A.: A luminescent solar concentrator with 7.1% power conversion efficiency.
Phys Status Solidi – R. 2(6), 257–259 (2008)
57. Goetzberger, A., Greubel, W.: Solar energy conversion with fluorescent collectors. Appl Phys.
14, 123–139 (1977)
58. Zhou, Y., Benetti, D., Fan, Z., Zhao, H., Ma, D., Govorov, A.O., Vomiero, A., Rosei, F.: Near
infrared, highly efficient luminescent solar concentrators. Adv Energy Mater. 6(11), 1501913
(2016)
59. Coropceanu, I., Bawendi, M.G.: Core/shell quantum dot based luminescent solar concentra-
tors with reduced reabsorption and enhanced efficiency. Nano Lett. 14(7), 4097–4101 (2014)
60. Meinardi, F., Colombo, A., Velizhanin, K.A., Simonutti, R., Lorenzon, M., Beverina,
L., Viswanatha, R., Klimov, V.I., Brovelli, S.: Large-area luminescent solar concentrators
based on ‘Stokes-shift-engineered’ nanocrystals in a mass-polymerized PMMA matrix. Nat
Photonics. 8(5), 392–399 (2014)
61. Tytus, M., Krasnyj, J., Jacak, W., Chuchmala, A., Donderowicz, W., Jacak, L.: Differences
between photoluminescence spectra of type-I and type-II quantum dots. J Phys Conf Ser.
104, 012011 (2008)
312 G. Liu et al.

62. Kim, S., Fisher, B., Eisler, H.-J., Bawendi, M.: Type-II quantum dots: CdTe/CdSe (core/shell)
and CdSe/ZnTe (core/shell) heterostructures. J Am Chem Soc. 125(38), 11466–11467 (2003)
63. Gheshlaghi, N., Pisheh, H.S., Karim, M.R., Malkoc, D., Ünlü, H.: Interfacial strain effect on
type-I and type-II core/shell quantum dots. Superlattice Microst. 97, 489–494 (2016)
64. Reiss, P., Protiere, M., Li, L.: Core/shell semiconductor nanocrystals. Small. 5(2), 154–168
(2009)
65. De Geyter, B., Justo, Y., Moreels, I., Lambert, K., Smet, P.F., Van Thourhout, D., Houtepen,
A.J., Grodzinska, D., de Mello Donega, C., Meijerink, A., Vanmaekelbergh, D., Hens, Z.:
The different nature of band edge absorption and emission in colloidal PbSe/CdSe core/shell
quantum dots. ACS Nano. 5(1), 58–66 (2010)
66. Vasudevan, D., Gaddam, R.R., Trinchi, A., Cole, I.: Core-shell quantum dots: properties and
applications. J Alloys Compd. 636, 395–404 (2015)
67. Dabbousi, B.O., Rodriguez-Viejo, J., Mikulec, F.V., Heine, J.R., Mattoussi, H., Ober, R.,
Jensen, K.F., Bawendi, M.G.: (CdSe)ZnS core-shell quantum dots: synthesis and charac-
terization of a size series of highly luminescent nanocrystallites. J Phys Chem B. 101(46),
9463–9475 (1997)
68. Zhao, H., Chaker, M., Ma, D.: Effect of CdS shell thickness on the optical properties of water-
soluble, amphiphilic polymer-encapsulated PbS/CdS core/shell quantum dots. J Mater Chem.
21(43), 17483 (2011)
69. Pal, B.N., Ghosh, Y., Brovelli, S., Laocharoensuk, R., Klimov, V.I., Hollingsworth, J.A.,
Htoon, H.: ‘Giant’ CdSe/CdS core/shell nanocrystal quantum dots as efficient electrolumi-
nescent materials: strong influence of shell thickness on light-emitting diode performance.
Nano Lett. 12(1), 331–336 (2012)
70. Zhu, J., Wang, S.-N., Li, J.-J., Zhao, J.-W.: The effect of core size on the fluorescence emission
properties of CdTe@CdS core@shell quantum dots. J Lumin. 199, 216–224 (2018)
71. Itzhakov, S., Shen, H., Buhbut, S., Lin, H., Oron, D.: Type-II quantum-dot-sensitized solar
cell spanning the visible and near-infrared spectrum. J Phys Chem C. 117(43), 22203–22210
(2013)
72. Verma, S., Kaniyankandy, S., Ghosh, H.: Charge separation by indirect bandgap transitions
in CdS/ZnSe type-II core/shell quantum dots. J Phys Chem C. 117, 10901–10908 (2013)
73. Park, Y.S., Bae, W.K., Padilha, L.A., Pietryga, J.M., Klimov, V.I.: Effect of the core/shell
interface on auger recombination evaluated by single-quantum-dot spectroscopy. Nano Lett.
14(2), 396–402 (2014)
74. Selopal, G.S., Zhao, H., Tong, X., Benetti, D., Navarro-Pardo, F., Zhou, Y., Barba, D., Vidal,
F., Wang, Z.M., Rosei, F.: Highly stable colloidal “giant” quantum dots sensitized solar cells.
Adv Funct Mater. 27(30), 1701468 (2017)
75. Tong, X., Kong, X.-T., Zhou, Y., Navarro-Pardo, F., Selopal, G.S., Sun, S., Govorov, A.O.,
Zhao, H., Wang, Z.M., Rosei, F.: Near-infrared, heavy metal-free colloidal “giant” core/shell
quantum dots. Adv Energy Mater. 8(2), 1701432 (2018)
76. Brovelli, S., Schaller, R.D., Crooker, S.A., Garcia-Santamaria, F., Chen, Y., Viswanatha, R.,
Hollingsworth, J.A., Htoon, H., Klimov, V.I.: Nano-engineered electron-hole exchange inter-
action controls exciton dynamics in core-shell semiconductor nanocrystals. Nat Commun. 2,
280 (2011)
77. Meinardi, F., McDaniel, H., Carulli, F., Colombo, A., Velizhanin, K.A., Makarov, N.S.,
Simonutti, R., Klimov, V.I., Brovelli, S.: Highly efficient large-area colourless luminescent
solar concentrators using heavy-metal-free colloidal quantum dots. Nat Nanotechnol. 10(10),
878–885 (2015)
78. Tong, X., Zhou, Y., Jin, L., Basu, K., Adhikari, R., Selopal, G.S., Tong, X., Zhao, H., Sun, S.,
Vomiero, A., Wang, Z.M., Rosei, F.: Heavy metal-free, near-infrared colloidal quantum dots
for efficient photoelectrochemical hydrogen generation. Nano Energy. 31, 441–449 (2017)
79. Erickson, C.S., Bradshaw, L.R., McDowall, S., Gilbertson, J.D., Gamelin, D.R., Patrick, D.L.:
Zero-reabsorption doped-nanocrystal luminescent solar concentrators. ACS Nano. 8, 3461–
3467 (2014)
Core/Shell Quantum-Dot-Based Luminescent Solar Concentrators 313

80. Bradshaw, L.R., Knowles, K.E., McDowall, S., Gamelin, D.R.: Nanocrystals for luminescent
solar concentrators. Nano Lett. 15(2), 1315–1323 (2015)
81. Zhou, J., Zhu, M., Meng, R., Qin, H., Peng, X.: Ideal CdSe/CdS core/shell nanocrystals
enabled by entropic ligands and their core size-, shell thickness-, and ligand-dependent
photoluminescence properties. J Am Chem Soc. 139(46), 16556–16567 (2017)
82. Ghosh, Y., Mangum, B.D., Casson, J.L., Williams, D.J., Htoon, H., Hollingsworth, J.A.: New
insights into the complexities of shell growth and the strong influence of particle volume in
nonblinking “giant” core/shell nanocrystal quantum dots. J Am Chem Soc. 134(23), 9634–
9643 (2012)
83. Michalska, M., Aboulaich, A., Medjahdi, G., Mahiou, R., Jurga, S., Schneider, R.: Amine
ligands control of the optical properties and the shape of thermally grown core/shell
CuInS2 /ZnS quantum dots. J Alloys Compd. 645, 184–192 (2015)
84. Hanifi, D.A., Bronstein, N.D., Koscher, B.A., Nett, Z., Swabeck, J.K., Takano, K.,
Schwartzberg, A.M., Maserati, L., Vandewal, K., van de Burgt, Y., Salleo, A., Alivisatos, A.P.:
Redefining near-unity luminescence in quantum dots with photothermal threshold quantum
yield. Science. 363, 1199–1202 (2019)
85. Zhao, H., Chaker, M., Wu, N., Ma, D.: Towards controlled synthesis and better understanding
of highly luminescent PbS/CdS core/shell quantum dots. J Mater Chem. 21(24), 8898 (2011)
86. Huang, B., Yang, H., Zhang, L., Yuan, Y., Cui, Y., Zhang, J.: Effect of surface/interfacial
defects on photo-stability of thick-shell CdZnSeS/ZnS quantum dots. Nanoscale. 10(38),
18331–18340 (2018)
87. Yang, X., Zhao, D., Leck, K.S., Tan, S.T., Tang, Y.X., Zhao, J., Demir, H.V., Sun, X.W.:
Full visible range covering InP/ZnS nanocrystals with high photometric performance and
their application to white quantum dot light-emitting diodes. Adv Mater. 24(30), 4180–4185
(2012)
88. Tong, X., Kong, X.T., Wang, C., Zhou, Y., Navarro-Pardo, F., Barba, D., Ma, D., Sun, S.,
Govorov, A.O., Zhao, H., Wang, Z.M., Rosei, F.: Optoelectronic properties in near-infrared
colloidal heterostructured pyramidal “giant” core/shell quantum dots. Adv Sci (Weinh). 5(8),
1800656 (2018)
89. Navarro-Pardo, F., Zhao, H., Wang, Z.M., Rosei, F.: Structure/property relations in “giant”
semiconductor nanocrystals: opportunities in photonics and electronics. Acc Chem Res.
51(3), 609–618 (2018)
90. Tan, L., Zhou, Y., Ren, F., Benetti, D., Yang, F., Zhao, H., Rosei, F., Chaker, M., Ma, D.:
Ultrasmall PbS quantum dots: a facile and greener synthetic route and their high performance
in luminescent solar concentrators. J Mater Chem A. 5(21), 10250–10260 (2017)
91. Chatten, A.J., Barnham, K.W.J., Buxton, B.F., Ekins-Daukes, N.J., Malik, M.A.: Proc. of 3rd
World Conf. on photovoltaic energy conversion. IEEE Osaka. 3, 2657 (2003)
92. Xu, L., Yao, Y., Bronstein, N.D., Li, L., Alivisatos, A.P., Nuzzo, R.G.: Enhanced photon col-
lection in luminescent solar concentrators with distributed bragg reflectors. ACS Photonics.
3, 278–285 (2016)
93. Connell, R., Ferry, V.E.: Integrating photonics with luminescent solar concentrators: optical
transport in the presence of photonic mirrors. J Phys Chem C. 120(37), 20991–20997 (2016)
94. Connell, R., Pinnell, C., Ferry, V.E.: Designing spectrally-selective mirrors for use in
luminescent solar concentrators. J Opt. 20(2), 024009 (2018)
95. Song, H.J., Jeong, B.G., Lim, J., Lee, D.C., Bae, W.K., Klimov, V.I.: Performance limits of
luminescent solar concentrators tested with seed/quantum-well quantum dots in a selective-
reflector-based optical cavity. Nano Lett. 18(1), 395–404 (2018)
96. Mateen, F., Oh, H., Jung, W., Binns, M., Hong, S.-K.: Metal nanoparticles based stack
structured plasmonic luminescent solar concentrator. Sol Energy. 155, 934–941 (2017)
97. Zhao, H., Benetti, D., Tong, X., Zhang, H., Zhou, Y., Liu, G., Ma, D., Sun, S., Wang, Z.M.,
Wang, Y., Rosei, F.: Efficient and stable tandem luminescent solar concentrators based on
carbon dots and perovskite quantum dots. Nano Energy. 50, 756–765 (2018)
314 G. Liu et al.

98. Liu, G., Zhao, H., Diao, F., Ling, Z., Wang, Y.: Stable tandem luminescent solar concentrators
based on CdSe/CdS quantum dots and carbon dots. J Mater Chem C. 6(37), 10059–10066
(2018)
99. Needell, D.R., Ilic, O., Bukowsky, C.R., Nett, Z., Xu, L., He, J., Bauser, H., Lee, B.G., Geisz,
J.F., Nuzzo, R.G., Alivisatos, A.P., Atwater, H.A.: Design criteria for micro-optical tandem
luminescent solar concentrators. IEEE J Photovolt. 8(6), 1560–1567 (2018)
100. Tamang, S., Lincheneau, C., Hermans, Y., Jeong, S., Reiss, P.: Chemistry of InP nanocrystal
syntheses. Chem Mater. 28, 2491–2506 (2016)
101. Li, L., Reiss, P.: One-pot synthesis of highly luminescent InP/ZnS nanocrystals without
precursor injection. J Am Chem Soc. 130, 11588–11589 (2008)
102. Lim, J., Bae, W.K., Lee, D., Nam, M.K., Jung, J., Lee, C., Char, K., Lee, S.: InP@ZnSeS,
core@composition gradient shell quantum dots with enhanced stability. Chem Mater. 23(20),
4459–4463 (2011)
103. Kim, S., Kim, T., Kang, M., Kwak, S.K., Yoo, T.W., Park, L.S., Yang, I., Hwang, S., Lee, J.E.,
Kim, S.K., Kim, S.W.: Highly luminescent InP/GaP/ZnS nanocrystals and their application
to white light-emitting diodes. J Am Chem Soc. 134(8), 3804–3809 (2012)
104. Karatum, O., Jalali, H.B., Sadeghi, S., Melikov, R., Srivastava, S.B., Nizamoglu, S.: Light-
emitting devices based on type-II InP/ZnO quantum dots. ACS Photonics. 6(4), 939–946
(2019)
105. Nagamine, G., Nunciaroni, H.B., McDaniel, H., Efros, A.L., de Brito Cruz, C.H., Padilha,
L.A.: Evidence of band-edge hole levels inversion in spherical CuInS2 quantum dots. Nano
Lett. 18(10), 6353–6359 (2018)
106. Li, C., Chen, W., Wu, D., Quan, D., Zhou, Z., Hao, J., Qin, J., Li, Y., He, Z., Wang, K.: Large
stokes shift and high efficiency luminescent solar concentrator incorporated with CuInS2 /ZnS
quantum dots. Sci Rep. 5, 17777 (2015)
107. Liu, H., Li, S., Chen, W., Wang, D., Li, C., Wu, D., Hao, J., Zhou, Z., Wang, X., Wang,
K.: Scattering enhanced quantum dots based luminescent solar concentrators by silica
microparticles. Sol Energ Mater Sol C. 179, 380–385 (2018)
108. Liu, G., Sun, B., Li, H., Wang, Y., Zhao, H.: Integration of photoelectrochemical devices
and luminescent solar concentrators based on giant quantum dots for highly stable hydrogen
generation. J Mater Chem A. 7, 18529–18537 (2019)
Index

A Conduction band (CB), 259, 291


photon to current conversion efficiency Conduction band minimum (CBM), 32
(APCE), 231 Core-multishell, see Ternary chalcogenides
Absorption coefficient, 124, 144, 146, 147 Core/shell quantum dot
Alloy interface, 207–208 atoms and molecules, 123
Ammonia, 85 cylindrical, 159
Amplified spontaneous emission (ASE), 201 density-matrix formalism, 127
Anode, 224–226 electronic structure, 124
Atomic force microscopic (AFM) images, 91, energy gap, 124
92 fabrication technology, 127
Auger recombination (AR) process, 198 high quantum yield, 159
impurity states
binding energy, 136
B donor impurity, 134
Band alignment, 221, 224, 226, 232, 233, 240, electron binding energy, 137, 138
242, 245, 246, 248 electron energy levels, 137, 138
Biomass conversion, 260 electronic properties, 134
Biosensing, 21 electron wave function, 135
Building-integrated PVs (BIPVs), 288 linear and nonlinear optical absorption,
134
non-integer quantum number, 136
C parameters, 134
Cadmium telluride (CdTe), 14–15 probability density distribution,
Carrier dynamics, 231, 232, 235, 236, 238, 136–137
242, 244–246, 248 radial wave function, 135–136
Carrier separation, 229 Schrodinger equation, 135
Carrier transfer, 242 infrared devices, 128
Carrier transport, 229–230 interband optical absorption, 144–147
Charge extraction (CE), 231 Kratzer confinement potential, 125, 126
Charge transfer (CT), 267 layered geometry, 125
Chemical bath deposition (CBD) process, 222 one particle states
Chemical stability, 202, 226, 227, 235, 245, boundary conditions, 129
258, 273–275, 290 charge carriers, 132
Coffee-ring effect, 215 density distribution, 131

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 315
Springer Nature Switzerland AG 2020
X. Tong, Z. M. Wang (eds.), Core/Shell Quantum Dots, Lecture Notes in Nanoscale
Science and Technology 28, https://doi.org/10.1007/978-3-030-46596-4
316 Index

Core/shell quantum dot (cont.) transformations, 142


electron-hole interaction, 133 two-dimensional Schrodinger equation,
energy spectrum dependence, 130 140
hyper-geometric function, 134 types, 125
inner and outer interfaces, 128 Winternitz-Smorodinsky confinement
mathematical model, 128 potential, 125, 126
optical characteristics, 132 zero-dimensional systems, 124
radial distribution, 131 Counter electrode (CE), 221, 227
radial equation, 129 Crystal structure, 45–46
radial wave function, 131 Cylindrical nanolayer
Schrodinger equation, 128 electron-electron interaction, 150
spherical, 128, 129 electron states, 150
spherical oscillator, 133 energy spectrum, 151–153
spherical symmetry, 129 orbital current, 153–156
transcendental equation, 130 quantum tunneling, 150
Winternitz-Smorodinsky parameters, ring-shaped systems, 150
132, 133, 134 spherical symmetry, 150
optical properties, 124 spin magnetic moment, 150, 156–158
optical transitions, 127 spin-orbit interaction, 150
optoelectronic devices, 159 wave functions, 151–153
photoluminescence, 124
physical processes, 127
physical properties, 123–124, 128 D
polymer-coated magnetite core, 127 Density functional theory (DFT), 139
quadrupole moment, 148–150 Distributed feedback (DFB), 213
quantum nanostructures, 125
rectangular and shifted parabolic
confinement potentials, 125, 126 E
semiconductor nanocrystals, 124 Electrical generation, 260
single-particle states, 125 Electrochemical impedance spectroscopy
spherical, 127, 159 (EIS), 231
Stark shift, 127 Electrodes, 227
thin spherical nanolayer, 159 Electroluminescence (EL), 20
two-electron sates Electrolysis, 260
Coulomb interaction, 144 Electron and hole wave function, 207
effective radius, 143 Electron diffusion process, 229
electrons, 140 Electron–hole separation, 229
energy spectrum, 139 Electron injection efficiency, 231
ground-state energy, 141 Electron transport layers (ETLs), 19, 224
Hamiltonian, 140 Electrophoretic deposition (EPD) approach,
harmonic approximation, 143 274
harmonic oscillations, 139 Emission peak efficiency, 36
Mathieu equation, 139 Energy-dispersive X-ray spectrometer (EDX),
oscillation, 143 39
perturbation energy, 144 Energy-dispersive X-ray spectroscopy (EDS),
perturbation theory, 143 61
radial direction, 140 Energy gap
radial wave function, 140 AF parameter, 173
Schrodinger equation, 141 Bose–Einstein factor, 172
spherical QD, 140 electron charge, 173
spherical surface, 138 electronic subsystem, 171
spherium model, 139 electron–phonon interaction, 172–174
Index 317

exciton–phonon interaction, 172 High-resolution transmission electron


Huang–Rhys factor, 174 microscope (HRTEM) images, 53,
linear model, 172 245, 266
linear–quadratic relation, 172 Hole transport layers (HTLs), 19
phonons, 174 Homogeneous broadening, 186–187
photoluminescence bands, 174 Hydrogen, 260–261
physical interpretation, 172
second-order perturbation theory, 172
semiconductor materials, 171 I
single-phonon approximation, 172 Impurity states, 134–138
spectrum, 171 Incident photon to current conversion
temperature-dependent factors, 174 efficiency (IPCE), 231, 273
temperature-independent parameters, 173 Indium-doped tin oxide (ITO), 224
thermal dilation, 173 Inhomogeneous broadening, 186–187
thermal expansion and electron–phonon Intensity-modulated photocurrent spectroscopy
interaction, 173 (IMPS), 231
Ensemble, 182–183 Interfacial defects, 245
Exciton dynamics process, 206–208 Interfacial layers, 245, 247
Exciton–phonon interaction Internal quantum efficiency (IQE), 231
LakeShore Model, 174 Inverted type-I core/shell QDs, 237
linear model, 177
longitudinal acoustic vibrations (LA), 178
nonlinear character, 178 L
optical density, 175 Laser devices, 213
optical transitions, 176, 178 Laser interference ablation (LIA), 214
parameters, 177–178 Light absorption, 228
phonon modes, 178 Light harvester, 226
square-shaped symbols, 176 Light-harvesting efficiency (LHE), 231
temperature behavior, 178 Luminescence, 31
temperature dependencies, 174, 175, 177 Luminescent properties
temperature evolution, 176 absorption and emission spectra, 70
temperature-induced change, 176 AIS and AIS/ZnS QDs
temperature ranges, 178 absorption spectra, 76, 77, 79
External quantum efficiency (EQE), 20, 231 aggregation stability, 76
bandgap, 77
broad distribution, 83
F CdS LO phonon energy, 85
Flash hot injection, 203–204 core/shell AIS/ZnS-GSH QDs, 79
Fluorine-doped tin oxide (FTO), 224, 274 core/shell CIS/ZnS QDs, 81
Förster resonance energy transfer (FRET), 21 D-A model, 82
Full width at half maximum (FWHM), 36, 43 distribution of distances, 83
electron and hole traps, 82
electron-hole distances, 82
G electron-hole recombination
Giant volume, 207 mechanisms, 75
Gibbs free energy, 263 electron-phonon interaction, 84, 86
Glutathione, 72 high-resolution photoelectron spectra,
79
hot-injection/heating-up methods, 76
H Huang-Rhys factor, 84, 85
Half-width, 179–181 lattice imperfections, 81
Heavy metals, 281–283 lattice vibrations, 84
Hexadecylamine (HDA), 6 mercapto-and amino-groups, 85
318 Index

Luminescent properties (cont.) biomedical labeling, 68


molar AIS concentration, 78 biosensing, 94
normalized absorption, 86, 87 broad absorption bands, 68
optical density, 78 broadband character, 68
phonon energy, 84 D-A model, 69, 71
phonon peak, 85 distributions and sources, 70
PL characteristics, 80 donor and acceptor, 70
PL quantum yields (QYs), 80 electrons and holes, 71
PL spectra, 76, 86, 87 emission of photons, 69
radiative and non-radiative “energy depth,” 71
recombination, 83 energy factors, 112
radiative recombination, 82, 84 high thermal sensitivity and photostability,
semiconductors, 76 94
size distribution, 82 lattice defects, 68, 69, 113
size-selection method, 76, 82–83 light-emitting technologies, 94
size-selective precipitation, 81 measurements, 112
spectral parameters, 83 multicolored cell tracking, 68
spectral QD characteristics, 77 optical and physicochemical properties, 68
spectrum, 83 optical properties, 70
STE model, 84 optoelectronic systems, 68
S-to-In ratio, 80 physical model, 113
surface defects, 81 radiative and non-radiative processes, 94
surface-to-volume ratio, 85 reports of experimental evidence, 70
thermal lattice energy, 83 single-particle PL measurements, 91–93
trapped carriers, 83 size-selective precipitation, 112
2-propanol, 77 spectral emission parameters, 95
vibrational modes, 84 spectral parameters, 112
vibrational relaxation, 86 temperature (see Temperature)
analogy, 68 temperature-dependent variation, 71, 113
aqueous synthesis and optical properties tunneling mechanism, 69
absorption, 74 variability of application-relevant features,
Ag/In ratio, 73 68
AIS QDs, 72 vibrational modes, 113
bandgap, 73 Luminescent solar concentrators (LSCs)
biosensing applications, 71 bandgap alignment and optical properties,
carboxyl anions, 72 290
efficient phase transfer and stabilization, core/shell quantum dots
71 exciton dynamic, 290–292
FWHM, 75 optical efficiency, 293
glutathione, 72 optoelectronic applications, 293
heating-up and hot-injection protocols, shell materials, 291, 293
71 SILAR approach, 293–294
heat treatment, 75 stability, 294–295
lattice reconstruction, 74 Stokes shift, 292–293
mercaptoacids, 73 water transfer, 294
PL efficiency, 72 cu-based ternary/quaternary eco-friendly
PL spectra, 74 core/shell QD-based LSCs, 304–307
spectral parameters, 73 eco-friendly InP/ZnO Core/Shell
thermal treatment, 73, 75 QD-Based LSCs, 303–304
topographic surface, 73 efficiency, 288
water-soluble multifunctional ligands, fluorescent materials, 290
72 fluorophores, 288
Zn2+ ions, 75 fossil energy, 288
ZnS deposition, 75 internal quantum efficiency, 289
Index 319

near-infrared PbS/CdS Core/Shell homogeneous and inhomogeneous


QD-based LSCs, 295–298 broadening, 186–187
optical efficiency, 289, 290 measurements, 167
optical performance, 290 noise level, 169
PCE, 288 optical density, 167, 168
photons traveling, 289 particle-size distribution, 168
polymer matrix/emissions, 289 QD-1 and QD-2 InP/ZnS, 167
PV devices, 288 QD size distribution, 185–186
QD-based LSC, 288, 289 quantum-mechanical approaches, 171
short-circuit current, 290 simulation, 187–189
solar energy, 288 spectral processing technique, 169
solar tracking system, 288 spin–orbit splitting, 169
visible CdSe/CdS Core/Shell QD-based temperature band broadening, 189–191
LSCs three-layer structure, 167
BIPVs and plastic greenhouses, 301 transition energy, 169
diffuse/specular mirror, 298 wavelength, 168
emitted photons, 298 Nanoimprint lithography, 214
external quantum efficiency (EQE), Near-infrared (NIR) emissions, 42
298, 299 Nonrenewable energy resources, 260
J-V characteristics, 300
normalized absorption, 296
optical efficiency, 299, 300 O
optical properties, 298 Octadecylamine (ODA), 265
photons, 298 One-dimensional (1D) nanostructures, 224,
PL spectra, 296, 297 226
polymer matrix, 298 One-electron current, 150
sandwich structure, 300 Open-circuit voltage decay (OCVD), 231
shell layer, 298 Optical absorption
tandem structure, 300–302 biomarkers, 166
transmission electron micrograph, 300 biosensors, 166
waveguide transparent/semitransparent chemical elements, 166
materials, 288 luminescent processes, 166
nanocrystal absorption (see Nanocrystal
absorption)
M photoluminescence (PL) spectra, 166
Mercaptoacetate (MA) anions, 85 quantum dots (QDs), 165
Mercaptoacetic acid (MAA), 69, 71 semiconductor nanocrystals, 165
Mercaptopropionic acid, 71 spatial dimensions, 165–166
Mesoporous carbon (MC), 223 transition energy, 166
Microring, 213 Optical gain performance, 210–213
Microsphere resonators, 213 Optical microstructures, 213
Optical property, 290, 298, 307–308
Optoelectronic properties, 245
N
Nanocrystal absorption
atomistic pseudopotential approach, 169 P
Beer–Lambert law, 168 Photoconversion efficiency (PCE), 220
derivative spectrophotometry (see Photoelectrochemical cell
Nanocrystal absorption) applications, 258
disorder effects, 189–191 band alignment, 259
electronic structure, 166 band structure, 259
half-width, 179–181 CB and VB, 259
hardware/numerical methods, 169 charge carriers, 259
320 Index

Photoelectrochemical cell (cont.) Polyethyleneimine (PEI), 85, 95


charge dynamics Power conversion efficiency (PCE), 19, 288
quasi-type II core/shell QDs, 271, 272
semiconductor films, 272–273
type I core/shell QDs, 271, 272 Q
type II core/shell QDs, 271, 272 Quantum-confined Stark effect (QCSE),
electronic band structure, 259 209–210
energy consumption, 257 Quantum dot sensitized solar cells (QDSCs)
fossil fuels, 257 architecture
heavy metals, 281–283 anode, 224–226
hydrogen, 258, 260–261 CE, 227
morphologies, 258 electrodes, 227
organic ligands, 258 light harvester, 226
PEC hydrogen generation, 261–263 carrier separation, 229
physical/chemical and optical properties, carrier transport, 229–230
258 conventional p-n junction solar cells, 228
physical properties, 259 core/shell QDs
quasi-type II core/shell QDs, 269–271 band alignment, 232
semiconductor colloidal quantum dots bandgap tuning, 232
(QDs), 258 classification, 232–233
semiconductor material, 258 effective strategy, 232
solar energy, 258, 260–261 non-radiative carrier recombination,
synthesis of core/shell QDs, 264–266 231–232
TiO2 , 258 optoelectronic properties, 231
type I core/shell QDs, 266–267 photogenerated electrons, 232
type II core/shell QDs, 267–269 synthesis, 233–235
Photoelectrolysis, 261 device architecture, 221
Photoexcited electrons, 261 environmental pollution and global
Photoluminescence (PL) analysis warming, 220
average radiative lifetime, 87, 88 fossil fuels, 220
bandgap energy, 43 growing energy demand and environment-
distance-dependent Coulomb interaction, related issues, 220
89 history, 221–224
emission rate, 89 hot electron extraction, 221
fluorescence emission spectrometer, 41 light absorption, 228
growing multishells, 42 nanomaterials and device architecture, 220
higher intensity, MnS shell, 42 optoelectronic properties, 221
influence of shell concentration, 43, 44 PCE, 220
non-radiative recombination, 88, 89 photoelectric effect, 220
phonons, 90 photoelectrochemical, 220
PL spectra, 42 photothermal, 220
radiative recombination, 88, 89 photovoltaic (PV), 220
self-trapped exciton model, 90 photovoltaic characterizations, 230–231
size-selected colloidal AIS/ZnS QDs, 87, photovoltaic performance
88 core and shell interface optimization,
spectral parameters, 90 245–248
valence band (VB), 43 quasi-type-II core/thick-shell QDs,
vibrational relaxation, 90 242–245
Photoluminescence quantum yields (PL QYs), reverse type-I core/shell QDs, 237–238
197–198 type-I core/shell QDs, 235–237
Photophysical/chemical stability, 235–237 type-II core/shell QDs (see Type-II
Photovoltage decay (TVD), 231 core/shell QDs)
Photovoltaic characterizations, 230–231 recombination, 230
Photovoltaic (PV) devices, 288 silicon solar cell technology, 220
Index 321

single-photon absorption, 221 photobleaching, 10


solar energy, 220 photoblinking, 10
thin-film technology, 220 quantum confinement, 2
third-generation PV devices, 220 semiconductor nanostructured materials, 1,
Quantum yield, 293–294 2
shell thickness, 7
SILAR technique, 9, 10
R surface effect, 2–4
Reactive oxygen species (ROS), 22 synthesis, 5, 14–17
Reference electrode, 263 TC-SP, 12
Renewable energy resources, 260 TOP-SILAR method, 8
Reverse type-I core/shell QDs, 237–238 UV/blue/green, 11
zinc acetate and octanthiol, 8
zinc ethylxanthate, 12
S zinc stearate, 12
Schrödinger equation, 242 Semiconductor film, 224
Selected area electron diffraction (SAED) Semiconductor nanocrystal, 165, 166, 190
patterns, 38, 54, 61 Semiconductors, see Semiconductor core/shell
Self-trapped excitons (STE), 69 quantum dots (QDs)
Semiconductor core/shell quantum dots (QDs) Shell layer, 245
absorption and emission spectra, 13, 14 Shell thickness, 242
applications Size distribution, 185–186
LEDs, 19–21 Soft lithography, 214
solar cells, 18–19 Solar cells, 18–19
band alignment, 8 Solar-driven PEC cells
biology, applications core/alloyed-shell/shell QDs
biosensing, 21 band structure and carrier transition,
gene and drug delivery, 22 278, 280
therapy, 22 device stability, 278
bulk semiconductor material, 2 hole transfer capability, 277
CdTe, 14 hole transport layer, 276
characterization, 14–17 interfacial layer, 277
chemical and photochemical stability, 11 photoanodes, 278–281
classification, 4 photocurrent density, 278, 280
dissolving dimethylcadmium and selenium self-oxidation, 276
shot, 8 stability measurements, 278, 279
electron-hole pair, 2 core/thin-shell QDs, 274–275
electronic properties, 1, 10 “giant” core/shell QDs (g-QDs), 276
electronic wave function, 9 photoanodes, 273
electron leaves, 2 photocurrent density, 274
emission wavelength, 11, 12 Solar energy, 260–261
epitaxial shell growth, 10 Solar energy conversion devices, 242
evolution, 11, 12 Spatial electron–hole overlap, 244
growth mechanism, 7, 9 Spatial separation, 242
hot injection method, 8 Spectral characteristics, 144
injection method, 5–6 Stability, 294–295
lattice strain and defect states, 7 Static and dynamic disorder, 182–183
material parameter, 7 Stimulated radiation, 213–216
monolayers, 8 Stokes shift, 290, 292–293
narrow emission peak, 10 Successive ionic layer adsorption and reaction
near-infrared spectrum, 11 (SILAR) method, 8, 202–203, 223,
non-injection method, 6 293–294
optical properties, 1, 10 Superior photophysical/chemical stability,
organometallic precursors, 12 245
322 Index

Synthesis, see Semiconductor core/shell metal-chalcogenide QDs, 105


quantum dots (QDs) non-radiative energy transfer processes,
Syringe pump injection, 203 104
non-radiative recombination, 104, 105
radiative recombination, 104, 105
T single-exponential functions, 104
Temperature radiative and non-radiative processes,
AIS and AIS/ZnS QDs, 108–109 107–108
bulk and nanometer compounds, 111 radiative recombination, 110
“classical”, 111 spectral PL parameters, 108
defect-related PL (DPL), 108 sustainable semiconductor nanomaterials,
electronic structure, 110 108
electron-phonon interaction, 108 temperature-dependent PL evolutions, 108
emission bands, 111 thermal sensitivity, 95
evolution variations, PL intensity, 96–99
behavior, 183–184 Ternary chalcogenides
static and dynamic disorder, 182–183 AgInS2 nanocrystals, 33
excitonic PL (EPL), 108 binary chalcogenides, 31
heating, 95 Cd2+-Doped Core and Core/Shell AgInS2
luminescent properties, 95 Nanocrystals
mechanisms/routes, 111 crystal structure analysis, 51–52
multifunctional ligands, 95 elemental compositions, 54
optical phonon decay, 110 photoluminescence properties, 50–51
PEI, 95 surface morphology analysis, 53–54
PL band maximum position UV-Vis absorption spectra, 48–50
absorption spectra, 100 cluster science, 30
behavior, 99 core and shell nanocrystals, 32
excitonic absorption of CdS-PEI QDs, core-multishell and alloyed
101 AgInS2/CdS/ZnS shell nanocrystals
GSH molecules, 101 crystal structure analysis, 59–60
heating/cooling, 100 morphological studies, 60–61
linear manner, 101 photoluminescence studies, 56–58
macromolecular PEI, 101 XPS analysis, 58–59
measurements, 100 core-multishell architecture, 32
parameters, 100 core-multishell CuInS2/MnS/ZnS
ratiometric sensor, 102 nanocrystals
sensor output, 103 absorption properties, 39–41
size-selected colloidal AIS QDs, 102 crystal structure analysis, 45–46
spectral PL parameters, 100 influence of shell concentration, 45–46
stabilization, 101 morphological analysis, 46–48
surface metal complexes, 102 photoluminescence analysis (see
temperature-dependent variation, 102 Photoluminescence analysis)
thermal dissociation of Cd(II)-PEI, 101 XPS analysis, 43–45
two-channel portable spectrometer, 103 core/shell nanoparticles, 33–34
PL decay curves core/shell type, 30, 31
characteristics, 106 electronic energy levels, 30
dependences, 105 energy levels, 30
electronic energy transfer, 105 epitaxial growth, 32
energy donor and acceptor, 105–106 growth of multishell over core nanocrystals,
energy transfer, 106 34
fitting parameters, 104, 106–107 Mn2+ Ion-Doped Core and Core/Shell
kinetic curves, 103 CuInS2 Nanocrystals
Kohlrausch-type stretched exponent crystal structure analysis, 38
model, 104 morphological properties, 38–39
Index 323

optical properties, 34–35 Transmission electron microscopy (TEM)


photoluminescence properties, 36–37 images, 91
XPS analysis, 37 Trioctylphosphine (TOP), 5, 240
properties of nanoclusters, 29 Trioctylphosphine oxide (TOPO), 5, 106
semiconductor materials, 30 Type-II core/shell QDs
shell materials, 30, 31 band engineering approach, 240, 241
Tetradecylphosphonic acid (TDPA), 240 electron injection, 238
Thermal cycling alternate ionic layer electron transport layer, 238, 240
adsorption and reaction (TC- light-harvesting materials, 240
SILAR) method, 202–203 QDSCs and optoelectronic devices, 239
Thermal cycling coupled single precursor solar cell application, 240
(TC-SP), 12 structural and optical characterization, 241
Thermolysis, 260 TEM images, 239, 240
Thick-shell core/shell quantum dots
AR process, 198
basic characteristics, 199–201 U
flash hot injection, 203–204 Ultraviolet (UV) light stability, 205
“giant” QDs, 198
monolayers (MLs), 198
nucleation and growth dynamics, 198
V
optical properties, 198
Valence band (VB), 228, 259, 291
exciton dynamics process, 206–208
Valence band maximum (VBM), 32
fundamental, 204–206
optical gain performance, 210–213
QCSE, 209–210
stimulated radiation, 213–216 W
optical stability, 198 Wide-bandgap semiconductor, 221
photophysical and photochemical stability,
198
semiconductor nanocrystals, 197 X
SILAR method, 202–203 X-ray diffraction (XRD), 38, 78
syringe pump injection, 203 X-ray photoelectron spectroscopic (XPS), 37,
ultrafast nonradiative process, 198 78
Thioglycolic acid (TGA), 14, 237
Total internal reflection (TIR), 288
Transient absorption (TA), 238, 244 Z
Transient photocurrent decay (TCD), 231 (Zero-phonon) emission line (ZPL), 69

You might also like