CDS QDs - T Mallick

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Physics and Chemistry of Solids 164 (2022) 110603

Contents lists available at ScienceDirect

Journal of Physics and Chemistry of Solids


journal homepage: www.elsevier.com/locate/jpcs

Carbazole-decorated fluorescent CdS quantum dots: A potential


light-harvesting material
Tamanna Mallick a, Abhijit Karmakar a, Moumita Kar b, Sourav Dutta c, Sudip Kumar Mondal b,
Debabrata Mandal c, Anup Pramanik d, Naznin Ara Begum a, *
a
Organic Chemistry Division, Department of Chemistry, Visva-Bharati (Central University), Santiniketan, 731 235, India
b
Physical Chemistry Division, Department of Chemistry, Visva-Bharati (Central University), Santiniketan, 731 235, India
c
Department of Chemistry, University College of Science and Technology, University of Calcutta, 92, Acharya Prafulla Chandra Road, Kolkata, 700 009, India
d
Department of Chemistry, Sidho-Kanho-Birsha University, Purulia, West Bengal, 723104, India

A R T I C L E I N F O A B S T R A C T

Keywords: An organic-inorganic nanohybrid system (AMC-CdS QD) comprised of a fluorescent carbazole analog, AMC (3-
CdS QDs amino-N-methyl carbazole), and CdS QDs has been synthesized, which shows promise as a sustainable energy
Carbazoles harvesting material. FT-IR and 1H NMR spectroscopy have ascertained the covalent grafting of the carbazole
Organic-inorganic nanohybrids
moieties onto the mercaptopropionic acid (MPA)-capped surface of the CdS QDs. Detailed photophysical char­
Fluorescence
Förster resonance energy transfer
acterization of this nanohybrid system have confirmed the occurrence of efficient Fӧrster resonance energy
Light-harvesting materials transfer (FRET) from its organic to inorganic counterpart. This phenomenon was also demonstrated using SCC-
DFTB-based quantum chemical calculations. As explored in the present work, the photoelectron transfer effi­
ciency of AMC-CdS QDs signifies their plausible future applications as light-harvesting materials.

1. Introduction to their exceptional size- and composition-dependent optoelectronic


properties, tuneable band gap, luminescence, high molar extinction
Research on the development of sustainable and green energy tech­ coefficient and large intrinsic dipole moments related to the quantum
nologies based on solar photovoltaics has witnessed a remarkable up­ confinement effect. It is noteworthy that cadmium chalcogenide QDs, e.
surge recently. Scientists are now focusing on third-generation solar g., CdSe, CdTe, and CdS, absorb light in the visible light region and
cells with higher photoconversion efficiency (PCE) per unit area than exhibit size-dependent narrow and tuneable band gaps, broad excitation
first and second-generation solar cells [1,2]. Examples of and narrow emission spectra and size, as well as surface chemistry and
third-generation solar cells are dye-sensitized solar cells (DSSCs), modulated fluorescence behaviour [1,15,16]. Furthermore, such QDs
organic/polymer-based solar cells, quantum dots (QDs) sensitized solar show higher photo-, thermal-, and moisture stability when compared to
cells, and perovskite solar cells [2–6]. dye molecules and perovskites [1,16]. Such properties open up their
Over the past two decades, DSSCs have attracted a tremendous applications as sensitizers in QD-sensitized solar cells (QDSSCs), which
amount of attention due to their eco-friendliness, low-cost operation and are emerging as a promising alternatives to DSSCs [17–19].
easy fabrication. However, DSSCs are not free from drawbacks such as Among these chalcogenides, CdS QDs are often studied as model QDs
low PCE, use of expensive dyes and long-term stability. These limit their due to their size-dependent emission in the visible light region, signifi­
use on a commercial/industrial basis [7]. To overcome these problems, cant photo-stability, and fluorescence quantum yield [1,15,16,20,21].
dyes have been replaced with other light-harvesting materials, such as On the other hand, stable CdS QDs can be easily synthesized under mild
semiconductor quantum dots (QDs), and thus, QD-sensitized solar cells conditions using economically viable protocols [22,23]. At the same
(QDSSCs) with effective charge separation and transport capacities have time, several surface modification techniques have been reported that
been developed, which have emerged as promising alternatives to DSSCs can tune the optoelectronic properties of CdS QDs [23,24]. CdX QDs (X
[2,8–14]. Colloidal semiconductor nanocrystals, i.e., quantum dots = S, Te) capped with photo-sensitive organic ligands, such as azo­
(QDs), have been extensively explored as light-harvesting materials due benzene derivatives and porphyrin moieties [19,25–27] or QDs

* Corresponding author.
E-mail address: [email protected] (N.A. Begum).

https://doi.org/10.1016/j.jpcs.2022.110603
Received 11 September 2021; Received in revised form 21 January 2022; Accepted 21 January 2022
Available online 31 January 2022
0022-3697/© 2022 Elsevier Ltd. All rights reserved.
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

embedded in a polymeric matrix with natural flexibility [28,29] have large number of carbazole analogs or polymeric nanoparticles contain­
been extensively studied for their diverse range of applications in the ing carbazole moieties have been explored as optoelectronic sensor
fields of energy conversion and organic light-emitting diodes (OLEDs) materials [34–40]. The significant photophysical properties of these
materials as well as highly accurate biomedical diagnostic tools. These molecular scaffolds, such as distinct chromophoric/fluorophoric nature,
organic-inorganic hybrid nanomaterials have the dual advantages of response to the Förster resonance energy transfer (FRET) phenomenon,
each component and show optimized performance derived from their and high quantum yield values, make them suitable for the development
components [19,25,26,30,31]. These studies have motivated our group of organic semiconducting, photo-sensitive, hole-transporting, and op­
to explore the utility of 3-amino-N-methyl carbazole (AMC), a carbazole toelectronic devices, solar cells and optoelectronic sensor materials
analog with characteristic chromophoric/fluorophoric properties as the [34–40]. However, carbazole may give poor device performance due to
organic counterpart in an CdS QDs based organic-inorganic nanohybrid its planar structure, which is highly susceptible to aggregation. Such
system. phenomena are also observed with porphyrin moieties, which are used
Carbazoles are N-heterocyclic bioactive compounds, which are in DSSCs [18]. These problems can be avoided in the case of a nano­
ubiquitous in plants as secondary metabolites (Scheme 1). At the same hybrid system consisting of an opto-electronically active simple carba­
time, carbazoles with varied structural patterns can also be synthesized zole moiety and CdS QDs, and thus, it can be used as a light-harvesting
in the laboratory [32,33]. These π-conjugated small molecules show material in solar cells.
tuneable HOMO-LUMO energy and photo- and thermal stability, and a This motivated us to synthesize an organic-inorganic nanohybrid

Scheme 1. Synthesis of (a) AMC: (i) Me2SO4, acetone, KOH, reflux; (ii) AcOH–HNO3; (iii) Sn/HCl, reflux and (b) AMC-CdS QDs: (i) 0.1 M NaOH, 30% H2O2, water
60 ◦ C, 2 h; (ii) 3-Amino-N-methyl carbazole (AMC), EDC, toluene, reflux, 24 h.

2
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

system, AMC-CdS QDs, by covalently anchoring fluorescent carbazole sets and the Perdew–Burke–Ernzerh (PBE) [47] exchange-correlation
(AMC) moieties onto the MPA-modified surface of CdS QDs [Scheme 1]. energy functional was employed throughout the calculations. More de­
Grafting AMC onto the MPA-capped CdS QDs was confirmed based on tails on the computational methodologies can be found in the SI.
FT-IR and 1H NMR spectroscopy. At the same time, XRD and HR-TEM
measurements were carried out to study the structural and morpho­ 3. Results and discussion
logical properties of the as-synthesized CdS QDs. The overlap of the
emission spectra of the AMC molecules with the excitation spectrum of 3.1. Covalent grafting of 3-amino-N-methyl carbazole (AMC) on the
the MPA-capped CdS QDs indicates the occurrence of FRET from the MPA-CdS QDs surface
carbazole molecules to the MPA-capped CdS QDs. The efficiency and
mechanism of FRET in the process involving the components in the As mentioned earlier, AMC molecules were anchored on the MPA-
present nanohybrid system were established using steady-state fluores­ CdS QDs surface via amide bond formation. FT-IR and 1H NMR spec­
cence studies and fluorescence lifetime measurements utilising the time- troscopy confirmed the attachment of the carbazole moieties on the
correlated single-photon counting (TCSPC) method. The experimental MPA-functionalized surface of the MPA-CdS QDs and the results shown
results were further validated using tight-binding density-functional in Fig. 1a-b. The QDs-surface attached organic functional groups give
theory based on quantum-chemical calculations. rise to the characteristic FT-IR spectral patterns of the AMC-CdS QDs and
The synthetic protocol used to fabricate the AMC-CdS QD nano­ MPA-CdS QDs. Therefore, to understand the efficacy of the present
hybrid system is economically viable and can be easily implemented. carbazole grafting protocol, we have compared the FT-IR spectrum of
The photoelectron transfer efficiency of this low-cost nanohybrid system the AMC-CdS QDs with spectra obtained for the MPA-CdS QDs and free
opens up their utility in the design and development of solar cells as well AMC (as shown by the shaded area in Fig. 1a). The FT-IR peak of MPA-
as FRET-based bio-medical diagnostic tools. CdS QDs at 3370 cm− 1 is associated with the –OH stretching vibrations
of the –COOH group. For both the MPA-CdS QDs and AMC-CdS QDs, the
2. Materials and methods FT-IR peak at 672 cm− 1 corresponds to the S–C bond stretching vibra­
tions (Fig. 1a). Fig. 1a shows the free AMC exhibits FT-IR peaks at 3401
2.1. Materials cm− 1 (primary amine), 1603 cm− 1 (bending N–H) and 1328 cm− 1 (C–N
stretching). Thus, AMC was linked with the MPA-CdS QDs through an
The precursors for the CdS QDs: cadmium sulphate (CdSO4⋅8/3H2O), amide bond.
3-mercaptopropionic acid (MPA), and hydrogen peroxide (H2O2), were This was confirmed by the appearance of the IR peaks corresponding
purchased from Sigma Aldrich, India. 1-(3-Dimethyl aminopropyl)-3- to the AMC-CdS QDs at 1656 and 1570 cm− 1, which represent the amide
ethyl carbodiimide hydrochloride (EDC) (Spectrochem) was used as carbonyl stretching and N–H bending vibrations of the amide group,
the coupling reagent. All other chemicals were of analytical grade and respectively. A shift in the C–N stretching vibration was also observed in
purchased from Merck, India. Milli-Q (Milli-Q Academic with 0.22 mm the AMC-CdS QDs (1303 cm− 1) when compared to the free AMC (1325
Millipak R-40) water was used as per the requirements. cm− 1). It was interesting to note that the IR spectrum of AMC-CdS QDs
closely resembles those of the free AMC and MPA-CdS QDs.
2.2. Synthesis of AMC-CdS QDs by covalent grafting of 3-amino-N- Thus, the FT-IR results demonstrate the successful covalent attach­
methyl carbazole (AMC) on MPA-CdS QDs ment of the AMC molecules onto the MPA-modified surface of the MPA-
CdS QDs via an amide linkage involving the carboxylic acid end groups
The synthesis of the MPA capped CdS (MPA-CdS) QDs was carried and primary amine group at the 3-position of AMC [Fig. 1a].
out using previously reported methods [41,42]. A brief description of This was further established by comparing the 1H NMR spectrum of
the synthesis and characterization protocols are given in the Supple­ the AMC-CdS QDs with the AMC and MPA-CdS QDs precursors, as
mentary Information (SI). 3-Amino-N-methyl carbazole (9-methyl-9­ shown in Fig. 1b. The 1H NMR spectrum obtained for the AMC-CdS QDs
H-carbazol-3-amine) (AMC) was synthesized, purified, and matches the characteristic features (as denoted by the similar colours in
characterized according to a previously reported method [43] (Scheme this figure) observed for AMC and the MPA-CdS QDs. Thus, this
1a). It was then grafted on the MPA-CdS QDs surface via an amide comparative 1H NMR spectral data further confirmed the grafting of the
linkage, as shown in Scheme 1b. In the present method, EDC was used to carbazole moieties on the MPA-capped surface of the CdS QDs. The 1H
activate the carboxyl groups of MPA attached to the surface of the CdS NMR spectra (with full range) of these systems are also shown in the SI.
QDs for their effective coupling with the 3-amino group of AMC to yield
amide bonds (Scheme 1b). Grafting of the AMC moieties on the 3.2. The formation, structures and morphologies of the MPA-CdS QDs
MPA-capped CdS surface was confirmed using Fourier and AMC-CdS QDs
transformed-Infrared (FT-IR) and 1H NMR spectroscopy. The method­
ologies and other details are provided in the SI. The UV–Vis absorption spectrum of the MPA-CdS QDs [Fig. 2 (a)]
was in good agreement with that reported for MPA-capped CdS QDs
2.3. Spectroscopic studies on the MPA-CdS QDs and AMC-CdS QDs [22–24,41,42,48]. Bulk CdS shows an absorption edge at 515 nm. In the
present case, the MPA-CdS QDs show an absorption peak (broad hump)
UV–Vis, steady-state, and time-dependent fluorescence spectroscopy at 386 nm. Such a blue shift confirms the formation of smaller-sized
of the AMC-CdS QDs, MPA-CdS QDs and AMC were carried out particles [22,49]. The UV–Vis absorption data obtained for the
following the methods described in the SI. Details of the fluorescence MPA-CdS QDs were also explored to determine their size with the help of
quantum yield (QY) calculations used for the AMC-CdS QDs and MPA- an effective mass model (Fig. S1 in the SI) and the size was determined to
CdS QDs in a water:ethanol (2:1) medium are given in the SI. be 2.52 nm. This was further confirmed based on TEM (discussed in a
later section).
2.4. Computational studies based on quantum mechanical calculations The XRD pattern obtained for the MPA-CdS QDs [Fig. 2 (b)] re­
sembles that of a zinc blende crystal structure [22–24,41,42]. The three
The self-consistent charge density-functional tight-binding (SCC- prominent peaks observed at 26.24, 44.44, and 52.44◦ correspond to the
DFTB) method [44,45] was used to study the electronic properties of the (111), (220) and (311) crystal of the CdS cubic phase (JCPDS
AMC-CdS QDs. We have used the previously derived SCC-DFTB No.10-0454) planes, respectively [42]. The size and morphology of the
parameter sets used for Cd–S and other pairs of elements involving C, MPA-CdS QDs were further elucidated with the help of TEM analysis.
N, O, and H [46]. The Slater-type orbitals (STOs) were chosen as basis The TEM images of the MPA-CdS QDs and AMC-CdS QDs are shown in

3
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

Fig. 1. (a) FT-IR and (b) 1H NMR spectra obtained for the MPA-CdS QDs, AMC-CdS QDs and AMC (indicated by red, blue and green colours, respectively).

Fig. 2c (i-ii) and (d) (i-ii), respectively. In both cases, mainly spheroidal appearance of bright spots in the SAED patterns of MPA-CdS QDs [inset
particles are visible. The particle size distribution in each case of the of Fig. 2c (ii)] are a clear indication of the crystalline nature of the QDs
MPA-CdS and AMC CdS QDs (as denoted by the bar diagrams) are shown [50].
in the insets of Fig. 2c (i) and (d)(i). The average diameter of the A similar phenomenon was observed in the SAED pattern of the
MPA-CdS QDs, as calculated by analysing 105 QDs, was found to be 2.4 AMC-CdS QDs [inset of Fig. 2d (ii)], which undoubted proved the con­
± 0.4 nm. This was in good agreement with the size calculated from the servation of the crystalline nature after the covalent grafting of AMC
UV–Vis spectral data, which was found to be 2.52 nm. In a similar onto the MPA-CdS QDs. The TEM analysis of the MPA-CdS QDs and
calculation based on the particle size distribution, the average diameter AMC-CdS QDs indicate that the anchoring of the carbazole moieties on
of the AMC-CdS QDs was obtained to be 3.1 ± 0.3 nm. On the other the MPA-modified surface of the CdS QDs did not drastically affect the
hand, from the UV–Vis spectral data, the particle size of the AMC-CdS primary surface structure (i.e., size and morphology) of the spheroidal
QDs was found to be 2.88–3.00 nm. Therefore, the particle size ob­ QDs.
tained from the TEM and UV–Vis data are in good agreement. The Energy dispersive X-ray analysis (EDX) was carried out to check the
detailed particle size calculations based on the UV–Vis spectral data are elemental composition and purity of the as-synthesized AMC-CdS QDs.
discussed in the SI. The EDX spectrum obtained for the AMC-CdS QDs is shown in Fig. 2d
Fig. 2c (ii) shows the HRTEM of the MPA-CdS QDs, where the d- (iii), which confirmed the presence of elemental C, N, O, S and Cd. These
spacing (indicated by a yellow arrow) was found to be 0.33 nm for the are the main constituents of the AMC-CdS QDs. The composition of these
(111) plane of the cubic CdS. The selected area electron diffraction elements in the AMC-CdS QDs was found to be 65.88% (C), 17.72% (N),
(SAED) pattern of these particles is shown in the inset of Fig. 2c (ii). The 10.39% (O), 3.36% (S) and 2.64% (Cd). Therefore, EDX measurements

4
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

3.3. A comparison of the UV–Vis spectra obtained for the MPA-CdS QDs
and AMC-CdS QDs and understanding their optical band gap energies

Fig. 3a shows the UV–Vis absorption spectra obtained for the MPA-
CdS QDs and AMC-CdS QDs, and Fig. 3b shows the UV–Vis spectrum
of AMC. The inset of Fig. 3a shows the band edge absorption of the MPA-
CdS QDs and AMC-CdS QDs. The absorption edge of the AMC-CdS QDs
was slightly red-shifted when compared to that of the MPA-CdS QDs.
This may be due to the carbazole moieties forming a thin shell around
the MPA-capped CdS QDs. This causes a relaxation of its quantum
confinement when compared to the MPA-CdS QDs. However, in both
cases the absorption range falls within 375–460 nm, which is typical of
CdS QDs. In bulk CdS, the UV–Vis absorbance maximum is found to be in
the range of 475–550 nm [23]. Therefore, the absorption band edge of
the MPA-CdS and AMC-CdS QDs was found to be blue-shifted when
compared to the bulk CdS.
The band gap energy (Eg) of MPA-CdS QDs and AMC-CdS QDs was
also calculated using the Tauc relationship [22,52]:

(αhv) = (hv − Eg )n (1)

where, α is the absorption coefficient, n is a constant with a value of ½


for direct band gap semiconductors, such as CdS QDs, h is the Plank’s
constant, and ν denotes the frequency of the incident photon.
The absorption data obtained for MPA-CdS and AMC-CdS QDs were
fitted using this relationship and the plots of (αhν)2 vs. hν for MPA-CdS
and AMC-CdS QDs are shown in Fig. 3c. These were further explored to
calculate the band gap of the QDs. The band gap energy (Eg) for MPA-
CdS QDs was determined to be 2.85 eV. However, this value was
found to be shifted to 2.65 eV for the AMC-CdS QDs. This shift can be
attributed to the covalent linkage of the carbazole moieties on the MPA-
capped surfaces of the CdS QDs. However, the calculated band gap en­
ergy of the QD system was found to be higher when compared to that of
the bulk CdS band gap (2.42 eV) [22]. Hence, the size of the AMC-CdS
QDs was still within the QDs region. However, the covalent linkage of
the chromophoric/fluorophoric carbazole moieties on the CdS QD sur­
face via MPA greatly influences their UV–Vis light absorption properties.
Therefore, we further compared their fluorescence behaviour to un­
derstand the role of surface-attached functionalities in controlling the
fluorescence behaviour of the as-synthesized CdS QD-based nanohybrid
system.

3.4. Fluorescence spectra obtained for the MPA-CdS QDs and AMC-CdS
QDs

We have measured the steady-state fluorescence spectra of the MPA-


CdS QDs and AMC-CdS QDs along with AMC in water/ethanol (2:1 v/v).
In AMC, a characteristic fluorescence emission spectrum was observed
Fig. 2. The (a) absorbance spectrum, (b) XRD pattern and (c) (i) TEM (inset: at an excitation wavelength (λex) of 290 nm (Fig. 4a). In addition, AMC
particle size distribution) and (ii) HRTEM (SEAD pattern in inset) images of the shows significant emission upon excitation over the wavelengths range
MPA-CdS QDs. (d) (i) TEM images of the AMC-CdS QDs (inset: particle size of 290–420 nm (Fig. S2 in the SI). However, a prominent fluorescence
distribution), (ii) corresponding close view (SEAD pattern in inset) and (iii) EDX
pattern was observed in the MPA-CdS QDs at λex = 400 nm (Fig. 4a). For
pattern of the AMC-CdS QDs.
the AMC-CdS QDs, in addition to the emissions observed at λex 290 and
400 nm (Fig. 4a), characteristic fluorescence emissions were also
confirmed that no other elemental impurities were present in the AMC- observed at λex 320, 340, 360 and 380 nm (Fig. 4b). The AMC-CdS QDs
CdS QDs and this undoubtedly proved its purity [51]. are the hybrid of MPA-CdS QD and AMC. Therefore, it responses to
In the case of metal chalcogenide QDs, the surface-attached ligands photoexcitation over a wide range of wavelengths i.e., λex 290–400 nm
play an important role in modulating the surface defects in such QDs, and exhibits emission spectra over a broad range (399–612 nm). Such
which in turn greatly influence their UV–Vis and photoluminescence characteristics indicate its potential future application as a white light-
behaviour [18,23]. In the present case, to understand the role of the CdS emitting diode.
QD-surface anchored fluorophoric AMC moieties, we compared the Fig. 4a shows that upon photoexcitation at 400 nm (λex), the AMC-
photophysical behaviour, i.e., steady-state UV–Vis and fluorescence CdS QDs exhibit a redshift of ~9 nm in the λem when compared to
spectra and time-dependent fluorescence dynamics of the MPA-CdS QDs that of the MPA-CdS QDs. A redshift (~50 nm) in the λem of AMC-CdS
and AMC-CdS QDs. QDs was also observed when compared to AMC at λex 290 nm
(Fig. 4a). At the same time, the fluorescence emission intensity of AMC
was found to be diminished (~3-fold) compared to the AMC-CdS QDs

5
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

Fig. 3. (a) UV–Vis absorption spectra obtained for the MPA-CdS QDs (red colour) and AMC-CdS QDs (blue colour). Inset: Band edge plot for the MPA-CdS QDs and
AMC-CdS QDs. (b) UV–Vis absorption spectrum of AMC. (c) Optical band gap plots of the MPA-CdS QDs (red colour) and AMC-CdS QDs (blue colour). (d) The visual
appearance of the MPA-CdS QDs (i), AMC (ii), and AMC-CdS QDs (iii).

(Fig. 4a). Similarly, we also compared the excitation spectra of the MPA- 3.5. Elucidating the Förster resonance energy transfer (FRET) in AMC-
CdS QDs and AMC-CdS QDs under the same experimental conditions CdS QDs
(Fig. 4c). A good overlap was observed between the spectra in the
wavelength range of 452–580 nm, i.e., towards the tail end of these The fluorescence quantum yield (QY) of the MPA-CdS QDs and AMC-
spectra, as shown in Fig. 4c. However, a redshift in the peak position (25 CdS QDs in water:ethanol (2:1 v/v) using Coumarin-153 as a reference
nm) was observed in the wavelength range of 332–357 nm. These results was determined to be 7.36 and 11.81%, respectively (see SI). Hence, the
confirm the covalent linking of AMC on the MPA-capped surface of the surface-anchored carbazole moieties were responsible for increasing the
CdS QDs without altering the basic QDs structure of the AMC-CdS QDs. QY of AMC-CdS QDs when compared to that of the MPA-CdS QDs.
The excitation–emission spectra of AMC-CdS QDs (λex 290, 320, 340, Notably, there is an appreciable spectral overlap between the
360, 380 and 400 nm) along with those of AMC (λex 290 nm) and MPA- normalized excitation spectrum of MPA-CdS QDs (acceptor) and fluo­
CdS QDs (λex 400 nm) are shown in Figs. S3 (a-c) in the SI. rescence emission spectra of AMC (donor), as shown in Fig. 4d. The
The MPA-carbazole moieties passivate the surface-exposed dangling FRET efficiency (ΦE) is strongly dependent on the extent of the overlap
bonds of the CdS QDs and this phenomenon can modulate the deep track between the donor’s emission spectrum and the absorption spectrum of
emission states of CdS QDs systems. Besides such structural effects, one the acceptor. At the same time, the distance between the donor and
of the obvious causes for the prominent fluorescence spectral changes in acceptor species also influences the ΦE [53]. Thus, the present func­
AMC may be its conjugation with the MPA-capped CdS QD system, tionalization process was found to be helpful in bringing the donor
which induces a Förster resonance energy transfer (FRET) between the moieties (AMC) in the proximity of the surface of acceptor species (CdS
AMC (donor) and MPA-capped CdS QDs (acceptor) [16,22], which are QDs), as confirmed by the Förster radius calculated from fluorescence
linked to each other via –NH (CO)CH2CH2–S- bonds (Scheme 1b). lifetime measurements (discussed later).
The spectral overlaps between the normalized excitation spectrum of
the MPA-CdS QDs (acceptor) and fluorescence emission spectra of AMC
(donor) in different solvents (EtOH and DMF) are show in Fig. S4 (SI). A
change in the solvent system has an effect on the FRET between these

6
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

Fig. 4. (a) (i) Fluorescence spectra obtained


for AMC (green, λex 290 nm), AMC-CdS QDs
(dark blue, λex 290 nm and blue, λex 400
nm) and MPA-CdS QDs (red, λex 400 nm).
The corresponding normalized fluorescence
spectra are shown in (ii). (b) Normalized
fluorescence spectra obtained for the AMC-
CdS QDs at different excitations (λex 290,
320, 340, 360, 380 and 400 nm). (c)
Normalized excitation spectra obtained for
the MPA-CdS QDs (red, λem 607 nm) and
AMC-CdS QDs (blue, λem 612 nm). (d) The
spectral overlap of the normalized excitation
spectrum of the MPA-CdS QDs (red) and
emission spectrum of AMC (green).

systems. Fig. S5 in the SI shows the spectral overlap behaviour of these are shown in Fig. 5a, while a comparison of the TCSPC decay curves of
systems in their powdered and thin-film states. The occurrence of FRET the MPA-CdS QDs and AMC-CdS QDs are shown in Fig. 5b. The corre­
in the AMC-CdS QDs was also observed in the solid-state, indicating its sponding results are summarized in Table 1.
device applicability. The average lifetime (τav) of AMC at λex = 290 nm was 5.50 ns. In this
Analysis of the fluorescence lifetime data based on TCSPC mea­ case, λem = 375 nm. However, τav for the AMC-CdS QDs was 3.10 and
surements can be an authentic and precise way to understand the FRET 0.50 ns when λex = 290 and 400 nm, respectively, and the corresponding
phenomenon between AMC and the MPA-capped CdS surface in the λex was 360 and 612 nm, respectively. The τav of MPA-CdS QDs at λex
AMC-CdS QDs because such dynamical studies are free from the limi­ 400 nm (λem 607 nm) was found as 0.20 ns. In both cases, it was evident
tations of steady-state techniques. that the average lifetime of AMC decreases in the AMC-CdS QDs con­
The TCSPC decay curves obtained for the donor (AMC) in the jugate system. Hence, it can be concluded that after the covalent
absence and presence of the acceptor species, i.e., MPA-capped CdS QDs, attachment of AMC on the surface of MPA-capped QDs, the τav of the

7
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

Fig. 5. Fluorescence decay curves obtained for (a) AMC (λex 290 nm, λem 375 nm) and the AMC-CdS QDs (λex 290 nm, λem 360 nm) and (b) the MPA-CdS QDs (λex
400 nm, λem 607 nm) and AMC-CdS QDs (λex 400 nm, λem 612 nm).

moiety, probably via the FRET mechanism. Furthermore, the average


Table 1
lifetime data indicate that comparatively more efficient energy transfer
Fitting parameters for the fluorescence lifetime profile of AMC, AMC-CdS QDs,
from AMC to CdS-MPA QD occurs at λex = 290 nm when compared to λex
and MPA-CdS QDs in water:ethanol (2:1, v/v).
= 400 nm.
Sample τ1/ a1 τ2/ns a2 τavg/ The FRET efficiency (ΦE) was calculated using the following equa­
ns ns
tion [16,44]:
AMC (λex 290 nm) (λem 375 nm) 2.50 0.76 15.20 0.24 5.50
AMC-CdS QDs (λex 290 nm) (λem 360 1.70 0.81 9.00 0.19 3.10 τ
ΦE = 1 − (2)
nm) τ′
MPA-CdS QDs (λex 400 nm) (λem 607 5.00 0.02 0.10 0.98 0.20
nm) where, τ and τ′ are the average lifetime of the AMC-CdS QDs (3.10 ns)
AMC-CdS QDs (λex 400 nm) (λem 612 6.80 0.04 0.20 0.96 0.50 and AMC (5.50 ns), respectively at λex = 290 nm. The calculated ΦE for
nm)
this system was obtained as 0.44 at λex = 290 nm. Similarly, the energy
τn and an are the lifetime (ns) and amplitude of the nth component; τavg is the transfer rate (kt) was calculated using equation (3) [54] and it was
amplitude-weighted average lifetime, calculated with the help of τavg = Ʃanτn. determined to be 0.14 ns− 1.
1 1
AMC-CdS QDs hybrid increases in comparison to MPA-CdS QDs. This is a kt = − ′ (3)
τ τ
clear indication that in the present AMC-CdS QDs nanohybrid system,
the QD-surface anchored AMC molecules act as a donor moiety, while We also calculated the r value, i.e., the approximate distance be­
the CdS QDs sereve the role of an acceptor. tween the donor and acceptor, using following equation (4) [16]:
The decrease in the lifetime of AMC further confirms that there was τ 1/6
r = R0 ( ) (4)
an effective energy transfer from our donor system to the acceptor τ′ − τ

8
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

where R0 is the Förster radius, i.e., the distance between donor and 3.6. Understanding the potential of AMC-CdS QDs as light-harvesting
acceptor moieties at which the 50% energy transfer occurs. materials with the help of quantum mechanical studies
The R0 value was calculated using equation (5) [16,53]:
The present nanohybrid AMC-CdS QDs involves an efficient energy
transfer phenomenon, observed in the case of potentially light-
1
kp2 φd J(λ) 6
R0 = 0.211( ) (5) harvesting materials, such as porphyrin-graphene nanohybrid or
ηd4
porphyrin-CdTe QDs used in solar cells [18,19,57]. Hence, the AMC-CdS
As shown in this equation, the Förster radius (R0) depends on the QDs can be explored for their potential as light-harvesting materials,
relative spatial orientation of the transition dipoles (kp) of the donor. Its which can be used to develop solar cells.
value was taken to be 2/3 assuming the random rotational diffusion of Our experimental results indicate the longer carrier lifetime in the
the donor in the excited state. R0 also depends on the refractive index of AMC-CdS QD nanohybrid system when compared to its precursor, the
the medium (ƞd = 1.353, 40% ethanol) [55] and the quantum yield of MPA-capped CdS QDs. To further understand the reason behind the
the donor, i.e., AMC (φd = 0.15) (see SI). In the above expression, J(λ) extended lifetime of the nano-hybrid system, we have performed theo­
denotes the overlap integral of the donor’s fluorescence emission spec­ retical calculations to understand the relative energy level alignments of
trum and the acceptor’s excitation spectrum. J(λ) can be expressed using the AMC-CdS QDs.
equation (6) [53,56]: Life-time fluorescence measurements and other spectral data of
∫∞ AMC-CdS QDs and their precursors, e.g., AMC and MPA-CdS QDs, reveal
FD (λ)εA (λ)λ4 dλ the occurrence of the FRET from the carbazole moieties (donor) to the
CdS QDs (acceptor). Furthermore, upon photoexcitation of the surface-
J(λ) = n ∫∞ (6) anchored carbazole moieties, FRET from these molecules to the CdS QDs
FD (λ) occurs, followed by back-transfer of the hole from the QDs to the car­
n bazoles [29]. Therefore, the electron and hole charge carriers are
generated in the AMC-CdS QD nanohybrid system upon light absorption.
Where FD (λ) represents the fluorescence intensity of the donor in the
To see the spatial position of electron and hole, we plotted the charge
absence of the acceptor, ƐA (λ) is the molar extinction coefficient of the
densities of the highest occupied molecular orbital (HOMO) and lowest
acceptor and λ is the wavelength in nm.
unoccupied molecular orbital (LUMO) of the hybrid system.
We have calculated the J(λ) value to obtain the R0 value or distance
From Fig. 6 (a and b), it can be seen that the HOMO was localized on
(r) between the acceptor (MPA-capped CdS QDs) and donor (AMC)
carbazole moiety, whereas the LUMO was localized on the CdS QDs. This
molecules. Fig. 4d shows the wavelength range representing the spectral
ensures a type-II band alignment of AMC-CdS QDs, where the electron
overlap of the emission spectra of AMC and the excitation spectra of the
and hole reside on two different counterparts of this hybrid system. The
MPA-capped CdS QDs was found to be 357–409 nm. An excel spread­
type-II band alignment was also reflected from the projected density of
sheet was used to sum up the total normalized fluorescence intensity of
states (PDOS), as shown in Fig. 6 (c and d). The relative positions of the
AMC in the wavelength range of 357–409 nm and using this data, we
PDOS of the CdS QDs and AMC reveal that HOMO was dominated by the
determined FD (λ) to be 37.40. The total value of the wavelength-
carbazole moiety, whereas the CdS QDs control the LUMO. Therefore,
dependent molar extinction co-efficient ƐA (λ) of the MPA-CdS QDs for
this nano-hybrid shows a clear spatial charge separation after photo­
the wavelength range of 357–409 nm was calculated to be 43165.81
excitation and the electron-hole recombination was hindered. Such a
M− 1cm− 1. The value of λ4 (357–409 nm) was 1.17 × 1012 nm4. Using
phenomenon prolongs the carrier lifetime, making AMC-CdS QDs
these parameters, the overlap integral J(λ) value for the present nano­
worthy of being explored as light-harvesting materials in the future.
hybrid system was estimated to be 5.05 × 1016 M− 1cm− 1nm4. Using this
J(λ) value, the R0 [equation (5)] and r [equation (4)] values were
4. Conclusions
calculated to be 7.0 and 7.40, respectively. So, in our system, the donor
and acceptor are situated within the range of the Förster radius. This
In the present work, we have covalently linked a structurally simple
confirms the occurrence of an efficient FRET between the surface-
carbazole moiety, 3-amino-N-methyl carbazole (AMC), with character­
anchored carbazole molecules (i.e., donor species) and the acceptor,
istic chromophoric/fluorophoric properties with MPA-capped CdS QDs
MPA-capped CdS QDs.
to obtain an organic–inorganic nanohybrid material (AMC-CdS QDs).
The rate of energy transfer (kt) was also calculated using equation (7)
The synthetic protocol used was very simple, cost-effective, and easily
[16,54]:
implementable in any laboratory set-up. Furthermore, the AMC-CdS
1 R0 QDs with light-harvesting potential show appreciable photo-, thermal
kt = ( ′ )( )6 (7)
τ r and moisture stabilities. Detailed photophysical characterization of this
nanohybrid system confirmed the occurrence of FRET from the carba­
where kt is the energy transfer rate and τ′ represents the average lifetime zole moieties (donor) to the MPA-capped CdS QD-surface (acceptor) in
of the donor. R0 and r indicate the Förster radius and distance between the solution phase as well as in the powdered/thin-film states.
the acceptor and donor, respectively. Furthermore, SCC-DFTB based quantum chemical calculations indicate
The kt value was determined to be 0.14 ns− 1, which was in good the type-II band alignment of the AMC-CdS QD nanohybrid, where the
agreement with the kt value obtained using equation (3), which is based electron and hole reside on its two different components. Therefore, the
on average lifetime values (as discussed earlier). electron–hole recombination process was hindered, indicating the effi­
All of the FRET parameters of the AMC-CdS QDs (Table 2) are in cacy of the AMC-CdS QDs as photovoltaic materials. In the future, the
favour of the AMC-CdS QDs nanohybrid system. energy transfer efficiency of such carbazole decorated fluorescent QDs
could be explored toward the development of FRET-based light-har­
vesting materials. The AMC-CdS QDs response to photoexcitation at
Table 2
FRET parameters for the AMC-CdS QDs conjugate system. several λexvalues and their emission over a broad range (399–612 nm)
have been observed. This property of AMC-CdS QDs also indicates their
τ τ’ J(λ)a R0 b rc ktd ΦEe
potential use as white light-emitting diodes in the future. On the other
φd
(ns) (ns) (M− 1cm− 1nm4) (nm) (nm) (ns− 1)
hand, the presence of bio-friendly and non-toxic carbazole moieties on
0.15 3.10 5.50 5.05 × 1016 7.10 7.40 0.14 0.44
fluorescent AMC-CdS QDs opens up their applications in the develop­
a b c
Calculations were carried out using equation (6), (5), (4), (3) and (2) . d e
ment of high-accuracy biomedical diagnostic tools.

9
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

Fig. 6. HOMO and LUMO charge densities


plots obtained for the (a) AMC-Cd17S17 QDs
and (b) AMC-Cd55S55 QDs nanohybrid sys­
tems. The red surface indicates the HOMO
and the blue surface indicates the LUMO.
The ochre, yellow, cyan, red, blue and
white balls denote cadmium, sulphur, car­
bon, oxygen, nitrogen, and hydrogen. The
projected density of states (PDOS) of the
AMC-CdS QD nanohybrid system for two
different QD sizes: Cd17S17 QD (c) and
Cd55S55 QD (c). The Fermi energy level was
set to zero.

Conflicts of interest References

The authors have no conflicts to disclose. [1] V.T. Chebrolu, H.-J. Kim, Recent progress in quantum dot sensitized solar cells: an
inclusive review of photoanode, sensitizer, electrolyte, and the counter electrode,
J. Mater. Chem. C 7 (2019) 4911–4933.
Data availability statement [2] Z. Pan, H. Rao, I. Mora-Seró, J. Bisquert, X. Zhong, Quantum dot-sensitized solar
cells, Chem. Soc. Rev. 47 (2018) 7659–7702.
[3] A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo, H. Pettersson, Dye-sensitized solar cells,
Data that support the findings of this study are available from the Chem. Rev. 110 (2010) 6595–6663.
corresponding author upon reasonable request. [4] M. Grätzel, Recent advances in sensitized mesoscopic solar cells, Acc. Chem. Res.
42 (2009) 1788–1798.
[5] G.H. Carey, A.L. Abdelhady, Z. Ning, S.M. Thon, O.M. Bakr, E.H. Sargent, Colloidal
Author statement quantum dot solar cells, Chem. Rev. 115 (2015) 12732–12763.
[6] M.M. Lee, J. Teuscher, T. Miyasaka, T.N. Murakami, H.J. Snaith, Efficient hybrid
solar cells based on meso-superstructured organometal halide perovskites, Science
Tamanna Mallick: Conceptualization, Methodology, Investigation,
338 (2012) 643–647.
Original draft preparation, Abhijit Karmakar.: Investigation, Data [7] I. Joseph, H. Louis, T. Unimuke, I. Etim, M. Orosun, J. Odey, An overview of the
curation, Moumita Kar: Software, Computation, Validation, Sourav operational principles, light harvesting and trapping technologies, and recent
Dutta: Experiment, Sudip Kumar Mondal: Data analysis, Debabrata advances of the dye sensitized solar cells, Appl. Sol. Energy 56 (2020) 334–363.
[8] I.J. Kramer, E.H. Sargent, Colloidal quantum dot photovoltaics: a path forward,
Mandal: Data analysis, Validation, Anup Pramanik: Computation, Data ACS Nano 5 (2011) 8506–8514.
Validation, Naznin Ara Begum: Supervision, Reviewing and Editing. [9] P.V. Kamat, K. Tvrdy, D.R. Baker, J.G. Radich, Beyond photovoltaics:
semiconductor nanoarchitectures for liquid-junction solar cells, Chem. Rev. 110
(2010) 6664–6688.
Declaration of competing interest [10] I.J. Kramer, E.H. Sargent, The architecture of colloidal quantum dot solar cells:
materials to devices, Chem. Rev. 114 (2014) 863–882.
[11] L. Li, X. Yang, J. Gao, H. Tian, J. Zhao, A. Hagfeldt, L. Sun, Highly efficient CdS
The authors declare that they have no known competing financial quantum dot-sensitized solar cells based on a modified polysulfide electrolyte,
interests or personal relationships that could have appeared to influence J. Am. Chem. Soc. 133 (2011) 8458–8460.
[12] A. Kay, M. Grätzel, Low cost photovoltaic modules based on dye sensitized
the work reported in this paper.
nanocrystalline titanium dioxide and carbon powder, Sol. Energy Mater. Sol. Cell.
44 (1996) 99–117.
Acknowledgments [13] H. Wang, Y. Bai, H. Zhang, Z. Zhang, J. Li, L. Guo, CdS quantum dots-sensitized
TiO2 nanorod array on transparent conductive glass photoelectrodes, J. Phys.
Chem. C 114 (2010) 16451–16455.
This work was supported by DSTBT, GoWB, India (Grant No. 1934 [14] M. Grätzel, Photoelectrochemical cells, Nature 414 (2001) 338–344.
(Sanc.)/ST/P/S&T/15G-20/2019 to NAB), DST-PURSE Program [SR/ [15] X. Lan, M. Chen, M.H. Hudson, V. Kamysbayev, Y. Wang, P. Guyot-Sionnest, D.
V. Talapin, Quantum dot solids showing state-resolved band-like transport, Nat.
PURSE Phase 2/42(G) & Phase 2/42(C)] of Visva-Bharati and DST-FIST Mater. (2020) 1–7.
programs of the Dept. Of Chemistry, Visva-Bharati are acknowledged for [16] A. Diac, M. Focsan, C. Socaci, A.-M. Gabudean, C. Farcau, D. Maniu, E. Vasile,
instrumental facilities. TM, AK and SD thank UGC for their fellowships A. Terec, L.M. Veca, S. Astilean, Covalent conjugation of carbon dots with
rhodamine B and assessment of their photophysical properties, RSC Adv. 5 (2015)
[MANF and UGC-CSIR (NET), respectively]. We sincerely thank Prof. P.
77662–77669.
Sarkar, Department of Chemistry, Visva-Bharati, for his support in [17] P.K. Santra, P.V. Kamat, Tandem-layered quantum dot solar cells: tuning the
computational studies. We also thank Mr. P. Majee, Department of photovoltaic response with luminescent ternary cadmium chalcogenides, J. Am.
Chemistry, Visva-Bharati, for his help in analysing the fluorescence Chem. Soc. 135 (2013) 877–885.
[18] B. Rajbanshi, P. Sarkar, Optimizing the photovoltaic properties of CdTe quantum
lifetime data. Finally, we acknowledge CRNN, University of Calcutta, dot-porphyrin nanocomposites: a theoretical study, J. Phys. Chem. C 120 (2016)
India for the HRTEM facility. 17878–17886.
[19] B. Mandal, S. Sarkar, P. Sarkar, Theoretical studies on understanding the feasibility
of porphyrin-sensitized graphene quantum dot solar cell, J. Phys. Chem. C 119
Appendix A. Supplementary data (2015) 3400–3407.
[20] B.S. Rao, B.R. Kumar, V.R. Reddy, T.S. Rao, G. Chalapathi, Preparation and
characterization of CdS nanoparticles by chemical co-precipitation technique,
Supplementary data to this article can be found online at https://doi.
Chalcogenide Lett. 8 (2011) 177–185.
org/10.1016/j.jpcs.2022.110603.

10
T. Mallick et al. Journal of Physics and Chemistry of Solids 164 (2022) 110603

[21] P. Jadhav, G.R. Bhand, K. Mohite, N. Chaure, CdS quantum dots synthesized by [39] C. Zhang, J. Luo, Y. Yu, S. Yang, C. Pan, G. Yu, Building carbazole-decorated
low-cost wet chemical technique, in: AIP Conference Proceedings, AIP Publishing styrene–acrylic copolymer latexes and films for iron (III) ion detection, Colloids
LLC, 2017, p. 50146. Surf. A Physicochem. Eng. Asp. 629 (2021) 127487.
[22] A. Razzaq, J.Y. Lee, B. Bhattacharya, J.-K. Park, Surface treatment properties of [40] J. Keyvan Rad, A.R. Mahdavian, H. Salehi-Mobarakeh, A. Abdollahi, FRET
CdS quantum dot-sensitized solar cells, Appl. Nanosci. 4 (2014) 745–752. phenomenon in photoreversible dual-color fluorescent polymeric nanoparticles
[23] A.K. Bansal, F. Antolini, S. Zhang, L. Stroea, L. Ortolani, M. Lanzi, E. Serra, based on azocarbazole/spiropyran derivatives, Macromolecules 49 (2016)
S. Allard, U. Scherf, I.D.W. Samuel, Highly luminescent colloidal CdS quantum dots 141–152.
with efficient near-infrared electroluminescence in light-emitting diodes, J. Phys. [41] N. Mahapatra, S. Panja, A. Mandal, M. Halder, A single source-precursor route for
Chem. C 120 (2016) 1871–1880. the one-pot synthesis of highly luminescent CdS quantum dots as ultra-sensitive
[24] J. Zhang, Y. Guo, H. Fang, W. Jia, H. Li, L. Yang, K. Wang, Cadmium sulfide and selective photoluminescence sensor for Co 2+ and Ni 2+ ions, J. Mater. Chem. C
quantum dots stabilized by aromatic amino acids for visible light-induced 2 (2014) 7373–7384.
photocatalytic degradation of organic dyes, New J. Chem. 39 (2015) 6951–6957. [42] P. Kumar, D. Kukkar, A. Deep, S. Sharma, L. Bharadwaj, Synthesis of
[25] M. Budyka, O. Chashchikhin, Spectral and photochemical properties of hybrid mercaptopropionic acid stabilized CDS quantum dots for bioimaging in breast
organic-inorganic nanosystems based on CdS quantum dots and a styrylquinoline cancer, Adv Mat Lett 3 (2012) 471–475.
ligand, High Energy Chem. 50 (2016) 349–355. [43] T. Mallick, A. Karmakar, D. Mandal, A. Pramanik, P. Sarkar, N.A. Begum,
[26] I.L. Medintz, S.A. Trammell, H. Mattoussi, J.M. Mauro, Reversible modulation of Harnessing carbazole based small molecules for the synthesis of the fluorescent
quantum dot photoluminescence using a protein-bound photochromic fluorescence gold nanoparticles: a unified experimental and theoretical approach to understand
resonance energy transfer acceptor, J. Am. Chem. Soc. 126 (2004) 30–31. the mechanism of synthesis, Colloids Surf. B Biointerfaces 172 (2018) 440–450.
[27] S. Saeed, P.A. Channar, A. Saeed, F.A. Larik, Fluorescence modulation of CdTe [44] S. Sarkar, S. Pal, P. Sarkar, A. Rosa, T. Frauenheim, Self-consistent-charge density-
nanowire by azobenzene photochromic switches, J. Photochem. Photobiol. Chem. functional tight-binding parameters for Cd–X (X= S, Se, Te) compounds and their
369 (2019) 159–165. interaction with H, O, C, and N, J. Chem. Theor. Comput. 7 (2011) 2262–2276.
[28] M. Zorn, W.K. Bae, J. Kwak, H. Lee, C. Lee, R. Zentel, K. Char, Quantum Dot− block [45] S. Sarkar, S. Pal, P. Sarkar, Electronic structure and band gap engineering of CdTe
copolymer hybrids with improved properties and their application to quantum dot semiconductor nanowires, J. Mater. Chem. 22 (2012) 10716–10724.
light-emitting devices, ACS Nano 3 (2009) 1063–1068. [46] S. Sarkar, S. Saha, S. Pal, P. Sarkar, Electronic structure and bandgap engineering
[29] F.S. Morgenstern, A. Rao, M.L. Böhm, R.J. Kist, Y. Vaynzof, N.C. Greenham, of CdTe nanotubes and designing the CdTe nanotube–fullerene hybrid
Ultrafast charge-and energy-transfer dynamics in conjugated polymer: cadmium nanostructures for photovoltaic applications, RSC Adv. 4 (2014) 14673–14683.
selenide nanocrystal blends, ACS Nano 8 (2014) 1647–1654. [47] S. Sarkar, S. Saha, S. Pal, P. Sarkar, Electronic structure of Thiol-capped CdTe
[30] V.K. Komarala, Y.P. Rakovich, A. Bradley, S.J. Byrne, Y.K. Gun’ko, N. Gaponik, quantum dots and CdTeQD–carbon nanotube nanocomposites, J. Phys. Chem. C
A. Eychmüller, Off-resonance surface plasmon enhanced spontaneous emission 116 (2012) 21601–21608.
from CdTe quantum dots, Appl. Phys. Lett. 89 (2006) 253118. [48] R.S. Yadav, P. Mishra, R. Mishra, M. Kumar, A.C. Pandey, Histidine functionalised
[31] J. Li, D. Bao, X. Hong, D. Li, J. Li, Y. Bai, T. Li, Luminescent CdTe quantum dots and biocompatible CdS quantum dots synthesised by sonochemical method, J. Exp.
nanorods as metal ion probes, Colloids Surf. A Physicochem. Eng. Asp. 257 (2005) Nanosci. 5 (2010) 348–356.
267–271. [49] P. Thangadurai, S. Balaji, P. Manoharan, Surface modification of CdS quantum dots
[32] H.-J. Knölker, K.R. Reddy, Isolation and synthesis of biologically active carbazole using thiols-structural and photophysical studies, Nanotechnology 19 (2008)
alkaloids, Chem. Rev. 102 (2002) 4303–4428. 435708.
[33] T. Watanabe, S. Oishi, N. Fujii, H. Ohno, Palladium-catalyzed direct synthesis of [50] R.F. Egerton, PHYSICAL PRINCIPLES of ELECTRON MICROSCOPY : an
carbazoles via one-pot N-arylation and oxidative biaryl coupling: synthesis and Introduction to TEM, SEM, and AEM, SPRINGER, 2018.
mechanistic study, J. Org. Chem. 74 (2009) 4720–4726. [51] B.N. Ganguly, S. Dutta, S. Roy, J. Röder, K. Johnston, M. Martin, ISOLDE-
[34] L. Kortekaas, F. Lancia, J.D. Steen, W.R. Browne, Reversible charge trapping in bis- Collaboration, Investigation on structural aspects of ZnO nano-crystal using radio-
carbazole-diimide redox polymers with complete luminescence quenching active ion beam and PAC, Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact.
enabling nondestructive read-out by resonance Raman spectroscopy, J. Phys. Mater. Atoms 362 (2015) 103–109.
Chem. C 121 (2017) 14688–14702. [52] D.K. Gupta, M. Verma, K. Sharma, N. Saxena, Synthesis, characterization and
[35] J.V. Grazulevicius, I. Soutar, L. Swanson, Photophysics of carbazole-containing optical properties of CdSe/CdS and CdSe/ZnS core-shell nanoparticles, Indian J.
systems. 3 Fluorescence of carbazole-containing oligoethers in dilute solution, Pure Appl. Phys. 55 (2017) 113–121.
Macromolecules 31 (1998) 4820–4827. [53] J.R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer Science &
[36] N.A. Begum, N. Roy, S. Mandal, S. Basu, D. Mandal, Fluorescence spectroscopy of a Business Media, 2013.
naturally occurring carbazole alkaloid: murrayanine, J. Lumin. 129 (2009) [54] C. Biskup, T. Zimmer, L. Kelbauskas, B. Hoffmann, N. Klöcker, W. Becker,
158–163. A. Bergmann, K. Benndorf, Multi-dimensional fluorescence lifetime and FRET
[37] C. Zhang, J. Luo, L. Ou, Y. Lun, S. Cai, B. Hu, G. Yu, C. Pan, Fluorescent porous measurements, Microsc. Res. Tech. 70 (2007) 442–451.
carbazole-decorated copolymer monodisperse microspheres: facile synthesis, [55] S. Andher, Z. Gadhawala, S. Vyas, K. Goswami, Comparative study of specific
selective and recyclable detection of iron (III) in aqueous medium, Chemistry-A refraction of potassium salts solutions in binary (Ethanol+ water) mixtures, Res. J.
European Journal 24 (2018) 3030–3037. Phys. Sci. 3 (2015) 1.
[38] J.K. Rad, A.R. Mahdavian, Preparation of photoresponsive functionalized acrylic [56] D. Ghosh, N. Chattopadhyay, Gold nanoparticles: acceptors for efficient energy
nanoparticles containing carbazole groups for smart cellulosic papers, Iran. J. transfer from the photoexcited fluorophores, Opt Photon. J. 3 (2013) 18–26.
Polym. Sci. Technol. 30 (2017) 435–446. [57] J.B. Sambur, T. Novet, B. Parkinson, Multiple exciton collection in a sensitized
photovoltaic system, Science 330 (2010) 63–66.

11

You might also like