Suprabha S. Sahoo,1 Dattatray Chadar,1
Manilal Murmu,2, 3 Priyabrata
Banerjee,2,3 Sunita Salunke-Gawali1,*
and R. J. Butcher4
1 Department of
Chemistry, Savitribai Phule Pune University, Pune 411007, India.
2 Surface Engineering
& Tribology Group, CSIR-Central Mechanical Engineering Research Institute,
M.G. Avenue, Durgapur 713209, West Bengal, India.
3 Academy of Scientific
and Innovative Research (AcSIR), AcSIR
Headquarters CSIR-HRDC Campus, Sector-19, Kamla Nehru
Nagar, Ghaziabad, 201002, Uttar Pradesh, India.
4 Department of
Chemistry, Howard University, Washington, D.C, 20059, USA.
*E-mail: sunita.salunke@unipune.ac.in (S. Salunke-Gawali)
Abstract
Three benzo[a]phenoxazine based
photosensitizers known as 10-Chloro-6-methyl-5H benzo[α]phenoxazin-5-one
(BPO-Cl), 6-methyl-5H-benzo[α]phenoxazin-5-one (BPO), and
6-methyl-5H-benzo[α]phenothiazin-5-one (BPT) synthesized from provitamin
K3 were used as photosensitizers to fabricate a DSSC device. The inbuilt
intra-molecular charge transfer (ICT), favorable
π–π interaction, planar shape, and compatible redox nature with
appropriate frontier molecular orbitals (HOMO and LUMO) alignment are the
significant physicochemical characteristics of benzo[α]phenoxazine moiety
that rationalize to become an efficient photosensitizer. Thus, the bare benzo[a]phenoxazine dyes without any efficient anchor and
donor group are evaluated here with their intrinsic chemical behavior for efficient precursors in various energy
transferring material. The intramolecular charge transfer occurring through
aromatic rings bridged by two heteroatoms (O/S and N) is tunable
with DFT analysis and single-crystal XRD data. The characteristic solar cell
parameters show that chlorine substitution enhances the open-circuit voltage (Voc) from 0.53 V in BPO to 0.56 V in
BPO-Cl. A noticeable enhancement in short-circuit current density (Jsc) nearly fivefold is observed for BPO
dye compared to its thio-substituted BPT unit. The
substitution of chlorine reduces the recombination rate in the device studied
from electrochemical impedance spectroscopy. Thus BPO
template is an excellent precursor to tailor various D/A groups to produce
effective photosensitizer evaluated by physicochemical properties tunable with several optoelectrical, structural,
morphological, and electrochemical analyses.
Table of Content
Keywords: Benzo[α]phenoxazine; Recombination resistance;
Physicochemical properties; Naphthoquinone; Photosensitizer;
DSSC.
Introduction
The
application of organic dyes in dye-sensitized solar cells (DSSC),[1] organic solar cells (OSCs),[2] OLEDs, sensors,[3,4] storage devices,[5] to attain achievable synthetic targets with desired
moieties is one of the highly explored and demanded areas nowadays. Large
extinction coefficient, non-toxicity, cost-effectiveness, tunable
optical properties, adequate redox
nature, and facile molecular engineering are the
effective positive attributes of organic dyes over metal-complex based dyes.[6,7] Hence
nontoxic, efficiently synthesized organic precursors has been engineered these
days to produce several types of effective D-π-A, D-D-π-A, porphyrin,
squaraine architecture[8]
of dyes by tailoring different donor(D) and acceptor(A) moieties to the organic
framework as a central stem. In this context, numerous publications are
reported with various organic frameworks like xanthenes[9,10] coumarins,[11-14] squaraine,[15,16] anthracene,[17,18] bodipy,[19-21]etc.
used in DSSC. However, benzo[α]phenoxazine dyes are significantly less
explored despite their planar shape, favorable
π–π interaction with good intra-molecular charge transfer (ICT); for
an efficient DSSC photosensitizer.[22]
Photosensitizers are the primary controlling unit of
DSSC, so numerous reports are available by varying the functionality of
different organic moieties in the photosensitizer. Like
benzo[α]phenoxazine molecule, this can be functionalized as both donor and
π-linker. However, very few reports are available with phenoxazine
moieties, which can perform a dual function as donor and π-linker in
specific D-π-A dyes such as Tian and his co-workers reported phenoxazine
as an efficient donor (D) for DSSC dyes.[23] Similarly, Tan and his team employed two phenoxazine-based
dyes where phenoxazine and substituted phenoxazine were used as donor molecule;[24] Wenjun et al. reported phenothiazine as a hole
transporting material in organic semiconductor devices and used them as a
π-linker as well as a donor molecule.[25] Further, Tian and his team used phenoxazine unit as
both linker and donor moiety,[26] whereas benzo[α]phenoxazine moieties were used
as chromophores for DSSC by Schroder et al.[22] and benzo[a]phenothiazine was used by Bhand et al. as a photosensitizer with Zirconium
semiconducting metal oxide.[27] However, benzo[α]phenoxazine dyes have not yet
been explored as photosensitizers with or without any other tailored group. The
positive attributes of benzo[α]phenoxazine over phenoxazine moiety is it’s; i) extended absorbance,
which can help to harness more solar radiation, ii) extended conjugation with
planarity, iii) adequate redox nature, iv) proper alignment of molecular
orbitals (HOMO, LUMO), v) inbuilt internal charge transport through the moiety
(ICT) and vi) prone to agile molecular engineering with flexible sites for
various D/A substitution.
Herein, two benzo[α]phenoxazine derivatives,
namely, 6-methyl-5H-benzo[α]phenoxazin-5-one (BPO), chlorine substituted benzo[a]phenoxazine that is
10-Chloro-6-methyl-5H-benzo[α]phenoxazin-5-one (BPO-Cl), and one
benzo[α]phenothiazine molecule such as
6-methyl-5H-benzo[α]phenothiazin-5-one (BPT) were utilized as
photosensitizers with nanoporous TiO2 photoanode
prepared through doctor blade technique. A comparative study of all the dyes'
physicochemical properties was studied to find out the impact of substitution
on the moiety. The detailed morphological, photo-physical, electrochemical, and
solar cell characterizations were analyzed, showing
chlorine substituted dye (BPO-Cl) is the potential candidate for photosensitization.
Furthermore, to evaluate the dye's inherent chemical potentiality, single
crystal XRD and DFT analysis were tuned with electrochemical analysis. The
primary reason that impedes solar cell parameters like VOC and JSC
in the device is the aggregation and lack of efficient donor and anchoring
moiety because the charge collection efficiency of these dyes is more than ~
99% means the dyes are quite efficient to absorb the solar radiation and can
produce a high amount of charge carrier but are not efficient to separate the
charge carriers to generate electricity. Herein, in this study, the addition of
simple chlorine substitution manages the recombination mechanism in the device
to many folds. This exemplifies an efficient D/A group at the particular sites in the moiety that will enhance the dyes
solar cell effectiveness by controlling its recombination dynamics. So to get an efficient photosensitizer with phenoxazine
functionality, benzo[α]phenoxazine is a suitable precursor to being work
upon with an adequate D/A group for which all the mandatory physicochemical
properties are studied in this investigation.
2. Experimental section and computational Methods
2.1 Materials
The
Fluorine doped tin oxide (FTOs) were used as a source of conducting glass having
sheet resistance 13-18 Ω/cm2 purchased from Sigma Aldrich used
in the photoanode preparation. P25 Degussa (Nanoshell
LLC, US) was used as a source of titanium nanopowder.
The chemicals associated with photoanode and photosensitizer preparation like
ethanol (Merck), acetonitrile (Merck), acetylacetone (SRL), ethylcellulose
(SRL), and α-terpineol (kemphasol) were used as
received. Iodine/triiodide solution used as an electrolyte in the device
composed of Lithium iodide (SRL, India), iodine(Sigma
Aldrich), tert butyl pyridine (ACROS Organics,
Belgium) also used directly without any
further purification. The synthetic procedure and the characterization of
photosensitizer dye with benzo[α]phenoxazine and
benzo[α]phenothiazine template are available in the previous report.[28]
2.2 Device fabrication
The
device was constructed using a nanocomposite mesoporous TiO2 photoanode
as a working electrode prepared by the doctor blade technique with FTO
substrate. A compact layer designed from TiCl3 and NaOH with basic
pH was prepared by the Chemical Bath Deposition (CBD) method. After calcination
with 450 °C, a mesoporous uniform layer of TiO2 paste was applied on
the compact layer and again annealed at 450 °C to get the desired photoanode's
specific morphology. The detailed synthetic method was reported in our previous
paper.[28]
The synthetic procedure used for the preparation of BPO, BPO-Cl, and BPT is
taking place by condensation of provitamin K3 (2-methyl-1,4-naphthoquinone)
with derivatives of 2-aminophenol and 2-aminothiophenol to produce
benzo[α]phenoxazine and benzo[α]phenothiazine moiety. Specifically,
carbonyl group present on the quinonoid ring,
heteroatoms (O/S, N) in the form of oxo/thio, and
imine group entangled in the moiety can be used extra anchoring sites. In
addition to this, the total molecule acts as a p-donor and a p-linker due to its very good redox property and efficient intramolecular charge
transfer (ICT) through the heteroatoms. One way electron donation property
(switching on nature) and π-π stacking interactions are these dyes'
significant characteristics. (I¯/I3¯) used as
electrolyte composed of lithium iodide 0.5 M, 0.05 M iodine, and 0.15 M
tertbutyl pyridine in 100 ml acetonitrile as per laboratory reported method[29] followed by platinum-coated FTOs as the counter
electrode. Both the working and counter electrodes were separated through the
spacer and connected through the electrolyte to regenerate the dye molecule and
maintain electron neutrality in the device through charge transport via the
electrolyte.
2.3 Solar cell device characterization
The
absorption and emission spectra were measured with SHIMADZU UV 1650
spectrophotometer from 200 nm to 600 nm and JASCO spectrofluorometer FP-8300
between 450 and 700 nm. To get a better insight into the excited state
lifetime, the fluorescence lifetime of the samples was measured with a
time-correlated single-photon counter (TCSPC) method using a light-emitting
diode (nano-LED) excited with 450 nm by Horiba Fluorolog
FL3 (FL-1057) instrument with fast photomultiplier tubes (PMTs). The decay data
were analyzed using commercial software given by the
Horiba instrument. FT-IR spectra were obtained from BRUKER VERTEX by SHIMADZUFT
8400 Spectrometer between 4000 - 400 cm−1 by ATR technique.
The electrochemical analysis was performed with CHI 6054E Electrochemical
Analyser (CHI 660) at a scan rate of 0.1 – 0.5 Vs-1 in acetonitrile
against Ag/AgNO3 (BAS) as the reference electrode, Platinum disc
electrode (CHI102, surface area 0.025 cm2) as working electrode and
Platinum wire electrode (CHI115) as the counter electrode at room temperature
(26 °C). Before commencing measurement each time, the analyte was deoxygenated by
nitrogen purging. Field emission scanning electron microscopy, FESEM, (Carl
Zeiss, Merlin Compact), was conducted to study the corresponding photoanode's
morphology and elemental composition. Solar cell parameters are checked with a
solar simulator in association with Keithley 2420 source meter for
current-voltage measurements under 1 Sun.
Electrochemical impedance spectroscopy (EIS) was carried out using a
two-electrode system between -1 Hz to -10 kHz frequency range to study the
electron dynamics in the device under dark conditions.
2.3.1 The single-crystal XRD data collection
X-ray
quality red crystals of BPO were obtained
after evaporation of pure red fraction from the column. Crystals of the
appropriate size were chosen for data collection. Data was collected through D8
Venture PHOTON 100 CMOS diffractometer using graphite monochromatized Mo-Ka radiation (l = 0.7107 Å) with exposure/frame = 10 sec for molecule
A. The X-ray generator was operated at 50 kV and 30 mA for Mo-Ka
radiation. An initial set of cell constants and orientation matrix were
calculated from 24 frames and 60 frames for Mo and Cu source, respectively. The
optimized strategy used for data collection consisted of different j
and w scans with 0.5° steps in j/w.
Crystal to detector distance was 5.00 cm with 512 x 512 pixels/frame,
Oscillation/frame -0.5°, maximum detector swing angle was –30.0°, beam centered at (260.2, 252.5) and in-plane spot width of 1.24
unit. Bruker SAINT Program and Bruker SADABS were employed for data integration
and empirical absorption correction for intensity data. The program was
integrated into the APEX II package. The data were corrected for Lorentz and
polarization effect. The structure was solved by Direct Method using SHELX-97.[30] The final refinement of the structure was performed
by full-matrix least-squares techniques with anisotropic thermal data for
non-hydrogen atoms on F2. The non-hydrogen atoms were refined anisotropically, whereas the hydrogen atoms were refined at
calculated positions as riding atoms with isotropic displacement parameters.
Molecular diagrams were generated using ORTEP 3[31] and Mercury programs.[32]
2.3.2 Computational Studies
All
the geometrical optimizations were performed by the ORCA program package
(Version 2.7.0) employing hybrid B3LYP functional with TZV(P) and SV(P) basis
set.[33]Following geometry optimization, the time dependant –
density functional theory (TD-DFT) calculation was performed using Turbomole (V.7.0) with TmoleX
interface 4.1.1.[34-37] The COSMO model was used to incorporate the effect of
the solvent. Molecular orbitals were plotted by Molekel
4.3 software. In this experimentation, acetonitrile has been used as a solvent.
The ORCA program was used to calculate the
contributions of molecular orbitals for the organic entities. The
highest occupied molecular orbital (HOMO) and lowest unoccupied molecular
orbital (LUMO) values were calculated using B3LYP functional to evaluate the
electron density at the ground state and excited state. The plots of molecular orbitals
were visualized and generated using Molekel 4.3
software. Similarly, to obtain the data's precision and reliability, TD-DFT
computations were carried out with oscillatory strength calculations.
3. Results and discussion
3.1 Photosensitizers Characterization
3.1.1 Single Crystal XRD analysis of BPO Dye
ORTEP
diagram of BPO (CCDC 2018650) is shown in Fig.1(a). Single crystal X-ray structure data is
presented in Table S1 and hydrogen bonding
interaction in Table S2. BPO crystallizes in
tetragonal space group P42/n. There are two molecules in the asymmetric
unit labelled in Fig.1(b)
(A and B). They are deferred primarily by bond distances as well as close
contacts with other molecules. Carbonyl distances C(14B)-O(2B) is 1.251 Å while
it is longer in C(14A)-O(2A) (1.249 Å) (Fig.1 (b),
Table S3); however, in the range of oxidized
form of the naphthoquinones. Longer C(1A)-O(1A) distance ~1.416 Å is observed
in one of the asymmetric molecules A. Fig. S1(Supplementary
Fig. S1) showed the atoms involved in hydrogen
bonding interaction and p-p stacking interaction. The asymmetric molecules are
joined by the C-H∙∙∙O hydrogen bonding interaction, where the
C(3)-methyl is involved in this interaction. Both the
asymmetric molecules also showed p-p stacking interaction; the only difference is the
number of carbons taking part in this interaction. Asymmetric molecule A
(green) is in the vicinity of four similar molecules (Fig.1(c)),
while molecule B is in the vicinity of three similar molecules via p-p
stacking and C-H∙∙∙O interaction. p-p
stacking of two equivalent symmetry molecules is shown in Fig.1(c), and Fig. S2
is provided in the supplementary information file. Alternate p-p
stacked chain of type A-A and B-B is observed; chains are joined by
C-H∙∙∙O interaction. The molecular packing of BPO is shown in
Fig.1(d). The tetragonal axis lies at the center of the square. The single-crystal XRD data of BPO-Cl
and BPT are reported previously.[28] The herring borne architecture of BPO-Cl and
butterfly architecture of BPT molecule facilitates reduce aggregation and
intermolecular excimer formation[25] of dye molecule and hence capable of enhancing more
solar cell efficacy.
(a)
Molecule B Molecule A
(b)
(c)
(d)
Fig. 1 (a)
ORTEP of BPO dye, the ellipsoids were drawn with a 50% probability, (b) Bond
distances in asymmetric molecules of BPO, (c) p-p
stacking interaction of symmetry equivalent molecules of BPO, (d) Molecular
packing of BPO down c-axis.
Fig.2 UV-Visible spectra of dye, TiO2 film, and
dye loaded TiO2 film, (a) BPO, (b) BPO-Cl, (c) BPT, (d) shift in
conduction band after dye loading, (e) excitation and emission spectra
together, (f) Excited state lifetime data of BPO, BPO-Cl and BPT dye.
3.1.2 Optical study
Experimentally
measured absorption spectra of BPO, BPO-Cl, and BPT sensitizers in acetonitrile
show two distinctive peaks near 350 nm and 450 nm, as shown in Fig.2 (a, b, c). In all the three dyes, the
absorption band near 440 nm was attributed to n-π* transition,
and the band ~ 350 nm was assigned to π-π* transition. In
BPT, the n-π* transition is observed at a more extended wavelength
region near 474 nm, as shown in Fig.2(c). In Fig.2(a, b, c), a comparison between dye loaded film,
blank TiO2 film, and basic dye shows a bathochromic shift of the dye
molecule after absorption into
the
film, confirming a metal to ligand charge transfer (MLCT). The bathochromic
shift towards the visible region of the solar spectrum and hyperchromic shift
observed in Fig. 2 (a, b, c) declares
complexation between TiO2 and dye molecule, followed by an adequate
amount of dye loading in the film, respectively. The optical band gap of TiO2
photoanode was found to be 3.2 eV. After dye loading, the Fermi level
shift was very prominent in all three dyes, as shown in Fig.2(d). The compound's optical bandgap after dye loading is 3.04
eV for BPO-Cl, 3.166 eV for BPO, and 2.13 eV for BPT concerning TiO2 (3.25
eV) before dye loading.
3.1.3 Emission Spectra
To
study the emission behavior, fluorescence spectra
have been recorded between 450 nm –700 nm, as shown in Fig.2(e).
It shows emission maxima at 558 nm, 556 nm, 571 nm with Stokes
shift of 118 nm, 105 nm, and 97 nm for BPO, BPO-Cl, and BPT dyes. The
electrical bandgap denoted by E0-0 was calculated to be 2.52 eV for
BPO and 2.49 eV and 2.33 eV for BPO-Cl, and BPT respectively, which are less
than the estimated HOMO-LUMO gap value by TD-DFT calculations showed the later
part of this paper. To get a more in-depth insight into photoluminescence
spectra (PL spectra), time-dependent fluorescence spectra with TCSPC
(time-correlated single-photon count) method was recorded, which shows most
extended excited-state lifetime of BPO dye is 1.42 ns followed by 1.35 ns, 1.37
ns for BPO-Cl, and BPT respectively in Fig.2(f) and
Table 1. The data were fitted with double
exponential decay by DAS (Data Analysis software) provided by the same
instrument.
3.1.4 Electrochemical studies
The
electrochemical analysis unveils the dyes' redox nature, which is the most
fundamental parameter to study chemical reactivity and electron transfer
parameters. The three-electrode system performed cyclic voltammetry (CV)
experiments with a different scan rate of 100 – 500 mVs-1. To find out
either of the highest occupied molecular orbital (HOMO) and lowest unoccupied
molecular orbital (LUMO) positions as depicted in Fig.3(a).
The HOMO and LUMO values were calculated with E0-0 values from emission studies vide Fig.2 (e). It can be noticed that BPO possesses the
highest bandgap of 2.52 eV among all the three dyes, whereas BPT shows the
lowest bandgap of 2.33eV. The optoelectrical study
detailed data and electrochemical analysis have been presented in Fig.3(b) and Table 1.
It shows the LUMO level of all the three dyes above the fermi level of TiO2
semiconducting metal oxide. The HOMO level is below the work-function of
the standard redox potential of iodine/triiodide (0.35 V versus Normal Hydrogen
Electrode NHE).[38] The difference of 0.9–1.31 eV provides an adequate
driving force for dye regeneration in the device. The HOMO levels have been
calculated using the formula HOMO = LUMO–Eg (optical bandgap).
Again, it is in agreement with chlorine's substitution,
and sulfur atom comparatively lowers the HOMO level.[39] In BPO-Cl, the electron injection efficiency is the
lowest, facilitating the fastest electron transfer from LUMO to the conduction
band (CB) of the semiconducting metal oxide and hence produces more (Voc and Jsc) solar power
conversion efficiency among the three devices.
Table1 Photophysical and
electrochemical data of dyes and dye loaded films.
Compound Name |
λmax (nm) |
λ PL
(nm) |
Stokes shift(nm) |
aE0-0 (ev) |
bEHOMO (ev) |
cELUMO (ev) |
Excited-state
lifetime (nm) |
BPO |
440 |
558 |
118 |
2.52 |
-5.7 |
-3.18 |
1.426×10-9 |
BPO- Cl |
451 |
556 |
105 |
2.49 |
-6.11 |
-3.62 |
1.349×10-9 |
BPT |
474 |
571 |
97 |
2.33 |
-5.94 |
-3.61 |
1.373×10-9 |
(a) (b)
Fig. 3 a)
Cyclic voltammograms of BPO, BPO-Cl, and BPT dye, b) HOMO-LUMO comparison plot
with CB of TiO2 and redox potential of I¯/I3¯.
FMO |
BPO |
BPO-Cl |
BPT |
Optimized
Structure |
|
|
|
HOMO-1 |
-6.86 eV |
-6.96 eV |
-6.72 eV |
HOMO |
-6.06 eV |
-6.13 eV |
-5.96 eV |
LUMO |
-3.01 eV |
-3.11 eV |
-3.03 eV |
LUMO-1 |
-1.37 eV |
-1.44 eV |
-1.43 eV |
Fig.
4 Optimized geometrical structure and electron density
distribution on Frontier molecular orbital plots of BPO, BPO-Cl, and BPT dye.
HOMO and LUMO values are depicted at an isovalue of 0.06.
3.1.5 Computational studies
A
density functional theory (DFT) study has been carried out to study the
structural property, geometry, and probability of electron density on each
constituent element. The ease of electron flow in the BPO, BPO-Cl, and BPT
moieties through the heteroatoms is screened from the frontier molecular
orbitals (HOMO and LUMO) geometries. The isodensity
plots were shown in Fig. 4 with the optimized
geometry of the dyes. The HOMO and LUMO orbitals' energy level values are very similar to the values obtained from experimental
calculations. A minute difference of + 0.5 eV is observed only in LUMO
values of BPO-Cl and BPT dye, whereas other values are exceptionally tunable (+ 0.1 eV) with the experimental values. An
in-depth investigation on HOMO and LUMO values shows electron cloud is highly
diffused on the quinonoid and heteroatom-infused aromatic rings in HOMO levels.
After absorption of radiation in excitation mode, the electron cloud is moved
towards the benzenoid ring and highly dense on the quinonoid ring near carbonyl
moiety in the LUMO orbitals. However, in both cases, the electron cloud is
highly populated near carbonyl moiety. The DFT analysis revealed the dyes can be functionalized as both donor and
π-linker for photo-sensing activity. The energy values associated with the
moiety and its higher HOMO and LUMO levels are presented in the supplementary
file vide Fig. S4 and Table
S4.
(a) (b)
Fig. 5 (a) XRD pattern of TiO2 film (b)
Crystallite size calculation.
Fig. 6 Morphology
of (a) Compact TiO2 layer, (b) Nanocomposite mesoporous TiO2 and (c) Thickness of the film by FE-SEM technique.
3.2 Photoanode
Characterisation
3.2.1 Structural study
The powder XRD analysis of the mesoporous
nanocomposite TiO2 films reveals its anatase phase and crystalline
nature. The prototypic peak pattern of (101), (004), and (105) planes, as shown
in Fig. 5(a), signifies the percent dominance
of anatase form over the rutile phase. This condition is very conducive for the
charge carrier's propagation through the photoanode material, which is very
advantageous for DSSC application. The crystallite size calculation was done
with Full-width half maxima (FWHM) by Scherer's equation[40] as shown in Fig. 5(b) and
was found to be 35.53 nm.
3.2.2 Morphological Analysis
The elemental and morphological analysis has been
carried out with FE-SEM characterization, which shows the mesoporous nature of the
TiO2 working electrode's surface. The compact layer TiO2
grew by the chemical bath deposition (CBD) technique exhibit the floral
morphology with a highly dense and closely packed
layer below the mesoporous layer Fig. 6(a). The
transverse section of the film was studied to obtain the thickness of the 10.33
μm, as shown in Fig.
6(c). The uniformity and porosity are displayed in Fig. 6(b), which is crucial for dye loading. The
elemental composition of the TiO2 photoanode has been carried out by
Energy Dispersive X-Ray analysis (EDX) presented in supplementary file Fig. S3.
Fig. 7 FT-IR
spectra of basic dye and dye loaded TiO2 film.
3.2.3 FTIR study of dye and dye loaded film
In Fig. 7, the FT-IR
spectra of the dye and dye loaded TiO2 films were compared, which
shows the bonding interaction between the dye and film along with few minor
changes presented in Table 2. In BPO, the
intensity of carbonyl stretching frequency in basic dye became highly intense
in dye loaded film, confirming TiO2 and carbonyl moiety's bonding
interaction with a minute peak shift of 15 cm-11as depicted in Fig. 7(a). Similarly, in BPO-Cl and BPT, the carbonyl
stretching frequency νC=O at
1629 cm-1[41] is shifted to ~1626 cm-1
with an enhanced intensity, which validates the bonding interaction between the
photosensitizer and photoanode as shown in Fig. 7(b) and Fig. 7(c). In all
the dyes, the stretching frequency nearer to 1277 cm-1, 1034 cm-1
in BPO and 1268 cm-1, 1042 cm-1 in BPO-Cl is
shifted to 1239 cm-1, 1072 cm-1, and 1266 cm-1,
1085 cm-1 respectively in dye loaded films ascribed to carbonyl νC-O modes,[28] which confirm the added
interaction between the dye and TiO2 films. However, in BPT, the
peaks near 1235 cm-1 and 1033 cm-1 in the dye molecule
are not distinguished in the dye-loaded film probably is the primary reason for
lower solar cell efficiency due to the lower extent of dye loading. Peaks near
1376 cm-1, 920 cm-1 in the dye molecule are very
prominent in the dye loaded films near 1315 cm-1 and 957 cm-1.
The C-H vibrational frequency near 2900-3000 cm-1 in the dye
molecule is undetectable in the dye-loaded films, whereas only visible in the
BPO molecule. The aromatic protons near 2200-2300 cm-1 in BPO-Cl dye
perhaps involved in the internal hydrogen bonding formation and very prominent
in the BPO-Cl dye-loaded films, as depicted in Fig.
7(b). In BPO-Cl, the νC-Cl
bond frequency was found to be near 737 cm-1. In all the dye
loaded films, the peak near 1574 + 22 cm-1 due to νC=C
stretching frequency is very prominent, confirming the nitrogen atom's
involvement in bonding interaction between the dye and photoanode material.
Thus, the FT-IR spectrum ensures various functional groups present in the dye
molecule and its binding mode with the semiconducting metal oxide TiO2
surface.
Table 2 FT-IR frequencies of BPO,
BPO-Cl, and BPT after and before dye loading
Name |
nC=O (cm-1) |
nC=C (cm-1) |
nC-O (cm-1) |
|||
Dye |
Dye Loaded Film |
Dye |
Dye Loaded Film |
Dye |
Dye Loaded Film |
|
BPO |
1612 |
1627 |
1505 |
1520 |
1277 1034 |
1239 1072 |
BPO-Cl |
1629 |
1626 |
- |
- |
1268 1042 |
1266 1085 |
BPT |
1629 |
1627 |
1596 |
1596 |
1235 1033 |
- - |
3. 3 Solar cell characterization
3.3.1 J-V
Characterisation
The
photovoltaic performance of the dyes was studied by J-V measurement using iodine/tri-iodide liquid electrolyte, the
platinum counter electrode with solar simulator under 60 mW/cm2
light intensity. The solar cell characteristic parameters like
short-circuit current density (Jsc), open-circuit voltage (Voc) are presented in Fig. 8 and Table 3. The precursor dyes without any efficient
donor and anchoring group can produce an open-circuit voltage (Voc) of 0.35 V
with a current density of 85 μA/ cm2
with a low light intensity 60 mW/cm2. To
study the impact of input power source on these dyes, irradiation of 80 mW/cm2 power source was introduced, which
produces enhanced solar cell characteristics. Nearly fivefold improvement for
BPO and BPO-Cl dyes in current density is observed, where as
for BPT dye, the enhancement is nearly ten times more. We observed a noticeable
increment for all the three dyes with enhanced VOC from 0.35 V to 0.53V and JSC from 0.085 mA/cm2 to 0.44 mA/cm2
for BPO dye, VOC from 0.34
V to 0.56 V and JSC from
0.110 mA/cm2 to 0.38 mA/cm2 for BPO-Cl and in BPT dye
open circuit voltage increased from 0.22 V to 0.49 V and short circuit voltage
from 0.047 mA/cm2 to 0.24 mA/cm2 with change in input
power source from 60 mW/cm2 to 80 mW/cm2. Among all the dyes, BPO-Cl possesses the
highest VOC.
Electrochemical impedance spectroscopy is carried out to determine the effect
of electrochemical changes in the device behind this enhancement. It is well
understood that the increase or decrease in VOC
is governed by the potential difference between the quasi-fermi-level (EF,n) of semiconducting metal oxide and the
reduction potential of electrolyte.[42-45]
It is directly influenced by the fluctuating electron density of the dye
molecule's HOMO and LUMO energy levels displayed in computational studies.
Though the exact correlation between VOC,
JSC, and dye architecture
is not detected, the incorporation of chlorine atom greatly impacts solar cell
characteristics.
Fig. 8 Current
density-voltage plot at (a) 60 mW/cm2 (b)
80 mW/cm2.
Table 3 Solar
cell parameters with a different input power source
|
60 mW/cm2 |
80 mW/cm2 |
||
Dyes
Name |
VOC (V) |
JSC (mA/cm2) |
VOC (V) |
JSC (mA/cm2) |
BPO |
0.35 |
0.085 |
0.53 |
0.440 |
BPO-Cl |
0.34 |
0.110 |
0.56 |
0.430 |
BPT |
0.22 |
0.047 |
0.49 |
0.240 |
3.3.2 Electrochemical Impedance Spectroscopy
Again
to unfold the electron transfer in the device between various interfaces,
electrochemical impedance spectroscopy (EIS) has been carried out as shown in Figs.9 (b) and (c)
under the dark condition at a forward bias of -0.8 V. All the crucial
parameters calculated from Nyquist and Bode plot vide Fig. 9 are presented in Table 4 after fitting with an equivalent circuit
diagram, as shown in Fig. 9(a). A glance
towards EIS spectra in Fig. 9 (b), Nyquist plot
illustrates the smaller semicircle towards higher frequency range produces a
resistance between electrolyte and counter electrode interface designated by RCE, a greater semicircle in
the middle-frequency range termed as recombination resistance (RREC) at
photoanode/dye/electrolyte interface and Rs
is the total series resistance present in the device due to all the wired
connections. As the difference between RREC
and RCE is very high, the first semicircle between the electrolyte and
counter electrode interface (RCE)
is not visible as a single semicircle in the middle-frequency range. It is well
known that a higher RREC
value produces a lower recombination process and hence enhances the device's Voc.[46] Again, higher the series resistance lowers the solar
cell effectiveness.[47]
Fig. 9(a) Equivalent circuit diagram, (b)Nyquist plot, (c) Bode
Plot of the BPO, BPO-Cl, and BPT loaded device.
Table 4 EIS parameters of BPO, BPO-Cl, and BPT dye loaded
film.
Dye |
RS (Ω) |
RCE (Ω) |
RREC (Ω) |
Ʈe (ms) |
Ln (μm) |
ηcc% |
BPO |
25.89 |
4.37 |
1332.62 |
0.057 |
17.46 |
99.67 |
BPO-Cl |
27.74 |
3.94 |
2930.38 |
0.059 |
27.24 |
99.86 |
BPT |
25.00 |
5.48 |
317.879 |
0.050 |
7.62 |
98.27 |
Considering the three dyes, the BPO-Cl dye possesses
the highest RREC value.
The lowest RCE value gives
maximum charge collection efficiency and hence produces maximum Voc in the
device. On the contrary, BPT shows the lowest RREC value and highest RCE; therefore, it has a lower open-circuit voltage and
low efficiency. The Bode plot in Fig. 9(c)
explains the device's electron lifetime, directly affecting the charge
collection efficiency and the current density, Jsc, of the device.[47] The electron lifetime can be calculated using the
equation τeff = 1/2πfmax, where fmax is the maximum frequency of the Bode plot's intermediate
frequency arc and is found to be 0.059 ms for BPO-Cl,
0.057 ms for BPO, and lowest for BPT that is 0.050 ms. Again more considerable
diffusion length in BPO-Cl is also a vital parameter to produce higher charge
collection efficiency (ηcc)
in the device, which is directly related to the device's Jsc. A trend of higher Voc and Jsc in BPO-Cl
followed by BPO and then BPT is fine-tuned with diffusion length and hence
charge collection efficiency in the device. Thus all
the EIS parameters are fine-tuned with solar cell entities. Higher Voc and JSC in BPO-Cl dye are credited
to lesser recombinations, higher electron lifetime,
and high charge collection efficiency in the device. All the formulae required
to calculate the EIS parameters are as per the literature.[48-50] It is clear from EIS analysis that the simple
incorporation of chlorine in the BPO skeleton reduces recombination chances a
lot as recombination resistance becomes double in BPO-Cl dye, thus produces
more VOC and JSC in the device. Hence,
tailoring of efficient anchoring and donating groups will enhance the solar
cell efficiency to many folds.
In this investigation, the electron percolation and
recombination mechanism in the DSSC device are schematically presented in Fig. 10. A glance at evaluating the physicochemical
properties of photosensitizers on solar cell effectiveness is illustrated here
through the pictorial form. When photosensitizers (BPO, BPO-Cl, and BPT) are
exposed to solar radiation; they absorb the photon of light and excited from
the ground state (S0/S+) to the excited state (S+/S*)
where the excited dye molecule injects the electrons to the conduction band
(CB) of TiO2 semiconducting metal oxide coated on fluorine-doped tin
oxide (FTO) substrate.[51] As the HOMO level is below the work function of
electrolyte; they get adequate driving force for dye regeneration. The LUMO
level above the CB provides good electron injection efficiency.[52] Further,
the planar shape of the moiety percolates the
electron through the aromatic ring of the photosensitizers.
Moreover, the chlorine atom addition reduces the
recombination by reducing the HOMO-LUMO gap, as reflected in Fig. 3, which agrees with the literature.[39] Again, these electrons are transferred from TiO2
photoanode to platinum counter electrode (Pt) and further reduces the oxidized
dye molecule by iodine/triiodide electrolyte. Along with this, some recombinations are also taking place at several interfaces
in the device, which can also be controlled by modifying the photosensitizer behavior and device fabrication techniques.
Fig. 10 Schematic presentation
of DSSC device with Benzo[α]phenoxazine moiety as a photosensitizer with
electron percolation and recombination paths.
4.
Conclusions
A detailed evaluation of physicochemical properties of
three benzo[α]phenoxazine dyes was studied and implemented as
photosensitizers successfully. The application of benzo[α]phenoxazine
moiety as a photosensitizer was demonstrated through a DSSC device with
optical, electrochemical, structural, and solar cell (current-density)
characterizations. The unsubstituted benzo[α]phenoxazine moiety produces
noticeable VOC. Even the
open-circuit voltage (VOC)
is good enough to glow the small LEDs efficiently. Hence, tailoring an
efficient D/A group on benzo[α]phenoxazine produces more solar cell
efficiency than the phenoxazine precursors. A simple chlorine substitution on
the benzo[α]phenoxazine skeleton reduces the recombination more than twice
in BPO-Cl. It enhances the open-circuit voltage (VOC) as well as current density. Thus
a targeted D/A group will generate enhanced solar cell output. With the
increase in incident light source, there is a steady enhancement of VOC and JSC in the device, which is very conducive for DSSC
fabrication. Out of this experiment, it is evident that
benzo[α]phenoxazines have the inherent potential to become a
photosensitizer and an efficient precursor for dyes. Resourceful dyes can be
obtained with adequate D/A substitution on the BPO template at targeted sites.
Moreover, the BPO-Cl loaded device's effectiveness is credited to its proximity
of LUMO level to the conduction band of TiO2 semiconducting metal
oxide, reduced recombination, high diffusion length, and comparatively longer
electron lifetime. So tailoring efficient donor and
anchoring moiety to the benzo[α]phenoxazine template can be used as a new
platform for producing useful photosensitizers.
Acknowledgments
We
acknowledge the various instrumental support from the Department of Physics,
and the Department of Chemistry Savitribai Phule Pune
University. We also grateful to the Department
of Science and Technology, Government of India, for financial support vide
Sanction order DST/TMD/SERI/S173(G). PB is very thankful to the Department of
Higher Education, Science and Technology and Biotechnology, Govt. of West
Bengal, India, for providing financial assistance to carry out this research
work [vide sanction order no. 78(Sanc.)/ST/P/S&T/6G-1/2018
dated 31.01.2019 and project no. GAP-225612].
Supporting information
Applicable
Conflict of interest
There are no conflicts
to declare.
[1] B. O'regan and M. Grätzel, Nature, 1991, 353, 737-740, doi: 10.1038/353737a0.
[2] Q. Chang, H. Chen. J. Yuan, Y. Hu, J. Hai, W. Liu, F.
Cai, J. Hong, X. Xiao and Y. Zou, J. Energy Chem., 2020, 51, 7-13, doi: 10.1016/j.jechem.2020.03.036.
[3] A. A. Yadav, V. C. Lokhande, R. N. Bulakhe and C. D.
Lokhande, Microchim. Acta, 2017, 184, 3713-3720, 10.1007/s00604-017-2364-3.
[4] S. Patil, R. Bulakhe, P. Deshmukh, N. Shinde and C.
Lokhande, Sens. Actuators A Phys., 2013, 201, 387-394, doi: 10.1016/j.sna.2013.07.019.
[5] R. N. Bulakhe, S. A. Arote, B. Kwon and S. Park, I.
In, Mater. Today Chem., 2020, 15, 100210, doi: 10.1016/j.mtchem.2019.100210.
[6] G. A. Sewvandi,
C. Chen, T. Ishii, T. Kusunose, Y.
Tanaka, S. Nakanishi, and Q. Feng, J. Phys. Chem. C, 2014,118, 20184-20192, doi: 10.1021/jp5058397.
[7] A. Mishra,
M. K. Fischer and P. Bäuerle, Angew. Chem. Int. Ed., 2009, 48, 2474-2499, doi: 10.1002/anie.200804709.
[8] N. Mariotti, M.
Bonomo, L. Fagiolari, N. Barbero, C. Gerbaldi, F. Bella and C. Barolo, Green Chem., 2020, 22, 7168-7218, doi:10.1039/D0GC01148G.
[9] E. Guillen, F. Casanueva and J. A. Anta, J. Photochem. Photobiol. A, 2008, 200, 364-370. doi: 10.1016/j.jphotochem.2008.08.015.
[10] G. Sharma, P.
Balraju, M. Kumar and M. Roy, Mater. Sci. Eng. B, 2009, 162, 32-39, doi: 10.1016/j.mseb.2009.01.033.
[11] R.
Sánchez-de-Armas, M. A. San Miguel, J. Oviedo and J. F. Sanz, Phys. Chem. Chem.
Phys., 2012, 14, 225-233, doi: 10.1039/C1CP22058F.
[12] K. Hara, Z.
Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y. Dan-oh, C. Kasada, A. Shinpo and S. Suga, J. Phys. Chem. B, 2005, 109, 15476-15482, doi:10.1021/jp0518557.
[13] M. Lončarić, D. Gašo-Sokač, S.
Jokić and M. Molnar, Biomolecules, 2020, 10, 151, doi: 10.3390/biom10010151.
[14] A. B.
Athanas, S. Thangaraj and S. Kalaiyar, Chem. Phy. Lett., 2018, 699, 32-39, doi: 10.1016/j.cplett.2018.03.033.
[15] Y. Shi R. B. M. Hill, J. Yum, A. Dualeh, S. Barlow,
M. Grätzel, S. R. Marder, and M. K.Nazeeruddin,
Angew. Chem. Int. Ed., 2011, 50, 6619-6621, doi:10.1002/anie.201101362.
[16] X. Wang, J.
Xu, M. Li, D. Fang, B. Chen, L. Wang
and W. Xu, RSC Adv., 2013, 3, 5227-5237, doi: 10.1039/C3RA40193F.
[17]Y. Chen, V.
Nguyen, H. Chou, Y. S. Tingare, T. Wei, and C. Yeh, ACS Appl. Energy
Mater., 2020, 3, 5479-5486, doi: 10.1021/acsaem.0c00464.
[18] W.J. Fan, W.
Fan, Y. Chang, J. Zhao, Z. Xu, D. Tan, and Y.
Chen, New J. Chem., 2018, 42, 20163-20170, doi:10.1039/C8NJ03592J
[19] S.
Kolemen, O. A. Bozdemir, Y. Cakmak, G.
Barin, S. Erten-Ela, M. Marszalek, J. Yum, S. M. Zakeeruddin, M. K. Nazeeruddin, M. Grätzel and E.
U. Akkaya, Chem. Sci., 2011, 2, 949-954, doi: 10.1039/C0SC00649A.
[20] H. Klfout, A.
Stewart, M. Elkhalifa and H. He, ACS Appl. Mater. Interfaces, 2017, 9, 39873-39889, doi: 10.1021/acsami.7b07688.
[21] V. Saravanan, S.
Ganesan and P. Rajakumar, RSC Adv., 2020, 10, 18390-18399, doi: 10.1039/D0RA01672A.
[22] F. A. Schröder, J. M. Cole, P. G. Waddell and S.
McKechnie, Adv. Energy Mater., 2015, 5, 1401728, doi: 10.1002/aenm.201570047.
[23] H. Tian, X.
Yang, R. Chen, A. Hagfeldt and L. Sun, Energy Environm. Sci., 2009, 2, 674-677, doi:10.1039/B901238A.
[24] H. Tan, C.
Pan, G Wang, Y. Wu, Y. Zhang, G. Yu, and M. Zhang, Dyes Pigm., 2014, 101, 67-73, doi:10.1016/j.dyepig.2013.09.039.
[25] W. Wu, J.
Yang, J. Hua, J. Tang, L. Zhang, Y. Long
and H. Tian, J. Mat. Chem., 2010, 20, 1772-1779, doi:10.1039/B918282A.
[26] H. Tian,
I. Bora, X. Jiang, E. Gabrielsson, K.
M. Karlsson, A. Hagfeldt and L.
Sun, J. Mater. Chem., 2011, 21, 12462-12472, doi:10.1039/C1JM12071A.
[27] S. Bhand, D. Chadar, K. Pawar, M. Naushad, H.
Pathan and S. Salunke-Gawali, J. Mater. Sci: Mater. Electron., 2018, 29, 1034-1041, doi:10.1007/s10854-017-8003-2.
[28] D. Chadar, S. S. Rao, A. Khan, S. P. Gejji, K. Sideeq
Bhat, T. Weyhermüller and S. Salunke-Gawali, RSC Adv., 2015, 5, 57917-57929, doi:10.1039/C5RA08496B.
[29] S. S. Khadtare, A. P. Ware, S. Salunke-Gawali, S.
R. Jadkar, S. S. Pingale and H. M. Pathan, RSC Adv., 2015, 5, 17647-17652, doi: 10.1039/C4RA14620D.
[30] G. P. Sarmiento, F. Martini, R. G. Vitale, L. E.
Fabian, J. Afeltra, D. Vega, G. Y. Moltrasio and A. G. Moglioni, Arab. J. Chem., 2019, 12, 21-32, doi:10.1016/j.arabjc.2017.11.019.
[31] L. J. Farrugia, J. Appl. Cryst., 2012, 45, 849-854, doi:10.1107/S0021889812029111.
[32] C. F. Macrae, I. J. Bruno, J. A. Chisholm, P. R. Edgington, P. McCabe, E.
Pidcock, L. Rodriguez-Monge, R. Taylor, J. van de Streek and P. A. Wood, J. Appl. Cryst., 2008, 41, 466-470, doi:10.1107/S0021889807067908.
[33] S. K. Saha,
A. Hens, N. C. Murmu and P. Banerjee, J. Mol. Liq., 2016, 215, 486-495, doi:10.1016/j.molliq.2016.01.024.
[34] P. Ghosh, B.
G. Roy, S. Jana, S. K. Mukhopadhyay and P. Banerjee, Phys. Chem. Chem.
Phys., 2015, 17, 20288-20295, doi:10.1039/C5CP02525G.
[35] P. Ghosh and P. Banerjee, Phys. Chem. Chem.
Phys., 2016, 18, 22805-22815, doi:10.1039/C6CP01620K.
[36] R. Ahlrichs, M. Bär, M. Häser, H. Horn and C.
Kölmel, Chem. Phys. Lett., 1989, 162, 165-169, doi:10.1016/0009-2614(89)85118-8.
[37] A. Schäfer, H. Horn and R. Ahlrichs, J. Chem. Phys., 1992, 97, 2571-2577, doi: 10.1063/1.463096.
[38] G. Boschloo and A. Hagfeldt, Acc. Chem. Res., 2009, 42, 1819-1826, doi: 10.1021/ar900138m.
[39] R. Zhou, J.
Feng, X. Peng, L. Zhong and R. Sun, J. Energy Chem., 2020, doi: 10.1016/j.jechem.2019.09.010.
[40] P. M. Kibasomba, S. Dhlamini, M. Maaza, C. P. Liu,
M. M. Rashad, D. A. Rayan and B.W. Mwakikung, Results Phys., 2018, 9, 628-635, doi:10.1016/j.rinp.2018.03.008.
[41] D. Chadar, D.
Chakravarty, D. N.Lande, S.P. Gejji, S. Sahoo and S. Salunke-Gawali, J.
Mol. Struct., 2017, 1149, 84-91, doi:
10.1016/j. molstruc.2017.07.091.
[42] S. Huang, G. Schlichthörl, A. Nozik, M. Grätzel and
A. Frank, J. Phys. Chem. B, 1997, 101, 2576-2582, doi:10.1021/jp962377q2.
[43] H. Paulsson, L. Kloo, A. Hagfeldt and G. Boschloo, J. Electroanal. Chem., 2006, 586, 56-61, doi:10.1016/j.jelechem.2005.09.011.
[44] T. Privalov, G. Boschloo, A. Hagfeldt, P. H.
Svensson and L. Kloo, J. Phys. Chem. C, 2009, 113, 783-790, doi:10.1021/jp810201c.
[45] S. Borbón, S. Lugo, D. Pourjafari, N. P. Aguilar, G. Oskam, and I. López, ACS Omega, 2020, 5, 10977-10986, doi:10.1021/acsomega.0c00794.
[46] T. Zhang, C.
An, Q. Lv, J. Qin, Y. Cui, Z. Zheng, B. Xu, S. Zhang, J. Zhang, C. He and
J.Hou, J. Energy Chem., 59, 30-37, doi:10.1016/j.jechem.2020.11.021.
[47] M. Adachi, M. Sakamoto, J. Jiu, Y. Ogata and S.
Isoda, J. Phys. Chem. B, 2006, 110, 13872-13880, doi:10.1021/jp061693u.
[48] L. Zhang, X. Yang, W. Wang, G. G. Gurzadyan, J. Li,
X. Li, J. An, Z Yu, H. Wang, B Cai, A. Hagfeldt and L. Sun, ACS Energy Lett., 2019, 4, 943-951, doi:10.1021/acsenergylett.9b00141.
[49] C. V. Jagtap, V. S. Kadam, S. R. Jadkar and H. M.
Pathan, ES Energy Environ., 2019, 3, 60-67, doi:10.30919/esee8c220.
[50] A. R. Bredar, A. L. Chown, A. R. Burton and B. H.
Farnum, ACS Appl. Energy Mater., 2020, 3, 66-98, doi:10.1021/acsaem.9b01965.
[51] D. Cahen, G. Hodes, M. Grätzel, J. F. Guillemoles
and I. Riess, J. Phys. Chem. B, 2000, 104, 2053-2059, doi:10.1021/jp993187t.
[52] D. P. Hagberg, T. Marinado, K. M. Karlsson, K.
Nonomura, P. Qin, G. Boschloo, T. Brinck, A. Hagfeldt, and L. Sun, J. Org. Chem., 2007, 72, 9550-9556, doi:10.1021/jo701592x.
Author
information
Suprabha Soumya Sahoo, perceiving her Ph.D. under the supervision of Professor Sunita Salunke – Gawali and Dr. Habib M.
Pathan at Department of Chemistry in Savitribai Phule Pune University, Pune
(India). She was awarded her master's degree in Organic Chemistry in 2012 from
the same university and did her Bachelors of Education
(B.Ed.) in 2013. Her research interests are synthesizing and engineering
photosensitizers, preparing various photoanodes for dye-sensitized solar cells,
and device fabrication techniques.
Dattatray Ashruba
Chadar completed his M.Sc. in Analytical Chemistry at the Department of
Chemistry, Dr. Babasaheb Ambedkar Marathawada
University, Aurangabad (India). He received his Ph.D. in Chemistry from the
Department of Chemistry, Savitribai Phule Pune University, India, under the
supervision of Professor Sunita Salunke-Gawali. The
topic of his Ph.D. study was on synthesis, characterization, metal
complexation, and biological activity of amino/imino
derivatives of Vitamin K3 (2-Methyl-1,4-naphthoquinone). He worked as a Research
Associate in CSIR-Central Salt and Marine Chemical Research Institute,
Bhavnagar in India. His research interest mainly focuses on coordination
chemistry, organic synthesis, and carbohydrate-based biopolymers.
Manilal Murmu received
his M.Sc. degree in Chemistry in 2014 from Visva
Bharat University, Santiniketan, Bolpur,
West Bengal, India. At present, he is pursuing a Ph.D. in Chemical Sciences
under the Academy of Scientific and Innovative Research (AcSIR)
under the joint supervision of Dr. Priyabrata
Banerjee, Principal Scientist, and Dr. Naresh Chandra Murmu,
Senior Principal Scientist, Surface Engineering and Tribology Group,
CSIR-Central Mechanical Engineering Research Institute (CMERI), Durgapur. His
research is primarily focused on developing and characterizing organic
corrosion inhibitors and anticorrosive coating to protect metallic materials.
His research is also in the field of theoretical and computational chemistry.
Priyabrata Banerjee is a
Principal Scientist at Surface Engineering & Tribology Group, CSIR-Central
Mechanical Engineering Research Institute (CMERI), Durgapur, 713209, West
Bengal, India & Associate Professor at Academy of Scientific and Innovative
Research (AcSIR), CSIR-HRDC Campus, Ghaziabad,
201002, Uttar Pradesh, India. Dr. Banerjee received his Ph.D. degree in 2007
from the Indian Association for Cultivation of Science, Jadavpur, Kolkata,
India. He did his Post Doctoral research at Max
Planck Institute for Bio-Inorganic Chemistry, Müelheim,
Germany, during 2007-2010. He was a visiting fellow at the University of Water
Science, HTWD, Dresden, Germany, and presently he has been selected as
CSIR-Raman Research Fellow (2019-2020) at Ghent University, Belgium. He has
published 104 research papers (3711 citations) in several International SCI
journals (h-index: 33, i-index: 78), 10 book
chapters, 10 magazines, and several review articles in world reputed forums.
His current research interest broadly covers selective bio-relevant
cation-anion detection, corrosion science, theoretical and computational
chemistry, solid waste management, wastewater treatment, Metal mediated
C-heteroatom bond fusion metal-organic complexes, and their hitherto unexplored
radical chemistry development.
Sunita Salunke-Gawali, received
her M.Sc. (1993) in Inorganic Chemistry and Ph.D. (1999) from Pune University.
As a professional experience she worked as Post-doctoral Research Associate at Laboratoire de Magnétisme et d'Optique, Versailles France (Prof. F. Varret,
2001-2002), Department of Chemistry, IIT Bombay, India (Prof. C. P. Rao, 2002
and 2004), Universidade do Porto, Portugal,
supervised by Prof. Eulália Pereira (2004-2007) and
Max-Planck-Institut für Bioanorganische Chemie, Mülheim an der Ruhr, Germany (Dr. Eckhard
Bill, 2007- 2008). She joined as a
Reader in the Department of Chemistry, Savitribai Phule Pune University, in
2008, where she serves now as a Professor. Her research interests include
coordination and bioorganic Chemistry of naphthoquinone ligands, developing
photosensitizer for DSSC, HPLC method development for anticancer drugs, and separation of tautomers, chemosensors, and metallosurfactants.
She is the author of more than 87 articles in international journals.
Ray J. Butcher, Born in
Greymouth, New Zealand, on 11th October 1945. Educated at the University of
Canterbury in New Zealand. Appointed Instructor 1974-76 at the University of
Virginia, Charlottesville, Virginia and Post-Doctoral Fellow 1976-77,
Georgetown University, Washington D.C. Joined the Chemistry Department at
Howard University as an Assistant Professor in 1977 and promoted to Associate
Professor in 1982 and Professor in 1997. Has had many visiting appointments,
including a Navy Distinguished Summer Faculty Fellow at the Naval Research
Laboratory, a Visiting Senior Scientist at Los Alamos National Laboratory, Los Alamos, a
distinguished visiting professor at the Indian Institute of Technology Bombay,
in Mumbai, India. Has held two Fulbright-Nehru Fellowships to India (1989,
2009) and a Distinguished Chair Fulbright-Nehru Fellowship to India (2019). Was
elected a Fellow of the Royal Society of Chemists (FRSC) in 2018. Has published
over 1340 papers in refereed journals since 1975.
Publisher’s Note: Engineered Science
Publisher remains neutral with regard to jurisdictional claims in published
maps itand instutional affiliations.