You are on page 1of 12

Materials Chemistry and Physics 287 (2022) 126312

Contents lists available at ScienceDirect

Materials Chemistry and Physics


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

Star-type melamine based conjugated carboxy functionalized porphyrin


trimer for DSSCs: An efficient approach to clean, aggregation free and true
energy generation
Darpan V. Bhuse a, Vijaykumar M. Bhuse b, Pundlik R. Bhagat a, *
a
Department of Chemistry, School of Advanced Sciences, Vellore Institute of Technology, Vellore, 632014, Tamilnadu, India
b
Energy Materials Laboratory, Chemistry Department, Institute of Science, Nagpur, Maharashtra, 440001, India

H I G H L I G H T S

• Design and synthesis of star-shaped, metal-free, multiple D-π-A type porphyrin small molecule sensitizer for DSSC.
• The highest nine carboxymethyl groups give excellent anti-aggregation stability and electronic coupling.
• An electron-deficient planar triazine core sets in-plane- π stacking over TiO2 surface.
• A perfect energy alignment of dye frontiers with TiO2 and redox couple and the shortest charge injection time (0.2 ms).
• A DSSC evolved as additive-free, stable, true energy-generating with a record 7.1% PCE and 0.451 eV Eloss.

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

Keywords: A novel, π-conjugated melamine-based trimeric porphyrin dye with serving nine carboxylate linker groups has
Porphyrins been prepared to use as a panchromatic dye for application in dye-sensitized solar cells (DSSCs). The star-shaped
Melamine trimer was characterized by using Proton Nuclear Magnetic Resonance (NMR) Spectroscopy, Carbon Nuclear
Dye-sensitized solar cells
Magnetic Resonance Spectroscopy (CMR), Fourier Transformed Infrared Spectroscopy (FT-IR), UV-VIS Absorp­
Cyclic voltammetry
Photosensitizers
tion Spectroscopy, Thermo-gravimetric analysis (TGA), Cyclic Voltammetry (CV) and Electrochemical Imped­
ance Studies (EIS). CV studies revealed frontiers energy levels at − 5.654 (HOMO) and − 3.879 eV (LUMO) versus
vacuum. The optical band gap evolved via absorption spectra measurements was found to be 1.49 eV. A planar
perfect conjugation among three porphyrins sets up multiple push-pull communication and lowers the bandgap,
while nine carboxylate linkers cover-up TiO2 surface effectively leading to better electronic dynamics. The DSSC
set up using iodide-based redox couples forms a simple, stable, additive-free, truly energy generating device with
7.1% conversion efficiency, the highest among trimeric porphyrin category.

1. Introduction (DSSC), a third energy generation technology among the four, has
emerged as a cost-efficient and eco-friendly way with features like the
The cosmopolitan issues to tackle in near future are the hovering wide choice of device components, design flexibility and easy setup of
energy demands and curtailing of environmental pollution judiciously to the DSSC devices, [3–6]. A standard DSSC consists of a dye-sensitized
minimize carbon footprints. Till today, conventional fossil fuels are semiconductor as photoanode, the redox couple in a suitable solvent
satisfying the majority of energy demands. Sustainable renewable en­ as electrolyte and the inert conducting material as the counter electrode
ergy resources, in the context of depleting fossil reserves, have attracted (cathode) [7]. The dye adsorbed on the semiconductor, upon light
more attention as they are eco-friendly and abundant [1]. Harnessing irradiation generates and separates the charge carriers (electrons and
solar energy using photovoltaic technology (PV) is the leading way to holes). Further, electrons are injected into the semiconductor to travel
zero carbon emission and has registered a benchmark of 29% share in toward the cathode via an external circuit, while holes are trapped by
total energy generation in 2020 [2]. The dye-sensitized solar cell the redox component in the electrolyte. At the cathode, electrons are

* Corresponding author.
E-mail address: drprbhagat111@gmail.com (P.R. Bhagat).

https://doi.org/10.1016/j.matchemphys.2022.126312
Received 28 January 2022; Received in revised form 6 May 2022; Accepted 22 May 2022
Available online 3 June 2022
0254-0584/© 2022 Elsevier B.V. All rights reserved.
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

gained by another redox component in the electrolyte to complete the linked porphyrin dimers tethered with 2-propionic acids or 2,4-penta­
circuit [8]. dienoic acids as dye on nanocrystalline TiO2 solar cells to deliver
A dye is the integral component of the DSSC serving as a light- 4.2% PCE [42]. A push-pull with multiple electron-donating groups on
harvesting antenna. Wide varieties of the dyes are available under the dimeric porphyrin reported excellent light-harvesting ability in the
synthetic and natural based categories [9]. Among them, the near-infrared region exhibiting PCE of 6.21% under AM 1.5 condition
nature-based porphyrins are gaining much importance because of their [29]. Our previous work on multi-push pulls porphyrin dimer linked via
panchromatic and high light absorption ability, flexibility in molecular a conjugated aromatic spacer and functionalized with multiple carbox­
designing [10–13], good charge separation capabilities [14], better ylate anchoring cum acceptor arms has reported excellent stability with
electronic tuning with TiO2 semiconductor [15] and low cost, non-toxic a record PCE of 6.9% in a truly energy-generating state, without using
nature [16]. In addition porphyrin exhibit good chemical inertness [17], any co-adsorbent, efficiency enhancers etc in an aqueous-based iodide
excellent stability in both aqueous and nonaqueous solutions [18,19] redox couple [20].
making it as a powerful sensitizer for true energy-generating DSSC de­ A trimer with three chromophores in a star -architecture using a
vices [20–23]. With these features, porphyrin-based DSSC has registered 1,3,5-triazine (melamine) heterocyclic aromatic core has shown excel­
a remarkably fast improvement in PCE as compared to other photovol­ lent molecular symmetry, spatial co-planarity, good anti-aggregation,
taics [24]. On the contrary, other emerging solar cell technologies uti­ faster electron transport and extra stability [50,51]. The lone pairs on
lizing a perovskite-based hybrid (PSC) and bulk heterojunction (BHJ) each of the N in the melamine core can participate in electronic com­
have recorded lab-scale PCE of 27% [25], however, they need to address munications of the dye facilitating the process of electron separation,
the issues such as proper layer communication, stability against ion injection and transportation [52]. Melamine based dyes being planar
migrations and defects, substitute for Pb and cost immediately for exhibit preferably horizontal orientation over the TiO2 surface and thus
commercialization in near future [26,27]. establish a strong π type stacking between the dye and semiconductor
Porphyrin offers a unique feature of flexibility in designing its molecule [53]. This facilitates the rapid electronic couplings between
structure via modifying its core, for example, by attaching one or more the chromophore groups and eliminates the possibility of π-π stacking
donor and acceptor groups (called push-pull mechanisms) [28,29], among dye molecules to give admirable anti-aggregation stability [20,
adding conjugation [30,31], adding one or more binding groups to 54–56]. A literature search, however, reveals only a few reports on the
acceptor moieties and attaching long alkyl chains to have some flexi­ melamine cored trimeric porphyrin category, for example, an ethynyl
bility in binding [32–34]. The engineering of porphyrin substantially linked Zn doped porphyrin trimer dye synthesized by Hamamura et al.
modifies the electrochemical, and photophysical properties, and also reported a PCE of 3.17% [57], A a propeller-shaped, triazine-linked
demonstrates endurance against thermal and photo-decomposition [12, porphyrin triad synthesized by Zervaki et al. was used as a sensitizer for
35,36]. Like other organic dyes, the porphyrin dye is prone to exhibit DSSC exhibiting the PCE of 5.56%. The higher efficiency was observed at
aggregation over the semiconductor surface which hampers the per­ the cost of using chenodeoxy-cholic acid (CDCA) as a co-sensitizer and
formance and the stability of the DSSC device [37]. This prevailing issue zinc metal doping [58]. In another attempt, the porphyrin triad linked to
can be addressed by incorporating a multiple, electron acceptor cum the ester-containing C70 group has shown a PCE of 3.93% in a BHJ cell
strong binding organic groups on its periphery to serve simultaneously [59]. To the best of our knowledge, no reports are available on
as super-linker to semiconductor surface and electron acceptor [38]. A metal-free- triazine cored trimeric porphyrins entangled with multiple
carboxylate group is considered an ideal linker because it can execute push-pulls and multiple carboxylate binder cum acceptor arms.
multi-binding modes (monodentate, bidentate type, H-bonding) giving Considering the inherent stability and the excellent optoelectronic
strong chemisorption of dye over the TiO2 surface [39]. features of the trimeric porphyrins in star architecture as compared to
The monomeric porphyrin has been used as a dye in numerous DSSC monomeric and dimeric porphyrin dyes, we undertake to synthesize a
assemblies [3,35,40,41]. Mathew et al. synthesized a molecularly engi­ new, symmetric, metal-free, multiple D-π-A type, melamine based
neered porphyrin as a sensitizer for DSSC application and reported 13% porphyrin trimer as a dye for applications in DSSC. The conjugated
PCE using cobalt bipyridyl redox-based electrolyte and the co-sensitizer. melamine (π system) was linked to three porphyrins via cyanamide
The highest efficiency observed was the joint effect of the engineered groups, while the periphery of all the porphyrins (Donor) was entangled
porphyrin and the co-sensitizer (chromophore) [33]. It was realized that with multiple carboxylates functionalized benzimidazole moieties to act
a single porphyrin, even after core engineering exhibits a high bandgap, as linker cum acceptor groups (Acceptors). This new architecture is
and poor and tapered absorption ability, especially in the red and NIR expected to show a low bandgap, better DSSC performance and extra
part of the spectral region, limiting the overall performance of DSSC anti-aggregation stability. The trimeric dye will be used in a simple,
devices. To remedy this, a single porphyrin is often coupled with a additive-free, TiO2 based DSSC device using aqueous iodide redox
suitable co-sensitizer as supportive chromophores to encompass the electrolyte. The DSSC devoid of any additives (such as efficiency en­
missing spectral regions [10]. hancers, anti-aggregating agents, stabilizers and co-sensitizer) will be
In an attempt to make the porphyrin wide spectral range absorber truly energy-generating as no additive-mediated degradations are
(and to lower the band gap), two or more porphyrins are linked together involved accounting for true performance.
chemically either directly or through π-conjugation [42–44] or fused This manuscript reports the synthesis, characterization, electro­
together [45] to establish good electronic communication between the chemical and photovoltaic applications of the triazine based porphyrin
porphyrins. The dimerization reshuffles the electron densities and thus trimer as a dye in a true energy-generating DSSC.
lowers the frontier energy and the optical band gap so that the porphyrin
can encompass the NIR portion of the spectral region [46]. Additionally, 2. Experimental section
vesting multiple push-pulls on dimer further improves charge transfer
dynamics [17,29,47–49]. Any aromatic conjugation present between 2.1. Materials
two porphyrins of a dimer (at meso-positions) helps to set up a perfect
core-planarity so that the dye molecule can cover a substantial portion of Melamine, Sodium sulphate, Sodium chloride, Sodium hydride
the TiO2 surface and establish a π stacking avoiding the detrimental (57–63% in oil), Glyoxal (40 wt%) HCl acid (36%), Methanol, Ethyl
charge recombination which otherwise occurs at the surface of TiO2 acetate, Formaldehyde (37%), Acetic acid, Diethyl ether and 2-Chloro­
degrading the performance drastically [30,35]. The dimers are thus acetic acid were availed from Loba Chemie. Pyrrole, (99%), Dry THF,
preferred over a single porphyrin congruent because of their panchro­ CDCl3, Dimethyl sulfoxide-d6, Chloroform, DMSO, n-methyl pyrrolidone
matic nature, better charge separation capabilities and additional sta­ (NMP) used were all synthetic/AR grade. The F.T.O plates and TiO2 of
bility. With these strategies, Park et al. reported the use of meso-meso 30–50 nm particle size (99.9%) were obtained from Ultra- Nanotech,

2
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Scheme 1. Synthesis of Star-shaped porphyrin trimer.

Bangalore. Tetrabutylammonium hexafluoro-phosphate (98%) Ferro­ obtained for the final dye.
cene (99.8%) were acquired from Sigma Aldrich. Deionized water was
employed for aqueous solution preparation. All chemicals were 2.3.1. Preparation of Intermediate (A): (2E,2′ E,2′′ E)-2,2′ ,2’’-((1,3,5-
employed as received. triazine-2,4,6-triyl)tris (azaneylylidene)) triacetaldehyde
An intermediate (A) was synthesized by reacting melamine (1.26g)
with glyoxal (2.92 g) (taken in 10 mmol: 50 mmol ratios) under reflux at
2.2. Characterization techniques
65 ◦ C for 44–48 h. Methanol was used as solvent and ZnCl2 was used as a
catalyst. Any unreacted glyoxal from the product was eliminated by
The proton NMR and CMR spectra of the products were acquired
using 400 MHz and 100 MHz Bruker spectrometer at room temperature. successive washing with methyl alcohol and water. Yield- 1.80 g (70%).
The solvent used was DMSO-d6,CDCl3 and all the peaks were referred to The following analytical data corroborates the formation of (A).
1
as TMS (δ = 0.00). The absorption profile and the energy gap of the dye H NMR in DMSO‑d6: δ 9.373 (s, 3H), 8.764 (s, 3H) (*Supplementary
were calculated using Jasco UV–Vis Spectrophotometer, infrared spectra Information, Fig. S1, page no 3).
were acquired from FT-IR spectrophotometer (IR affinity Shimadzu). FTIR-(KBr, v/cm− 1): 3213.41, 1546.91, 1448.54, 1159.22, 1039.63,
The SEM and EDAX analysis of the trimer were carried out using Carl 810.10, 576.72, 507.28 (*Fig. S2, page no 3).
Zeiss EVO/18SH, UK. The thermal stability was scanned in the range of
50–900 ◦ C with the rate of 10 ◦ C/min in N2 on Thermoanalyzer SII, 7200 2.3.2. Preparation of Intermediate (B): 3-(chloromethyl)-2-
(TGA). The CV (cyclic voltammetry) and the LSV (linear sweep vol­ hydroxybenzaldehyde
tammetry) parameters were acquired using an Auto-Lab electrochemical A chloromethylation of salicylaldehyde was done in this step. To a
workstation (PGSTAT204). 150 ml of concentrated HCl acid, cooled to 0 ◦ C–4 ◦ C, added 20 ml
formaldehyde with magnetic stirring. After an interval of 10 min, 15 ml
of salicylaldehyde was added and stirred magnetically for the next
2.3. Synthesis of star-shaped trimer 44–48 h. Then the excess acid was decanted, and the product was
washed with deionized water for 20 min and recrystallized using n-
A star-shaped melamine-based porphyrin trimer with nine carbox­ hexane. Yield- 12.8 g (73%). The following analytical data corroborates
ylic groups has been produced using a reaction Scheme 1. The step-by- the formation of (B).
step reactions are displayed along with the characteristic NMR signals

3
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Fig. 1. The 1H NMR spectrum of Porphyrin trimer.

1 1
H NMR: δ 10.92 (s, 1H), 10.27 (s, 1H), 7.72 (d, 1H), 7.58 (dd, 1H), H NMR: δ 10.27 (s, 1H), 10.23 (s, 1H), 9.97 (s, 1H), 9.49 (s, 1H),
7.02 (d, 1H), 4.76 (s, 2H). (*Fig. S3, page no 4) 9.00 (s, 1H), 8.04 (t,1H), 7.84 (d,1H), 7.72 (t,1H), 7.52 (t,1H), 7.18
13
C NMR (100 MHz in DMSO‑d6): 191.19, 161.21, 137.42, 129.58, (d,1H), 7.11 (d, 1H), 5.80 (s, 2H), 5.52 (s, 2H). (*Fig. S9, page no 8)
13
129.25, 122.64, 118.22, 46.19 (*Fig. S4, page no 4). C NMR: 191.02, 169.36, 161.30, 135.50, 130.85, 129.34, 125.10,
FTIR-(KBr, v/cm− 1): 3201.83, 2875.86, 1581.61, 1483.26, 1379.10, 124.98, 124.86, 124.47, 122.82, 59.95, 49.73, 48.50, 48.18, 46.82
1278.81, 1240.23, 1188.15, 1145.27, 1114.86, 1008.77, 900.76, (*Fig. S10, page no 8).
846.75, 767.67, 715.59, 667.37, 572.86, 497.63, 449.41 (*Fig. S5, page FTIR-(KBr, v/cm− 1): 3350.35, 3039.81, 2845.00, 1728.22, 1680.00,
no 5). 1612.49, 1564.27, 1485.19, 1444.68, 1369.46, 1278.81, 1249.87,
1192.01, 1153.43, 1087.85, 933.55, 831.32, 744.52, 597.93, 551.64,
2.3.3. Preparation of Intermediate (C): 3-((1H-benzo[d]imidazole-1-yl) 507.28. (*Fig. S11, page no 9).
methyl)-2-hydroxybenzaldehyde
The hydrogen of benzimidazole was abstracted in this protocol. An 2.3.5. Synthesis of porphyrin trimer, E
air dried NaH (2.63 g, 114 mmol) was mixed with benzimidazole (13.80 The intermediate compounds (A) (1 mmol, 0.246 g), (D) (9 mmol,
g, 116 mmol) in THF solvent at 60 ◦ C temperature. To this, intermediate 3.120 g), and pyrrole (12 mmol, 0.805 g) were stirred in AcOH at 75 ◦ C
(B) dissolved in THF (19.6 g, 116 mmol) was added. The reaction was for 50 min. Thereafter, heating was stopped and magnetic stirring was
continued for the next 48 h. After the reaction, the remaining solvent continued overnight. The product of the reaction was vacuum filtered
was rota-evaporated. The obtained sticky product was extracted in ethyl and washed successively with water and acetone. Yield- 2.079 (59%).
acetate by shaking with an equal volume mixture of ethyl acetate and The following analytic data confirm the synthesis of (E).
1
brine. The ethyl acetate layer was separated and dried over anhydrous H NMR: δ 400 MHz, DMSO‑d6: δ 12.50 (s, 9H), 8.69 (s, 9H), 8.43 (s,
Na2SO4 for 4 h and finally, the product was obtained after evaporating 3H), 8.20 (s, 3H), 7.70 (s, 12H), 7.56 (d, 12H), 7.29 (q,8H), 7.21 (t,
solvent. The product was further washed with diethyl ether. Yield- 5.6 g 12H), 7.13 (d,7H), 7.08 (d, 9H), 7.02 (d, 9H), 6.93 (d, 8H), 6.24 (s, 18H),
(52%). The following analytical data corroborates the synthesis of (C). 5.97 (s, 9H), 5.44 (s, 18H), − 0.061 (s, 6H).
1
H NMR: δ 10.79 (s, 1H), 10.23 (s, 1H), 8.41 (s, 1H), 7.65 (d, 2H), FT-IR: (KBr, ν/cm− 1): 3169.04, 1610.56, 1485.19, 1369.46,
7.51 (q, 1H), 7.17 (q, 3H), 6.98 (d, 1H), 5.45 (s,2H). (*Fig. S6, page no 6) 1259.52, 1188.15, 999.13, 744.52, 667.37, 532.35, 464.84. (*Fig. S12,
13
C NMR: 191.25, 160.82, 144.58, 144.05, 136.11, 133.98, 128.47, page no 9)
128.05, 122.89, 122.67, 122.07, 119.96, 118.19, 111.15, 47.23
(*Fig. S7, page no 6).
FTIR-(KBr, v/cm− 1): 3089.96, 2873.94, 2540.25, 1815.02, 1674.21, 2.4. Electrode Fabrications and electrochemical measurements
1606.70, 1500.62, 1435.04, 1365.60, 1294.24, 1246.02, 1199.72,
1145.72, 1105.21, 1002.98, 954.76, 904.61, 823.60, 732.95, 628.60, 2.4.1. CV and EIS measurements
551.64, 507.28, 568.70, 418.55(*Fig. S8, page no 7). The CV measurements were done by using 10 mM trimer dye solu­
tion in DMSO solvent containing 100 mM NBu4PF6 electrolyte. A trimer
2.3.4. Preparation of Intermediate (D): 3-(carboxymethyl)-1-(3-formyl-2- dye solution was formulated by sonicating dye in 25 ml of DMSO solvent
hydroxybenzyl)-1H-benzo[d]imidazole-3-ium chloride for 45 min. The three electrodes used were the working electrode (Pt
Intermediate (D) was formed by the reaction of 2-chloroacetic acid square, area 0.5 cm2), counter electrode (Pt wire) and non-aqueous 0.01
(6.4 mmol, 0.604 g) and C (6.4 mmol, 1.630 g) in chloroform at 60 ◦ C for M Ag/AgNO3 as the reference electrode. Before measurements, all the
44–48 h. The sticky yellowish solid product was obtained after evapo­ solutions were deoxygenated by purging dry nitrogen gas. For CV
rating the solvent. Yield 1.02 g (80%). The following analytical data measurements, a potential window of − 2.2 to +1.5 V with varying scan
confirm the synthesis of (D). rates of 0.005, 0.010, 0.015 V/s was used. The measurements were
carried out by using a Ferrocene couple as an internal standard. The

4
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Fig. 2. The comparative FTIR spectrum of a) intermediates A and D and b) trimer dye (E).

E0Redox redox potentials of the ferrocene and the iodine-based redox


shuttle (aqueous solution of 100 mM M KI and 20 mM I2) used for the
DSSC were also measured. The EIS was measured for the DSSC cell was
set up in dark at the bias potential of 0.780 V in the 0.01Hz to 106 Hz
frequency range.

2.4.2. DSSC electrodes preparation


FTO glass plate (2Ω) was used as conducting base to prepare Dye/
TiO2/FTO (photoanode) and CNT coated FTO as counter electrodes. The
glass coated with FTO was sonicated sequentially in isopropyl alcohol,
acetone and deionized water, each for 40 min. The Doctor’s blade
technique was used to coat TiO2 suspension on FTO (the TiO2 suspension
was prepared by sonicating 150 mg TiO2 in 10 ml NMP for 50 min) The
TiO2/FTO was heated to 200 ◦ C in an oven under a vacuum for 2 h. The
deposited TiO2 layer was approximately 4 μm thick. This film was
further dipped in the 0.1 M dye solution prepared in DMSO solvent for
5–10 h followed by oven-drying at 150 ◦ C for 1 h to obtain Dye/TiO2/
FTO photoanode. The counter electrode was constructed by drop-casting
Fig. 3. TGA thermogram of trimer dye.
dispersion of CNT onto a clean FTO glass slide followed by drying at
180 ◦ C in a vacuum oven (the dispersion of CNT was prepared by son­
Fig. 1 have been identified as the peaks corresponding to the trimeric
icating 80 mg of CNT in 20 ml NMP for 1 h). A workable area of 0.5 ×
porphyrin having a fairly symmetrical environment. A singlet observed
0.5 cm2 of the working electrode was kept unmasked during DSSC
in the up-field region at − 0.061 ppm due to six protons confirms the
measurements.
formation of symmetric porphyrin trimeric core (− 0.061 ppm, s, 6H).
The singlet due to 9H from symmetric nine benzimidazole –CH bond was
3. Results and discussion
observed at 5.996 ppm. The singlet observed at 8.195 ppm due to 3H has
been assigned conveniently to three symmetrical –N=CH bonds between
3.1. Characterization of star shaped trimer dye
melamine and glyoxal modified porphyrin. The –OH peak of nine sym­
metric salicylaldehyde entangled benzimidazole moiety (from ionic
NMR and FT-IR studies were performed to confirm trimer dye for­
liquid-D) gave singlet at 8.426 ppm due to 9H, while the –CH2 protons
mation. The intermediates and the products of scheme- 1 were charac­
(s, 18) attached to benzimidazole- N and the phenyl carbon of salicy­
terized with the NMR, CMR and FT-IR and the data are included in the
laldehyde moiety (from ionic liquid-D) showed a peak at 5.443 ppm.
*Supplementary Information. The intermediate (A) was obtained by
The pyrrolic and aromatic protons of the porphyrin trimer molecule
condensation of the melamine and glyoxal. The evidence of the suc­
resonate in the 6.5–7.5 ppm range accounting to 105 hydrogens, while
cessful condensation was the formation of the N=CH bond. The reso­
the nine terminal –OH protons from carboxylic groups at the periphery
nance peak of these protons (3H) has appeared at 8.764 ppm (s, 3H) in
of the three porphyrins in trimer were found to resonate at 12.50 ppm (s,
NMR (Fig. S1) evidencing the successful condensation.
9H). confirming the formation of the porphyrin trimer.
The NMR spectrum of the final product of the scheme- 1 (star trimer)
The FT-IR of all the intermediates were recorded to compare and
is shown in Fig. 1. The porphyrin shows a characteristic large diamag­
confirm the presence of the specific functional groups on the in­
netic shift for the inner N–H proton signals giving negative chemical
termediates and the final trimer dye. The FT-IR of trimer alone is dis­
shifts (high field) and a large positive shift (low field) for outer pyro­
played in Fig. S12, Fig. 2 a) depicts the IR of the intermediates (A) and
lytic/substituents proton because of the strong shielding effects of
(D) (scheme- 1), while Fig, 2b) for the trimer dye (E). The intermediate
macrocyclic aromatic ring-current on inner proton as opposed to outer
(A) showed the C–N stretching due to the aromatic amine group at 1266
ones. A negative chemical shift of N–H proton is a piece of strong evi­
cm− 1 with a new peak at 1625 cm− 1 corresponding to the bending
dence for the formation of the porphyrin moieties [60]. All the peaks in

5
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Fig. 4. Normalized absorption spectrum obtained from DRS spectra of bare


TiO2 nanoparticle (red), dye adsorbed TiO2 nanoparticles(black), calculated Fig. 5. Calculation of Bandgap (Eg); the plot of [E(R)hν]2 Vs (hν).
dye (blue dotted) from DRS and in DMSO solution (green) for trimer dye. (For
interpretation of the references to colour in this figure legend, the reader is porphyrin via Q bands reshuffling to encompass the red and the NIR
referred to the Web version of this article.) regions. This makes porphyrin more panchromatic as the red and the
NIR portions of the spectrum corroborate with the maximum solar
vibration of the –NH bond. A broad peak in the 3000-3400 cm− 1 was photon flux. The metal-free trimeric dye having N, N′ , N’’-(triazine-
attributed to the CHO group evidenced by the successful condensation of 2,4,6-triyl) conjugated core linked to three porphyrins (D) via trime­
melamine with glyoxal. The intermediate D (an ionic liquid moiety) thanimine, while each porphyrin core entangled with benzimidazole
showed the aldehydic stretching in the 1661 - 1736 cm− 1 region. The affixed methyl carboxylate moiety as acceptor group (A) set up a mul­
peaks at 3069 and 3375 cm− 1 can be assigned as the vibration bands tiple D-π-A system is expected to exhibit a strong, synergic push-pull
originating from OH and COOH respectively. The FTIR spectrum of with substantial lowering of the frontiers energy and the bandgap. To
trimer dye (E) (Fig. 2b) displays peaks at 999 and 1188 cm− 1 due to investigate the symmetry and to evolve the optoelectronic feature, we
stretching of C–N and C=N respectively. The other peaks observed are at obtained the absorption spectra of dye in solution (0.1 mM in DMSO),
1259, 1369, 1485, 1610, 2984 and 3169 cm− 1 can be identified as peaks and the diffused reflectance spectra (DRS) of the TiO2 and dye adsorbed
due to HC=CH, OH, C–C, C=O, meso = CH and NH stretching vibrations TiO2. The normalized absorption spectra of the dye in the adsorbed state
respectively confirming the formation of the dye [61]. were obtained by subtracting the normalized absorption data of bare
The thermal stability of the timer was analyzed from TGA analysis TiO2 from dye absorbed on the TiO2. Fig. 4 shows the normalized ab­
(Fig. 3). The amount of dye loaded in the pan was 2.12 mg. In the course, sorption spectrum of dye adsorbed TiO2, bare TiO2, and the calculated
an initial 5% loss in weight (0.106 mg) can be accounted for due to the absorption spectrum of the dye. The corresponding optical data is
loss of adsorbed water molecules. The decomposition begins from exhibited in Table 1.
280 ◦ C rapidly up to 615 ◦ C accounting for 50% loss of mass (0.933 mg) A trimer shows a typical porphyrin absorption profile composed of
following a systematic degradation of the trimer dye, thereafter, a one strong Soret band and four Q bands confirming the formation of the
continuous but slow loss was observed up to 900 ◦ C (8%). The study metal-free porphyrin trimer with perfect electronic linking. The Soret
revealed good thermal stability up to 280 ◦ C probably due to strong band is strong and unsplit suggesting no intermolecular excitonic
linking with the triazine moiety. coupling among porphyrin units in their excited states reporting good
symmetry features [58]. The molar extinction (ε) values of dye in the
3.2. Absorption studies absorbed state were found smaller than the those in solution. The var­
iations are probably due to the interaction of the dye with DMSO [62].
The porphyrin exhibits a striking absorption feature that makes it The Q bands were broad with large molar absorption coefficients.
highly advantageous in photosynthesis, photocatalysis and DSSC ap­ broadening was observed due to S1 excited states interaction of all
plications. Porphyrin is a 26 π electron aromatic system out of which 18 porphyrin units. The optical bandgap of the trimer corresponds to the
π electrons are delocalised. A metal-free porphyrin exhibits a charac­ lowest energy peak in an absorption spectrum can be roughly estimated
teristic absorption spectrum consisting of the weak N, L, M bands in the as 1.5 eV = 1240/λ(nm).
UV regions, a strong (S0 →S2) B(Soret) band in the 380–500 nm range The accurate values of the bandgap energy are usually determined
and four Q bands in 500–750 nm range. Porphyrin engineering (intro­ from diffuse reflectance spectra (DRS). The measured reflectance spectra
duction of a single/multiple push-pull, oligomerization and conjugation can be transformed into the corresponding absorption spectra by
extension) at beta and meso positions often alter the absorbing ability of applying the Kubelka− Munk function [63].

Table 1
Absorption properties of the synthesized dye.
Dye Data Soret Band Q bands Bandgap (in eV)

λmax εmax I II III IV

λmax εmax λmax εmax λmax εmax λmax εmax


in DMSO 392 11500 542 6789 631 7100 743 6800 811 7280 1.45
Calculated 399 11224 545 6747 635 7088 748 6366 816 7356 1.49

Units: λmax - nm; εmax -M− 1cm− 1,calculated data for dye after subtracting TiO2 data from Dye + TiO2.

6
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Where, F(R) is the kubelka-Munk function (equivalent to absorption


coefficient α), ν - light photon energy, Eg-the bandgap of the material,
and n is a variable that takes value ½ or 2 for a direct allowed and in­
direct allowed type of transitions. Accordingly, we evaluated the
bandgap by plotting [F(R)hν]2 Vs (hν) plot (Fig. 5) to give the value 1.49
eV. The observed Eg was found to be lower than the bandgap of free
porphyrin [64] indicating the formation of tailored dye with a broader
absorption profile, as expected. A substantial lowering of the bandgap as
compared to a single porphyrin was observed. The optical band gap
obtained from solution spectra absorption spectra has shown a little
smaller (1.45 eV) than in the adsorbed state because of the solvent ef­
fects [62].

3.3. Cyclic voltammetry studies

CV is used to evaluate the E0Redox of a reversible/quasi reversible


reaction and to determine the magnitude of the frontier orbitals energies
(HOMO, LUMO) [65,66]. The CV of a 0.1 mM trimer dye solution in
DMSO containing 0.1 mol/L tetrabutylammonium hexafluorophosphate
(NBu4PF6), and ferrocene as internal standard was evaluated using a
three-electrode system at 25 ◦ C under purging of N2 gas. The scan rates
used were 0.01, 0.015, 0.020 mV/s, and the potential window was set as
− 2.2 to +1.5 V. (Fig. 6). As the E0Redox(Fc/Fc+) is highly solvent depen­
Fig. 6. CV of Dye in DMSO with Fc couple as internal standard. dent, its accurate value was evaluated by performing a similar experi­
ment in absence of the dye. The value of E0Redox(Fc/Fc+) was estimated as
− 16 mV vs. Ag/Ag+ (10 mM AgNO3) in DMSO which was found to agree
F(R) = K/S … (1) well with those exhibited in Fig. 6. The fact confirms that the ferrocene
does not exhibit reactions with the trimer dye [67]. Fig. 6 displays a pair
2
where K=(1-R) and S = 2R are the absorption and scattering co­ of quasi-reversible oxidation peaks (anodic peak, Ei = 0.611 V vs
efficients, respectively and R is reflectance. Ferrocene) and quasi-reversible reduction peak (cathodic peak, Ei =
The nature of transitions involved and the accurate bandgap is -1.146 V vs Ferrocene) corresponding to trimer dye, while the Fc couple
evaluated by using the Kubelka-Munk modified Tauc equation; exhibits an oxidation = 0.213 V and reduction peaks at − 0.162 V
respectively. All the redox peak potential has shown dependence on scan
[F(R)hν] 1/n
= B(hν- Eg) … (2)
rate as usual [68]. The absolute LUMO and HOMO energies of the trimer

Fig. 7. SEM of a) TiO2/FTO b) dye/TiO2/FTO, c) A star-shaped dye molecule over the TiO2 surface.

7
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

dye have been calculated by using [69];

E HOMO = - [Eox vs Fc + 5.043] eV … (3)

E LUMO = - [Ered vs Fc + 5.043] eV … (4)

Where Eox and Ered were respectively the oxidation and the reduction
potential of trimer dye. The EHOMO and ELUMO of the trimer were
calculated as – 5.654 eV and – 3.897 eV respectively. The exhibition of
quasi reversible behaviour in nonaqueous DMSO solvent suggests slow
electron transfer kinetics [70] The electrochemical bandgap obtained
from CV was 1.757 eV (ELUMO-EHOMO) which is larger than that optical
band gap obtained from the DRS spectrum of the adsorbed dye (1.49
eV), a common features exhibited by almost all organic dye in
nonaqueous solvents (the solvent effect) [38,71].

3.4. Scanning electron micrographs of photo-electrode

The morphology and the composition of TiO2/FTO and Dye/TiO2/


FTO electrodes have been investigated using an SEM equipped with
EDAX. The SEM image of TiO2/FTO (Fig. 7a) displays a typical meso­ Fig. 8. The Current-Voltage (I–V) plot in iodine redox electrolyte at 100
porous morphology consisting of TiO2 nano-colloids. While a Dye/TiO2/ mW/cm2.
FTO (Fig. 7b) reveals almost similar morphology and texture to TiO2/
FTO except for a little increase in the colloidal size. The images show a
uniform dye loading on the TiO2 surface without any appreciable ag­ Vmax*Jmax*100
η= (5)
gregation on the surface. The fact indicates the firm chemisorption of 100 mW
dye over the TiO2 surface. The vesting of nine –COOH groups on the
where Vmax and Jmax are the maximum voltage and maximum current
periphery with a planar core structure (*Fig S 15) might have led to
shown under the curve corresponding to maximum power.
multiple binding of trimer with TiO2, much similar to a three-pointed
The fill factor was evaluated using the equation;
starfish lending over a spherical object (Fig. 7c) [38,72]. The
star-shaped, planar dye molecules orient themselves horizontally to the Vmax*Jmax
ff = (6)
TiO2 surface favoring a strong π stacking with TiO2, which besides Voc*Jsc
strong binding executes excellent electronic communication [54–56,
73]. where Voc, Jsc refers to open-circuit voltage, short circuit current under
Another important electrode used in DSSC was Carbon Nano-Tubes no-load conditions.
(CNT) deposited over the FTO layer. This is considered a promising Five DSSC devices were evaluated under similar conditions. The
alternative to expensive Pt because of its highly conducting nature and stability of the DSSC device was tested once in a consecutive ten days to
excellent electrochemical activity because of the large surface, and know any changes to register. Fig. 8 shows the IV characteristics of the
nano-structure [74]. The composition of the dye and dye loaded over two cells showing very similar results (out of five). The IV curve showed
TiO2 has been evaluated by performing an EDAX analysis. EDAX of dye VOC = 0.800 V, JSC = 16.8 mA/cm2, Vmax = 0.560V, and Imax = 12.6
(Fig. S13) and dye loaded over TiO2 (Fig. S14) exhibits the peaks cor­ mA. Using equations (5) and (6), we evaluated efficiency (η) = 7.1% and
responding to the elements present besides confirming the weight a fill factor (ff) = 52.5%, respectively. The performance measurements
percent as per their compositions. carried out in a ten days program revealed no appreciable loss in the
performance (within experimental error range ±1%) revealing good
anti-aggregation and stability against the degrading photo-physical or
3.5. Photoelectrochemical studies chemical processes. Similarly, the IV performance of the photoanodes
having different dye sensitization times (4–12 h), evaluated to investi­
The dye engineering and reshuffling of ECB of semiconductors make gate the anti-aggregation stability did not exhibit any variations (within
the dye thermodynamically more capable for a DSSC. The addition of the 1% experimental-environmental error) accounting for the excellent
additives like hole-scavengers and other compensating dyes, and anti- anti-aggregation ability of the dye [76]. An excellent dye loading over
aggregating agents although enhancing the device performance was the TiO2 surface because of the firm bidentate linking [41] has also been
not long-lasting as the device shows too many chemical degradation corroborated by the SEM images, which almost show a similar texture
paths that deteriorate the performance and hence stability [75]. with a small increase in the grain size. The observation that unannealed
To analyze the performance and the stability of the trimer dye, we set triad/TiO2/FTO photoanodes registered poor performance over
up a DSSC cell using Dye/TiO2/FTO/Glass as photoanode, CNT/FTO/ annealed ones underlines the necessity of the annealing energy to
Glass as counter electrode and an aqueous 0.1 M KI and 0.02 M I2 based establish strong binding of carboxylate to the semiconductor.
redox couple as electrolyte. The I–V characteristics of the DSSC were We obtained the highest efficiency of 7.1% for multiple push-pull
measured on Potentiostat using linear scan voltammetry at the scan rate trimeric porphyrin dye under metal and additive-free, and energy-
of 1 mV/s in the 0 to +0.800 V potential range. The measurements were generating conditions with good stability. The literature survey re­
done at room temperature (300K) under the illumination of 100 mW/ veals no report on the porphyrin -triazene based trimers with multiple
cm2 intensity from the Xe source. push-pulls and anchors to compare with. Table 2 displays the compar­
To determine the energy parameters, the values of E0Redox = 0.502V ison of the performance of star-shaped dye with literature reported di­
for the iodide-iodine couple as determined from CV was used. The mers and trimers under various conditions. It was observed that
photoanode was annealed before use. The heating of the electrolytic literature cited dimeric and trimeric porphyrins dyes in the star archi­
solution was controlled by imposing a water barrier between the DSSC tecture, having a push-pull and a single or two carboxylate type
and the light source. anchoring hand exhibited poor performance even after using various
The efficiency of the DSSC was evaluated using the equation;

8
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Table 2
A table showing the comparison of the key parameters of the literature reported triazine based porphyrin dyes (monomer, dimer & trimers) with the trimer dye in this
study.
Dye Device Additive/ D-π-A Anchoring Redox couple, Band gap HOMO- Recombin- Efficiency Ref
Sensitizer groups Solvent used (eV) LUMO ation time (life (η), Voc &
(eV) time) Jsc

2,4,6-triphenyl-1,3,5- DSSC-Zn TBP, single 1-COOH I− /I−3 , 1.985* – -(2.04) 0.97% [79]
triazine-cored doped GNCS Acetonitrile + 1.029 V 514.6 V
porphyrin monomer valeronitrile vs NHE 2.5 mA
1,3,5-triazine-cored DSSC-Free TBP single 1-COOH I− /I−3 , 2.10* − 5.13 20 (23) #5.17% [76]
porphyrin monomer base Acetonitrile − 2.46 680 mV
(with 2 bodipy units) 10.86 mA
triazine-cored Porphyrin DSSC-Zn TBP, GNCS single 1-COOH I− /I−3 , 1.9 – – 0.13% [19]
dimer doped Acetonitrile 0.78V vs 501 mV
NHE 0.39 mA
unsymmetrical triazine- DSSC- TBP single 1-COOH I− /I−3 , 2.02* − 5.56 8 (24) 4.46 [43]
cored Triphenyl Partially Zn Acetonitrile (2.57) − 3.51 0.660 V
Porphyrin dimer doped 9.94 mA
triazine-cored Porphyrin DSSC-Zn TBP single 2- COOH I− /I−3 , 1.91 − 5.440 16.4 (27.8) 3.45% [38]
dimer doped Acetonitrile + − 3.320 640 mV
valeronitrile 5.93 mA
triazine-cored Porphyrin DSSC-Zn CDCA single 2-COOH I− /I−3 - 2.41* − 5.069 18 (24) 4.56% [80]
dimer doped − 2.349 680 mV
1.23 mA
triazine-cored Porphyrin DSSC-Zn CDCA single 1-COOH I− /I−3 , 2.08 − 5.237 2.43 (17.47) 5.56% [58]
trimer doped CH2Cl2 (2.694*) − 2.542 0.640V
– 12.42 mA
triazine-cored Porphyrin BHJ-two PP – single 1-COOH – 1.94 − 5.63 0.14 3. 93% [59]
trimer Zn doped – − 3.55 – 0.92 V
8.06 mA
Triazine core Porphyrin DSSC- No triple 9-COOH I− /I−3 Aqueous 1.74* − 5.654 0.2 (5.9) #7.056% This
trimer Freebase − 3.897 800 mV study
16.8 mA

BHJ- Bulk heterojunction cell, TBP- tert-butyl pyridine, GNCS- Guanidinium thiocyanate.
*Electrochemical bandgap #Free base porphyrin sensitizer.

Fig. 9. Nyquist plot (Z′′ vs Z′ ) showing frequency for charge transfer, recom­
bination in terms of a middle-range semicircle and a Warburg component. The
inset shows an equivalent impedance circuit.

Fig. 10. A plot of Z vs log F(Bode plot): Showing inflections points corre­
additive/sensitizers. sponding to various impedances (time constants).
The maximum obtainable Voc (Ef or ~ ECB -E0Redox) from the DSSC is
1.039V, which is the highest among all literature reported on triazine
into the ECB band of the TiO2 was found to be sufficiently positive
based porphyrin dimer and trimer. The loss in VOC was limited to 0.239
+0.381 eV on an absolute scale (ECB-ELUMO = − 3.879 + 4.26 eV) and is
V. The Eloss, defined as Eloss = Eg− eVOC is another parameter used to
enough high to inject charge from dye to TiO2 [77,78]. The other
evaluate the performance of solar cell (Eg-the bandgap of the dye and
thermodynamic driving parameter, ΔGreg (EHOMO-ERedox) computes the
eVoc - Voc expressed in eV). Most organic-based DSSC suffers from high
energy required to regenerate the redox couple and was calculated as
Eloss values up to 0.7–0.8 eV. The DSSC showing Eloss > 0.6 eV often
+0.355 eV. The positive driving force ensures efficient regeneration of
exhibits a dramatic drop in quantum efficiencies and ISC. A highly
the redox species at the counter electrode-solution interface.
desired value of <0.5 has been suggested for high performing porphyrin-
The as-synthesized trimer dye although exhibited good performance
based DSSC [41]. The Eloss evaluated for our trimer was 0.451 eV.
has still potential for further modifications, by the way of doping or side-
The driving force (ΔGinj) required to inject electron from dye LUMO
chain engineering to improve the performance further.

9
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

Table 3
EIS parameters of trimer based DSSC: Various time constants and other important cell parameters.
Rs (Ω) Cdl (F) Rct(Ω) Rt(Ω) τ1 (ms) τ2 (ms) τ3 (s) W(Ω) Dn(cm/s) d(μm) Ln(μm)
− 6 − 8
41.3 8.89 × 10 110 36.1 0.2 5.9 0.37 59 7.95 × 10 4.0 2.17

3.6. Electrochemical impedance (EIS) studies leading to a 7.1% PCE under AM 1.5G irradiation accounting to the
smallest ELoss of 0.451 V The firm binding of dye via nine carboxymethyl
A DSSC involves Faradiac charge transfers and therefore behaves as a linkers and a planar covering of dye over semiconductor surface
non-polarizable electrode system when in contact with electrolytes. The exhibited excellent anti-aggregation tendency and electronic linking of
measurements of EIS could provide important electrochemical data on the dye. A metal-free trimer dye was found stable and efficient sensitizer
the DSSC. The EIS of the DSSC was taken under the bias of 0.800 V using for a true energy-generating DSSC under additive-free conditions.
a two-electrode system. The impedances were measured in the
0.01–1000 kHz frequency range. The obtained experimental data were Compliance with ethical standards
fitted to a Randles equivalent circuit containing the parameters such as
solution resistance (Rs), charge transfer resistance (Rct), layer capaci­ All the authors are aware of ethical standard and hereby declare that
tance (Cdl), and the Warberg parameter [81]. Figs. 9 and 10 show EIS the research proposed fully comply with journal’s ethical research
data in terms of the Nyquist plot and the Bode plots respectively [82]. policies.
Nyquist plot (Fig. 9) exhibits a single semicircle (middle frequency)
concentering to a line of 45◦ slope towards the lowest frequency region. Funding
This is typical behaviour shown by iodide-triiodide-based DSSC
involving a kinetically controlled process coupled with mass transfer All the authors declare that no funding/grants or any other financial
phenomena across the interface. A middle semicircle is associated with support were received for this research work.
three-time constants (or frequencies) associated with processes in the
DSSC, viz. electron injection (diffusion of photogenerated electron into Research data policy and data availability statements
TiO2 electrode, ωed=1/τ1), charge recombination (lifetime of charges
generated at photoelectrode, ωk = 1/τ2) and ion diffusion (ωf = 1/τ3). The data generated and or analyzed in this research study are pro­
The frequencies (ω) of these processes were evaluated from the Bode vided in this manuscript and or included in the supplementary infor­
plot as inflections in the high, middle, and low-frequency regions, mation file. This is also available with the corresponding author and can
respectively [83]. The evaluated data have been tabulated in Table 3. A be obtained on request.
much smaller charge injection time (τ 1, 0.2 ms) than the lifetime of the
photogenerated charge (τ2, 5.9 ms) indicates a rapid injection of charge CRediT authorship contribution statement
to TiO2 than any recombination representing excellent charge separa­
tion capabilities [38]. The star-shaped triad showed much faster charge Darpan V. Bhuse: Synthetic part, data collection and, Data curation,
injection time and lifetimes than most of the triazine based dimers and Writing – original draft. Vijaykumar M. Bhuse: Data curation, and,
trimers [58,59,84]. It follows that Rt < Rct for working of the DSSC as an Formal analysis, and Interpretation. Pundlik R. Bhagat: Conceptuali­
energy generator. The trimer dye exhibits higher values of Rct (110 Ω) zation, Writing – review & editing, final editing, reviewing and pre­
indicating lesser recombination tendency (or more lifetime of the charge sentation of the manuscript. All authors commented on previous
carriers). The higher value is associated with the D− π− A nature of the versions read and approved the final version of the manuscript.
dye, which disturbs the dispersion forces on the dye by trapping I−3 near
the porphyrin more effectively, preventing the approach of I−3 to the Declaration of competing interest
linker group and the surface region [30,43,85]. A higher ionic diffusion
time constant (τ3, 0.37 s) observed indicates that the DSSC system is The authors declare that they have no known competing financial
under diffusion control. The electron diffusion coefficient (Dn) and the interests or personal relationships that could have appeared to influence
charge diffusion length (Ln) in the thin film based photo-anode were the work reported in this paper.
evaluated using equations; Dn = d2/τ1 and Ln = (Dn τ1) 1/2 (d ~ 4 μm, the
thickness of the TiO2 film) and were found to be 7.95 × 10− 8 cm/s, and Acknowledgement
2.17 μm respectively. The Ln was found, almost half of the thickness of
TiO2 film used indicating an effective collection of photogenerated The authors are thankful to VIT Vellore for providing ‘VIT SEED
charge carriers (Table 3). An equivalent Randles electronic circuit for GRANT’ to perform this research work. We also acknowledge the fa­
the device is shown as an inset in Fig. 9. It involves the combination of cilities provided by SIF DST-VITFIST, SEM SBST, RGEMS and Smart
Rct, Cdl in series with Rs, and the Warberg resistance (W).The Rs account Materials Laboratory for Bio-Sensing and Catalysis for providing basic
for the total resistance of solution including those of FTO sheet and facilities at VIT, Vellore. The thanks are also extended to the Institute of
electrical wirings. Science, Nagpur for providing CV, EIS measuring facilities.

4. Conclusions Appendix A. Supplementary data

A new, D-π-A type, trimeric triazine cored porphyrin small molecule Supplementary data to this article can be found online at https://doi.
dye has been designed. In all nine carboxymethyl functionalized org/10.1016/j.matchemphys.2022.126312.
hydroxy-phenyl-benzimidazolium moieties were attached to three por­
phyrins (three each at meta position). An N, N′ , N’’-(1,3,5-triazine-2,4,6- References
triyl) trimethanimine core was used as a stable, electron-deficient spacer
to link functionalized porphyrins. The star-shaped dye exhibits a planar [1] B. Parida, S. Iniyan, R. Goic, A review of solar photovoltaic technologies, Renew.
geometry with a narrow bandgap of (Eg, 1.49 eV) and frontier energies Sustain. Energy Rev. 15 (2011) 1625–1636, https://doi.org/10.1016/j.
rser.2010.11.032.
of − 5.654 eV (HOMO) and − 3.879 eV (LUMO) and exhibited a perfect [2] IEA, Global, Energy review: CO2 emissions in 2020, Glob. Energy Rev (2021) 1–15.
alignment with iodide-based redox couple and TiO2 semiconductor https://www.iea.org/articles/global-energy-review-co2-emissions-in-2020.

10
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

[3] M.E. Yeoh, K.Y. Chan, Recent advances in photo-anode for dye-sensitized solar [28] J. Chen, S. Ko, L. Liu, Y. Sheng, H. Han, X. Li, The effect of different alkyl chains on
cells: a review, Int. J. Energy Res. 41 (2017) 2446–2467, https://doi.org/10.1002/ the photovoltaic performance of D-π-A porphyrin-sensitized solar cells, New J.
er.3764. Chem. 39 (2015) 3736–3746, https://doi.org/10.1039/c4nj02263g.
[4] X. Zhang, Y. Zhu, X. Wu, H. He, G. Wang, Q. Li, Meso-Schiff-base substituted [29] T. Higashino, K. Sugiura, Y. Tsuji, S. Nimura, S. Ito, H. Imahori, A push pull
porphyrin dimer dyes for dye-sensitized solar cells: synthesis, electrochemical, and porphyrin dimer with multiple electron-donating groups for dye-sensitized solar
photovoltaic properties, Res. Chem. Intermed. 41 (2015) 4227–4241, https://doi. cells: excellent light-harvesting in near-infrared region, Chem. Lett. 45 (2016)
org/10.1007/s11164-013-1525-1. 1126–1128, https://doi.org/10.1246/cl.160591.
[5] T.W. Hamann, J.W. Ondersma, Dye-sensitized solar cell redox shuttles, Energy [30] S. Chaurasia, C.J. Liang, Y.S. Yen, J.T. Lin, Sensitizers with rigidified-aromatics as
Environ. Sci. 4 (2011) 370–381, https://doi.org/10.1039/c0ee00251h. the conjugated spacers for dye-sensitized solar cells, J. Mater. Chem. C. 3 (2015)
[6] M. Freitag, J. Teuscher, Y. Saygili, X. Zhang, F. Giordano, P. Liska, J. Hua, S. 9765–9780, https://doi.org/10.1039/c5tc02356d.
M. Zakeeruddin, J.E. Moser, M. Grätzel, A. Hagfeldt, Dye-sensitized solar cells for [31] J. Lu, B. Zhang, H. Yuan, X. Xu, K. Cao, J. Cui, S. Liu, Y. Shen, Y. Cheng, J. Xu,
efficient power generation under ambient lighting, Nat. Photonics 11 (2017) M. Wang, D-π-A porphyrin sensitizers with π-extended conjugation for mesoscopic
372–378, https://doi.org/10.1038/nphoton.2017.60. solar cells, J. Phys. Chem. C 118 (2014) 14785–14794, https://doi.org/10.1021/
[7] E. Kouhestanian, M. Ranjbar, S.A. Mozaffari, H. Salaramoli, Investigating the jp5030608.
effects of thickness on the performance of ZnO-based DSSC, Prog. Color. Color. [32] A.S. Hart, C.B. Kc, H.B. Gobeze, L.R. Sequeira, F. D’Souza, Porphyrin-sensitized
Coatings. 14 (2021) 101–112. solar cells: effect of carboxyl anchor group orientation on the cell performance,
[8] R.S. Shelke, S.B. Thombre, S.R. Patrikar, Status and perspectives of dyes used in ACS Appl. Mater. Interfaces 5 (2013) 5314–5323, https://doi.org/10.1021/
dye sensitized solar cells, Int. J. Renew. Energy Resour. 3 (2013) 12–19. am401201q.
[9] N.A. Ludin, A.M. Al-Alwani Mahmoud, A. Bakar Mohamad, A.A.H. Kadhum, [33] S. Mathew, A. Yella, P. Gao, R. Humphry-Baker, B.F.E. Curchod, N. Ashari-Astani,
K. Sopian, N.S. Abdul Karim, Review on the development of natural dye I. Tavernelli, U. Rothlisberger, M.K. Nazeeruddin, M. Grätzel, Dye-sensitized solar
photosensitizer for dye-sensitized solar cells, Renew. Sustain. Energy Rev. 31 cells with 13% efficiency achieved through the molecular engineering of porphyrin
(2014) 386–396, https://doi.org/10.1016/j.rser.2013.12.001. sensitizers, Nat. Chem. 6 (2014) 242–247, https://doi.org/10.1038/nchem.1861.
[10] H. Song, Q. Liu, Y. Xie, Porphyrin-sensitized solar cells: systematic molecular [34] G. Di Carlo, S. Caramori, L. Casarin, A. Orbelli Biroli, F. Tessore, R. Argazzi,
optimization, coadsorption and cosensitization, Chem. Commun. 54 (2018) A. Oriana, G. Cerullo, C.A. Bignozzi, M. Pizzotti, Charge transfer dynamics in β- and
1811–1824, https://doi.org/10.1039/c7cc09671b. meso-substituted dithienylethylene porphyrins, J. Phys. Chem. C 121 (2017)
[11] M.G. Walter, A.B. Rudine, C.C. Wamser, Porphyrins and phthalocyanines in solar 18385–18400, https://doi.org/10.1021/acs.jpcc.7b05823.
photovoltaic cells, J. Porphyr. Phthalocyanines 14 (2010) 759–792, https://doi. [35] A. Carella, F. Borbone, R. Centore, Research progress on photosensitizers for DSSC,
org/10.1142/S1088424610002689. Front. Chem. 6 (2018) 1–24, https://doi.org/10.3389/fchem.2018.00481.
[12] T. Higashino, H. Imahori, Porphyrins as excellent dyes for dye-sensitized solar [36] J. Kesters, P. Verstappen, M. Kelchtermans, L. Lutsen, D. Vanderzande, W. Maes,
cells: recent developments and insights, Dalton Trans. 44 (2015) 448–463, https:// Porphyrin-based bulk heterojunction organic photovoltaics: the rise of the colors of
doi.org/10.1039/c4dt02756f. life, Adv. Energy Mater. 5 (2015) 1–20, https://doi.org/10.1002/
[13] W.M. Campbell, K.W. Jolley, P. Wagner, K. Wagner, P.J. Walsh, K.C. Gordon, aenm.201500218.
L. Schmidt-Mende, M.K. Nazeeruddin, Q. Wang, M. Grätzel, D.L. Officer, Highly [37] L. Zhang, J.M. Cole, Dye aggregation in dye-sensitized solar cells, J. Mater. Chem.
efficient porphyrin sensitizers for dye-sensitized solar cells. https://doi.org/ A. 5 (2017) 19541–19559, https://doi.org/10.1039/c7ta05632j.
10.1021/jp0750598, 2007. [38] G.E. Zervaki, P.A. Angaridis, E.N. Koukaras, G.D. Sharma, A.G. Coutsolelos, Dye-
[14] T. Lazarides, G. Charalambidis, A. Vuillamy, M. Réglier, E. Klontzas, G. Froudakis, sensitized solar cells based on triazine-linked porphyrin dyads containing one or
S. Kuhri, D.M. Guldi, A.G. Coutsolelos, Promising fast energy transfer system via an two carboxylic acid anchoring groups, Inorg. Chem. Front. 1 (2014) 256–270,
easy synthesis: bodipy-porphyrin dyads connected via a cyanuric chloride bridge, https://doi.org/10.1039/c3qi00095h.
their synthesis, and electrochemical and photophysical investigations, Inorg. [39] E. Galoppini, Linkers for anchoring sensitizers to semiconductor nanoparticles,
Chem. 50 (2011) 8926–8936, https://doi.org/10.1021/ic201052k. Coord. Chem. Rev. 248 (2004) 1283–1297, https://doi.org/10.1016/j.
[15] J. Gong, K. Sumathy, Q. Qiao, Z. Zhou, Review on dye-sensitized solar cells ccr.2004.03.016.
(DSSCs): advanced techniques and research trends, Renew. Sustain. Energy Rev. 68 [40] K. Gao, Y. Kan, X. Chen, F. Liu, B. Kan, L. Nian, X. Wan, Y. Chen, X. Peng, T.
(2017) 234–246, https://doi.org/10.1016/j.rser.2016.09.097. P. Russell, Y. Cao, A.K.Y. Jen, Low-bandgap porphyrins for highly efficient organic
[16] A. Mohammad Bagher, Comparison of organic solar cells and inorganic solar cells, solar cells: materials, morphology, and applications, Adv. Mater. 32 (2020) 1–19,
Int. J. Renew. Sustain. Energy 3 (2014) 53, https://doi.org/10.11648/j. https://doi.org/10.1002/adma.201906129.
ijrse.20140303.12. [41] L. Xiao, T. Lai, X. Liu, F. Liu, T.P. Russell, Y. Liu, F. Huang, X. Peng, Y. Cao, A low-
[17] F. Arkan, M. Izadyar, Structural modification as the pioneer strategy in competition bandgap dimeric porphyrin molecule for 10% efficiency solar cells with small
of the porphyrin dye and perovskite solar cells: from dynamics to kinetics of the photon energy loss, J. Mater. Chem. A. 6 (2018) 18469–18478, https://doi.org/
photovoltaic processes, Appl. Phys. Lett. 115 (2019), https://doi.org/10.1063/ 10.1039/c8ta05903a.
1.5113901. [42] J.K. Park, J. Chen, H.R. Lee, S.W. Park, H. Shinokubo, A. Osuka, D. Kim, Doubly
[18] A. Hagfeldt, U.B. Cappel, G. Boschloo, L. Sun, L. Kloo, H. Pettersson, E.A. Gibson, β-functionalized meso - meso directly linked porphyrin dimer sensitizers for
Dye-sensitized Photoelectrochemical Cells, Elsevier Ltd, 2018, https://doi.org/ photovoltaics, J. Phys. Chem. C 113 (2009) 21956–21963, https://doi.org/
10.1016/B978-0-12-809921-6.00014-8. 10.1021/jp905675x.
[19] A. Luechai, N. Pootrakulchote, T. Kengthanomma, P. Vanalabhpatana, [43] G.E. Zervaki, M.S. Roy, M.K. Panda, P.A. Angaridis, E. Chrissos, G.D. Sharma, A.
P. Thamyongkit, Photosensitizing triarylamine- and triazine-cored porphyrin G. Coutsolelos, Efficient sensitization of dye-sensitized solar cells by novel triazine-
dimers for dye-sensitized solar cells, J. Organomet. Chem. 753 (2014) 27–33, bridged porphyrin-porphyrin dyads, Inorg. Chem. 52 (2013) 9813–9825, https://
https://doi.org/10.1016/j.jorganchem.2013.12.025. doi.org/10.1021/ic400774p.
[20] D. V Bhuse, P.R. Bhagat, Synthesis and characterization of a conjugated porphyrin [44] T. Zhang, X. Qian, P.F. Zhang, Y.Z. Zhu, J.Y. Zheng, A meso-meso directly linked
dyad entangled with carboxyl functionalized benzimidazolium : an efficient metal porphyrin dimer-based double D-π-A sensitizer for efficient dye-sensitized solar
free sensitizer for DSSCs, New J. Chem. 45 (2021) 1430–1445, https://doi.org/ cells, Chem. Commun. 51 (2015) 3782–3785, https://doi.org/10.1039/
10.1039/d0nj05387b. c4cc09640a.
[21] J. Min Park, J.H. Lee, W.D. Jang, Applications of porphyrins in emerging energy [45] Y. Kurumisawa, T. Higashino, S. Nimura, Y. Tsuji, H. Iiyama, H. Imahori,
conversion technologies, Coord. Chem. Rev. 407 (2020), 213157, https://doi.org/ Renaissance of fused porphyrins: substituted methylene-bridged thiophene-fused
10.1016/j.ccr.2019.213157. strategy for high-performance dye-sensitized solar cells, J. Am. Chem. Soc. 141
[22] Y. Matsuo, K. Ogumi, Y. Matsuo, I. Jeon, T. Nakagawa, Y. Matsuo, H. Wang, (2019) 9910–9919, https://doi.org/10.1021/jacs.9b03302.
K. Ogumi, Recent progress in porphyrin- and phthalocyanine-containing perovskite [46] W. Auwärter, D. Écija, F. Klappenberger, J.V. Barth, Porphyrins at interfaces, Nat.
solar cells, RSC Adv. 10 (2020) 32678–32689, https://doi.org/10.1039/ Chem. 7 (2015) 105–120, https://doi.org/10.1038/nchem.2159.
d0ra03234d. [47] M.M. Al Mogren, N.M. Ahmed, A.A. Hasanein, Molecular modeling and
[23] L.L. Li, E.W.G. Diau, Porphyrin-sensitized solar cells, Chem. Soc. Rev. 42 (2013) photovoltaic applications of porphyrin-based dyes: a review, J. Saudi Chem. Soc.
291–304, https://doi.org/10.1039/c2cs35257e. 24 (2020) 303–320, https://doi.org/10.1016/j.jscs.2020.01.005.
[24] NREL, Best, Research-cell efficiency Chart.pdf, nrel. https://www.nrel.gov/pv/ [48] G. Magna, D. Monti, C. Di Natale, R. Paolesse, M. Stefanelli, The assembly of
assets/pdfs/best-research-cell-efficiencies.20190802.pdf, 2019, 1. porphyrin systems in well-defined nanostructures: an update, Molecules 24 (2019),
[25] F. Sahli, J. Werner, B.A. Kamino, M. Bräuninger, R. Monnard, B. Paviet-Salomon, https://doi.org/10.3390/molecules24234307.
L. Barraud, L. Ding, J.J. Diaz Leon, D. Sacchetto, G. Cattaneo, M. Despeisse, [49] S. Chen, L. Yan, L. Xiao, K. Gao, W. Tang, C. Wang, C. Zhu, X. Wang, F. Liu,
M. Boccard, S. Nicolay, Q. Jeangros, B. Niesen, C. Ballif, Fully textured monolithic X. Peng, W.K. Wong, X. Zhu, A visible-near-infrared absorbing A-π2-D-π1-D-π2-A
perovskite/silicon tandem solar cells with 25.2% power conversion efficiency, Nat. type dimeric-porphyrin donor for high-performance organic solar cells, J. Mater.
Mater. 17 (2018) 820–826, https://doi.org/10.1038/s41563-018-0115-4. Chem. A. 5 (2017) 25460–25468, https://doi.org/10.1039/c7ta06217f.
[26] F. Zhang, K. Zhu, Additive engineering for efficient and stable perovskite solar [50] T. Altalhi, A.A. Gobouri, M.S. Refat, M.M. El-Nahass, A.M. Hassanien, A.A. Atta, H.
cells, Adv. Energy Mater. 10 (2020) 1–26, https://doi.org/10.1002/ A. Saad, A.N. Alhazaa, Structural, electrochemical and optical properties of 1,2,4-
aenm.201902579. triazine derivative, Appl. Phys. Mater. Sci. Process 126 (2020) 1–9, https://doi.
[27] S. Rafique, S.M. Abdullah, K. Sulaiman, M. Iwamoto, Fundamentals of bulk org/10.1007/s00339-020-03995-4.
heterojunction organic solar cells: an overview of stability/degradation issues and [51] P. Puthiaraj, Y.R. Lee, S. Zhang, W.S. Ahn, Triazine-based covalent organic
strategies for improvement, Renew. Sustain. Energy Rev. 84 (2018) 43–53, https:// polymers: design, synthesis and applications in heterogeneous catalysis, J. Mater.
doi.org/10.1016/j.rser.2017.12.008. Chem. A. 4 (2016) 16288–16311, https://doi.org/10.1039/c6ta06089g.

11
D.V. Bhuse et al. Materials Chemistry and Physics 287 (2022) 126312

[52] H. Zhong, H. Lai, Q. Fang, New conjugated triazine based molecular materials for electroanalytical tools for studying reaction mechanisms, Chem. Sci. 10 (2019)
application in optoelectronic devices: design, synthesis, and properties, J. Phys. 6404–6422, https://doi.org/10.1039/c9sc01545k.
Chem. C 115 (2011) 2423–2427, https://doi.org/10.1021/jp109806m. [69] M. Namazian, C.Y. Lin, M.L. Coote, Benchmark calculations of absolute reduction
[53] J. Chen, T. Zhang, S. Wang, R. Hu, S. Li, J.S. Ma, G. Yang, Intramolecular potential of ferricinium/ferrocene couple in nonaqueous solutions, J. Chem. Theor.
aggregation and optical limiting properties of triazine-linked mono-, bis- and tris- Comput. 6 (2010) 2721–2725, https://doi.org/10.1021/ct1003252.
phthalocyanines, Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 149 (2015) [70] Y.H. Chiang, H.H. Chou, W.T. Cheng, Y.R. Li, C.Y. Yeh, P. Chen, Porphyrin dimers
426–433, https://doi.org/10.1016/j.saa.2015.04.093. as hole-transporting layers for high-efficiency and stable perovskite solar cells, ACS
[54] Z. Lu, C. Li, T. Fang, G. Li, Z. Bo, Triindole-cored star-shaped molecules for organic Energy Lett. 3 (2018) 1620–1626, https://doi.org/10.1021/
solar cells, J. Mater. Chem. A. 1 (2013) 7657–7665, https://doi.org/10.1039/ acsenergylett.8b00607.
c3ta11182b. [71] S. Admassie, O. Inganäs, W. Mammo, E. Perzon, M.R. Andersson, Electrochemical
[55] M. Guerrini, E.D. Aznar, C. Cocchi, Electronic and optical properties of protonated and optical studies of the band gaps of alternating polyfluorene copolymers, Synth.
triazine derivatives, J. Phys. Chem. C 124 (2020) 27801–27810, https://doi.org/ Met. 156 (2006) 614–623, https://doi.org/10.1016/j.synthmet.2006.02.013.
10.1021/acs.jpcc.0c08812. [72] K. Sharma, V. Sharma, S.S. Sharma, Dye-sensitized solar cells: fundamentals and
[56] T. Kengthanomma, P. Thamyongkit, J. Gasiorowski, A.M. Ramil, N.S. Sariciftci, On current status, Nanoscale Res. Lett. 13 (2018), https://doi.org/10.1186/s11671-
the potential of porphyrin-spiked triarylamine stars for bulk heterojunction solar 018-2760-6.
cells, J. Mater. Chem. A. 1 (2013) 10524–10531, https://doi.org/10.1039/ [73] S.Y. Shiau, C.H. Chang, W.J. Chen, H.J. Wang, R.J. Jeng, R.H. Lee, Star-shaped
c3ta11095h. organic semiconductors with planar triazine core and diketopyrrolopyrrole
[57] T. Hamamura, J.T. Dy, K. Tamaki, J. Nakazaki, S. Uchida, T. Kubo, H. Segawa, branches for solution-processed small-molecule organic solar cells, Dyes Pigments
Dye-sensitized solar cells using ethynyl-linked porphyrin trimers, Phys. Chem. 115 (2015) 35–49, https://doi.org/10.1016/j.dyepig.2014.12.007.
Chem. Phys. 16 (2014) 4551–4560, https://doi.org/10.1039/c3cp55184a. [74] S. Hwang, M. Batmunkh, M.J. Nine, H. Chung, H. Jeong, Dye-sensitized solar cell
[58] G.E. Zervaki, E. Papastamatakis, P.A. Angaridis, V. Nikolaou, M. Singh, counter electrodes based on carbon nanotubes, ChemPhysChem 16 (2015) 53–65,
R. Kurchania, T.N. Kitsopoulos, G.D. Sharma, A.G. Coutsolelos, A propeller-shaped, https://doi.org/10.1002/cphc.201402570.
triazine-linked porphyrin triad as efficient sensitizer for dye-sensitized solar cells, [75] G. Hodes, Photoelectrochemical cell measurements: getting the basics right,
Eur. J. Inorg. Chem. (2014) 1020–1033, https://doi.org/10.1002/ejic.201301278. J. Phys. Chem. Lett. 3 (2012) 1208–1213, https://doi.org/10.1021/jz300220b.
[59] G.D. Sharma, G.E. Zervaki, P.A. Angaridis, T.N. Kitsopoulos, A.G. Coutsolelos, [76] Z.E. Galateia, N. Agapi, N. Vasilis, G.D. Sharma, C.G. Athanassios, Scorpion-shaped
Triazine-bridged porphyrin triad as electron donor for solution-processed bulk mono(carboxy)porphyrin-(BODIPY)2, a novel triazine bridged triad: synthesis,
hetero-junction organic solar cells, J. Phys. Chem. C 118 (2014) 5968–5977, characterization and dye sensitized solar cell (DSSC) applications, J. Mater. Chem.
https://doi.org/10.1021/jp500090h. C. 3 (2015) 5652–5664, https://doi.org/10.1039/c4tc02902j.
[60] V.B. Khajone, P.R. Bhagat, Synthesis of polymer-supported Brønsted acid- [77] I. Chung, B. Lee, J. He, R.P.H. Chang, M.G. Kanatzidis, All-solid-state dye-
functionalized Zn–porphyrin complex, knotted with benzimidazolium moiety for sensitized solar cells with high efficiency, Nature 485 (2012) 486–489, https://doi.
photodegradation of azo dyes under visible-light irradiation, Res. Chem. Intermed. org/10.1038/nature11067.
46 (2020) 783–802, https://doi.org/10.1007/s11164-019-03990-2. [78] D.A. Panayotov, S.P. Burrows, J.R. Morris, Infrared spectroscopic studies of
[61] S.U. Raut, P.R. Bhagat, Efficient photocatalytic acetalization of furfural to biofuel conduction band and trapped electrons in UV-photoexcited, H-Atom n-doped, and
components using carboxyl-functionalized porphyrin photocatalyst, under visible thermally reduced TiO 2, J. Phys. Chem. C 116 (2012) 4535–4544, https://doi.
light irradiations, Biomass Convers. Biorefinery (2021), https://doi.org/10.1007/ org/10.1021/jp2053103.
s13399-021-01658-9. [79] J. Chen, S. Ko, L. Liu, Y. Sheng, H. Han, X. Li, The effect of porphyrins suspended
[62] V. Venkatraman, A.E. Yemene, J. De Mello, Prediction of Absorption Spectrum with different electronegative moieties on the photovoltaic performance of
Shifts in Dyes Adsorbed on Titania, 2019, pp. 1–13. monolithic porphyrin-sensitized solar cells with carbon counter electrodes, New J.
[63] P. Makuła, M. Pacia, W. Macyk, How to correctly determine the band gap energy of Chem. 39 (2015) 2889–2900, https://doi.org/10.1039/c4nj02186j.
modified semiconductor photocatalysts based on UV-vis spectra, J. Phys. Chem. [80] G.E. Zervaki, V. Tsaka, A. Vatikioti, I. Georgakaki, V. Nikolaou, G.D. Sharma, A.
Lett. 9 (2018) 6814–6817, https://doi.org/10.1021/acs.jpclett.8b02892. G. Coutsolelos, A triazine di(carboxy)porphyrin dyad versus a triazine di(carboxy)
[64] M. Dayer, A. Moosavi-Movahedi, M. Dayer, Band Assignment in hemoglobin porphyrin triad for sensitizers in DSSCs, Dalton Trans. 44 (2015) 13550–13564,
porphyrin ring spectrum: using four-orbital model of gouterman, Protein Pept. Lett. https://doi.org/10.1039/c5dt01141h.
17 (2010) 473–479, https://doi.org/10.2174/092986610790963645. [81] A. Lasia, Electrochemical impedance spectroscopy and its applications. https://doi.
[65] M. Al-Ibrahim, H.K. Roth, M. Schrödner, A. Konkin, U. Zhokhavets, G. Gobsch, org/10.1007/978-1-4614-8933-7, 2014.
P. Scharff, S. Sensfuss, The influence of the optoelectronic properties of poly(3- [82] A. Sacco, Electrochemical impedance spectroscopy: fundamentals and application
alkylthiophenes) on the device parameters in flexible polymer solar cells, Org. in dye-sensitized solar cells, Renew. Sustain. Energy Rev. 79 (2017) 814–829,
Electron. 6 (2005) 65–77, https://doi.org/10.1016/j.orgel.2005.02.004. https://doi.org/10.1016/j.rser.2017.05.159.
[66] E.M. Espinoza, J.A. Clark, J. Soliman, J.B. Derr, M. Morales, V.I. Vullev, Practical [83] S. Sarker, A.J.S. Ahammad, H.W. Seo, D.M. Kim, Electrochemical impedance
aspects of cyclic voltammetry: how to estimate reduction potentials when spectra of dye-sensitized solar cells: fundamentals and spreadsheet calculation, Int.
irreversibility prevails, J. Electrochem. Soc. 166 (2019) H3175–H3187, https:// J. Photoenergy 2014 (2014), https://doi.org/10.1155/2014/851705.
doi.org/10.1149/2.0241905jes. [84] G.D. Sharma, P.A. Angaridis, S. Pipou, G.E. Zervaki, V. Nikolaou, R. Misra, A.
[67] V.V. Pavlishchuk, A.W. Addison, Conversion constants for redox potentials G. Coutsolelos, Efficient co-sensitization of dye-sensitized solar cells by novel
measured versus different reference electrodes in acetonitrile solutions at 25◦ C, porphyrin/triazine dye and tertiary aryl-amine organic dye, Org. Electron. 25
Inorg. Chim. Acta. 298 (2000) 97–102, https://doi.org/10.1016/S0020-1693(99) (2015) 295–307, https://doi.org/10.1016/j.orgel.2015.06.048.
00407-7. [85] M. Berhe Desta, S. Chaurasia, J.T. Lin, Reversed Y-shape di-anchoring sensitizers
[68] C. Sandford, M.A. Edwards, K.J. Klunder, D.P. Hickey, M. Li, K. Barman, M. for dye sensitized solar cells based on benzimidazole core, Dyes Pigments 140
S. Sigman, H.S. White, S.D. Minteer, A synthetic chemist’s guide to (2017) 441–451, https://doi.org/10.1016/j.dyepig.2017.01.068.

12

You might also like