Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

pubs.acs.

org/cm Article

Molecular Engineering of Robust Starburst-Based Organic


Photosensitizers for Highly Efficient Photocatalytic Hydrogen
Generation from Water
Yan-Yi Kwok, Po-Yu Ho, Ying Wei, Zhong Zheng, Sze-Chun Yiu, Cheuk-Lam Ho,*
and Shuping Huang*
Cite This: Chem. Mater. 2022, 34, 5522−5534 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via HONG KONG POLYTECHNIC UNIV on September 22, 2023 at 05:35:16 (UTC).

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Six donor−donor−π−acceptor (D−D−π−A) triphenyl-


amine-based starburst organic dyes with different electron-donating
moieties and thiophene-based π-linkers were synthesized and characterized.
Their photophysical and electrochemical properties, together with their
photocatalytic hydrogen evolution performance as photosensitizers (PSs),
were investigated. Distinctive and prolonged hydrogen evolution perform-
ances of these PSs under visible-light irradiation from water in their
platinized TiO2 composites were demonstrated: a turnover number (TON)
of up to 24 900 (252 h) with 1560 μmol (37.6 mL) hydrogen produced, an
initial turnover frequency (TOFi) of 1130 h−1, initial activity (Activityi) of
705 mmol g−1 h−1, and initial apparent quantum yield (AQYi) of 12.1%. To
the best of our knowledge, according to the TOF and TON values, the
designed PS system is one of the most efficient and robust photocatalytic
H2 generation systems adopting a TiO2-anchoring molecular PS in the literature. The results showed that the starburst triarylamine
donor moiety with phenothiazine functionality and the alkyl chain on the thiophene π-linker could boost the performance and
longevity of the photocatalytic system, providing a powerful strategy for the development of starburst-based highly efficient and
robust D−D−π−A organic photosensitizers in the future.

1. INTRODUCTION than the Eg of semiconductors could facilitate the process by


With the surging demand for energy and resources, the extending the capacity of spectral response and improving the
development of sustainable, renewable, and clean energy has visible-light-induced activities. Sacrificial electron donors
been in the limelight of research in recent decades. (SEDs) are often employed as additives in the photocatalytic
Photocatalytic hydrogen evolution has emerged as a promising water-splitting reactions to scavenge the photogenerated holes
method to exploit alternative and sustainable energy resources. (the radical cation PS•+ return to PS, regeneration of dye) and
The hydrogen generated is an ideal energy carrier that is suppress the recombination of charge carriers.6−10
carbon footprint-free and inexhaustible.1,2 The water-splitting Different types of light-harvesting chromophores, including
process involves two redox half-reactions, namely, proton metal complexes and organic compounds, have been exploited
reduction into a hydrogen molecule at the cathode and water as PSs for the water-splitting process. An ideal PS should be
oxidation at the anode. Our research focuses on hydrogen able to resist water corrosion and have broad spectral
production efficiency, which can be assessed by investigating absorption and high charge carrier mobility to achieve effective
the reduction part of the water-splitting process independ- charge separation.8,11,12 Metalated PSs anchoring on TiO2,
ently.3 especially ruthenium and iridium complexes, have been used in
The cleavage of water into hydrogen and oxygen using photocatalysis due to their attractive performance and effective
platinum nanoparticles combined with TiO2 has been receiving metal-to-ligand charge transfer.9,13−15 From the viewpoint of
much attention since 1972, as this is a simple way to utilize
visible light for photocatalytic hydrogen generation.4,5 It works
when the energy of the absorbed photon is larger than the Received: February 21, 2022
band energy of the semiconductor photocatalyst. However, as Revised: May 27, 2022
the band gap (Eg) of TiO2 is larger than 3 eV, its light- Published: June 9, 2022
harvesting region is restricted to the UV region. Introducing
PSs with the highest occupied molecular orbital (HOMO)−
lowest unoccupied molecular orbital (LUMO) gap smaller

© 2022 American Chemical Society https://doi.org/10.1021/acs.chemmater.2c00556


5522 Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Figure 1. Molecular structures of dyes P1−P6.

reducing the energy costs that are required to fabricate solar starburst donor moieties were recorded as effective blockers to
water-splitting devices, metal-free organic dyes have emerged prevent direct contact between electrolyte/aqueous media and
as an attractive cost-effective alternative over metal complexes TiO2 through steric hindrance, hence decelerating charge
for photocatalysis. Organic dyes offer structural versatility to recombination in the charge-separated state.23−25 The large
tune the absorption wavelengths with superior molar extinction surface area occupied by the starburst donor moieties on TiO2
coefficients (ε) and charge-transfer ability as compared with could improve dyes’ regeneration rate and was proven to
the conventional metalated dyes.13,16−18 In particular, π- improve the photovoltaic efficiency in DSSCs with the positive
conjugated molecules with electron donor−acceptor moieties charge localized on the dye.7,23,26 On the other hand,
for effective intramolecular charge transfer (ICT) exhibit triphenylamine (TPA)-based dyes are known for their high
noticeable photovoltaic conversion efficiencies (PCEs) for dye- stabilities and photovoltaic activities.27 Attributed to TPA’s
sensitized solar cell (DSSC) applications. Due to such nonplanarity and exceptional hydrophobicity, the aggregation
convenient ICT modulation, large capacity and cross section of dyes on the semiconductors’ interface and charge
for absorption, and efficient charge separation upon photo- recombination of the dye layer interface could be inhibited
excitation, organic PSs are also applicable to water split- by blocking out water molecules from approaching to the
ting.9,19−21 However, most of the reported rod-shaped organic TiO2−carboxylate interface, while the molar extinct coefficient
PSs are prone to aggregate on the TiO2 surface and induce and broadness of the absorption spectrum of PS could be
intermolecular quenching.22 To further strengthen the stability enhanced. Incorporating the electron-donating substituents at
against aggregation and prevent charge recombination, the the para-position of the phenyl groups of TPA, which are
dyes discussed herein were designed with a starburst donor electrochemically active, would enhance the electron richness
moiety with conjugated linkers. The bulky and nonplanar and energy/transfer efficiency, stabilize the cationic radicals of
5523 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

TPA, and decrease its oxidation potential, such that the distillation with respective drying agents. All of the reactions were
stability of the PSs could be enhanced drastically upon the monitored by thin-layer chromatography (TLC) using silica-gel-
electron injection.7,16,28−31 Phenothiazine, fluorene, and coated aluminum plates from Merck. Column chromatography with
carbazole, which were selected to incorporate as electron- silica gel (230−400 mesh) from DAVISIL was used for material
purification.
donating functionalities on TPA in the present study, are all 2.2. Instrumentation. 1H and 13C nuclear magnetic resonance
redox-active chromophores, with significant redox cyclability, (NMR) spectra were measured in methanol-d4, DMSO-d6, or
persistent radical cation formation, and electron-donating chloroform-d, with tetramethylsilane (TMS) used as an internal
ability.9,32−34 The electron-rich sulfur and nitrogen atoms in standard for chemical shift calibration on a Bruker Avance-III 400
phenothiazine can enhance the delocalization of electrons from MHz FT-NMR system. The integration and labeling of peaks were
the donor to the acceptor and are highly stable in the cationic performed by a Bruker NMR software Suite-TopSpin. Mass spectra
radical form; hence, an extra stability would be provided for were collected by an Agilent 6540 liquid chromatography-electrospray
the photo-induced electron transfer to the semiconductor. ionization quadrupole-time-of-flight (ESI-QTOF) mass spectrometer.
Phenothiazine’s nonplanarity can also prevent dyes from Ultraviolet−visible spectra were obtained by a Varian Cary 4000 of a
UV−visible spectrophotometer at 293 K, with the samples dissolved
aggregation.35−38 in dichloromethane. Photoluminescence spectra were acquired by an
Cyanoacetic acid, which has an electron-withdrawing cyano Agilent G9800AA Cary Eclipse (type II) fluorescence spectropho-
group, is the most commonly used acceptor. The carboxylate tometer at 293 K, with the samples dissolved in dichloromethane.
group in cyanoacetic acid forms an ester linkage with the Fluorescence lifetimes of the PSs were determined by a Photon
semiconductor to provide effective charge injection as its Technology International (PTI) QuantaMaster 50 steady-state
energy level overlaps well with the CB of TiO2.39−41 The spectrofluorometer at 293 K. The Marquardt-based nonlinear least-
selected thiophene and its derivatives are usually used as squares fitting routine was adopted for analyzing the decay curves.
conjugated electron-donating linkers that can extend electronic Those curves were presented in a biexponential function, regarding
conjugation and facilitate the charge transfer over the donor the equation I(t) = A1 exp (−t/τ) + A2 exp (−t/τ 2). Cyclic
voltammetry was performed on a CHI 630C electrochemical
and acceptor cores of PSs.42−44 The hydrophobic alkyl linkers analyzer/workstation, with a scan rate of 100 mV s−1. The
can defend the interface from the attack by water and prevent experiments were performed in dichloromethane solution, using a
dye aggregation, thus stabilizing dye−TiO2 nanocomposites three-electrode cell with a 3 mm glassy carbon working electrode, an
under aqueous conditions.45−47 Meanwhile, the hexyl group on aqueous Ag/AgCl reference electrode, and a platinum counter
the thiophene unit may impair the molecular coplanarity to electrode. Tetrabutylammonium hexafluorophosphate (0.1 M) was
prevent the π-aggregation on the TiO2 surface and, more used as the supporting electrolyte, and ferrocene was applied as an
importantly, to restrain the charge recombination. Moreover, internal standard, so that the corresponding one-electron oxidation
the hexylthiophene group was also introduced to evaluate the potential could be measured. Electrochemical impedance spectrosco-
effect of molecular planarity on optical, electrochemical, and py (EIS) was operated in a three-electrode system, in which the
carbon rod and Ag/AgCl electrode were used as counter and
photocatalytic properties. reference electrodes, respectively, and 0.1 M KCl solution was used as
Based on such structural consideration, starburst-based an electrolyte. The samples were dispersed in a mixture solvent of 1
metal-free organic PSs P1−P6 that exploited the D−D−π−A mL of methanol and 0.1 mL of Nafion solution with a mass of 1 mg.
framework with different starburst triarylamine donor moieties The prepared samples were dropped on the platinum carbon
were synthesized in this study (Figure 1). These molecules electrode, and EIS data were collected by applying a sine wave with
matched well with the prerequisites of PSs in photocatalytic H2 an amplitude of 50 mV over a frequency range from 100 000 to 0.01
generation. Light-driven hydrogen generation studies from Hz on a CHI 760 electrochemical workstation.
aqueous media were carried out. The results revealed that the 2.3. Preparation of Platinized Titanium (IV) Oxide (Pt-TiO2).
starburst framework could effectively prevent the π-aggregation The 0.5 wt% platinized TiO2 was prepared by adding 1.6 g of TiO2
nanopowder (anatase, 99.7% trace metal basis, particle size <25 nm
on the TiO2 surface and restrain the charge recombination. from Sigma-Aldrich) and 0.1 mL of H2PtCl6 (8 wt %) solution into 40
Remarkably, P4 attained an active and robust H2 generation mL of methanol. The gray slurry was stirred vigorously under the
system with a turnover number (TON) of up to 24 900 in 252 radiation from a mercury-coated lamp (300 W, HF300PD, EYE
h with 1560 μmol (37.6 mL) hydrogen produced. The initial Lighting) for 24 h. The grayish crude product was purified by washing
turnover frequency (TOFi) was 1130 h−1, the initial activity with methanol and centrifuged (3500 rpm, 5 min) thrice. Pt-TiO2 was
(Activityi) was 710 mmol g−1 h−1, and the initial apparent then dried in a vacuum oven at 60 °C for 8 h.
quantum yield (AQY i ) was 12.1% under visible-light 2.4. Preparation of the Dye-Loaded Platinized Titanium(IV)
irradiation from water. The performance of our photocatalytic Oxide. The prepared Pt-TiO2 (20 mg) was added to 2.5 mL (50 μM)
of PS-dissolved dichloromethane solution, followed by sonication
water-splitting systems is, to the best of our knowledge, one of
thoroughly at 20 °C for 30 min. TiO2 nanoparticles would turn from
the highest reported thus far for three-component systems gray to pale purple-red and would be distributed in the colored milky
utilizing PS-anchoring Pt-TiO2 nanoparticle composites solution. The solution mixture was then centrifuged at 3500 rpm for
(Tables S3−S6).9,34,48,49 This report provides a logical strategy 10 min to reclaim the nanoparticles by removing the colorless
for the development of efficient PSs for water splitting and supernatant and drying under vacuum in the dark overnight. The
other photocatalytic applications. dried pellet would be directly used for the photocatalytic hydrogen
evolution experiment. The collected supernatant was used for
estimating the dye-loading percentage for each PS separately by
2. EXPERIMENTAL SECTION comparing absorbance differences before and after the dye adsorption
2.1. Materials and Reagents. All chemical reactions were of their low-energy peaks. These carboxylate-anchoring dyes were
conducted under a nitrogen atmosphere with Schlenk link techniques. favorably attached onto the TiO2 surfaces and yielded close to 100%
All glassware were cleaned and dried in an oven before use. The adsorption. The corresponding UV/vis absorption spectra of P1−P6
commercially available reagents were used without further purifica- in CH2Cl2 before and after dye adsorption are illustrated in Figure
tion, unless otherwise specified. These reagents were purchased from S52 for dye-loading percentage calculation.
Sigma-Aldrich, Acros Organics, Tokyo Chemical Industry Co. Ltd., 2.5. Photocatalytic Hydrogen Evolution Studies. The experi-
Dieckmann, J&K Scientific, and Macklin. Solvents were dried by ments were conducted by adding the dye-loaded TiO2-Pt pellet (∼20

5524 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Scheme 1. Synthetic Scheme of P1−P3a

a
(i) Pd(PPh3)4, THF/H2O, K2CO3; (ii) N-bromosuccinimide, dry THF; (iii) Pd(PPh3)4, THF/H2O, K2CO3; and (iv) piperidine, cyanoacetic
acid, dry chloroform.

mg) and 5 mL of aquatic ascorbic acid (AA, 0.8M) solution at pH 4.0, sample dye. The water-splitting performance was also presented in
the latter of which acted as the SED for regenerating the PSs, into a terms of turnover number (TON), turnover frequency (TOF), initial
25 mL pear-shaped flask. AA was chosen because of its well-studied turnover frequency (TOFi), and initial photocatalytic activity
relationship between hydrogen generation efficiency and pH and its (Activityi).
illustrious redox reactions.50 The sealed pear-shaped flask was then 2.6. Theoretical Calculation. To gain insights into the energy
purged with a gas mixture of nitrogen/methane (80:20 mol %) for 10 levels and electronic transition processes upon photoexcitation and
min. The presence of methane was important and served as the conformational relaxation, density functional theory (DFT) and time-
internal standard for the gas chromatography (GC) analysis for every dependent density functional theory (TD-DFT) calculations and the
trial. The reaction mixture would be stirred rapidly at 19 °C with the molecular geometries of five dyes at the first excited singlet states were
blue (ca. 472 nm) or green (ca. 520 nm) light-emitting diodes optimized by Gaussian 09 program51 at the B3LYP/6-311G(d,p) and
(LEDs) under a just-fit container, which blocked out the environ- MPWPW91/6-311G(d,p) levels, respectively. The PBE0 hybrid
mental stray light and prevented the loss of scattered light from the functional and the 6-31G* basis set were used. The solvent effects
LEDs. Powers of the LEDs were measured by a power meter (model: in CH2Cl2 media were taken into account by performing the self-
BIM-7001, Hangzhou Brolight Technology Co., Ltd.) with a thermal consistent reaction field (SCRF) calculations using the polarizable
sensor (model: BIM-7203-0100F). The measured power was ∼80 continuum model (PCM).52,53 The optimized geometries were
mW for each reaction flask. At each time point, the headspaces of confirmed with all real frequencies.
each flask were detected by GC (Shimadzu GC-8A with 5 Å
molecular sieve column and a thermal conductivity detector) to 3. RESULTS AND DISCUSSION
determine the volume of hydrogen produced based on the peak area
ratio of hydrogen/methane with reference to the calibration curve
3.1. Synthesis and Characterization. The molecular
(Figure S48). The time points taken were 5, 21, 26, 42, 47, 63, 68, 84, structures of the newly synthesized organic PSs are displayed
89, 105, 126, 147, 168, 189, 210, 231, and 252 h from 14 independent in Figure 1, while their synthetic routes are illustrated in
reaction flasks (duplicate for each sample) for P1−P6 and blank (Pt− Schemes 1 and 2. The key dibromo intermediates, 5-(4-(bis(4-
TiO2). The blue/green LED irradiation was assumed to be bromophenyl)amino)phenyl)thiophene-2-carbaldehyde (3)
monochromatic with a maximum emission intensity at 472/520 nm and 5-(4-(bis(4-bromophenyl)amino)phenyl)-4-hexylthio-
for calculating the initial apparent quantum yield (AQYi%) for each phene-2-carbaldehyde (10), were prepared by palladium-
5525 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Scheme 2. Synthetic Scheme of P4−P6a

a
(i) n-Butyllithium, dimethylformamide, dry THF; (ii) N-bromosuccinimide, dimethylformamide; (iii) Pd(PPh3)4, THF/H2O, K2CO3; (iv) N-
bromosuccinimide, dry THF; (v) Pd(PPh3)4, THF/H2O, K2CO3; and (vi) piperidine, cyanoacetic acid, dry chloroform.

catalyzed Suzuki coupling reactions, followed by dibromination characterized by electrospray ionization quadrupole-time-of-
with N-bromosuccinimide (NBS). The Suzuki coupling flight mass spectrometry and 1H NMR spectroscopy. The
reaction was then applied again with corresponding aromatic specific downfield singlet peaks observed at around δ 8.5 ppm
boronic acids, followed by Knoevenagel condensation with on their 1H NMR spectra imply the presence of the CC
cyanoacetic acid to yield the final products P1−P6. All of the double bond of cyanoacetic acid. The success of the
organic precursors were characterized by 1H and 13C NMR aforementioned final Knoevenagel condensation could be
spectroscopies. The singlet peak found at around δ 10.0 ppm supported by the disappearance of the aldehyde peaks at δ
in the aldehyde precursors corresponds to the proton of the 10.0 ppm and the rise of those at δ 8.5 ppm regarding the
aldehyde functional group. All of the target organic PSs were conversion of aldehyde to cyanoacrylic acid. However,
5526 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

acquisition of good quality of the 13C NMR spectra of P1−P6 Table 1. Data of UV/Vis Absorption, Emission, and
was difficult due to their partial solubilities in common organic Lifetimes of P1−P6 in CH2Cl2
solvents. All spectra could be found in Figures S1−S47.
λmax [nm] λem [nm] (excited at
3.2. Photophysical Properties. The UV−visible absorp- dye (ε[ M−1 cm−1)]) 400 nm) τ [ns] (χ2)
tion and photoluminescence spectra of P1−P6 were measured
P1 360 (37 000) 588 2.87 (1.12)
in dichloromethane and are presented in Figure 2. The
500 (23 500)
P2 359 (43 800) 641 1.54 (1.29)
488 (23 000)
P3 345 (45 700) 626 2.62 (1.27)
494 (21 900)
P4 363 (77 100) 619 3.06 (1.17)
471 (30 400)
P5 363 (70 200) 598 3.93 (1.03)
466 (24 200)
P6 340 (88 800) 646 2.42 (1.25)
474 (28 500)

ruthenium dyes.56 In general, the absorption bands of the


π−π* transition maxima are influenced by the nature of
electron-donating moieties on the TPA rings in P1−P6, with a
bathochromic shift following the order of carbazole (λmax =
345 and 340 nm for P3 and P6, respectively) < fluorene (λmax
= 359 and 363 nm for P2 and P5, respectively) ∼
phenothiazine (λmax = 360 and 363 nm for P1 and P4,
respectively). Besides, the optical properties of P1−P6 were
also affected by the π-conjugation bridge to a considerable
extent. The peak absorption wavelengths of the ICT bands
showed substantial hypsochromic effects when comparing P1
(λmax = 500 nm) to P4 (λmax = 471 nm), P2 (λmax = 488 nm)
to P5 (λmax = 466 nm), and P3 (λmax = 494 nm) to P6 (λmax =
474 nm), after the substitution of a hexyl moiety into the
thiophene linker. This was probably due to the planarity
difference of the dyes. The introduction of a hexyl unit in the
thiophene ring increased the dihedral angles between π-bridge
and triphenylamine planes, which were calculated at their
optimized geometries at the ground singlet states by the
Gaussian 09 program. The calculated dihedral angles of P1−
P6 were 20.11, −20.69, 19.63, 39.36, −39.57, and 38.26°,
respectively (Table S1). This meant that the introduction of
the hexyl unit twisted the π-conjugation system and disrupted
the planarity to induce a light blue shift on the absorption band
and hinder the ICT process from the donor to the acceptor.57
On the other hand, the electron-donating strength of 3-
hexylthiophene was obviously larger than thiophene, which
should induce a stronger ICT process within the dye. The
Figure 2. (a) UV/vis absorption spectra and (b) photoluminescence altered planar conjugation of thiophene by the vibrations of the
spectra (excitation wavelength = 400 nm) of P1−P6 in CH2Cl2 at 293 malleable hexyl chain involved a trade-off among the
K. characteristics of their light absorption.27 In addition, a
significant increment of almost twice was observed in ε of
corresponding data are tabulated in Table 1. All of the organic the PSs in π−π* transitions with hexylthiophene groups than
PSs display two major broad absorption bands from 300 to 650 those with thiophene, implying that the hexyl chain at the β-
nm. The absorption bands with a shorter wavelength (centered position of thiophene was beneficial to the light-harvesting
at around 350 nm) are attributed to the localized π−π* capability in the near UV region. The calculated absorption
transitions throughout the electron-rich aromatic rings, while spectra for P1−P6 (Figure S51) were obtained by the PBE0
those with a longer wavelength (centered at around 470−500 functional. The positions of the calculated absorption peaks of
nm) are attributed to the intramolecular charge transfer (ICT) the dyes were consistent with the experimental ones.
from the electron-donating moieties to the electron-with- All of the dyes demonstrated photoluminescence in
drawing anchoring group. The molar extinction coefficients (ε) dichloromethane at room temperature, showing structureless
(25 000−90 000 mol −1 cm −1 ) of the target dyes are broad peaks with emission maxima (λem) ranging from 588 to
comparable to those of reported D−π−A starburst donor- 646 nm. The emission spectra over the measured spectral
based PSs26,30,48,54,55 and much higher than those typical window presented no band from the triplet excited state, and
5527 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

the lifetimes (τ) of P1−P6 were measured in 1.54−3.93 ns LUMO energy level differences could be calculated (ranging
(Table 1), thus fluorescent in nature.58 from 2.12 to 2.22 eV). The values of those dyes with
3.3. Electrochemical Properties. To figure out the hexylthiophene were larger than those with thiophene as the π-
impact of molecular engineering on energy levels, cyclic linker, proving that changes in the π-linker in a D−π−A
voltammetry (CV) was used to investigate the electrochemical framework were influential to the separation between the
properties of these starburst PSs, including the optimal energy HOMO and LUMO, while a trend showed that the donor
levels of the highest occupied molecular orbital (HOMO) and moieties would have little effect to the electrochemical band
the energy offsets of the synthesized PSs. P1−P6 were added gap (dihexylfluorene > hexylcarbazole > hexylphenothiazine).
into the supporting electrolyte (0.1 M tetrabutylammonium Owing to the better planarity of thiophene, the E0‑0 values of
hexafluorophosphate) in dichloromethane solution at a P1−P3 were determined to be 2.12, 2.14, and 2.13 V,
scanning rate of 100 mV s−1. The reference electrode was respectively, which were slightly narrower than those of P4−
Ag/AgCl calibrated with ferrocene−ferrocenium (Fc+/Fc) as P6 (2.21, 2.22, and 2.20 V, respectively). Besides, the PSs’
an internal reference. The cyclic voltammograms in Figure S49 ELUMO were calculated to be ranged from −2.63 to −2.94 eV,
showed that the dyes displayed either reversible or quasi- which were more positive than the conduction band of TiO2
reversible electron transfer at the oxidative potentials (E vs Ag/ (−4.0 eV).15 The much larger negativity of TiO2 than that of
Ag+ > 0 V). Referring to the results illustrated in Table 2 and the LUMO of the dyes would be favorable to the rapid
electron injection because of the larger charge separation, thus
Table 2. Electrochemical Data and Energy Levels for P1−P6 the stronger driving force.16,61 It is worthy to note that the
dye Eox, Va EHOMO, eVb E0−0, eVc Eox*, Vd ELUMO, eVe LUMO energy level of P4 is more positive than the conduction
P1 0.05 −4.85 2.12 −2.07 −2.73
band of TiO2, compared to the other four compounds,
P2 0.28 −5.08 2.14 −1.86 −2.94
indicating a stronger driving force for electron injection. This
P3 0.25 −5.05 2.13 −1.88 −2.92
was the reason for the outstanding performance of P4. The
P4 0.04 −4.84 2.21 −2.17 −2.63
HOMO levels of the dyes with the same donor moieties were
P5 0.28 −5.08 2.22 −1.94 −2.86 measured, and the results showed that they had to be almost
P6 0.27 −5.07 2.20 −1.93 −2.87 the same values (P1 vs P4, P2 vs P5, and P3 vs P6) (Figure
3). This means the planarity adjustment of the π-linker could
a
Onset of first oxidation potentials was measured by cyclic
voltammetry in dichloromethane (CH2Cl2) solution containing 0.1 effectively tune the LUMO level nearly without any influence
M of NBu4PF6 against the Ag/Ag+ reference electrode, with a glassy of the HOMO level of the dye, showing a quite promising
carbon working electrode and platinum wire auxiliary electrodes. strategy to selectively optimize energy levels of the dye. The
b
Calculated from −(Eox + 4.36), as reversible oxidation of ferrocene experimental energy levels for P1−P6 (Table 2) generally
was E1/2 = 0.44 V and the EHOMO of ferrocene is equal to −4.80 eV vs agreed with the calculated values (Table S2).
to vacuum level. cE0−0 was determined from the onset of the Electrochemical impedance spectroscopy (EIS) was per-
absorption spectrum at the corresponding lowest energy absorption formed to evaluate the charge recombination and the
maxima. dEox* = Eox + E0−0. eELUMO = EHOMO + E0−0. resistance of charge transfer of the selected photosensitizers
based on the line model of typical transmission.62,63 The
Figure 3, the first oxidation potentials (Eox) of the PSs P1−P6, Nyquist plots of dyes P1−P4 and P6 are shown in Figure S50.
which coincided with the HOMO energy levels, were found to The arc radii of P2 > P3 > P1 > P6 ∼ P4 further confirm that
be 0.05, 0.28, 0.25, 0.04, 0.28, and 0.27 V, respectively. the incorporation of hexylthiophene as the spacer could
Therefore, the calculated EHOMO values of P1−P6 were found effectively reduce the charge transport resistance and enhance
to be ranged from −4.84 to −5.08 eV, which were more the interfacial electron transfer.63−66 This also supports the
negative than the redox potential energy level of ascorbic acid performance of dyes in the photocatalytic hydrogen
(−4.65 eV at pH 4).59 This confirms that the dye regeneration production.
from the oxidation state could occur expeditiously by the 3.4. Theoretical Calculation. Investigation of frontier
thermodynamic driving force during the photocatalytic water molecular orbitals of dye molecules is essential for studying the
splitting.48,60 The absorption spectra showed the zero−zero charge-separated states. The molecular orbital diagram in
excitation energies (E0−0) of the dyes, and thus the HOMO− Figure 4 shows that the lowest unoccupied molecular orbital

Figure 3. Illustration of the energy levels (HOMO and LUMO) of P1−P6.

5528 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

substantial contribution to the LUMO, which facilitated a


significant electronic coupling with the TiO2 surface and
electron injection efficiency. On the contrary, the highest
occupied molecular orbital (HOMO) for the compounds was
delocalized along the donor units, mainly at the triphenylamine
for all compounds. It is worth noting that the HOMOs of P1
and P4 were more localized at one of the phenothiazine units
and they were not found at all in the electron-accepting units,
while those of P3 and P6 were distributed among the carbazole
and triphenylamine units with little HOMOs located at the
thiophene unit. The difference in HOMO and LUMO orbitals
for P4 > P1 > P6 > P3 > P5 > P2 indicated that the electrons
in TiO2 had the last tendency to recombine with the holes in
phenothiazine of P4.67 Although they had different electron
donors extended from the triphenylamine center, their
HOMOs were still distributed along triphenylamine, resulting
in the almost the same HOMO energy levels (around −5.15
eV) except P2 (−5.38 eV) and P5 (−5.34) from Table S2.
The LUMO energy levels of the six compounds varied from
−2.44 to −2.60 eV (Table S2), resulting in different energy
gaps (P2 (2.86 eV) > P5 (2.80 eV) > P3 (2.72 eV) > P1 ∼ P6
(2.60 eV) > P4 (2.55 eV)).
3.5. Photocatalytic Hydrogen Production. On account
of the outstanding light-harvesting ability of the PSs in the
400−600 nm region, it is foreseeable that P1−P6 would have
great potential for effective photocatalytic water splitting. It has
been previously reported by other works that platinum-
containing dyads used as PSs enabled unprecedentedly
efficient and steady H2 production from water after attaching
to platinized TiO2 nanoparticles.16,68 Herein, we carry out the
light-driven H2 generation studies by adopting the same
photocatalytic system for our organic PSs. In short, the H2
generation experiments were conducted in aqueous solutions
at pH 4.0, with AA (0.8 M) serving as the hole scavenger.
Similar to previous reports, GC analysis based on a methane
internal standard calibration method was used to measure the
amount of H2 produced at the end of the irradiation. The
percentage of dye loading on TiO2 was measured by changes in
absorbance of the dyes’ solution after the adsorption (Figure
S52), which are inversely related to the bulkiness of the PSs
because of the size-matching effect.69,70 The dye-loading
percentages found for our dyes laid between 96 and 98%
(Table 3) due to their similar molecular volumes. The detailed
procedures of platinizing TiO2 and dye loading onto TiO2 are
stated in Section 2.
The hydrogen generation fitting curves for the five dyes are
presented in Figure 5, and the corresponding data (TON,
TOF, TOFi, Activityi, and AQYi%) are tabulated in Table 3.
P1−P6 PSs facilitated the photocatalytic hydrogen evolution
(ranging from 5.0 to 37.6 mL) when compared with the
control bare Pt−TiO2 (4.9 mL) under blue-light irradiation
(472 nm) for more than 105 h (Figure 5a). The H2 production
increment was quite drastic for all PSs in the first 21 h, except
for P4, which had the ability to sustain that to up to 47 h. The
system with P4 had the highest and most remarkable
photocatalytic activity, with 37.6 mL of hydrogen and TON
of 24 900 over 252 h, while the TON values of P1−P3, P5,
Figure 4. Surface plots of P1−P6 estimated from DFT calculations and P6 were 11 500, 3320, 6560 (over 169 h), 6980 (over 171
using the B3LYP/6-311G(d,p) basis set. h), and 15 800 (over 105 h), respectively, under blue light
irradiation. To compare the performances equitably with other
(LUMO) of all dyes was mainly localized at the π-bridge and reported dyes, the initial hydrogen production activity
acceptor units (thiophene and 2-cyanoacrylic acid). The (activityi) and the initial apparent quantum yield (AQYi) of
carboxylic acid anchoring group of the five dyes had a our PSs were also calculated. It is worthy to note that P4
5529 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Table 3. Light-Driven H2 Generation Data With and Without PSs P1−P6 on Pt−TiO2
dye DL%a H2 (mL) TONb TOF (h−1)c TOF (h−1) optimal values at plateau TOFi (h−1)d activityi (mmol g−1 h−1)e AQYi%f
Irradiation under Blue LED (472 nm) for 169 h
P1 97 17.3 11 500 68.1 124 (84 h) 1020 636.1 11.1
P2 98 5.01 3320 19.7 35.3 (84 h) 315 197.0 3.53
P3 98 9.90 6560 38.8 73.0 (84 h) 607 379.1 6.29
P4 97 37.6 (252 h) 24 900 98.9 218 (105 h) 1130 704.5 12.1
P5 96 10.5 (171 h) 6980 40.8 221 138.0 2.36
P6 97 23.9 (105 h) 15 800 151 333 (47 h) 943 589.6 10.1
Pt-TiO2 4.90 106.6 1.77
Irradiation under Green LED (520 nm) for 169 h
P1 97 6.53 4330 25.6 33.0 (84 h) 151 94.3 2.41
P2 98 1.34 888 5.26 6.74 (84 h) 51.2 32.0 0.936
P3 98 5.17 3430 20.3 27.4 (84 h) 134 83.9 2.38
P5 96 5.19 (171 h) 3440 20.1 132 82.8 1.42
Pt-TiO2 1.03 20.9 0.478
210 h
P4 97 6.18 4090 19.5 26.7 (147 h) 170 106.3 2.80
P6 97 3.78 2510 11.9 16.3 (147 h) 121 75.6 1.99
a
Dye-loading percentage. bTurnover number of H2 is calculated as 2 × number of moles of H2 produced divided by the number of moles of PS
attached to platinized TiO2. cTurnover frequency is calculated per hour. dInitial turnover frequency in the first 5 h. eInitial photocatalytic activity of
the system is defined as the number of micromoles of H2 evolved per gram of platinum loaded per hour. fInitial apparent quantum yield percentage
(AQYi%) of the system.

achieved exceptionally high Activityi of 705 mmol g−1 h−1 and could have more effective excitation. Hexylphenothiazine (in
AQYi of 12.1%, and P6 came the second with decent Activityi P1) and hexylcarbazole (in P3) are, therefore, better electron-
of 590 mmol g−1 h−1 and AQYi of 10.1% after irradiating for 5 donating moieties than dihexylfluorene (in P2). Although the
h. By comparing their TOF, TON, and Acitivtyi values, the energy gap of P1 (and P4) is similar to P3 (and P6), the
system with P4 is certainly one of the most capable and robust ELUMO of P1 (and P4) is much smaller than that of P3 (and
photocatalytic water-splitting systems using organic-PS/TiO2/ P6), which is more favorable for expeditious electron injection
Pt under visible light. 3,48,71 The durability was also to TiO2;71−73 thus, better performance of P1 (and P4) than
considerably strong, as the H2 evolution reactivity and TOF P3 (and P6) under both blue- and green-light irradiation could
of P4 could sustain beyond 250 h, suggesting superior be observed.
photostability of the P4-sensitized system. Most of the During the photocatalytic process, decomposition of PS may
metal-free PSs for photocatalytic H2 generation reported in occur upon light irradiation due to the formation of the
the literature suffer from instability (only active within 10 h) reduced dyes (PS−) via the reductive quenching pathway,9,75
and/or poor activity (TOF < 100 h−1),9 and our results show which is commonly observed in most of the reported
that P4 has the ability to retard dye degradation or desorption metallated76 and metal-free PSs.34,71,77 To strengthen their
from TiO2 over the prolonged irradiation.72−74 Besides, a longevity, therefore, a hydrophobic, bulky, and electron-rich
comparison between the reaction mixture prior and after the starburst donor structure was adopted in the PS, so that the
irradiation (Figure S54) showed that the aquatic AA solution excited species PS* would be unfavorably reduced by the SED
turned from colorless to orange-yellow due to the oxidation of and any unfavorable chemical quenchers, thus shielding the
ascorbic acid to dehydroascorbic acid to regenerate PS during formation of PS− on the TiO2 surface.46 The large surface area
the water splitting. provided by the starburst donor in our dyes could bolster the
On the other hand, after the green-light irradiation (520 dye regenerating process, such that the decomposition of the
nm) for 169 h, P1 remained active as shown in Figure 5b. With metastable oxidized PS+ species was hindered.78−80
6.5 mL of hydrogen generated by P1 with TON of 4330, the By their steric hindrance, adding alkyl groups on the
performances of our dyes in green light were in the order of P1 electron-rich moieties and thiophene rings could prevent dye
> P4 > P3 ∼ P5 > P6 > P2 in terms of the final amount of H2 aggregation and modify the orientation of the dyes to become
produced. However, the Activityi and AQYi of P5 were favorable for electron injection to TiO2 and suppress charge
significantly higher than those of P2, while those of P4 were recombination.44,81−84 The adventurous effect of dye regener-
slightly higher than those of P1. Such a phenomenon was not ation and precluding undesirable charge recombination could
observed between P3 and P6. It should be noted that the six be further enhanced by introducing the hydrophobic alkyl
PSs exhibited a different pattern of activity under different light group at the central moieties of the scaffold, for instance,
irradiation. The higher TON and volume of H2 produced by hexylthiophene in the π-spacers of P4−P6, as the SED can
P4−P6 than P1−P3 under blue-light irradiation (472 nm) contact with PS better and more surface of TiO2 could be
could be attributed to the blue shift and much more covered, respectively, compared to adding the alkyl chains only
pronounced absorbance of P4−P6 in the shorter wavelength to the donors.85−88 This structural design helped P1−P6 to get
region. More blue light photons could be absorbed by P4−P6 better endurance as compared with other PSs that were
for dyes’ excitation, so more electrons could be transferred to without the starburst donor moiety, and they could stay active
Pt−TiO2 for water reduction. With the narrower HOMO− over the long period of light irradiation time (over 100 h in
LUMO energy gap of P1 and P3 than that of P2, P1, and P3 general and over 200 h under blue-light irradiation for P4).
5530 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

TON of 24 900, TOFi of 1130 h−1, and AQYi% of 12.1.


Notably, the hydrophobic and electron-rich starburst-based
structure, with the introduction of the alkyl group(s) to the
thiophene π-spacer, could potentially boost the stability and
robustness and suppress the charge recombination of the
organic PSs in the photocatalytic hydrogen evolution system.
We expect that further structural optimization for the
absorption of longer wavelengths would provide an oppor-
tunity for the organic dyes to exhibit much higher performance
in photocatalysis.


*
ASSOCIATED CONTENT
sı Supporting Information

The Supporting Information is available free of charge at


https://pubs.acs.org/doi/10.1021/acs.chemmater.2c00556.
Synthesis and characterizations of intermediates and
P1−P6, computational studies and all other spectra data
(PDF)

■ AUTHOR INFORMATION
Corresponding Authors
Cheuk-Lam Ho − Department of Applied Biology and
Chemical Technology, The Hong Kong Polytechnic
University, Hong Kong, P. R. China; PolyU Shenzhen
Research Institute, Shenzhen 518057, P. R. China;
orcid.org/0000-0001-8596-0307; Email: cheuk-
lam.ho@polyu.edu.hk
Shuping Huang − College of Chemistry, Fuzhou University,
Fuzhou 350108 Fujian, P. R. China; orcid.org/0000-
0003-4815-1863; Email: huangshp@gmail.com
Authors
Yan-Yi Kwok − Department of Applied Biology and Chemical
Technology, The Hong Kong Polytechnic University, Hong
Kong, P. R. China; PolyU Shenzhen Research Institute,
Shenzhen 518057, P. R. China
Figure 5. Kinetic traces of hydrogen generation with respect to Po-Yu Ho − Department of Applied Biology and Chemical
different PSs (P1−P6 and Pt−TiO2) under the irradiation of (a) blue Technology, The Hong Kong Polytechnic University, Hong
LED (472 nm) and (b) green LED (520 nm) at 50 mW. Each sample Kong, P. R. China; PolyU Shenzhen Research Institute,
flask contained 0.8 M ascorbic acid (5 mL) in water (pH 4.0) and 20 Shenzhen 518057, P. R. China; Present Address: Biological
mg of PS−TiO2−Pt composite sample. Inorganic Chemistry Laboratory, Francis Crick Institute, 1
Midland Road, London NW1 1AT, United Kingdom;
4. CONCLUSIONS Present Address: Department of Chemistry, King’s
A series of novel D−D−π−A metal-free triphenylamine College London, Britannia House, Trinity Street, London
starburst donor-based PSs with different donating moieties SE1 1DB, United Kingdom
on triphenylamine and thiophene/hexylthiophene as the π- Ying Wei − College of Chemistry, Fuzhou University, Fuzhou
linkers were designed and synthesized in this work. This design 350108 Fujian, P. R. China
with different donating substituents and combinations would Zhong Zheng − Department of Applied Biology and Chemical
allow the structural studies on some essential properties like Technology, The Hong Kong Polytechnic University, Hong
the HOMO−LUMO gap and absorbance on the water- Kong, P. R. China; PolyU Shenzhen Research Institute,
splitting performances. Introducing the hexyl chain onto the Shenzhen 518057, P. R. China; Department of Materials
thiophene π-spacer could significantly enhance the absorption Science and Engineering, City University of Hong Kong, Hong
properties of P4 and P6 and double molar extinction Kong, P. R. China
coefficients at around 350 and 470 nm. Sze-Chun Yiu − Department of Applied Biology and Chemical
P1−P6-sensitized Pt−TiO2 photocatalysts showed impres- Technology, The Hong Kong Polytechnic University, Hong
sive hydrogen productivity in aqueous ascorbic acid at pH 4.0 Kong, P. R. China; PolyU Shenzhen Research Institute,
under blue and green lights, compared to the blank TiO2−Pt, Shenzhen 518057, P. R. China; orcid.org/0000-0002-
over the exceptionally long irradiation. Specifically, the most 3001-365X
remarkable one is attributed to P4, yielding 1560 μmol (37.6 Complete contact information is available at:
mL) of hydrogen produced over 252 h under blue light with a https://pubs.acs.org/10.1021/acs.chemmater.2c00556
5531 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Notes (18) Yen, Y.-S.; Chou, H.-H.; Chen, Y.-C.; Hsu, C.-Y.; Lin, J. T.
The authors declare no competing financial interest. Recent developments in molecule-based organic materials for dye-


sensitized solar cells. J. Mater. Chem. 2012, 22, 8734−8747.
(19) Chang, Y. J.; Chou, P.-T.; Lin, Y.-Z.; Watanabe, M.; Yang, C.-J.;
ACKNOWLEDGMENTS Chin, T.-M.; Chow, T. J. Organic dyes containing oligo-phenothiazine
C.-L.H. thanks the Hong Kong Research Grants Council for dye-sensitized solar cells. J. Mater. Chem. 2012, 22, 21704−21712.
(PolyU 123021/17P), the Science, Technology, and Innova- (20) Koumura, N.; Wang, Z.-S.; Mori, S.; Miyashita, M.; Suzuki, E.;
tion Committee of Shenzhen Municipality Hara, K. Alkyl-Functionalized Organic Dyes for Efficient Molecular
(JCYJ20180306173720000), Environment and Conservation Photovoltaics. J. Am. Chem. Soc. 2006, 128, 14256−14257.
(21) Kitamura, T.; Ikeda, M.; Shigaki, K.; Inoue, T.; Anderson, N.
Fund (ECF 79/2020) from the Government of HKSAR, and
A.; Ai, X.; Lian, T.; Yanagida, S. Phenyl-Conjugated Oligoene
the Hong Kong Polytechnic University (ZVVU and ZVXU) Sensitizers for TiO2Solar Cells. Chem. Mater. 2004, 16, 1806−1812.
for their financial support. (22) Hua, Y.; Chang, S.; Huang, D.; Zhou, X.; Zhu, X.; Zhao, J.;

■ REFERENCES
(1) Armaroli, N.; Balzani, V. The Hydrogen Issue. ChemSusChem
Chen, T.; Wong, W.-Y.; Wong, W.-K. Significant Improvement of
Dye-Sensitized Solar Cell Performance Using Simple Phenothiazine-
Based Dyes. Chem. Mater. 2013, 25, 2146−2153.
(23) Lo, C.-Y.; Kumar, D.; Chou, S.-H.; Chen, C.-H.; Tsai, C.-H.;
2011, 4, 21−36.
Liu, S.-H.; Chou, P.-T.; Wong, K.-T. Highly Twisted Dianchoring
(2) Lewis, N. S.; Nocera, D. G. Powering the planet: Chemical
challenges in solar energy utilization. Proc. Natl. Acad. Sci. U.S.A. D−π−A Sensitizers for Efficient Dye-Sensitized Solar Cells. ACS Appl.
2006, 103, 15729−15735. Mater. Interfaces 2016, 8, 27832−27842.
(3) Ahmad, H.; Kamarudin, S. K.; Minggu, L. J.; Kassim, M. (24) Zhang, F.; Fan, J.; Yu, H.; Ke, Z.; Nie, C.; Kuang, D.; Shao, G.;
Hydrogen from photo-catalytic water splitting process: A review. Su, C. Nonplanar Organic Sensitizers Featuring a Tetraphenylethene
Renewable Sustainable Energy Rev. 2015, 43, 599−610. Structure and Double Electron-Withdrawing Anchoring Groups. J.
(4) Balzani, V.; Credi, A.; Venturi, M. Photochemical conversion of Org. Chem. 2015, 80, 9034−9040.
solar energy. ChemSusChem 2008, 1, 26−58. (25) Tiwari, A.; Pal, U. Effect of donor-donor- π -acceptor
(5) Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at architecture of triphenylamine-based organic sensitizers over TiO 2
a Semiconductor Electrode. Nature 1972, 238, 37−38. photocatalysts for visible-light-driven hydrogen production. Int. J.
(6) Lu, J.; Liu, S.; Li, H.; Shen, Y.; Xu, J.; Cheng, Y.; Wang, M. Hydrogen Energy 2015, 40, 9069−9079.
Pyrene-conjugated porphyrins for efficient mesoscopic solar cells: the (26) Ning, Z.; Zhang, Q.; Wu, W.; Pei, H.; Liu, B.; Tian, H.
role of the spacer. J. Mater. Chem. A 2014, 2, 17495−17501. Starburst Triarylamine Based Dyes for Efficient Dye-Sensitized Solar
(7) Ho, P.-Y.; Wang, Y.; Yiu, S.-C.; Yu, W.-H.; Ho, C.-L.; Huang, S. Cells. J. Org. Chem. 2008, 73, 3791−3797.
Starburst Triarylamine Donor-Based Metal-Free Photosensitizers for (27) Zhang, Y.; Cheng, J.; Deng, W.; Sun, B.; Liu, Z.; Yan, L.; Wang,
Photocatalytic Hydrogen Production from Water. Org. Lett. 2017, 19, X.; Xu, B.; Wang, X. Theoretical study of D-A′-π-A/D-π-A′-π-A
1048−1051. triphenylamine and quinoline derivatives as sensitizers for dye-
(8) Zhang, X.; Peng, T.; Song, S. Recent advances in dye-sensitized sensitized solar cells. RSC Adv. 2020, 10, 17255−17265.
semiconductor systems for photocatalytic hydrogen production. J. (28) Abbotto, A.; Manfredi, N.; Marinzi, C.; De Angelis, F.;
Mater. Chem. A 2016, 4, 2365−2402. Mosconi, E.; Yum, J.-H.; Xianxi, Z.; Nazeeruddin, M. K.; Grätzel, M.
(9) Cecconi, B.; Manfredi, N.; Montini, T.; Fornasiero, P.; Abbotto, Di-branched di-anchoring organic dyes for dye-sensitized solar cells.
A. Dye-Sensitized Solar Hydrogen Production: The Emerging Role of Energy Environ. Sci. 2009, 2, 1094−1101.
Metal-Free Organic Sensitizers. Eur. J. Org. Chem. 2016, 2016, 5194− (29) Hagberg, D. P.; Jiang, X.; Gabrielsson, E.; Linder, M.;
5215. Marinado, T.; Brinck, T.; Hagfeldt, A.; Sun, L. Symmetric and
(10) Maeda, T.; Nitta, S.; Sano, Y.; Tanaka, S.; Yagi, S.; Nakazumi, unsymmetric donor functionalization. comparing structural and
H. Near-infrared squaraine sensitizers bearing benzo[c,d]indolenine spectral benefits of chromophores for dye-sensitized solar cells. J.
as an acceptor moiety. Dyes Pigm. 2015, 122, 160−167. Mater. Chem. 2009, 19, 7232−7238.
(11) Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for (30) Ning, Z.; Tian, H. Triarylamine: a promising core unit for
water splitting. Chem. Soc. Rev. 2009, 38, 253−278. efficient photovoltaic materials. Chem. Commun. 2009, 45, 5483−
(12) Watanabe, M. Dye-sensitized photocatalyst for effective water 5495.
splitting catalyst. Sci. Technol. Adv. Mater. 2017, 18, 705−723. (31) Zhang, F.; Luo, Y.; Song, J.; Guo, X.; Liu, W.; Ma, C.; Huang,
(13) Mishra, A.; Fischer, M. K.; Bauerle, P. Metal-free organic dyes Y.; Ge, M.; Bo, Z.; Meng, Q. Triphenylamine-based dyes for dye-
for dye-sensitized solar cells: from structure: property relationships to sensitized solar cells. Dyes Pigm. 2009, 81, 224−230.
design rules. Angew. Chem., Int. Ed. 2009, 48, 2474−2499. (32) Tang, J.; Chen, Y.; Cong, L.; Lin, B.; Sun, Y. Novel tri-carbazole
(14) Shen, Q.; Ogomi, Y.; Park, B.-w.; Inoue, T.; Pandey, S. S.; modified fluorene host material for highly efficient solution-processed
Miyamoto, A.; Fujita, S.; Katayama, K.; Toyoda, T.; Hayase, S. blue and green electrophosphorescent devices. Tetrahedron 2014, 70,
Multiple electron injection dynamics in linearly-linked two dye co- 3847−3853.
sensitized nanocrystalline metal oxide electrodes for dye-sensitized (33) Do, T. T.; Hong, H. S.; Ha, Y. E.; Yoo, S. I.; Won, Y. S.; Moon,
solar cells. Phys. Chem. Chem. Phys 2012, 14, 4605−4613. M.-J.; Kim, J. H. Synthesis and characterization of conjugated
(15) Yiu, S. C.; Ho, P. Y.; Kwok, Y. Y.; He, X.; Wang, Y.; Yu, W. H.; oligoelectrolytes based on fluorene and carbazole derivative and
Ho, C. L.; Huang, S. Development of Strong Visible-Light-Absorbing application of polymer solar cell as a cathode buffer layer. Macromol.
Cyclometalated Iridium(III) Complexes for Robust and Efficient Res. 2015, 23, 367−376.
Light-Driven Hydrogen Production. Chem. − Eur. J. 2022, 28, (34) Huang, J. F.; Lei, Y.; Luo, T.; Liu, J. M. Photocatalytic H 2
No. e202104575. Production from Water by Metal-free Dye-sensitized TiO 2
(16) Ho, P.-Y.; Siu, C.-H.; Yu, W.-H.; Zhou, P.; Chen, T.; Ho, C.-L.; Semiconductors: The Role and Development Process of Organic
Lee, L. T. L.; Feng, Y.-H.; Liu, J.; Han, K.; Lo, Y. H.; Wong, W.-Y. Sensitizers. ChemSusChem 2020, 13, 5863−5895.
Molecular engineering of starburst triarylamine donor with (35) Kim, S. H.; Kim, H. W.; Sakong, C.; Namgoong, J.; Park, S. W.;
selenophene containing π-linker for dye-sensitized solar cells. J. Ko, M. J.; Lee, C. H.; Lee, W. I.; Kim, J. P. Effect of Five-Membered
Mater. Chem. C 2016, 4, 713−726. Heteroaromatic Linkers to the Performance of Phenothiazine-Based
(17) Hara, K.; Miyamoto, K.; Abe, Y.; Yanagida, M. Electron Dye-Sensitized Solar Cells. Org. Lett. 2011, 13, 5784−5787.
Transport in Coumarin-Dye-Sensitized Nanocrystalline TiO2 Electro- (36) Park, J. H.; Ko, K. C.; Kim, E.; Park, N.; Ko, J. H.; Ryu, D. H.;
des. J. Phys. Chem. B 2005, 109, 23776−23778. Ahn, T. K.; Lee, J. Y.; Son, S. U. Photocatalysis by Phenothiazine

5532 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

Dyes: Visible-Light-Driven Oxidative Coupling of Primary Amines at Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.;
Ambient Temperature. Org. Lett. 2012, 14, 5502−5505. Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J.
(37) Lee, J.; Kwak, J.; Ko, K. C.; Park, J. H.; Ko, J. H.; Park, N.; Kim, Gaussian 16, Rev. C.01, Gaussian Inc.: Wallingford, CT, 2016.
E.; Ryu, D. H.; Ahn, T. K.; Lee, J. Y.; Son, S. U. Phenothiazine-based (52) Barone, V.; Cossi, M.; Tomasi, J. A new definition of cavities
organic dyes with two anchoring groups on TiO2 for highly efficient for the computation of solvation free energies by the polarizable
visible light-induced water splitting. Chem. Commun. 2012, 48, continuum model. J. Chem. Phys. 1997, 107, 3210−3221.
11431−11433. (53) Cossi, M.; Scalmani, G.; Rega, N.; Barone, V. New
(38) Park, J. M.; Jung, C. Y.; Wang, Y.; Choi, H. D.; Park, S. J.; Ou, developments in the polarizable continuum model for quantum
P.; Jang, W.-D.; Jaung, J. Y. Effect of additional phenothiazine donor mechanical and classical calculations on molecules in solution. J.
and thiophene π-bridge on photovoltaic performance of quinoxaline Chem. Phys. 2002, 117, 43−54.
cored photosensitizers. Dyes Pigm. 2019, 170, 107568−107584. (54) Hagberg, D. P.; Yum, J.-H.; Lee, H.; De Angelis, F.; Marinado,
(39) Pan, B.; Zhu, Y.-Z.; Ye, D.; Zheng, J.-Y. Improved conversion T.; Karlsson, K. M.; Humphry-Baker, R.; Sun, L.; Hagfeldt, A.;
efficiency in dye-sensitized solar cells based on porphyrin dyes with Grätzel, M.; Nazeeruddin, M. K. Molecular Engineering of Organic
dithieno[3,2-b:2′,3′- d]pyrrole donor. Dyes Pigm. 2018, 150, 223− Sensitizers for Dye-Sensitized Solar Cell Applications. J. Am. Chem.
230. Soc. 2008, 130, 6259−6266.
(40) Brown, D. G.; Schauer, P. A.; Borau-Garcia, J.; Fancy, B. R.; (55) Hua, Y.; Jin, B.; Wang, H.; Zhu, X.; Wu, W.; Cheung, M.-S.;
Berlinguette, C. P. Stabilization of Ruthenium Sensitizers to TiO 2 Lin, Z.; Wong, W.-Y.; Wong, W.-K. Bulky dendritic triarylamine-based
Surfaces through Cooperative Anchoring Groups. J. Am. Chem. Soc. organic dyes for efficient co-adsorbent-free dye-sensitized solar cells. J.
2013, 135, 1692−1695. Power Sources 2013, 237, 195−203.
(41) Hara, K.; Kurashige, M.; Dan-oh, Y.; Kasada, C.; Shinpo, A.; (56) Chen, C.-Y.; Wang, M.; Li, J.-Y.; Pootrakulchote, N.; Alibabaei,
Suga, S.; Sayama, K.; Arakawa, H. Design of new coumarin dyes L.; Ngoc-Le, C.-H.; Decoppet, J.-D.; Tsai, J.-H.; Grätzel, C.; Wu, C.-
having thiophene moieties for highly efficient organic-dye-sensitized G.; Zakeeruddin, S. M.; Grätzel, M. Highly Efficient Light-Harvesting
solar cells. New J. Chem. 2003, 27, 783−785. Ruthenium Sensitizer for Thin-Film Dye-Sensitized Solar Cells. ACS
(42) Hung, W.-I.; Liao, Y.-Y.; Hsu, C.-Y.; Chou, H.-H.; Lee, T.-H.; Nano 2009, 3, 3103−3109.
Kao, W.-S.; Lin, J. T. High-Performance Dye-Sensitized Solar Cells (57) Wu, Z.; Li, X.; Li, J.; Ågren, H.; Hua, J.; Tian, H. Effect of
Based on Phenothiazine Dyes Containing Double Anchors and bridging group configuration on photophysical and photovoltaic
Thiophene Spacers. Chem. − Asian J. 2014, 9, 357−366. performance in dye-sensitized solar cells. J. Mater. Chem. A 2015, 3,
(43) Arteaga, D.; Cotta, R.; Ortiz, A.; Insuasty, B.; Martin, N.; 14325−14333.
Echegoyen, L. Zn(II)-porphyrin dyes with several electron acceptor (58) Wong, W.-Y.; Wang, X.-Z.; He, Z.; Chan, K.-K.; Djurišić, A. B.;
groups linked by vinyl-fluorene or vinyl-thiophene spacers for dye- Cheung, K.-Y.; Yip, C.-T.; Ng, A. M.-C.; Xi, Y. Y.; Mak, C. S. K.;
sensitized solar cells. Dyes Pigm. 2015, 112, 127−137. Chan, W.-K. Tuning the Absorption, Charge Transport Properties,
(44) Wu, Y.; Zhang, X.; Li, W.; Wang, Z.-S.; Tian, H.; Zhu, W. and Solar Cell Efficiency with the Number of Thienyl Rings in
Hexylthiophene-Featured D-A-π-A Structural Indoline Chromo- Platinum-Containing Poly(aryleneethynylene)s. J. Am. Chem. Soc.
phores for Coadsorbent-Free and Panchromatic Dye-Sensitized 2007, 129, 14372−14380.
Solar Cells. Adv. Energy Mater. 2012, 2, 149−156. (59) Li, X.; Yu, J.; Low, J.; Fang, Y.; Xiao, J.; Chen, X. Engineering
(45) Song, J.; Wu, S.; Xing, P.; Zhao, Y.; Yuan, J. Di-branched heterogeneous semiconductors for solar water splitting. J. Mater.
triphenylamine dye sensitized TiO2 nanocomposites with good Chem. A 2015, 3, 2485−2534.
photo-stability for sensitive photoelectrochemical detection of (60) Haque, S. A.; Palomares, E.; Cho, B. M.; Green, A. N. M.;
organophosphate pesticides. Anal. Chim. Acta 2018, 1001, 24−31. Hirata, N.; Klug, D. R.; Durrant, J. R. Charge Separation versus
(46) Kroeze, J. E.; Hirata, N.; Koops, S.; Nazeeruddin, M. K.; Recombination in Dye-Sensitized Nanocrystalline Solar Cells: the
Schmidt-Mende, L.; Grätzel, M.; Durrant, J. R. Alkyl Chain Barriers Minimization of Kinetic Redundancy. J. Am. Chem. Soc. 2005, 127,
for Kinetic Optimization in Dye-Sensitized Solar Cells. J. Am. Chem. 3456−3462.
Soc. 2006, 128, 16376−16383. (61) Bessho, T.; Yoneda, E.; Yum, J.-H.; Guglielmi, M.; Tavernelli,
(47) Tiwari, A.; Mondal, I.; Pal, U. Visible light induced hydrogen I.; Imai, H.; Rothlisberger, U.; Nazeeruddin, M. K.; GräTzel, M. New
production over thiophenothiazine-based dye sensitized TiO2 Paradigm in Molecular Engineering of Sensitizers for Solar Cell
photocatalyst in neutral water. RSC Adv. 2015, 5, 31415−31421. Applications. J. Am. Chem. Soc. 2009, 131, 5930−5934.
(48) Huang, J. F.; Lei, Y.; Xiao, L. M.; Chen, X. L.; Zhong, Y. H.; (62) Bisquert, J. Theory of the Impedance of Electron Diffusion and
Qin, S.; Liu, J. M. Photocatalysts for H 2 Generation from Starburst Recombination in a Thin Layer. J. Phys. Chem. B 2002, 106, 325−333.
Triphenylamine/Carbazole Donor-Based Metal-Free Dyes and (63) Yu, J.; Jin, J.; Cheng, B.; Jaroniec, M. A noble metal-free
Porous Anatase TiO2 Cube. ChemSusChem 2020, 13, 1037−1043. reduced graphene oxide−CdS nanorod composite for the enhanced
(49) Zani, L.; Melchionna, M.; Montini, T.; Fornasiero, P. Design of visible-light photocatalytic reduction of CO2 to solar fuel. J. Mater.
dye-sensitized TiO2 materials for photocatalytic hydrogen produc- Chem. A 2014, 2, 3407−3416.
tion: light and shadow. J. Phys.: Energy 2021, 3, No. 031001. (64) Xiao, X.; Wei, J.; Yang, Y.; Xiong, R.; Pan, C.; Shi, J.
(50) Natali, M. Elucidating the Key Role of pH on Light-Driven Photoreactivity and Mechanism of g-C3N4 and Ag Co-Modified
Hydrogen Evolution by a Molecular Cobalt Catalyst. ACS Catal. Bi2WO6 Microsphere under Visible Light Irradiation. ACS
2017, 7, 1330−1339. Sustainable Chem. Eng. 2016, 4, 3017−3023.
(51) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; (65) Zhang, J.; Yu, J.; Jaroniec, M.; Gong, J. R. Noble Metal-Free
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, Reduced Graphene Oxide-ZnxCd1−xS Nanocomposite with En-
G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; hanced Solar Photocatalytic H2-Production Performance. Nano Lett.
Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. 2012, 12, 4584−4589.
V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams; Ding, F.; Lipparini, (66) Fan, J.; Liu, S.; Yu, J. Enhanced photovoltaic performance of
F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; dye-sensitized solar cells based on TiO2 nanosheets/graphene
Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; composite films. J. Mater. Chem. 2012, 22, 17027−17036.
Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; (67) Zhang, J.; Kan, Y.-H.; Li, H.-B.; Geng, Y.; Wu, Y.; Su, Z.-M.
Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; How to design proper π-spacer order of the D-π-A dyes for DSSCs? A
Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; density functional response. Dyes Pigm. 2012, 95, 313−321.
Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, (68) Zheng, B.; Sabatini, R. P.; Fu, W.-F.; Eum, M.-S.; Brennessel,
V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; W. W.; Wang, L.; McCamant, D. W.; Eisenberg, R. Light-driven
Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; generation of hydrogen: New chromophore dyads for increased

5533 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534
Chemistry of Materials pubs.acs.org/cm Article

activity based on Bodipy dye and Pt(diimine)(dithiolate) complexes. (86) Liu, J.; Zhang, J.; Xu, M.; Zhou, D.; Jing, X.; Wang, P.
Proc. Natl. Acad. Sci. U.S.A. 2015, 112, E3987−E3996. Mesoscopic titania solar cells with the tris(1,10-phenanthroline)cobalt
(69) Zhang, M.; Li, Y.; Bai, C.; Guo, X.; Han, J.; Hu, S.; Jiang, H.; redox shuttle: uniped versus biped organic dyes. Energy Environ. Sci.
Tan, W.; Li, S.; Ma, L. Synthesis of Microporous Covalent 2011, 4, 3021−3029.
Phosphazene-Based Frameworks for Selective Separation of Uranium (87) Zeng, W.; Cao, Y.; Bai, Y.; Wang, Y.; Shi, Y.; Zhang, M.; Wang,
in Highly Acidic Media Based on Size-Matching Effect. ACS Appl. F.; Pan, C.; Wang, P. Efficient Dye-Sensitized Solar Cells with an
Mater. Interfaces 2018, 10, 28936−28947. Organic Photosensitizer Featuring Orderly Conjugated Ethylenediox-
(70) Wang, T.; Wang, Z.; Wang, C.; Xiao, Y.-J. The size-matching ythiophene and Dithienosilole Blocks. Chem. Mater. 2010, 22, 1915−
effect in 0.1Na1/3Ca1/3Bi1/3Cu3Ti4O12-xBa(Fe0.5Nb0.5)O3- 1925.
(0.9-x)PVDF composites. Ceram. Int. 2017, 43, S239−S243. (88) Dessì, A.; Monai, M.; Bessi, M.; Montini, T.; Calamante, M.;
(71) Lazarides, T.; McCormick, T.; Du, P.; Luo, G.; Lindley, B.; Mordini, A.; Reginato, G.; Trono, C.; Fornasiero, P.; Zani, L.
Eisenberg, R. Making Hydrogen from Water Using a Homogeneous Towards Sustainable H2 Production: Rational Design of Hydro-
System Without Noble Metals. J. Am. Chem. Soc. 2009, 131, 9192− phobic Triphenylamine-based Dyes for Sensitized Ethanol Photo-
9194. reforming. ChemSusChem 2018, 11, 793−805.
(72) Le, T. T.; Akhtar, M. S.; Park, D. M.; Lee, J. C.; Yang, O. B.
Water splitting on Rhodamine-B dye sensitized Co-doped TiO2
catalyst under visible light. Appl. Catal., B 2012, 111−112, 397−401.
(73) Watanabe, M.; Hagiwara, H.; Iribe, A.; Ogata, Y.; Shiomi, K.;
Staykov, A.; Ida, S.; Tanaka, K.; Ishihara, T. Spacer effects in metal-
free organic dyes for visible-light-driven dye-sensitized photocatalytic
hydrogen production. J. Mater. Chem. A 2014, 2, 12952−12961.
(74) Kaur, S.; Singh, V. Visible light induced sonophotocatalytic
degradation of Reactive Red dye 198 using dye sensitized TiO2.
Ultrason. Sonochem. 2007, 14, 531−537.
(75) Eckenhoff, W. T.; Eisenberg, R. Molecular systems for light
driven hydrogen production. Dalton Trans. 2012, 41, 13004−13021.
(76) Tinker, L. L.; McDaniel, N. D.; Curtin, P. N.; Smith, C. K.;
Ireland, M. J.; Bernhard, S. Visible Light Induced Catalytic Water
Reduction without an Electron Relay. Chem. − Eur. J. 2007, 13,
8726−8732.
(77) Zhang, P.; Wang, M.; Dong, J.; Li, X.; Wang, F.; Wu, L.; Sun, L.
Photocatalytic Hydrogen Production from Water by Noble-Metal-
Free Molecular Catalyst Systems Containing Rose Bengal and the
Cobaloximes of BFx-Bridged Oxime Ligands. J. Phys. Chem. C 2010,
114, 15868−15874.
(78) Li, F.; Jennings, J. R.; Wang, Q. Determination of Sensitizer
Regeneration Efficiency in Dye-Sensitized Solar Cells. ACS Nano Recommended by ACS
2013, 7, 8233−8242.
(79) Daeneke, T.; Mozer, A. J.; Uemura, Y.; Makuta, S.; Fekete, M.; Subphthalocyanines for Visible-Light-Driven Hydrogen
Tachibana, Y.; Koumura, N.; Bach, U.; Spiccia, L. Dye Regeneration Evolution: Tuning Photocatalytic Performance with
Kinetics in Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2012, 134, Molecular Design
16925−16928.
(80) Zhao, L.; Wagner, P.; Barnsley, J. E.; Clarke, T. M.; Gordon, K. Buket Güntay, Mine Ince, et al.
C.; Mori, S.; Mozer, A. J. Enhancement of dye regeneration kinetics in JUNE 09, 2023
ACS APPLIED ENERGY MATERIALS READ
dichromophoric porphyrin−carbazole triphenylamine dyes influenced
by more exposed radical cation orbitals. Chem. Sci. 2016, 7, 3506−
3516. Influence of Redox Couple on the Performance of ZnO Dye
(81) Lee, J.; Kwak, J.; Ko, K. C.; Park, J. H.; Ko, J. H.; Park, N.; Kim, Solar Cells and Minimodules with Benzothiadiazole-Based
E.; Ryu, D. H.; Ahn, T. K.; Lee, J. Y.; Son, S. U. Phenothiazine-based Photosensitizers
organic dyes with two anchoring groups on TiO2 for highly efficient Carlos A. Gonzalez-Flores, Gerko Oskam, et al.
visible light-induced water splitting. Chem. Commun. 2012, 48, NOVEMBER 08, 2022
11431−11433. ACS APPLIED ENERGY MATERIALS READ
(82) Cheng, M.; Yang, X.; Zhao, J.; Chen, C.; Tan, Q.; Zhang, F.;
Sun, L. Efficient Organic Dye-Sensitized Solar Cells: Molecular Sensitizers of Metal Complexes with Sulfur Coordination
Engineering of Donor-Acceptor-Acceptor cationic dyes. ChemSu- Achieving a Power Conversion Efficiency of 12.89%
sChem 2013, 6, 2322−2329.
(83) Qu, S.; Qin, C.; Islam, A.; Wu, Y.; Zhu, W.; Hua, J.; Tian, H.; Yinfeng Ma, Chaofan Zhong, et al.
Han, L. A novel D−A-π-A organic sensitizer containing a JULY 11, 2023
ACS APPLIED MATERIALS & INTERFACES READ
diketopyrrolopyrrole unit with a branched alkyl chain for highly
efficient and stable dye-sensitized solar cells. Chem. Commun. 2012,
48, 6972−6974. Panchromatic NIR-Absorbing Sensitizers with a
(84) Nguyen, W. H.; Bailie, C. D.; Burschka, J.; Moehl, T.; Grätzel, Thienopyrazine Auxiliary Acceptor for Dye-Sensitized Solar
M.; McGehee, M. D.; Sellinger, A. Molecular Engineering of Organic Cells
Dyes for Improved Recombination Lifetime in Solid-State Dye- William E. Meador, Jared H. Delcamp, et al.
Sensitized Solar Cells. Chem. Mater. 2013, 25, 1519−1525. MAY 09, 2023
(85) Ganesan, P.; Chandiran, A.; Gao, P.; Rajalingam, R.; Grätzel, ACS APPLIED ENERGY MATERIALS READ
M.; Nazeeruddin, M. K. Molecular Engineering of 2-Quinolinone
Based Anchoring Groups for Dye-Sensitized Solar Cells. J. Phys. Get More Suggestions >
Chem. C 2014, 118, 16896−16903.

5534 https://doi.org/10.1021/acs.chemmater.2c00556
Chem. Mater. 2022, 34, 5522−5534

You might also like