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

Research Article

Chemistry—A European Journal doi.org/10.1002/chem.202104575

www.chemeurj.org

Development of Strong Visible-Light-Absorbing


Cyclometalated Iridium(III) Complexes for Robust and
Efficient Light-Driven Hydrogen Production
Sze-Chun Yiu+,[a, b] Po-Yu Ho+,[a, b, d] Yan-Yi Kwok+,[a] Xiaojie He,[c] Yi Wang,[b] Wai-Hong Yu,[b]
Cheuk-Lam Ho,*[a, b] and Shuping Huang*[c]

Abstract: Weak light absorption of common Ir(III) complexes evolution performance of these cationic Ir(III) complexes.
(e. g., using phenylpyridine as the ligand) has hindered their Remarkably, Ir4 with 4,4’-dimethoxytriphenylamine achieved
applications in photocatalytic hydrogen generation from the highest turn-over number (TON) of 12 300 and initial
water as an efficient photosensitizer. To address this issue, a turnover frequency (TOFi) of 394 h 1, with initial activity
series of cyclometalated Ir(III) complexes (Ir1–Ir5), featuring (activityi) of 547 000 μmol g 1 h 1 and initial apparent quan-
different electron-donating substituents to enhance the tum yield (AQYi) of 9.59 %, under the illumination of blue
absorptivity, have been synthesized and studied as photo- light-emitting diodes (LEDs) for 105 hours, which demon-
sensitizers (PSs) for light-driven hydrogen production from strated a stable three-component photocatalytic system with
water. Ir6–Ir7 were prepared as fundamental systems for high efficiency. The TON (based on n(H2)/n(PSr)) in this study
comparisons. Electron donors, including 9-phenylcarbazole, is the highest value reported to date among the similar
triphenylamine, 4,4’-dimethoxytriphenylamine, 4,4’-di(N- photocatalytic systems using Ir(III) complexes with Pt nano-
hexylcarbazole)triphenylamine moieties were introduced on particles as catalyst. The great potential of using triphenyl-
6-(thiophen-2-yl)phenanthridine-based cyclometalating (C^N) amine-based Ir(III) PSs in boosting photocatalytic performance
ligands to explore the donor effect on the hydrogen has also been shown.

Introduction

Harnessing solar energy via photovoltaic devices has become


the main trend in developing sustainable energy attributed to
the surging demand of the electricity-based economy.[1] Hydro-
[a] Dr. S.-C. Yiu,+ Dr. P.-Y. Ho,+ Y.-Y. Kwok,+ Dr. C.-L. Ho
Department of Applied Biology and Chemical Technology gen is regarded as a suitable energy carrier because of its high
The Hong Kong Polytechnic University energy density and free of carbon footprint. Photocatalytic
Hung Hom, Hong Kong (P. R. China) hydrogen evolution (PHE) can be achieved by water-splitting,
and which is composed of two half-reactions, namely hydrogen
PolyU Shenzhen Research Institute
Shenzhen (P. R. China) evolution reaction (HER) and oxygen evolution reaction (OER).[2]
E-mail: cheuk-lam.ho@polyu.edu.hk Herein, attention will only focus on HER with three-component
[b] Dr. S.-C. Yiu,+ Dr. P.-Y. Ho,+ Y. Wang, W.-H. Yu, Dr. C.-L. Ho photocatalytic systems (sacrificial electron donor-photosensi-
Department of Chemistry tizer-catalyst (SED-PS-catalyst)).
Hong Kong Baptist University In the solar-driven HER, PS plays a crucial role in photon
Waterloo Road, Kowloon Tong, Hong Kong (P. R. China)
energy absorption for creating electron-hole pairs transferring
[c] X. He, Dr. S. Huang
College of Chemistry to photocatalyst. Several common types of PSs include metal-
Fuzhou University free organic dyes, metal complexes, quantum dots, and nano-
Fuzhou, Fujian 350108 (P. R. China) particles. Although metal-free organic dyes have the advan-
E-mail: huangshp@gmail.com
tages of being cost-effective and highly light-absorbing, their
[d] Dr. P.-Y. Ho+
long-term stabilities have been a concern.[3] Alternatively,
(1) Biological Inorganic Chemistry Laboratory
The Francis Crick Institute metal-based counterparts such as cyclometalated Ru(II) or Ir(III)
1 Midland Road, London NW1 1AT (United Kingdom) complexes, Zn(II) porphyrins, and Pt(II) diimine dithiolates, have
and shown promising results in light-harvesting applications.[4] In
(2) Department of Chemistry
particular, Ru(II) complexes are the most studied metal-based
King’s College London
Britannia House, 7 Trinity Street, London SE1 1DB (United Kingdom) PSs due to intense absorptions along with low energy and
[+] These authors contributed equally to this work. long-lived metal-to-ligand charge-transfer (MLCT) transitions.
Supporting information for this article is available on the WWW under However, using labile NCS ligands, narrow colour tunability,
https://doi.org/10.1002/chem.202104575 and poor thermal/photostability featuring from their low lying

Chem. Eur. J. 2022, e202104575 (1 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

triplet metal-centred (3MC) states have limited the applications results also provide insight into the importance of choosing
of the Ir(III) complexes.[5] On the other hand, with the appropriate electron-donating substituents on the C^N ligand
remarkable robustness and attractive tuneable photophysical of cationic iridium complexes, in which great potential of using
characteristics, cationic Ir(III) complexes with the formula [Ir- triphenylamine-based Ir(III) PSs in boosting the photocatalytic
(C^N)2(N^N)] +, have been reported as PSs in PHE for the first performance has been realized.
time by Bernhard et al.[6] However, most of the reported cationic
Ir(III) complexes exhibited relatively weak molar extinction
coefficient (ɛ < 10,000 M 1 cm 1) in the visible (MLCT/LLCT Results and Discussion
(ligand-to-ligand charge transfer)) region, hence poor photo-
catalytic efficiency.[4b] Several criteria for PSs’ modifications to Synthesis and structural characterization
improve PHE performance effectively have been suggested,
including (1) intense visible light absorption profile; (2) good The synthetic schemes of ligand precursors, cyclometalating
photostability; (3) long excited-state lifetime, and (4) optimal ligands L1–L5 and the corresponding Ir(III) complexes Ir1–Ir7
redox potential to catalyst/SED.[4d,7] were shown in Schemes 1 and 2. The intermediates 1–11 were
However, rare examples have been reported to fulfil these prepared by a combination of Ullmann condensation, bromina-
criteria to fully utilize Ir(III) complexes as efficient PS in the SED- tion by N-bromosuccinimide (NBS), Suzuki coupling reaction
PS-catalyst system.[8] The rapid decrement of catalytic activity and Buchwald-Hartwig amination. The four 2-aryl-thiophene-
within a few hours in some common solely Ir(III)-based photo- based derivatives were then treated with n-butyllithium
catalytic systems is attributed to their high dependence on solution at 78 °C, followed by trimethyl borate and subse-
diffusional controlled electron-transfer from unstable PS reduc- quent acid hydrolysis to generate corresponding boronic acids.
tive state (PS ) towards catalytic deactivation and the lifetime L2–L5 were achieved by reacting the respective boronic acids
of PS , thus hampering their photostability.[9] This could be with 6-chlorophenanthridine via Suzuki coupling reaction, while
eased by shortening the singly reduced state of PS with L1 was accomplished by coupling the commercially available 2-
increasing the electron transferring rate, for example, adapting thiopheneboronic acid with 6-chlorophenanthridine. Ir1–Ir7
a molecular design of adsorbing organic-inorganic hybrid were then synthesized in two-step synthesis protocols.[14] The μ-
materials on semiconductors to facilitate the directional elec- chloride-bridged dimers were firstly synthesized by reacting the
tron transport.[10] However, these systems are rarely explored in iridium(III) chloride hydrate with corresponding L1–L5. Subse-
the family of Ir(III)-based PS with limited success compared to quent silver cleavage of the dimers and co-ligation with the
semiconductor-free systems (Tables S6–S7).[4b] In terms of ɛ, 2,2’-bipyridyl-4,4’-dicarboxylic acid (H2dcbpy) afforded Ir1–Ir7
organic dyes exhibit relatively intense ɛ with broad absorption (Figures 1 and S1), in the presence of triethylamine. To ensure
bands arising from intramolecular charge transfer (ICT) in their the cationic species as hexafluorophosphate derivatives, the
donor-acceptor configurations.[3b,11] By judiciously selecting purified Ir(III) complexes were treated with HPF6 under room
appropriate functional “donor-acceptor” cyclometalating li- temperature. All complexes were found to be stable without
gands for Ir(III) complexes, their broadband visible light any measurable degradation under ambient laboratory con-
absorption is significantly enhanced, attributed to the excited ditions.
state energies with the combined effect of ICT in C^N ligand All the intermediates were characterized by 1H NMR
and MLCT/LLCT related to N^N ligand.[8,12] spectroscopy (Figures S2–S13). L1–L5 were fully characterized
In this study, a series of Ir(III) complexes (Ir1–Ir7) have been by a combination of techniques, such as 1H and 13C NMR
developed as light-harvesting materials for PHE (Figure 1). 6- spectroscopy (Figures S14–S23) and MALDI-TOF mass spectrom-
(Thiophen-2-yl)phenanthridine-based C^N ligands were em- etry (Figures S38–S42). The structural identities of Ir1–Ir7 were
ployed in these Ir(III) complexes to enhance the absorption confirmed through 1H and 19F NMR spectroscopy (Figures S24–
coverage via extended π-electron delocalization.[13] 9-Phenyl- S37) and ESI mass spectrometry (Figures S43–S49). However,
carbazole, triphenylamine, 4,4’-dimethoxytriphenylamine, 4,4’- acquisitions of good quality of the 13C NMR spectra were
di(N-hexylcarbazole)triphenylamine moieties were introduced difficult due to their partially solubilities in common organic
as electron-donating substituents on the C^N ligands to explore solvents.
their donor effect on PHE performance. TiO2 was employed as
an electron reservoir and a transporter to achieve a long-term
catalytic cycle. Combined with the intense UV-Vis absorption in Photophysical study
the Ir(III) complexes, highly stable water reduction systems have
been demonstrated (photocatalytic activities more than To investigate the effect of electron-donating properties of
105 hours). Remarkably, Ir4 bearing 4,4’-dimeth- different C^N ligands, UV-vis absorption spectra of Ir1–Ir7 were
oxytriphenylamine achieved the highest turnover number measured in the solution state (CH2Cl2 CH3OH solvent mixture)
(TON) of 12,300 and initial turnover frequency (TOFi) of 394 h 1, at room temperature. The resultant absorption parameters and
with initial activity (activityi) of 547,000 μmol g 1 h 1 and initial photophysical profile are summarized in Table 1. Figure 2
apparent quantum yield (AQYi) of 9.59 %, in which the TON reveals the overall features of the absorption spectra of these
(based on n(H2)/n(PS)) is the highest value reported to date in complexes. Some characteristic absorption bands of cationic
the systems of Ir(III) complexes with Pt NPs as the catalyst. The [Ir(C^N)2(N^N)] + were observed.[15] The intense and high energy

Chem. Eur. J. 2022, e202104575 (2 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Figure 1. Chemical structures of Ir(III) complexes Ir1–Ir7.

Table 1. UV/Vis absorption, photoluminescence and fluorescence lifetime


data for Ir1–Ir7 in CH2Cl2 CH3OH (1 : 1, v/v) at ambient temperature.
PS λabs (ɛ/M 1 cm 1)/nm λem/nm[a] τ/ns (χ2)

Ir1 324 (43033), 385 (21896), 464 (11548), 643 4370 (1.03)
495 (10078)
Ir2 321 (46083), 341 (42656), 430 (35692), 714 2021 (1.04)
516 (23580)
Ir3 309 (52789), 472 (44150), 529 (41367) 735 2893 (1.12)
Ir4 326 (53546), 538 (57615) 742 4489 (1.23)
Ir5 345 (106485), 535 (44576) 740 3984 (1.23)
Ir6 253 (42545), 377 (8780) 648 261 (0.93)
Ir7 282 (41708), 384 (10030) 657 258 (1.05)

[a] Excited at the corresponding lowest energy absorption maxima.

Figure 2. UV/Vis absorption spectra of Ir1–Ir7 in CH2Cl2 CH3OH (1 : 1, v/v) at


1 1 ambient temperature.
bands in 300–360 nm with ɛ ~ 40,000–100,000 M cm are
originated from spin-allowed π!π* ligand centred (LC) elec-
tronic transitions on both the C^N and N^N ligands. On the
other hand, the low energy bands above 400 nm (ɛ ~ 10,000– 60,000 M 1 cm 1) are ascribed to the mixture of MLCT and LLCT
transitions.

Chem. Eur. J. 2022, e202104575 (3 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Scheme 1. Synthetic schemes for intermediates 1–11.

Ir6 was prepared and known as one of the common iridium ligand-centred π!π* transition as compared to MLCT and LLCT
complex systems for photocatalytic studies. With the substitu- transitions in its absorption.[16] Besides, Ir4 exhibits broadband
tion of phenyl group by thiophene, there was a slight bath- visible light absorption with the highest ɛ of 5.8 × 104 M 1 cm 1
ochromic shift observed from Ir7. The increased π-conjugation at 538 nm due to the introduction of dimethoxytriphenylamine
of pyridine of the cyclometalating ligand in Ir1 demonstrated unit, which is a favourable characteristic for photocatalytic
red shifted absorption band with a slightly higher molar hydrogen generation from water. Notably, varying the electron-
absorptivity. To further enhance the absorptivity coefficient and donating strength of the electron donors in Ir2–Ir5 resulted in
broadening the light-harvesting spectra, the thiophenyl elec- red-shifted absorption peak maxima with enhanced absorbance
tron-donating moiety was coupled with bulky π-conjugates. as compared to Ir1. The lowest energy absorption bands,
Among all the complexes, Ir5 exhibited much higher ɛ in the presumed to be the mixing between intra-ligand transition on
high energy region as compared to visible region, which can be the C^N ligand and LLCT transition,[12a,c,17] shifted in the order of
attributed to the additional four 9-hexyl-9H-carbazole units in Ir6 < Ir7 < Ir1 < Ir2 < Ir3 < Ir4 ~ Ir5. This implies that the nature of
the molecular structures, results in a more contribution of electron-donating chromophores affects the electronic transi-

Chem. Eur. J. 2022, e202104575 (4 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Scheme 2. Synthetic schemes for ligands L1–L7 and Ir(III) complexes Ir1–Ir7.

tions of these [Ir(C^N)2(N^N)] + complexes. With respect to the


typical absorption profiles of Ir2–Ir5, they provide the possibility
of receiving light across a wider spectral range, which is a
favourable characteristic for their potential application in solar-
to-energy conversion.
Photoluminescence spectra of Ir1–Ir7 in CH2Cl2 CH3OH
mixture at room temperature were also recorded as shown in
Figure 3. Under photoexcitation, all the complexes exhibited
structured emission spectral profiles with emission band
maxima ranging from 643 to 742 nm with a shoulder peak at
their longer wavelengths, except that of Ir6 was insignificant,
resulting from the mixture of 3MLCT and 3LC characters.[11,13a,b]
These emissive complexes exhibit Stokes shifts that are
sufficiently large to minimize the spectral emission overlap with
Figure 3. Normalized photoluminescence spectra of Ir1–Ir7 in CH2Cl2 CH3OH
the lowest-energy MLCT absorption bands because of the weak
(1 : 1, v/v) at ambient temperature.
oscillator strength of the 3MLCT, which is indicative of
phosphorescence.[18] A dramatically red-shifted emissions in
these systems are achieved in complexes Ir2–Ir5 as compared
to Ir1, Ir6–Ir7, in which the electron-donating effect in Ir2–Ir5 character by the extended π-conjugated ligands in the excited
destabilizes their HOMOs. The photoluminescence lifetimes of states of Ir1–Ir5, thus slower triplet radiative decay[19]. All the
Ir1–Ir5 ranged from 2.02 to 4.49 μs while that of Ir6 and Ir7 results demonstrate that the photophysical profiles of PSs have
were around 260 ns (Figure S50). Comparing the modified Ir1– been effectively adjusted by the variation of C^N ligands.
Ir5 to the simple structured Ir6–Ir7, the Ir(III) complexes with
more bulky ligands showed a significantly longer emission
lifetime. This might be attributed to the weakened metal

Chem. Eur. J. 2022, e202104575 (5 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Electrochemical and computational studies calculated HOMO energy levels, the orbital energies vary from
5.96 to 5.13 eV, showing that different donor groups
In order to assess the thermodynamic feasibility of electron attached to thiophene moiety affect the composition of the
transfer, cyclic voltammetry (CV) focusing on oxidation poten- HOMO remarkably. Analysis of the orbital diagrams provided in
tials of Ir1–Ir5 were undertaken in deaerated acetonitrile Figures S53–S57 indicates that the HOMO is largely composed
solution. Despite the poor solubility hampering the analysis, the of π orbitals of C^N ligands, as well as the next highest orbitals
first oxidation wave were summarized in Table 2 and the cyclic (HOMO-1 and HOMO-2). The electron-donating ability of the
voltammogram is shown as Figure S51. All compounds undergo donor group on C^N ligand in Ir2–Ir4 definitely extends the
a single one-electron quasi-reversible oxidation processes. The conjugation range, which elevates the HOMO energy and
oxidation potentials (Eox) of Ir1–Ir5 appear in between 0.83 and lessens the HOMO-LUMO gap from 3.12 eV to 2.31 eV.
1.31 V versus Ag/AgCl reference electrode. In accordance with The TD-DFT calculations could also throw light on the
previous studies on cationic Ir(III) complexes, these potentials charge-transfer character in the process of electron transitions
can be ascribed to the one electron Ir(III)/Ir(IV) redox process, (Figure 4). In Ir1–Ir5, the electron transition from ground state
which are dependent on the electronic environment around (S0) to the first singlet excited state (S1) is mostly formed of
the iridium centre.[20] Substitution with more electron-donating HOMO!LUMO excitation (98.4 % for Ir1; 73.1 % for Ir2; 87.7 %
functionality on the C^N ligand results in shifting the first Eox to for Ir3; 92.5 % for Ir4; 77.4 % for Ir5). According to the TD-DFT
less positive side, subsequently destabilize the HOMO levels. results in Tables S1–S5 and the orbital diagrams in Figures S53–
This electronic effect consequently makes the metal complex to S57, the short wavelength absorption band (200–400 nm)
take oxidation easier. Based on the first Eox and E0-0, EHOMO and should be attributable to the intra-ligand π!π* transition of
ELUMO were calculated. EHOMO of Ir1–Ir5 were found in between C^N ligand with an admixture of relatively small (MLCT
5.19 and 5.67 eV, which are more negative than the redox involving the Ir and C^N ligand, which results in different
potential level of ascorbic acid (AA) ( 4.65 eV, pH ~ 4), which absorption wavelengths maxima for Ir1–Ir5. The charge-transfer
functions as SED in our water-splitting experiments. This allows character in long wavelength absorption bands can be assigned
effective dye regeneration from SED. On the other hand, the to IL transition of C^N ligand from electron-donating moieties
ELUMO (ranging from 3.16 to 3.36 eV) are less negative than to 2-thienyl-phenanthridine. The introduction of donor groups
the conduction band of TiO2 ( 4.0 eV), serving as a driving lessen the energy difference of related states and the maximum
force to inject electrons effectively for the light-driven H2 absorption wavelengths for Ir2–Ir5 (451.91 nm for Ir2;
generation.[21] The electron-donating effect of different chromo- 508.45 nm for Ir3; 540.16 nm for Ir4; 529.69 nm for Ir5) shifted
phores in Ir2–Ir5 decreases the Eg from 2.31 eV in Ir1 to 2.02 eV toward long wavelengths compared with that in Ir1
in Ir4. The results suggest that the electron donating group (441.70 nm), which is in accordance with experimental results.
imparts a marked effect on the EHOMO and subsequently energy
gaps of this series of Ir(III) complexes.
The time dependent density functional theory (TD-DFT) Thermal stability
calculations were carried out to elucidate the charge-transfer
character and electronic properties of Ir1–Ir5 and the data are To investigate the thermal stability of Ir1–Ir7, the thermogravi-
summarized in Tables S1–S5 and Figure S52. Generally, the metric analyses (TGA) for these Ir(III) complexes were performed
orbital contributions of the HOMOs and LUMOs for these and are reported in Figure S58. The thermolysis of materials
complexes are the same. The LUMOs of these complexes have was begun with an increasing rate of 10 °C min 1 in a flow of
predominantly composed of d orbitals of Ir and the π orbitals nitrogenous air, rising the temperature from 30 to 700 °C. After
of the 2,2’-bipyridine-4,4’-dicarboxylic acid character. As for the the initial slight loss of mass at 100–130 °C, the pyrolysis
occurred at around 300–350 °C by the major decomposition,
while most of them were thermally stable to weight loss below
300 °C. According to the temperature for reaching their 98 %
Table 2. Electrochemical data and energy levels for Ir1–Ir5.
weight as illustrated in Figure S58, the thermal stability of the
PS Eox/ EHOMO/ E0-0/ Eox*/ ELUMO/ Calculated Calculated Ir(III) complexes is in tandem with the bulkiness of their ligands
V[a] eV[b] eV[c] V[d] eV[e] EHOMO/eV ELUMO/eV
(i. e., Ir5 with the most bulky ligand showed the highest
Ir1 1.31 5.67 2.31 3.62 3.36 5.96 2.84 temperature to achieve the 98 % mass among all). Little charred
Ir2 1.15 5.51 2.22 3.37 3.29 5.71 2.85
black powders were remained in the crucible after the analysis.
Ir3 1.03 5.39 2.11 3.14 3.28 5.34 2.83
Ir4 0.83 5.19 2.02 2.85 3.17 5.13 2.82
Ir5 0.85 5.21 2.05 2.89 3.16 5.18 2.81

[a] Onset of first oxidation potentials were measured by cyclic voltamme- Photocatalytic hydrogen generation study
try in acetonitrile solution containing 0.1 M of NBu4PF6 against Ag/Ag +
reference electrode, with glassy carbon working electrode and platinum Different approaches have been proposed to improve the
wire auxiliary electrodes. [b] Calculated from (Eox + 4.36), as reversible
photostability of PS in the catalytic reaction, including the use
oxidation of ferrocene was E1/2 = 0.44 V and the EHOMO of ferrocene is equal
to 4.80 eV vs. to vacuum level. [c] E0-0 was determined from the onset of of sterically bulky substituents on N^N ligands,[9a] the use of
absorption spectrum at the corresponding lowest energy absorption orchestrated functional groups on PS to colloidal platinum
maxima. [d] EOX* = EOX + E0-0. [e] ELUMO = EHOMO + E0-0. nanoparticles,[9b] and replacement of labile N^N ligands by C^N

Chem. Eur. J. 2022, e202104575 (6 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Figure 4. Spatial plot of HOMO and LUMO distributions for complexes Ir1–Ir5.

moieties in neutral Ir(III) tricyclometalated complexes.[22] These adsorption on semiconductors was performed by sonicating PS
methods are particularly useful to stabilize the unstable and (50 μM) and platinized TiO2 in CH2Cl2 CH3OH mixture. The dye
reduced species of Ir(III) PS , which are generated in the loading of each PS was calculated from the percentage change
reductive quenching pathway. Recently, efficient electron of the lowest-energy peak absorbance in solution before and
injection from PSs to semiconductors with enhanced photo- after sonication. The dye adsorptions of all PSs were above 96 %
stability of PSs in light-driven H2 generation have been for Ir1–Ir5 and 93 % for Ir6–Ir7 (Figures S60–S66). The photo-
demonstrated by anchoring the PSs on the catalytic experiments were performed in acidic aqueous
semiconductors.[10a,23] In HERs, the semiconductors (TiO2 nano- solution (5 mL, pH 4.0), containing ascorbic acid (AA, 0.5 M) as
particles in this study) have dual functions: (i) immobilizing/ SED. Platinum nanoparticles (Pt NPs) and ascorbic acid (AA)
connecting PS and catalysts, and (ii) facilitating the fast and were used as hydrogen generation catalyst and SED to provide
directional electron transfer from PS to catalyst. Therefore, the reductive equivalents in water, respectively. The reaction flask
presence of two carboxylate groups on the bipyridine ligands in was then filled with nitrogen/methane (80 : 20 mol %) mixture
Ir1–Ir7 are expected to facilitate the direct linkage to TiO2, to gas and irradiated under either blue (~ 472 nm; 80 mW) or
increase their tolerance of water, and to ensure directional and green (~ 520 nm; 50 mW) LEDs. Gas chromatography (GC)
rapid electron transport between the excited PS and the Pt analysis was performed after certain irradiation period by
catalytic centre. In light of their broad and intense absorptions injecting 0.1 mL gas sample from headspace. The amount of H2
from 400 to 600 nm, we utilized Ir1–Ir7 as PSs for the light- produced was determined by the signal ratio of H2 to CH4
driven H2 generation using similar photocatalytic system (internal standard), with the aid of calibration curve (Fig-
published by Eisenberg and our groups previously.[24] Dye ure S59). The H2 generation data (including amount of H2

Chem. Eur. J. 2022, e202104575 (7 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

evolved, turnover number/frequency (TON/TOF), initial activity


Table 3. Light-driven H2 generation data with and without PSs.
(activityi) and apparent quantum yield (AQYi)) were summarized
Dye DL H2 TON[b] TOF[c] TOFi[d] Activityi[e] AQYi in Table 3 and the H2 evolution curves (total 105 hours) under
%[a] (mL) (h 1) (h 1) (μmol g 1 h 1) %[f]
blue and green LEDs were illustrated in Figures 5a and b,
Blue light (472 nm)
respectively.
Ir1 99.1 4.61 1630 15.5 94 111000 1.95 All PSs potently facilitated the generation of H2 upon light
Ir2 96.8 11.2 3520 33.4 143 189000 3.31
Ir3 97.0 26.8 8470 80.6 325 430000 7.55 irradiation (up to 40.8 mL and 11.1 mL for blue and green lights,
Ir4 98.2 40.8 12300 117 394 547000 9.59 respectively, for Ir4), except Ir6 and Ir7. We also performed the
Ir5 96.8 30.6 10800 103 408 484000 8.49 H2 evolution experiment with bare platinized TiO2 as control
Ir6 93.9 3.43 2270 21.6 88.1 55100 0.943
Ir7 93.2 2.44 1620 15.4 51.5 32200 0.551 experiment, only less than 3.3 and 1 mL of H2 were produced
Bare – 3.26 – – – 72000 1.26 for blue and green lights, respectively. Among all the PSs, Ir1
Pt@TiO2 showed less efficient hydrogen evolution, with a slightly higher
Green light (520 nm) activityi and AQYi than the control, while Ir6 and Ir7 just
Ir1 99.1 1.08 382 3.64 14 16700 0.425 performed similar to the control. This is mainly attributed to its
Ir2 96.8 2.73 861 8.20 15 20000 0.509 relatively low ɛ under the irradiation wavelength. By incorporat-
Ir3 97.0 8.20 2580 24.5 66 87200 2.22 ing the electron donors into C^N ligands, enhancement in the
Ir4 98.2 11.1 3340 31.8 85 118000 3.01
Ir5 96.8 9.33 3280 31.2 87 104000 2.64 H2 generation were observed in Ir2–Ir5. From Ir2, Ir3 to Ir4, the
Bare – 0.590 – – 5250 0.134 gradual increase in H2 generation is primarily attributed to the
Pt@TiO2 increasing ɛ with extended absorption wavelengths. In addition,
[a] Dye-loading percentage. [b] Turnover number of H2 is calculated as triphenylamine is a more steric hindrance donor than 9-phenyl-
number of mole of H2 produced divided by the number of mole of PS carbazole, which may result in more effective blocking of
attached to platinized TiO2. [c] Turnover frequency is calculated per hour.
charge recombination of PS.[25] As a result, Ir4 achieved the
[d] Initial turnover frequency in the first 2 h. [e] Initial photocatalytic
activity (in the first 2 h) of the system is defined as number of micromole highest TON of 12,300, with activityi of 547,000 μmol g 1 h 1 and
of H2 evolved per gram of platinum loaded per hour. [f] Initial apparent AQYi of 9.59 %, under the illumination of blue light LEDs after
quantum yield percentage (AQYi %) of the system in the first 2 h. 105 hours. Such TON (based on n(H2)/n(PS)) is the highest value
reported to date in the systems of Ir(III) complexes using Pt NPs
as catalyst (Tables S6–S7). In addition, by comparing the photo-
catalytic reaction mixture of Ir4 before and after the irradiation
(105 h), the TiO2 composite material still remained the original
colour and the AA aqueous solution changed from colourless to
pale yellow due to the presence of dehydroascorbic acid (DHA)
resulting from the dye-regeneration reaction. This indicated
that issues regarding dye desorption and photobleaching of PS
were not observed in our case. In combination with the
increasing trend of H2 generation curve, it is reasonable to
expect the photocatalytic system with Ir4 is still highly active
even under a longer duration of illumination. The high
stabilities of these PSs in photochemical reactions (photo-
catalytic activities > 105 hours) may be attributed to the
introduction of relatively bulky supporting groups on C^N
ligands, which provide steric protection to the metal coordina-
tion centre.[26] Furthermore, the incorporation on TiO2 is
expected to reduce the time spending on the singly reduced
PS with rapid electron injection to TiO2,[27] hence enhancing
the photostability of the Ir(III)-based PS.
Bulky substituent is expected to favour the dye regeneration
due to the large surface area occupied by the donor moiety,
and reduce the formation of unstable Ir(III) PS .[28] With the
encouraging result of our triphenylamine-based PS, we further
exploited Ir5 with more bulky substituents, 3-(9-hexyl-9H-
carbazole), on the triphenylamine. However, under the same
irradiation condition, Ir5 attained a TON of 10,800 with activityi
of 484,000 μmol g 1 h 1 and AQYi of 8.49 %, the values are in
between those of Ir3 and Ir4. Regarding the similar absorption
Figure 5. H2 evolution curves with respect to Ir1–Ir5 and platinized TiO2 profiles between Ir4 and Ir5, this may suggest the electronic
under the irradiation of (a) blue LED (472 nm) (with Ir6 and Ir7) and (b) green effect outweigh the steric effect of the donor substituents in
LED (520 nm). boosting the H2 evolution performance in our octahedral

Chem. Eur. J. 2022, e202104575 (8 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

cationic Ir(III) complexes, unlike the relatively planar organic The results indicate the importance of choosing an appropriate
PSs.[24b] electron donor on the C^N ligand of cationic iridium complexes
Under the irradiation of green light, a similar trend in the and show the great potential of using triarylamine-based Ir(III)
performance of the photocatalytic hydrogen generation is PSs in boosting the photocatalytic performance for efficient
observed. The TONs are generally smaller than those obtained hydrogen generation.
under blue light, this can be partially related to the lower power
of green LED used (50 mW at ~ 520 nm) compared to the blue
one (80 mW at ~ 472 nm). In addition, the reduced photon Experimental Section
absorptions in the green light region for Ir1 and Ir2 may also
Experimental details (i. e., synthetic details, structural and physical
attribute to this phenomenon. However, a more detailed study characterization, photocatalytic hydrogen production details and
is required for further analysis on this question. Due to the computational method) are described in the Supporting Informa-
absence of absorption spectra overlap of Ir6 and Ir7 to the tion.
green LED, therefore, Ir6 and Ir7 were not applicable for green
light irradiation.
General procedures for the synthesis of ligands L1–L5
Generally, it is expected that both processes of oxidative
and reductive quenching of Ir(III) PSs by water reduction Synthesis of thiophene-derived boronic acid was achieved by
dissolving the corresponding thiophene-derivatives (1 molar equiv-
catalyst and SED, respectively, are possible to happen in the
alent) in 50 mL anhydrous THF and cooled down to 78 °C. After
photocatalytic hydrogen evolution system.[29] The nature of 15 minutes, n-BuLi solution (1.5 molar equivalents) was added and
quenching is greatly depending on the energy levels of photo- the temperature was maintained at 78 °C for another 45 minutes.
excited PSs and the types of catalysts and SED applied. As Trimethylborate (2 molar equivalents) was then added. After
mentioned, TiO2 nanoparticles can facilitate the fast electron 16 hours, the reaction mixture was quenched with 2 M HCl and
transfer from excited PSs to catalyst. The ultrafast electron extracted using diethyl ether and water. The organic phase was
collected and the solvent was removed in vacuo. The crude product
injection (ps-scale) from N^N ligand on Ir(III) PSs to the
was purified through a wet column with silica gel eluting with
conduction band of TiO2 has been demonstrated upon hexane and CH2Cl2, followed by diethyl ether and CH2Cl2. The
immobilization.[27] Therefore, the excited Ir(III) PSs in our photo- boronic acid was then directly used in subsequent Suzuki coupling
catalytic system are proposed to be oxidative quenched. Under reaction. Boronic acid (1 molar equivalent) and 6-chlorophenanthri-
such mechanism, the excited Ir(III) PS* is thought to be dine (0.8 molar equivalent) were dissolved in 50 mL THF, Pd(PPh3)4
oxidative quenched by transferring electron into TiO2 to (0.05 molar equivalent) and 2 M aqueous solution of K2CO3 (5 mL)
were then introduced. The mixture was refluxed at 85 °C overnight.
generate Ir(III) PS +. The Ir(III) PS + was then accepted electron
The reaction mixture was then extracted with ethyl acetate and
from SED (AA in our system) to regenerate PS. water. The organic portion was collected and concentrated. The
In short, the above comparison of photocatalytic perform- target ligand was finally purified by a wet column with silica gel
ance from Ir2 to Ir5 clearly indicates the importance in choosing eluting with a mixture of hexane and CH2Cl2.
appropriate electron donors on the C^N ligand of cationic
iridium complexes, in order to enhance the photocatalytic
General procedures for the synthesis of Ir(III) complexes Ir1–
performance by optimizing the absorption profiles of the PSs. Ir7
These complexes are generally synthesized through a two-step
process. The dichloro-bridged Ir(III) dimers were firstly prepared by
Conclusion dissolving the corresponding ligand (2.2 molar equivalents) and
IrCl3 · 3H2O (1 molar equivalent) in 10 mL 2-ethoxyethanol and water
In this study, seven cationic Ir(III) complexes have been (3 : 1, v/v), the reaction mixture was then refluxed at 130 °C
synthesized and studied for their photocatalytic hydrogen overnight. After cooling down, the precipitate was collected and
generation on platinized TiO2 nanoparticles with AA as SED. washed by small amount of hexane, deionized (D.I.) water and
methanol. After drying in vacuum at 60 °C overnight, the dimer
Cyclometalated Ir(III) chromophores Ir2–Ir5 with various elec-
(1 molar equivalent) and silver hexafluorophosphate (4 molar
tron-donating functionalities exhibit intense absorption at equivalents) were dissolved in 50 mL CH2Cl2 and MeOH (1 : 1, v/v) in
wavelengths extending as far as 600 nm. This is presumed to be dark and refluxed at 50 °C. After an hour, the reaction mixture was
the mixing between intra-ligand transition on the C^N ligand cooled down and 2,2’-bipyridyl-4,4’-dicarboxylic acid (H2dcbpy)
and MLCT/LLCT transitions. This rational structural modification (2 molar equivalents, dissolved in 5 mL acetone with 2 mL triethyl-
on the C^N ligand makes these complexes attractive in photo- amine) was added to the mixture. The reaction mixture was
refluxed again at 60 °C overnight. Solvent was then removed in
catalytic hydrogen generation under visible light. Particularly,
vacuo and the crude product was purified through a wet column
Ir4 achieved the highest TON of 12 300 and TOFi of 394, with with silica gel eluting with acetone and CH2Cl2, followed by
activityi of 547 000 μmol g 1 h 1 and AQYi of 9.59 %, under the methanol and CH2Cl2. The Ir(III) complexes were then dissolved in
illumination of blue light LEDs after 105 hours. The TON (based minimal amount of methanol and CH2Cl2, and acidified with diluted
on n(H2)/n(PSr)) in this study is the highest value reported to hexafluorophosphoric acid solution. After few hours stirring, the
date in the system of Ir(III) with Pt NPs as the catalyst. Notably, Ir(III) complex was extracted with CH2Cl2 and distilled water, and
dried over anhydrous MgSO4. The filtrate was dried in vacuo to
the electron-rich and bulky triarylamine donor design has great
obtain the target Ir(III) complex.
potential in improving the robustness and visible light
absorption of the Ir(III)-based PSs for the photocatalytic system.

Chem. Eur. J. 2022, e202104575 (9 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Synthesis of Ir1 1.35 (m, 24H, alkyl), 0.88 (s, 12H, alkyl) ppm. 19F NMR (377 MHz,
DMSO-d6): ppm. 71.16 (d, JP-F = 711 Hz, PF6) ppm. HRMS (ESI-MS,
L1 (280 mg, 1.07 mmol) and IrCl3 · 3H2O (146 mg, 0.487 mmol) were m/z): [M + ] 2440.9410; calcd for (C154H130IrN10O4S2) 2440.9381.
used to synthesize the target complex following the general
procedure. A brown-orange powder was obtained as Ir1 in 18 %
yield (86 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.19 (m, 2H, Ar), 8.89 Synthesis of Ir6
(m, 2H, Ar), 8.79 (s, 2H, Ar), 8.62 (d, J = 8.0 Hz, 2H, Ar), 8.34 (d, J =
5.8 Hz, 2H, Ar), 8.07 (m, 4H, Ar), 7.99 (dd, J = 5.6, 1.3 Hz, 2H, Ar), 7.86 L6 (229 mg, 1.47 mmol) and IrCl3 · 3H2O (200 mg, 0.670 mmol) were
(d, J = 8.0 Hz, 2H, Ar), 7.43 (t, J = 7.8 Hz, 2H, Ar), 7.30 (d, J = 8.8 Hz, used to synthesize the target complex following the general
2H, Ar), 7.01 (t, J = 8.0 Hz, 2H, Ar), 6.43 (d, J = 4.9 Hz, 2H, Ar) ppm. 19F procedure. An orange powder was obtained as Ir6 in 31 % yield
NMR (377 MHz, DMSO-d6): ppm. 70.11 (d, JP-F = 712 Hz, PF6) ppm. (155 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.19 (s, 1H), 8.25 (d, J =
HRMS (ESI-MS, m/z): [M + ] 957.1177; calcd for (C46H28IrN4O4S2) 8.3 Hz, 1H), 8.11–7.82 (m, 4H), 7.63 (d, J = 5.8 Hz, 1H), 7.18–6.78 (m,
957.1181. 3H), 6.14 (d, J = 7.6 Hz, 1H). 19F NMR (466 MHz, DMSO-d6): δ 70.04
(d, JP-F = 708 Hz, PF6) ppm. HRMS (ESI-MS, m/z): [M + ] 745.1481; calcd
for (C34H24IrN4O4) 745.1427.
Synthesis of Ir2
L2 (630 mg, 1.25 mmol) and IrCl3 · 3H2O (170 mg, 0.569 mmol) were Synthesis of Ir7
used to synthesize the target complex following the general
procedure. A reddish orange powder was obtained as Ir2 in 14 % L7 (238 mg, 1.47 mmol) and IrCl3 · 3H2O (200 mg, 0.670 mmol) were
yield (120 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.31 (d, J = 7.8 Hz, used to synthesize the target complex following the general
2H, Ar), 8.90 (m, 2H, Ar), 8.61 (d, J = 7.8 Hz, 2H, Ar), 8.41 (d, J = procedure. An orange powder was obtained as Ir7 in 24 % yield
4.1 Hz, 4H, Ar), 8.22 (d, J = 7.8 Hz, 4H, Ar), 8.10 (m, 4H, Ar), 7.94 (d, (122 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.11 (s, 1H), 8.02 (d, J =
J = 8.4 Hz, 4H, Ar), 7.86 (d, J = 4.7 Hz, 2H, Ar), 7.59 (d, J = 8.4 Hz, 4H, 5.6 Hz, 1H), 7.91 (d, J = 5.8 Hz, 1H), 7.84–7.69 (m, 2H), 7.63 (d, J =
Ar), 7.50 (d, J = 8.4 Hz, 2H, Ar), 7.38 (m, 10H, Ar), 7.26 (t, J = 7.0 Hz, 4.7 Hz, 1H), 7.52 (d, J = 5.8 Hz, 1H), 6.90 (ddd, J = 7.4, 5.6, 1.7 Hz, 1H),
4H, Ar), 7.21 (s, 2H, Ar), 7.01 (t, J = 7.6 Hz, 2H, Ar) ppm. 19F NMR 6.13 (d, J = 4.7 Hz, 1H). 19F NMR (466 MHz, DMSO-d6): δ 70.08 (d,
(377 MHz, DMSO-d6): ppm. 71.15 (d, JP-F = 711 Hz, PF6) ppm. HRMS JP-F = 718 Hz, PF6) ppm. HRMS (ESI-MS, m/z): [M + ] 757.0590; calcd
(ESI-MS, m/z): [M + ] 1439.2973; calcd for (C82H50IrN6O4S2) 1439.2964. for (C30H20IrN4O4S2) 757.0555.

Synthesis of Ir3 Light-driven hydrogen production studies

L3 (800 mg, 1.55 mmol) and IrCl3 · 3H2O (210 mg, 0.703 mmol) were The photocatalytic reactions were carried out in a 25 mL pear-
used to synthesize the target complex following the general shaped flask sealed with rubber septa. To the flask, the dye-
procedure. A red powder was obtained as Ir3 in 23 % yield adsorbed platinized TiO2 was added into 5 mL aqueous ascorbic
(274 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.17 (m, 2H, Ar), 8.79 (s, acid (AA, 0.5 M) solution at pH 4.0. The flask was purged with
2H, Ar), 8.50 (d, J = 8.5 Hz, 2H, Ar), 8.41 (s, 2H, Ar), 8.29 (d, J = 5.5 Hz, mixture gas containing nitrogen and methane (80 : 20 mol %) for
2H, Ar), 8.00 (m, 4H, Ar), 7.80 (d, J = 5.1 Hz, 2H, Ar), 7.45 (d, J = 15 minutes. The methane gas was served as an internal stander for
8.5 Hz, 4H, Ar), 7.38 (d, J = 8.5 Hz, 2H, Ar), 7.25 (m, 10H, Ar), 7.04 (t, the GC analysis. The reaction flask was then stirred steadily under
J = 7.4 Hz, 4H, Ar), 6.96 (d, J = 7.5 Hz, 8H, Ar), 6.89 (s, 4H, Ar), 6.80 (d, the radiation of blue or green (ca. 472 and 520 nm, respectively)
J = 8.5 Hz, 4H, Ar) ppm. 19F NMR (377 MHz, DMSO-d6): ppm. 70.10 LEDs inside a just-fit container, which effective blocks the stray light
(d, JP-F = 711 Hz, PF6) ppm. HRMS (ESI-MS, m/z): [M + ] 1443.3280; from the environment. The light power was measured with a
calcd for (C82H54IrN6O4S2) 1443.3277. thermal sensor (Model: BIM-7203-0100F) and power meter (Model:
BIM-7001; Hangzhou Brolight Technology Co., Ltd.) and estimated
to be ~ 80 and 50 mW for blue and green light reaction,
Synthesis of Ir4 respectively. At each data point of the experiment, the headspaces
of the flasks were characterized by injecting 0.1 mL gas sample into
L4 (750 mg, 1.33 mmol) and IrCl3 · 3H2O (180 mg, 0.604 mmol) were GC (Agilent 7890B Gas Chromatography (GC) System with an
used to synthesize the target complex following the general Agilent HP-PLOT Molesieve 19091P-MS4 capillary column equipped
procedure. A dark red powder was obtained as Ir4 in 34 % yield with thermal conductivity detector). The amount of H2 produced
(340 mg). 1H NMR (400 MHz, DMSO-d6): δ 9.18 (m, 2H, Ar), 8.83 (m, was determined and quantified using a calibration plot of
2H, Ar), 8.54 (d, J = 8.2 Hz, 2H, Ar), 8.48 (s, 2H, Ar), 8.35 (m, 2H, Ar), integrated amount of H2 relative to CH4. In the course of 105 hours
8.04 (m, 4H, Ar), 7.87 (d, J = 5.0 Hz, 2H, Ar), 7.38 (m, 8H, Ar), 6.99 (d, radiation, eight GC samples were taken at different time points (5,
J = 8.9 Hz, 8H, Ar), 6.88 (m, 10H, Ar), 6.81(s, 2H, Ar), 6. 63 (d, J = 21, 26, 42, 47, 68, 89, 105 h) for each sample. The LED radiation is
8.9 Hz, 4H, Ar), 3.71 (s, 12H, OCH3) ppm. 19F NMR (377 MHz, DMSO- assumed to be monochromatic at emission intensity maximum
d6): ppm. 70.11 (d, JP-F = 711 Hz, PF6) ppm. HRMS (ESI-MS, m/z): (472 and 520 nm for blue and green light, respectively) and the
[M + ] 1563.3720; calcd for (C86H62IrN6O8S2) 1563.3700. corresponding apparent quantum yield values of each PS are
estimated according to the equation shown below.
Synthesis of Ir5
rate of H2 production � 2
L5 (1.10 g, 1.10 mmol) and IrCl3 · 3H2O (149 mg, 0.498 mmol) were AQY ð%Þ¼ �100 %
rate of incident photons
used to synthesize the target complex following the general
procedure. A red powder was obtained as Ir5 in 16 % yield
(192 mg). 1H NMR (400 MHz, CD2Cl2): δ 9.41 (s, 2H, Ar), 9.26 (s, 2H,
Ar), 8.56 (s, 2H, Ar), 8.42 (d, J = 5.7 Hz, 2H, Ar), 8.28 (s, 4H, Ar), 8.11 Computational methods
(d, J = 7.7 Hz, 4H, Ar), 7.91 (m, 6H, Ar), 7.66 (m, 12H, Ar), 7.45 (m,
16H, Ar), 7.21 (m, 14H, Ar), 7.05 (d, J = 8.5 Hz, 6 H, Ar), 6.85 (s, 4H, The 3D models of Ir1–Ir5 were designed using GaussView.[30] Then,
Ar), 6.71 (s, 2H, Ar), 4.30 (t, J = 7.4 Hz, 8H, NCH2), 1.87 (m, 8H, alkyl), the ground-state geometries optimizations and time dependent

Chem. Eur. J. 2022, e202104575 (10 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

density functional theory (TD-DFT) calculations of the above- [7] H. N. Kagalwala, D. N. Chirdon, S. Bernhard in Solar Fuel Generation, Vol.
mentioned compounds have been performed by the PBE0[31]//M06- 2 (Ed: E. Zysman-Colman), John Wiley & Sons Ltd., Hoboken, US, 2017,
2X[32] functional using Gaussian 09 program.[33] The 6–31G (d,p) p. 583–615.
[8] P. Wang, S. Guo, H.-J. Wang, K.-K. Chen, N. Zhang, Z.-M. Zhang, T.-B. Lu,
basis sets were used for C, H, N, O and S, while the Stuttgart-
Nat. Commun. 2019, 10, 3155.
Dresden (SDD) basis set and the effective core potentials (ECP) [9] a) D. R. Whang, K. Sakai, S. Y. Park, Angew. Chem. Int. Ed. 2013, 52,
were employed for Ir. The vertical excitation energies are computed 11612–11615; Angew. Chem. 2013, 125, 11826–11829; b) B. F. DiSalle, S.
for the first 60 singlet excited states because of computational Bernhard, J. Am. Chem. Soc. 2011, 133, 11819–11821.
limitations. The final UV-vis spectrum is obtained by a Gaussian [10] a) B. Zheng, R. P. Sabatini, W.-F. Fu, M.-S. Eum, W. W. Brennessel, L.
functional convolution with FWHM set at 0.66667 eV, which Wang, D. W. McCamant, R. Eisenberg, Proc. Nat. Acad. Sci. 2015, 112,
constitutes an average width for an absorption band observed in E3987–E3996; b) Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han, C. Li, Chem. Rev.
the UV-vis range. The molecular orbital diagrams were drawn using 2014, 114, 9987–10043.
[11] Y. Wu, W. Zhu, Chem. Soc. Rev. 2013, 42, 2039–2058.
Multiwfn and VMD.[34]
[12] a) Y.-J. Yuan, J.-Y. Zhang, Z.-T. Yu, J.-Y. Feng, W.-J. Luo, J.-H. Ye, Z.-G.
Zou, Inorg. Chem. 2012, 51, 4123–4133; b) S.-y. Takizawa, C. Pérez-
Bolívar, P. Anzenbacher Jr., S. Murata, Eur. J. Inorg. Chem. 2012, 2012,
3975–3979; c) A. Sinopoli, C. J. Wood, E. A. Gibson, P. I. P. Elliott, Eur. J.
Acknowledgements Inorg. Chem. 2016, 2016, 2887–2890; d) J. Yu, H. Tan, F. Meng, K. Lv, W.
Zhu, S. Su, Dyes Pigm. 2016, 131, 231–238.
[13] a) A. Tsuboyama, H. Iwawaki, M. Furugori, T. Mukaide, J. Kamatani, S.
C.-L. H. thanks the Hong Kong Research Grants Council (PolyU Igawa, T. Moriyama, S. Miura, T. Takiguchi, S. Okada, M. Hoshino, K.
123021/17P), the Science, Technology and Innovation Commit- Ueno, J. Am. Chem. Soc. 2003, 125, 12971–12979; b) M. Bandini, M.
Bianchi, G. Valenti, F. Piccinelli, F. Paolucci, M. Monari, A. Umani-Ronchi,
tee of Shenzhen Municipality (JCYJ20180306173720000), Envi- M. Marcaccio, Inorg. Chem. 2010, 49, 1439–1448; c) B. Jiang, Y. Gu, J.
ronment and Conservation Fund (ECF 79/2020) from the Qin, X. Ning, S. Gong, G. Xie, C. Yang, J. Mater. Chem. C 2016, 4, 3492–
Government of HKSAR and the Hong Kong Polytechnic 3498.
[14] a) A. M. Bünzli, E. C. Constable, C. E. Housecroft, A. Prescimone, J. A.
University (ZVVU and G-YBYZ) for their financial support. S. H.
Zampese, G. Longo, L. Gil-Escrig, A. Pertegás, E. Ortí, H. J. Bolink, Chem.
also acknowledges financial support from the National Natural Sci. 2015, 6, 2843–2852; b) A. Telleria, B. S. E. C. Kohlrausch, R.
Science Foundation of China (21703036). da C. Duarte, F. S. Rodembusch, J. Dupont, Z. Freixa, M. J. L. Santos,
ChemistrySelect 2016, 1, 2842–2848.
[15] S.-H. Wu, J.-W. Ling, S.-H. Lai, M.-J. Huang, C. H. Cheng, I. C. Chen, J.
Phys. Chem. A 2010, 114, 10339–10344.
Conflict of Interest [16] W. Tian, C. Yi, B. Song, Q. Qi, W. Jiang, Y. Zheng, Z. Qi, Y. Sun, J. Mater.
Chem. C 2014, 2, 1104–1115.
[17] a) C. Yang, F. Mehmood, T. L. Lam, S. L.-F. Chan, Y. Wu, C.-S. Yeung, X.
The authors declare no conflict of interest. Guan, K. Li, C. Y.-S. Chung, C.-Y. Zhou, T. Zou, C.-M. Che, Chem. Sci.
2016, 7, 3123–3136; b) Y. Kuramochi, O. Ishitani, Front. Chem. 2019, 7,
259.
[18] Y. Kawamura, J. Brooks, J. J. Brown, H. Sasabe, C. Adachi, Phys. Rev. Lett.
Data Availability Statement 2006, 96, 017404.
[19] C. Yang, F. Mehmood, T. L. Lam, S. L.-F. Chan, Y. Wu, C.-S. Yeung, X.
The data that support the findings of this study are available on Guan, K. Li, C. Y.-S. Chung, C.-Y. Zhou, T. Zou, C.-M. Che, Chem. Sci.
2016, 7, 3123–3136.
request from the corresponding author. The data are not [20] F. Neve, M. La Deda, A. Crispini, A. Bellusci, F. Puntoriero, S. Campagna,
publicly available due to privacy or ethical restrictions. Organometallics 2004, 23, 5856–5863.
[21] a) I. Chung, B. Lee, J. He, R. P. H. Chang, M. G. Kanatzidis, Nature 2012,
485, 486–489; b) J.-i. Fujisawa, T. Eda, M. Hanaya, Chem. Phys. Lett. 2017,
Keywords: iridium · photosensitizers · water splitting · 685, 23–26.
hydrogen generation · photocatalysis [22] Y.-J. Yuan, Z.-T. Yu, H.-L. Gao, Z.-G. Zou, C. Zheng, W. Huang, Chem. Eur.
J. 2013, 19, 6340–6349.
[23] a) E. Bae, W. Choi, J. Phys. Chem. B 2006, 110, 14792–14799; b) J. Zhang,
P. Du, J. Schneider, P. Jarosz, R. Eisenberg, J. Am. Chem. Soc. 2007, 129,
[1] a) V. Balzani, A. Credi, M. Venturi, ChemSusChem 2008, 1, 26–58; b) T. R. 7726–7727; c) Y.-J. Yuan, Z.-T. Yu, X.-Y. Chen, J.-Y. Zhang, Z.-G. Zou,
Cook, D. K. Dogutan, S. Y. Reece, Y. Surendranath, T. S. Teets, D. G. Chem. Eur. J. 2011, 17, 12891–12895.
Nocera, Chem. Rev. 2010, 110, 6474–6502. [24] a) P.-Y. Ho, B. Zheng, D. Mark, W.-Y. Wong, D. W. McCamant, R.
[2] a) A. Fujishima, K. Honda, Nature 1972, 238, 37–38; b) A. J. Bard, M. A. Eisenberg, Inorg. Chem. 2016, 55, 8348–8358; b) P.-Y. Ho, Y. Wang, S.-C.
Fox, Acc. Chem. Res. 1995, 28, 141–145. Yiu, W.-H. Yu, C.-L. Ho, S. Huang, Org. Lett. 2017, 19, 1048–1051; c) P.-Y.
[3] a) X. Zhang, T. Peng, S. Song, J. Mater. Chem. A 2016, 4, 2365–2402; b) B. Ho, M. F. Mark, Y. Wang, S.-C. Yiu, W.-H. Yu, C.-L. Ho, D. W. McCamant, R.
Cecconi, N. Manfredi, T. Montini, P. Fornasiero, A. Abbotto, Eur. J. Inorg. Eisenberg, S. Huang, ChemSusChem 2018, 11, 2517–2528.
Chem. 2016, 2016, 5194–5215. [25] a) T. N. Murakami, N. Koumura, M. Kimura, S. Mori, Langmuir 2014, 30,
[4] a) P. G. Bomben, K. C. D. Robson, B. D. Koivisto, C. P. Berlinguette, Coord. 2274–2279; b) W.-C. Chen, F.-T. Kong, R. Ghadari, Z.-Q. Li, X.-P. Liu, T. Yu,
Chem. Rev. 2012, 256, 1438–1450; b) I. N. Mills, J. A. Porras, S. Bernhard, Y. Huang, Y. Huang, T. Hayat, S.-Y. Dai, J. Phys. Chem. C 2017, 121, 8752–
Acc. Chem. Res. 2018, 51, 352–364; c) Y.-J. Yuan, Z.-T. Yu, D.-Q. Chen, Z.- 8759.
G. Zou, Chem. Soc. Rev. 2017, 46, 603–631; d) J.-H. Shon, T. S. Teets, ACS [26] A. Nakagawa, E. Sakuda, A. Ito, N. Kitamura, Inorg. Chem. 2015, 54,
Energy Lett. 2019, 4, 558–566. 10287–10295.
[5] a) S. Campagna, F. Puntoriero, F. Nastasi, G. Bergamini, V. Balzani in [27] S. Tschierlei, A. Neubauer, N. Rockstroh, M. Karnahl, P. Schwarzbach, H.
Photochemistry and Photophysics of Coordination Compounds: Ruthe- Junge, M. Beller, S. Lochbrunner, Phys. Chem. Chem. Phys. 2016, 18,
nium, (Eds.: V. Balzani, S. Campagna), Springer Berlin Heidelberg, Berlin, 10682–10687.
Heidelberg, 2007, p. 117–214; b) T. Bessho, E. Yoneda, J.-H. Yum, M. [28] a) L. Zhao, P. Wagner, J. E. Barnsley, T. M. Clarke, K. C. Gordon, S. Mori,
Guglielmi, I. Tavernelli, H. Imai, U. Rothlisberger, M. K. Nazeeruddin, M. A. J. Mozer, Chem. Sci. 2016, 7, 3506–3516; b) L. Miu, S. Yan, H. Yao, Q.
Grätzel, J. Am. Chem. Soc. 2009, 131, 5930–5934; c) L. L. Tinker, N. D. Chen, J. Zhang, Z. Wang, P. Cai, T. Hu, S. Ding, J. Chen, M. Liang, S.
McDaniel, S. Bernhard, J. Mater. Chem. 2009, 19, 3328–3337. Yang, Dyes Pigm. 2019, 168, 1–11.
[6] J. I. Goldsmith, W. R. Hudson, M. S. Lowry, T. H. Anderson, S. Bernhard, J. [29] a) L. L. Tinker, N. D. McDaniel, P. N. Curtin, C. K. Smith, M. J. Ireland, S.
Am. Chem. Soc. 2005, 127, 7502–7510. Bernhard, Chem. Eur. J. 2007, 13, 8726–8732; b) E. D. Cline, S. E.

Chem. Eur. J. 2022, e202104575 (11 of 12) © 2022 Wiley-VCH GmbH


Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104575

Adamson, S. Bernhard, Inorg. Chem. 2008, 47, 10378–10388; c) S. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi,
Panagiotakis, G. Landrou, V. Nikolaou, A. Putri, R. Hardré, J. Massin, G. N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, C. Bakken, C.
Charalambidis, A. G. Coutsolelos, M. Orio, Front. Chem. 2019, 7, 405. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin,
[30] R. Dennington II, T. Keith, J. Millam, GaussView 5.0.8, Semichem Inc., R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G.
Shawnee Mission, KS, 2009. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, A. D. Dapprich,
[31] J. P. Perdew, M. Ernzerhof, K. Burke, J. Chem. Phys. 1996, 105, 9982– A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox,
9985. AUSSIAN 09 (Revision D.01), Gaussian Inc.„ Wallingford, CT, 2009.
[32] Y. Zhao, D. G. Truhlar, Theor. Chem. Acc. 2008, 120, 215–241. [34] a) T. Lu, F. Chen, J. Comput. Chem. 2012, 33, 580–592; b) W. Humphrey,
[33] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. A. Dalke, K. Schulten, J. Mol. Graphics 1996, 14, 33–38.
Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H.
Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G.
Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J.
Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, Manuscript received: December 23, 2021
J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, Accepted manuscript online: February 16, 2022
E. N. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Version of record online: ■■■, ■■■■

Chem. Eur. J. 2022, e202104575 (12 of 12) © 2022 Wiley-VCH GmbH


RESEARCH ARTICLE
Intense and broadband UV-Vis ab- Dr. S.-C. Yiu, Dr. P.-Y. Ho, Y.-Y. Kwok, X.
sorption do matters in iridium(III)- He, Y. Wang, W.-H. Yu, Dr. C.-L. Ho*,
based photosensitizers for photocata- Dr. S. Huang*
lytic hydrogen generation. This study
1 – 13
reveals the importance of choosing an
appropriate electron donor on the Development of Strong Visible-
C^N ligands to improve the perform- Light-Absorbing Cyclometalated
ance. The use of triphenylamine- Iridium(III) Complexes for Robust
based Ir(III) PSs shows great potential and Efficient Light-Driven Hydrogen
in boosting the photocatalytic per- Production
formance.

You might also like