Next Article in Journal
Pd-Supported Co3O4/C Catalysts as Promising Electrocatalytic Materials for Oxygen Reduction Reaction
Previous Article in Journal
Catalytic Activity of Carbon Materials in the Oxidation of Minerals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Metal-Based Molecular Photosensitizers for Artificial Photosynthesis

School of Materials and Chemistry, University of Shanghai for Science and Technology, Shanghai 200093, China
Catalysts 2022, 12(8), 919; https://doi.org/10.3390/catal12080919
Submission received: 26 July 2022 / Revised: 8 August 2022 / Accepted: 17 August 2022 / Published: 19 August 2022
(This article belongs to the Special Issue Chiral Nanomaterials and Their Photo(Electro)catalytic Applications)

Abstract

:
Artificial photosynthesis (AP) has been extensively applied in energy conversion and environment pollutants treatment. Considering the urgent demand for clean energy for human society, many researchers have endeavored to develop materials for AP. Among the materials for AP, photosensitizers play a critical role in light absorption and charge separation. Due to the fact of their excellent tunability and performance, metal-based complexes stand out from many photocatalysis photosensitizers. In this review, the evaluation parameters for photosensitizers are first summarized and then the recent developments in molecular photosensitizers based on transition metal complexes are presented. The photosensitizers in this review are divided into two categories: noble-metal-based and noble-metal-free complexes. The subcategories for each type of photosensitizer in this review are organized by element, focusing first on ruthenium, iridium, and rhenium and then on manganese, iron, and copper. Various examples of recently developed photosensitizers are also presented.

1. Introduction

Over the past several decades, a large amount of fossil fuels have been exploited, especially during the Industrial Revolution [1,2,3]. The overexploitation of fossil fuels not only destroys the environment but has also led to serious global climate change and an energy crisis, which directly threaten the survival of humanity. To resolve these problems, thousands of green power sources are in development [4,5,6]. The Sun provides much more energy than current global demand; hence, it has become desirable in the development of novel sustainable energy. Artificial photosynthesis (AP), such as photocatalytic water splitting or CO2 reduction, can achieve the conversion of solar energy into chemical fuels, including carbon-free hydrogen as an environmentally friendly fuel, and industrial raw materials such as methanol [7,8,9,10]. Among various photosynthetic systems, molecular-based multi-component systems have been a research hotspot because of their properties, which can be controlled by the reasonable design of their chemical structures [11,12,13]. In a multi-component system, a photosensitizer plays an important role in light absorption and electron transfer; therefore, intense research has been devoted to photosensitizers, particularly from 2012 to 2022 as shown in Figure 1. This booming exploration of metal-complex-based photosensitizers has resulted in improving their performance and availability for commercial applications.

1.1. Photosynthesis in Nature

Natural photosynthesis in green plants and cyanobacteria is a critical reaction for the utilization of sunlight, which is crucial for providing sustainable energy on Earth. Photosynthesis in nature occurs through two unique photosystems, i.e., PSI and PSII, involving several electron transfer (ET) procedures [14]. As shown in Figure 2a, PSII is triggered by the photoexcitation of P680 and the release of electrons to give P680+, which is then involved in an oxygen-evolving reaction (OER). Simultaneously, PSI is initiated by the photoexcitation of P700, which then accepts electrons from the ET chain of PSII, resulting in the activation of the Calvin cycle, where CO2 is reduced into carbohydrates.

1.2. The Role of Photosensitizers in Artificial Photosynthesis

Inspired by nature, artificial photosynthesis (AP) simulates the function of critical components in a photocatalytic reaction [15]. Unlike natural photosynthesis, AP can not only convert CO2 into biomass, but it can also produce various chemical fuels such as hydrogen. There are three processes in a multi-component artificial photosynthesis reaction: (i) the light harvesting process in which a photosensitizer absorbs sunlight efficiently and then is converted into an excited state (PS*); (ii) the charge generation and separation step in which electron-transfer occurs between the PS* and an e acceptor/donor; (iii) a catalytic reaction. Clearly, a photosensitizer plays an important role in the absorption of solar photons and the injection of the photoexcited electrons into an acceptor in a photocatalytic reaction.

1.3. A General Scheme for Photosensitizer-Involved Photocatalysis

Figure 2b illustrates a general scheme for a photosensitizer-involved AP reaction. A ground-state photosensitizer (PS) is excited by sunlight and form PS* via a metal-to-ligand charge transfer (MLCT), i.e., electron-transfer occurs between the metal orbital and ligand orbital. PS* then undergoes electron transfer quenching (reductive or oxidative) in the presence of a donor or an acceptor, i.e., sacrificial agent (SA) resulting in the formation of a reduced photosensitizer (PS) or oxidized photosensitizer (PS+). Finally, PS or PS+ can be re-oxidized or be re-reduced back to PS in a catalytic reaction for CO2 reduction, hydrogen generation, or water oxidation [16,17,18]. However, it is worth noting that there are two competing side reactions that take place: (i) the PS* decays to PS by heat or a nonradiative pathway; (ii) back electron transfer occurs between SA and PS+ or between SA+ and PS. Therefore, to improve the overall photocatalytic efficiency, the lifetime of an excited state PS* must be long enough to overcome these side reactions. Meanwhile, it requires that PS or PS+ provides a suitable driving force to make catalysis occurs. In addition, because solar irradiation (AM1.5G) consists of only ~4% ultraviolet light (λ < 400 nm) and ~43% near-infrared light (λ > 800 nm) but also ~53% visible light (400 nm < λ < 800 nm), the ground-state PS should absorb as much visible light as possible to obtain a strong amount of energy from sunlight [19].
Due to the crucial role of photosensitizer(s) in AP, a comprehensive review on photosensitizers for solar hydrogen generation was published in 2017 [20]. In this review, some criteria of evaluating molecular photosensitizers were first described, and then recent achievements on metal-based photosensitizers and their importance in artificial photosynthesis were summarized and highlighted. The photosensitizers were categorized into two main families: noble-metal and noble-metal-free complexes. Noble-metal photosensitizers include ruthenium-, iridium-, and rhenium-based complexes, while noble-metal-free photosensitizers contain manganese-, iron-, copper-, and other metal-based complexes.

2. Parameters for Evaluating Photosensitizers

Some desirable characteristics for a photosensitizer include: (i) strong absorption of a distinguished range of the solar spectrum; (ii) a long excited-state lifetime (ns~µs); (iii) chemical stability in solution; (iv) reversible redox potentials. Therefore, several parameters are derived to evaluate the performance of photosensitizers.

2.1. UV-Vis Absorption

In metal coordinating complexes, intense absorptions may arise from the transfer of electronic charges between ligand orbitals and metal orbitals, either ligand-to-metal charge transfer (LMCT) or metal-to-ligand charge transfer (MLCT). Since ligands for MLCT usually have vacant, low-lying π* orbitals, such as py, bpy, phen, and other aromatic ligands, which are widely used in molecular metal photosensitizers, MLCT has been discussed in previous studies more often than LMCT in photocatalytic processes; hence, herein, MLCT was the focus. As mentioned previously, charge transfer transition is a critical process in artificial photosynthesis. For effective utilization of solar light, a basic requirement for a photosensitizer is absorbing sunlight over the visible wavelength and initiating electron transfer from or to its excited state. Using the description above, measuring UV-Vis absorption is usually the first and one of the most important steps in estimating the photocatalytic applications of a photosensitizer.
There are two main parameters that can be obtained by the UV-Vis spectrum:
(i) The charge transfer wavelength λ (in nm):
E MLCT   ( cm 1 ) =   10 7 λ   ( nm )
E MLCT   ( eV ) = E MLCT   ( cm 1 ) 8065
(ii) The molar extinction coefficient (or absorption), which is independent of concentration, and calculated by:
ε = A max cl   ( dm 3   mol 1   cm 1 )
where Amax is the maximum absorbance; c is measured in mol dm−3, and the cell path length, l, is in cm.

2.2. Ground-State Redox Potential

The ground-state redox properties of photosensitizers are important indicators for photocatalysis. A common method of measuring ground-state redox potentials for metal complexes is cyclic voltammetry (CV). Generally, CV can reflect the oxidation of the metal center and the reduction of ligands in a complex. Take the [M(L1)(L2)]n complex, owing to its MLCT character, as an example [21]. The electron transfer processes are:
(i) Oxidation:
[ M ( L 1 ) ( L 2 ) ] n     [ M ( L 1 ) ( L 2 ) ] n   + 1 +   e
(ii) Reductions:
[ M ( L 1 ) ( L 2 ) ] n +   e   [ M ( L 1 ˙ ) ( L 2 ) ] n   1
[ [ M ( L 1 ˙ ) ( L 2 ) ] n   1 +   e   [ [ M ( L 1 ˙ ) ( L 2 ˙ ) ] n   2
In addition to the oxidation and reduction abilities, the stability of a photosensitizer is another important consideration, since it must be capable of being regenerated over multiple turnovers in a photosynthesis reaction. As a result, the reversibility of the above redox processes, being demonstrated by cyclic voltammetry, is also important.
Since MLCT transition is accompanied by the occurrence of metal oxidation and ligand reduction, a correlation between the energies of the charge transfer absorptions and the electrochemical properties is established without considering the solvation and electron correlation effects:
E MLCT = | E ( M n   + 1 / M n ) | + | E ( L / L ) |
Evidently, the electrochemical data deconvolve important details, but the value of E MLCT alone is insufficient to determine whether a photosensitizer is suitable for a photocatalysis reaction.

2.3. Steady-State Emission

For a typical metal-based photosensitizer, the radiative quantum yield (Φ) can be calculated from the ratio between the number of emitted photons and the number of absorbed photons:
ϕ = I em I abs
where Iem can be obtained from the steady-state emission spectrum.
According to Lakowicz’s method in Principles of Fluorescence Spectroscopy [22], the relative quantum yield can be calculated using Equation (9):
Φ p = Φ ref × Int p Int ref × ( 1 10 A ref ) ( 1 10 A p ) × η p 2 η ref 2
where Int is the area under the emission peak; A is the absorbance at the excitation wavelength; η is the refractive index of the solvent. For the same type of photosensitizers in the same solvent, the absorbance value A is similar and η is the same, and the above equation can be simplified as Equation (10):
Φ p = Φ ref × Int p Int ref

2.4. Excited-State Redox Potential

For a multi-component photocatalytic process, the driving force of a photosensitizer must be adequate. As a result, the value of E([M]n + 1/*[M]n), which reflects the electron-donating ability, must be more negative than the reduction potential of the catalyst so that electron(s) can be transferred from the excited state to the catalyst. Conversely, the value of E([M]n/*[M]n−1), which reflects the electron-accepting ability, must be more positive than the reduction potential of the catalyst (usually in a high oxidation state) to ensure the electron transfer occurs to the excited state. In this way, the excited-state redox potentials for molecular photosensitizers are calculated by the following equations:
(i) Oxidative quenching:
E ( [ M ] n   + 1 / [ M ] n * ) =   E ( [ M ] n   + 1 / [ M ] n )   E 0
E ( L * / L ) =   E ( L 0 / L ) +   E 0
(ii) Reductive quenching:
E ( [ M ] n / [ M ] ( n   1 ) * ) =   E ( [ M ] n / [ M ] n   1 ) +   E 0
E ( L / L * ) =   E ( L / L )   E 0
where E ( [ M ] n   + 1 / [ M ] n ) and E ( [ M ] n / [ M ] n   1 ) are the ground-state potentials. E 0 can be estimated from one of the following methods: (i) the crossover point of the lowest-energy MLCT absorption band and the corresponding emission band [23]; (ii) the crossover point of the tangent line of the emission spectrum (left half) with a wavelength axis [24]; (iii) the first vibronic band in the low-temperature emission spectrum [25] (see details below).
Clearly, the excited-state redox potential is an important indicator for evaluating photosensitizers, but it should be noted that the calculation of the excited-state redox potential may have some error because of the uncertainty of E0.

2.5. Lifetime

To simplify our discussion, two states for a photosensitizer after being excited are shown in Figure 3: the singlet state 1MLCT (S1) and the triplet state 3MLCT (T1). The S1 state can decay to the ground state (S0) via a radiative way, i.e., fluorescence (kr) and/or a nonradiative way (knr). Afterwards, the singlet-triplet transition happens through intersystem crossing (ISC) with the association of energy loss as heat (kISC). Thereafter, the T1 state’s decay to the S0 state follows a similar way as the S1 state: a radiative way, i.e., phosphorescence ( k r ) and/or a nonradiative way ( k nr ). Therefore, these decay constants (hence, lifetime) are important criteria for evaluating a photosensitizer.
Generally, the lifetime ( τ 0 ) of a photosensitizer is measured by excited-state transient absorption (TA) spectra and the signal fit to single [24] or triple [26] (depends on the mechanism of photosensitizers) exponential decay kinetics. To complete a photocatalytic reaction, the excited-state lifetime must be long enough to overcome the decay process.
τ 0 = 1 k r + k nr
where k r = ϕ 0 ×   k 0 ; k nr = ( 1 ϕ 0 ) ×   k 0 .
Through measuring τ 0 , k r can be calculated by Equation (16):
k r = ϕ p τ 0
Therefore, knr can be calculated by combining Equations (9), (15) and (16) as shown below:
k nr = 1 ϕ p τ 0

2.6. Marcus Electron-Transfer Theory

Since photocatalytic reactions are essentially electron transfer (ET) processes, the kinetics of ET is governed by Marcus theory, shown in Equation (18).
k ET = ( k B T h ) exp [ ( λ + Δ G 0 ) 2 4 λ k B T ]
where k B refers to the Boltzmann constant; T is the temperature (K); h refers to Planck’s constant; λ refers to the reorganization energy. ΔG0 is the driving force of the electron-transfer reaction as shown below:
Δ G 0 = E h   ( electron   donor ) E h   ( electron   acceptor )
Apparently, increasing the driving force can improve the rate of electron transfer, i.e., the rate of catalysis as indicated by the excited-state potential equations and the Marcus relationship [27]. However, Marcus theory also demonstrates that there is a Marcus inverted region, where a stronger driving force leads to slower ET rates, hence, there is no normal standard criterion for photosensitizers for all photocatalytic reactions.

2.7. Stern-Volmer Quenching

Because of the occurrence of back-electron transfer between photosensitizers and catalysts, it is necessary to add a sacrificial reagent to most photocatalytic reactions. K SV is calculated from the emission quenching data as a function of the concentration of persulfate, which follows the Stern–Volmer relationship:
I 0 I =   k sv [ S 2 O 8 2 ] + 1
where
k sv =   τ 0 × k q
Using above discussions, the radiative quantum yield of a photosensitizer shows a dependent relationship with the concentration of a quencher, i.e., [Q]:
ϕ q = k r k 0 + k q [ Q ]
ϕ 0 ϕ q = 1 + k q [ Q ] k 0

2.8. Turnover Frequency (TOF)

TOF is an essential parameter of a catalyst, reflecting its catalytic activity and performance. A recent review by Xiujuan Wu and Licheng Sun outlined several calculation methods for evaluating catalysts [28]. It is less common to calculate the TOF for a photosensitizer than for a catalyst but, in some cases, when a standard catalyst is applied, TOF becomes an important criterion for comparing the performances of different photosensitizers. From the perspective of a definition, TOF refers to the amount of product converted from the reactant per mol of the effective photosensitizer per unit time.
TON = amount   of   transfered   electrons   for   product ( moles )   amount   of   photosensitizer   ( moles )
TOF = TON time   of   reaction
Since the plot of TON versus time does not usually follow a linear relationship, the TOF is taken from the pseudo-first-order rate constant, i.e., the maximum photocatalytic rate.

3. Noble-Metal-Based Photosensitizers

Noble metals, such as Ru [29], Ir [30], Pt [31,32,33,34], Os [35,36], and Re [37], have been extensively investigated as photosensitizers for artificial photosynthesis. These complexes have received wide attention due to the fact of their attractive photophysical electrochemical properties such as visible-light 1MLCT, long excited-state lifetime, high photocatalytic driving force, and high catalytic activities.

3.1. Ru-Based Photosensitizers

The ruthenium-polypyridine family is an early-developed photosensitizer [38], and its fundamental properties and mechanisms have been well investigated [39,40,41,42,43]. Current research efforts on this family of photosensitizer focus on improving its photochemical stability, antidecomposition, extreme pH value tolerance, and driving force at various environmental conditions.
Concepcion et al. developed a series of homoleptic Ru(II) polypyridyl complexes (Figure 4a), which showed decent photochemical water oxidation performance at pH 1 [24]. By modifying electron-withdrawing groups (i.e., −CF3 and −PO3H2) at the 4, 4′, or 5′ positions of bpy ligands, the redox potentials increased to as high as 1.6 V (vs. NHE) at pH 1. It is worth mentioning that unlike heteroleptic complexes, the redox potentials of which can be increased at the expense of decreasing the excited-state lifetimes [44], some of these homoleptic Ru(II) polypyridyl complexes presented increased lifetimes from 580 to 731 ns at pH 1. Their studies illustrated that replacing H with electron-withdrawing groups in homoleptic complexes provides a feasible strategy for designing and synthesizing high-performing photosensitizers at pH 1. The kinetic isotope effects (KIEs) of the PO3H2-containing complexes proved the concerted EPT between the photosensitizer and catalyst for the first time.
In addition to [Ru(bpy)3]2+-type photosensitizers, [Ru(phen)3]2+ complexes are another family of Ru photosensitizers, but they are more used in CO2 reduction reactions (CO2RR). However, the excited-state lifetime of raw [Ru(phen)3]2+ complexes is short (360 ns in CH3CN), resulting in a weakened photocatalytic performance for CO2RR. It has been proved that adding a conjugated group can improve the excited-state lifetime, but an improper introduction will make the excited oxidation potentials worse. To address this problem, Lu et al. creatively synthesized pyrene-modified [Ru(phen)2(L)]2+ photosensitizers (Figure 4b) and realized the fine-tuning of PS properties for efficient CO2 reduction. The conjugated group, pyrene, greatly improved the sensitizing ability of [Ru(phen)3]2+ complexes. Especially, [Ru(phen)2(3-pyrenylphen)]2+ (Ru-3) showed a moderate excited-state lifetime (68.2 µs) but a 17 times larger TOF than [Ru(phen)3]2+ with a large TON of 66,480 [45]. Their study clearly illustrated the importance of balancing lifetime and potential in improving the overall performance of photocatalytic systems.
In recent years, multielectron accumulation has been considered an effective strategy for improving photocatalytic performance. Several studies have shown that naphthalene [46], anthraquinone [47], and dipyrido-[3,2-a:2′,3′-c]phenazine (dppz) [48] own the electron-storage ability and can enhance the efficiency of light-driven processes. These bent-bridging-ligands involve ruthenium complexes; however, they usually require ingenious design of molecular structures and smart synthesis procedures for ligands. On the foundation of dppz-based complexes, Dietzek and Kerlidou et al. modified a pyridoquinolinone subunit on dppz (Figure 4c) [49] aimed at realizing electron accumulation by extending the π system. Various methods proved the 2e/2H+ mechanism of this photosensitizer and demonstrated its potential applications in producing H2O2.

3.2. Ir-Based Photosensitizers

Although Ru-based photosensitizers have shown vast potential in solar energy conversion, their poor photostability limit their application. In responding to this problem, Ir (III) complexes have drawn particular attention in replacing Ru (II) complexes over the past decade. In 2018, Bernhard et al. thoroughly reviewed Ir (III)-type photosensitizers [50]; hence, this review focused only on the development of Ir photosensitizers thereafter.
[Ir(ppy)2(bpy)]+ and its derivatives (Figure 5a) have been extensively studied as photosensitizers for energy conversion due to the fact of their decent photostability. Based on these studies, in 2020, Dietzek et al. developed and studied a series of iridium complex–polyoxometalate (POM) dyads, Ir-POMM, which have shown high efficiency in photocatalytic hydrogen evolution reactions [51,52]. By changing the central metal of POM from Mn3+ to Fe3+ to Co3+, the yields of charge-separated state Ir+-POMM (i.e., the rate-limiting intermediate) decreased, while the lifetime increased from 290 to 540 ps. A photoinduced electron transfer dynamics study illustrated that the catalytic capacity decreased in the order Ir-POMMn > Ir-POMCo > Ir-POMFe, which was in the same order of the yields of Ir+-POMM, demonstrating the high impact of yield on the catalytic capacity.
In 2021, Teets et al. prepared a series of β-diketiminate-modified (NacNac) iridium complexes as demonstrated in Figure 5b. The electron-rich NacNac ligands dramatically improved the photoinduced electron transfer rate, as indicated by the decreased excited redox potential from −2.1 to −2.5 V (vs. Fc+/Fc), making various challenging photoredox reactions occur under moderate conditions with high yield [53,54]. Later, they further improved the photocatalytic performance by replacing ppy with ptz and pmb, all of which presented significant excited redox potentials in the range of −2.4 to −2.8 V [55]. Their studies put forward practical proposals for designing novel photosensitizers with high efficiency and feasible photostability.

3.3. Re-Based Photosensitizers

Rhenium photosensitizers are usually found in the family of [Re(bpy)(CO)3Cl] complexes. Ishitani et al. developed a series of highly efficient rhenium(I) trinuclear photosensitizers (Figure 6a) for photocatalytic CO2 reduction [56]. As a consequence of the efficient reductive quenching of the Re-ring by triethanolamine and fast electron transfer, the CO2 reduction process showed high photocatalytic efficiencies. A summary of the characteristics and properties of Re(I) diimine complexes weree published by Ishitani et al. in 2017 [57]. In 2020, by modifying the [Re(bpy)(CO)3Cl] complex with an anchoring group, -PO3H2 (Figure 6b), Hamm et al. successfully loaded it with a cobalt catalyst, [Co(DMTPy-O-benzyl-3,5-bis(MePO3H)], on ZrO2, demonstrating the possibility of constructing heterogeneous photocatalytic systems using Re photosensitizers [58].
Recently, Fernandez-Terán et al. reported the feasibility of tuning the properties of rhenium(I) tricarbonyl complexes through changing the substituent with various groups (i.e., CN, CF3, Br, H, OMe, and NMe2) as shown in Figure 6c. It was found that the complex with the most electron-donating group, NMe2, showed a notable lifetime of 380 ns and a TON of over 2100 [59]. This is the first example of applying the 3ILCT state of a rhenium(I) tricarbonyl complex as a stable photosensitizer.
These studies implied that Re-based complexes are good candidates for replacing ruthenium as photosensitizers for stable photocatalytic reactions.
Table 1 summarized the properties of noble-metal-based photosensitizers in this review, as shown below.

4. Noble-Metal-Free Photosensitizers

Over the last decade, the investigation of photosensitizers based on nonprecious transition metals dramatically increased owing to their favorable (photo)physical properties, high earth abundancy, and low cost. Table 2 shows the earth abundancy of the metal elements mentioned in this review. However, due to the fact of these photosensitizers’ feature of unwanted nonradiative decay, they generally perform worse than noble-metal-based photosensitizers in terms of photophysical properties (i.e., MLCT and lifetime) and redox potential [60]. Nevertheless, nonprecious transition metal complexes have shown highly attractive potentials in photocatalytic hydrogen production and CO2 reduction [61]. In the subsequent sections, recent advances in noble-metal-free photosensitizers from low atomic number to high atomic number are summarized.

4.1. Mn-Based Photosensitizers

Manganese plays an essential role in absorbing light in natural photosynthesis process; hence, it has inspired the development of Mn-based nanomaterials for photocatalysis [63,64,65,66]. However, due to the fact of their intrinsic nonradiative decay of excited states, it is still a grand challenge to make Mn-based complexes luminescent and photoactive.
Wenger et al. recently reported a family of novel isocyanide Mn(I) complexes, [Mn(Lbi)3]+ and [Mn(Ltri)2]+, that exhibited the first example of manganese complexes with luminescent MLCT and photo-reactivity at room temperature [67,68] as shown in Figure 7. This series of Mn-based homoleptic photosensitizers is air-stable and nontoxic, and the redox potential of Mn(I/II) is reversible. Spectrum studies gave a 1MLCT absorption wavelength at ~390 nm and an emission wavelength at ~500 nm. These complexes have been successfully applied in energy conversion reactions. Unlike traditional energy transfer pathways (i.e., MLCT, LMCT, and MC excited states), their study provide the attractive possibility of realizing the triplet-energy-transfer photo-reactivity by a ligand-centered π–π* state.
Currently developed Mn-based complexes are not comparable with traditional noble-metal-based photosensitizers in terms of excited-state lifetime or driving force for photocatalytic reactions, but these studies open new avenues to exploring cheap bio-inspired molecules for artificial photosynthesis.

4.2. Fe-Based Photosensitizers

Iron is the most earth-abundant transition metal element [62]. Therefore, in recent years, much attention has been paid toward developing Fe-based photosensitizers considering their low cost. Inspired by Ru(II) polypyridyl complexes, which show decent photophysical properties [38], iron(II) polypyridyl photosensitizers have been widely studied [69,70,71,72,73,74].
Kunnus et al. recently reported a series of [Fe(N^N)(CN)4]2−-type photosensitizers, where N^N represents 2,2′-bipyridine (bpy); 2,3-bis(2-pyridyl)pyrazine (bpyz); 2,2′-bipyrimidine (bpym), respectively (Figure 8a), and studied the solvent effect on the chemical properties as well as the MLCT excitation energy dependence of the 3MLCT lifetime. Femtosecond transient absorption (TA) data and td-DFT calculations revealed the tunability of the excited-state relaxation mechanisms and charge transfer lifetimes, which increased from 0.22 to 16.9 ps with an increasing number of nitrogen in the ligand.
In addition to iron(II) polypyridyl complexes, iron (II) N-heterocyclic carbene (NHC) complexes (Figure 8b) have been extensively investigated as photosensitizers [75,76,77,78,79]. The strong bonding between ion and the carbene ligands leads to increased ligand-field splitting, resulting in the destabilization of the MC states. As a consequence, Fe(II)-NHC-type photosensitizers present prominent properties such as long excited-state lifetimes and high charge transfer abilities [77]. Bauer et al. studied the relationships between the number of NHC ligands and the photophysical properties of Fe photosensitizers. An increasing trend in the 3MLCT lifetime was found when there was an increase in the number of NHC ligands [78]. In their follow-up study, the excited-state dynamics of [FeII(tpy)(pyz-NHC)]2+ was investigated. The results of intersystem crossing and triplet dynamics further interpreted the excellent performance of Fe(II)-NHC photosensitizers.
At present, the photophysical properties and stability of Fe-based photosensitizers are not comparable to Ru-based complexes. However, its high earth-abundancy and low cost attract continuous attention. Future work on Fe-based photosensitizers will focus on extending their charge-separated state and improving their photostability.

4.3. Cu-Based Photosensitizers

Copper-based photosensitizers have been popular since the 1970s because of their MLCT nature and high earth abundancy [80,81,82,83,84]. With the extensive study of copper complexes, Cu-based photosensitizers possessing considerable properties and good performances have increasingly been developed. Unlike other abundant transition metals, copper complexes have a filled d subshell, which blocks the formation of a metal-centered excited state. Therefore, many copper (I) complexes show MLCT luminescence [60]. As a result of its sustainability and photophysical properties, in recent years, copper is replacing ruthenium for use as molecular photosensitizers in dye-sensitized solar cells (DSSCs) [85]. Housecroft and Constable have reviewed the progress of copper(I)-containing DSSCs [86].
Mulfort et al. developed and studied a series of copper(I) diimine photosensitizers based on the HETPHEN method [87,88] and immobilized these photosensitizers on the metal oxide surface via carboxylate groups [26]. The CuHETPHEN strategy provides many more possibilities for tuning the structures and properties of copper photosensitizers. These discoveries related to copper complexes promote important improvements in Cu-based photosensitizers and further increase their potential applications in the solar energy conversion industry. The photochemical processes of copper-based photosensitizers have been well explored by advanced technologies such as spectroscopics [89,90,91,92]. In general, ground-state copper complex is excited into a singlet MLCT state, which then undergoes a flattening distortion and becomes triplet MLCT state. Depending on the coordination ability of the solvent, the 3MLCT state complex may or may not bind the solvent molecule. After experiencing a strong interacting, the complex recovers to the ground state as shown in Figure 9. Chen et al. reviewed the photocatalytic mechanisms of copper(I) diimine complexes and their developments and applications in solar energy conversion in 2015 [93].
From a chelating element perspective, there are four types of copper (I) photosensitizers: Cu(N^N)(N^N), Cu(N^N)(P^P), Cu(N^N)(N^P), and Cu(N^P)(N^P) as shown in Figure 10a. In terms of Cu(N^N)(N^N)-type photosensitizers, Hadt et al. quantified the entatic effects of several [Cu(phen-X)2]+ complexes, where X are the 2,9-alkyl substitutions, and correlations were found between the lifetime, the excited-state relaxation energy, the reorganization energy, and the energy gap [94]. Castellano et al. successfully prepared various Cu-phen MLCT photosensitizers featuring bulky substituents. Impressive photophysical properties (τ = 1.5 μs, Φ = 2.6% in acetonitrile) were discovered when use dchtmp as the ligand. Their C−C radical coupling photochemistry strategy broke through the traditional limit of steric bulk possibility for homoleptic Cu(I) complexes [95,96].
In recent years, scientists have aimed at replacing nitrogen with phosphines in transition-metal-based photosensitizers. Generally, the introduction of phosphines usually brings about a higher stability and lower redox potentials [97,98,99,100,101,102]. Karnahl et al. prepared a hetero-bidentate P^N ligand, phox, and improved the performance of Cu-based photosensitizers [100]. Ishitani et al. synthesized [Cu(dmp)(DPEphos)]+ complexes, where DPEphos is 2-(diphenylphosphino)phenyl]ether, for photocatalytic CO2 reduction. By comparing the properties of a series of [Cu(N^N)(DPEphos)]+ photosensitizers, the influences of expending π-systems were discovered [103]. Dietzek et al. [91] and Karnahl et al. [104] recently studied the extended π-system effect on the properties of [Cu(xant)(dppz)]+ and [Cu(xant)(dmdppz)]+. The extended π-system provided unusual properties such as broadening and a red-shift of the MLCT absorption and a prolonged lifetime. Dietzek et al., in particular, studied the effect of the electronic structures of Cu(I) photosensitizers containing phosphines [105]. Their study on the Franck−Condon region of heteroleptic Cu(I) complexes provides new theoretical principles for designing Cu-based photosensitizers. Bruggeller et al. recently reviewed the development of photosensitizers with phosphines and summarized the influence of phosphines on the properties of photosensitizers [106].
Among the aforementioned noble-metal-free complexes, the copper family has the greatest potential to replace noble-metal photosensitizers as indicated by their promising photophysical properties such as long lifetimes. Nevertheless, their poor chemical stability and environmental sensitivity (to pH value, humidity, etc.) impede their broad applications in energy conversion. Efforts towards improving these properties will be the focus of future research.

4.4. Other Photosensitizers

Other types of metal-based photosensitizers are less common than manganese, iron, and copper. Inspired by natural photosynthesis, many of these photosensitizers (Figure 11) are based on porphyrin and its derivatives, which itself is a biomimetic light-harvesting molecule [107,108,109]. Poddutoori et al. reviewed unique aluminum (III) porphyrin complexes, which can act as either electron donor or electron acceptor in donor–acceptor systems because of their redox and optical properties [110]. Park et al. creatively designed positively charged Zn and neutral Sn porphyrins complexes, [ZnTMePyP4+]Cl4 and Sn(OH)2TPyP, Sn(Cl2)TPP-[COOMe]4, and Sn(Cl2)TPP-[PO(OEt)2]4, as photosensitizers for H2 evolution. Their study clearly demonstrated the influence of pH value as well as solution concentration. It also showed that the photocatalytic process can be realized by oxidative quenching of the photosensitizer [111]. The results of this study shed light on designing and synthesizing metal-containing photosensitizers based on porphyrin. Willner et al. developed an all-DNA system with Zn (II)-protoporphyrin IX (ZnIIPPIX) as a photosensitizer [112]. In this mimic PSI system, ZnIIPPIX presented effective photoinduced electron transfer (ET). Their report provides a versatile approach for designing supramolecular photosensitizers.
Detailed parameters for some of above noble-metal-free photosensitizers are summarized in Table 3.

5. Conclusions and Outlook

This review endeavored to summarize the evaluating parameters and recent work on metal-based molecular photosensitizers. Noble-metal-based photosensitizers, such as ruthenium, iridium, and rhenium, are the most investigated metal-based photosensitizers and many reviews on these photosensitizers are available; hence, here, recent advances in noble-metal-based photosensitizers were briefly reviewed. These photosensitizers, however, are scarce, expensive, and toxic to humans. Thus, noble-metal-free photosensitizers have become research hotspots in the solar energy conversion field. Due to the fact of their valuable photophysical properties and their charge-separation ability to induce photochemistry, manganese, iron, and copper, they have recently received particular attention. Some photosensitizers were found to be active, but further work in promoting their higher efficiency and stronger stability as well as a greater understanding of their mechanisms are needed for their large-scale industrial application. In addition to the metal-based photosensitizers that were highlighted in this review, biomolecular photosensitizers, such as from natural bacterial sources, are another attractive alternative to expensive metal-based photosensitizers.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

Contribution from Lijuan Wang who collected publication data in Figure 1. Support from Shanghai Outstanding Academic Leaders Program.

Conflicts of Interest

We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work, there is no professional or other personal interest of any nature or kind of in any product, service and/or company that could be construed as influencing the review entitled.

References

  1. Kharecha, P.A.; Hansen, J.E. Implications of “peak oil” for atmospheric CO2 and climate. Glob. Biogeochem. Cycles 2008, 22, GB3012. [Google Scholar] [CrossRef] [Green Version]
  2. Stern, D.I.; Kander, A. The role of energy in the industrial revolution and modern economic growth. Energy J. 2012, 33, 3. [Google Scholar] [CrossRef] [Green Version]
  3. Mayumi, K. Temporary emancipation from land: From the industrial revolution to the present time. Ecol. Econ. 1991, 4, 35–56. [Google Scholar] [CrossRef]
  4. Dincer, I. Renewable energy and sustainable development: A crucial review. Renew. Sust. Energ. Rev. 2000, 4, 157–175. [Google Scholar] [CrossRef]
  5. Vakulchuk, R.; Overland, I.; Scholten, D. Renewable energy and geopolitics: A review. Renew. Sust. Energ. Rev. 2020, 122, 109547. [Google Scholar] [CrossRef]
  6. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The role of renewable energy in the global energy transformation. Energy Strategy Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  7. Ren, D.; Loo, N.W.X.; Gong, L.; Yeo, B.S. Continuous production of ethylene from carbon dioxide and water using intermittent sunlight. ACS Sustain. Chem. Eng. 2017, 5, 9191–9199. [Google Scholar] [CrossRef]
  8. Kim, D.; Sakimoto, K.K.; Hong, D.; Yang, P. Artificial photosynthesis for sustainable fuel and chemical production. Angew. Chem. Int. Ed. 2015, 54, 3259–3266. [Google Scholar] [CrossRef]
  9. Ji, X.; Su, Z.; Wang, P.; Ma, G.; Zhang, S. Integration of artificial photosynthesis system for enhanced electronic energy-transfer efficacy: A case study for solar-energy driven bioconversion of carbon dioxide to methanol. Small 2016, 12, 4753–4762. [Google Scholar] [CrossRef]
  10. Jia, Y.; Xu, Y.; Nie, R.; Chen, F.; Zhu, Z.; Wang, J.; Jing, H. Artificial photosynthesis of methanol from carbon dioxide and water via a Nile red-embedded TiO2 photocathode. J. Mater. Chem. A 2017, 5, 5495–5501. [Google Scholar] [CrossRef]
  11. Stoll, T.; Castillo, C.E.; Kayanuma, M.; Sandroni, M.; Daniel, C.; Odobel, F.; Fortage, J.; Collomb, M.N. Photo-induced redox catalysis for proton reduction to hydrogen with homogeneous molecular systems using rhodium-based catalysts. Coord. Chem. Rev. 2015, 304, 20–37. [Google Scholar] [CrossRef]
  12. Puntoriero, F.; La Ganga, G.; Cancelliere, A.M.; Campagna, S. Recent progresses in molecular-based artificial photosynthesis. Curr. Opin. Green Sustain. Chem. 2022, 36, 100636. [Google Scholar] [CrossRef]
  13. Berardi, S.; Drouet, S.; Francàs, L.; Gimbert-Suriñach, C.; Guttentag, M.; Richmond, C.; Stoll, T.; Llobet, A. Molecular artificial photosynthesis. Chem. Soc. Rev. 2014, 43, 7501–7519. [Google Scholar] [CrossRef]
  14. Wang, C.; O’Hagan, M.P.; Willner, B.; Willner, I. Bioinspired artificial photosynthetic systems. Chem. Eur. J. 2022, 28, e202103595. [Google Scholar] [CrossRef]
  15. Barber, J.; Tran, P.D. From natural to artificial photosynthesis. J. R. Soc. Interface 2013, 10, 20120984. [Google Scholar] [CrossRef] [Green Version]
  16. Khalil, M.; Gunlazuardi, J.; Ivandini, T.A.; Umar, A. Photocatalytic conversion of CO2 using earth-abundant catalysts: A review on mechanism and catalytic performance. Renew. Sust. Energ. Rev. 2019, 113, 109246. [Google Scholar] [CrossRef]
  17. Puntoriero, F.; Sartorel, A.; Orlandi, M.; La Ganga, G.; Serroni, S.; Bonchio, M.; Scandola, F.; Campagna, S. Photoinduced water oxidation using dendrimeric Ru (II) complexes as photosensitizers. Coord. Chem. Rev. 2011, 255, 2594–2601. [Google Scholar] [CrossRef]
  18. Dogutan, D.K.; Nocera, D.G. Artificial photosynthesis at efficiencies greatly exceeding that of natural photosynthesis. Acc. Chem. Res. 2019, 52, 3143–3148. [Google Scholar] [CrossRef]
  19. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef]
  20. Yuan, Y.J.; Yu, Z.T.; Chen, D.Q.; Zou, Z.G. Metal-complex chromophores for solar hydrogen generation. Chem. Soc. Rev. 2017, 46, 603–631. [Google Scholar] [CrossRef]
  21. Shon, J.H.; Teets, T.S. Molecular photosensitizers in energy research and catalysis: Design principles and recent developments. ACS Energy Lett. 2019, 4, 558–566. [Google Scholar] [CrossRef]
  22. Lakowicz, J. Principles of Fluorescence Spectroscopy; Springer: New York, NY, USA, 2013. [Google Scholar]
  23. Islam, A.; Sugihara, H.; Arakawa, H. Molecular design of ruthenium (II) polypyridyl photosensitizers for efficient nanocrystalline TiO2 solar cells. J. Photochem. Photobiol. A 2003, 158, 131–138. [Google Scholar] [CrossRef]
  24. Wang, L.; Shaffer, D.W.; Manbeck, G.F.; Polyansky, D.E.; Concepcion, J.J. High-redox-potential chromophores for visible-light-driven water oxidation at low pH. ACS Catal. 2019, 10, 580–585. [Google Scholar] [CrossRef]
  25. Vlcek, A.; Dodsworth, E.S.; Pietro, W.J.; Lever, A. Excited state redox potentials of ruthenium diimine complexes; correlations with ground state redox potentials and ligand parameters. Inorg. Chem. 1995, 34, 1906–1913. [Google Scholar] [CrossRef]
  26. Eberhart, M.S.; Phelan, B.T.; Niklas, J.; Sprague-Klein, E.A.; Kaphan, D.M.; Gosztola, D.J.; Chen, L.X.; Tiede, D.M.; Poluektov, O.G.; Mulfort, K.L. Surface immobilized copper (i) diimine photosensitizers as molecular probes for elucidating the effects of confinement at interfaces for solar energy conversion. ChemComm 2020, 56, 12130–12133. [Google Scholar] [CrossRef]
  27. Marcus, R.A. Electron transfer reactions in chemistry. Theory and experiment. Rev. Mod. Phys. 1993, 65, 599. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, H.; Wu, X.; Sun, L. Copper-based homogeneous and heterogeneous catalysts for electrochemical water oxidation. Nanoscale 2020, 12, 4187–4218. [Google Scholar] [CrossRef]
  29. Tran, T.T.; Pino, T.; Ha-Thi, M.H. Watching intermolecular light-induced charge accumulation on naphthalene diimide by tris (bipyridyl) ruthenium (II) photosensitizer. J. Phys. Chem. C 2019, 123, 28651–28658. [Google Scholar] [CrossRef]
  30. Andreiadis, E.S.; Chavarot-Kerlidou, M.; Fontecave, M.; Artero, V. Artificial photosynthesis: From molecular catalysts for light-driven water splitting to photoelectrochemical cells. Photochem. Photobiol. 2011, 87, 946–964. [Google Scholar] [CrossRef]
  31. Chen, H.C.; Hetterscheid, D.G.; Williams, R.M.; van der Vlugt, J.I.; Reek, J.N.; Brouwer, A.M. Platinum (ii)–porphyrin as a sensitizer for visible-light driven water oxidation in neutral phosphate buffer. Energy Environ. 2015, 8, 975–982. [Google Scholar] [CrossRef] [Green Version]
  32. Yu, S.; Zeng, Y.; Chen, J.; Yu, T.; Zhang, X.; Yang, G.; Li, Y. Intramolecular triplet–triplet energy transfer enhanced triplet–triplet annihilation upconversion with a short-lived triplet state platinum (II) terpyridyl acetylide photosensitizer. RSC Adv. 2015, 5, 70640–70648. [Google Scholar] [CrossRef]
  33. Liu, Y.T.; Li, Y.R.; Wang, X.; Bai, F.Q. Theoretical investigation of NˆCˆN-coordinated Pt (II) and Pd (II) complexes for long-lived two-photon photodynamic therapy. Dyes Pigm. 2017, 142, 55–61. [Google Scholar] [CrossRef]
  34. Li, C.; Wang, Y.; Li, C.; Xu, S.; Hou, X.; Wu, P. Simultaneously broadened visible light absorption and boosted intersystem crossing in platinum-doped graphite carbon nitride for enhanced photosensitization. ACS Appl. Mater. 2019, 11, 20770–20777. [Google Scholar] [CrossRef]
  35. Irikura, M.; Tamaki, Y.; Ishitani, O. Development of a panchromatic photosensitizer and its application to photocatalytic CO2 reduction. Chem. Sci. 2021, 12, 13888–13896. [Google Scholar] [CrossRef]
  36. Lainé, P.P.; Campagna, S.; Loiseau, F. Conformationally gated photoinduced processes within photosensitizer–acceptor dyads based on ruthenium (II) and osmium (II) polypyridyl complexes with an appended pyridinium group. Coord. Chem. Rev. 2008, 252, 2552–2571. [Google Scholar] [CrossRef]
  37. Summers, P.A.; Calladine, J.A.; Ghiotto, F.; Dawson, J.; Sun, X.Z.; Hamilton, M.L.; Towrie, M.; Davies, E.S.; McMaster, J.; George, M.W. Synthesis and photophysical study of a [NiFe] hydrogenase biomimetic compound covalently linked to a Re-diimine photosensitizer. Inorg. Chem. 2016, 55, 527–536. [Google Scholar] [CrossRef]
  38. Juris, A.; Barigelletti, F.; Balzani, V.; Belser, P.; Von Zelewsky, A. New photosensitizers of the ruthenium-polypyridine family for the water splitting reaction. Isr. J. Chem. 1982, 22, 87–90. [Google Scholar] [CrossRef]
  39. Li, H.; Li, F.; Zhang, B.; Zhou, X.; Yu, F.; Sun, L. Visible light-driven water oxidation promoted by host–guest interaction between photosensitizer and catalyst with a high quantum efficiency. J. Am. Chem. Soc. 2015, 137, 4332–4335. [Google Scholar] [CrossRef]
  40. Li, H.; Li, F.; Wang, Y.; Bai, L.; Yu, F.; Sun, L. Visible-light-driven water oxidation on a photoanode by supramolecular assembly of photosensitizer and catalyst. ChemPlusChem 2016, 81, 1056–1059. [Google Scholar] [CrossRef]
  41. Limburg, B.; Bouwman, E.; Bonnet, S. Rate and stability of photocatalytic water oxidation using [Ru(bpy)3]2+ as photosensitizer. ACS Catal. 2016, 6, 5273–5284. [Google Scholar] [CrossRef] [Green Version]
  42. Sheridan, M.V.; Sherman, B.D.; Coppo, R.L.; Wang, D.; Marquard, S.L.; Wee, K.R.; Murakami Iha, N.Y.; Meyer, T.J. Evaluation of chromophore and assembly design in light-driven water splitting with a molecular water oxidation catalyst. ACS Energy Lett. 2016, 1, 231–236. [Google Scholar] [CrossRef]
  43. Pickens, R.N.; Neyhouse, B.J.; Reed, D.T.; Ashton, S.T.; White, J.K. Visible light-activated co release and 1O2 photosensitizer formation with Ru (II), Mn (I) complexes. Inorg. Chem. 2018, 57, 11616–11625. [Google Scholar] [CrossRef]
  44. Ashford, D.L.; Glasson, C.R.; Norris, M.R.; Concepcion, J.J.; Keinan, S.; Brennaman, M.K.; Templeton, J.L.; Meyer, T.J. Controlling ground and excited state properties through ligand changes in ruthenium polypyridyl complexes. Inorg. Chem. 2014, 53, 5637–5646. [Google Scholar] [CrossRef]
  45. Wang, P.; Dong, R.; Guo, S.; Zhao, J.; Zhang, Z.M.; Lu, T.B. Improving photosensitization for photochemical CO2-to-CO conversion. Natl. Sci. Rev. 2020, 7, 1459–1467. [Google Scholar] [CrossRef]
  46. Skaisgirski, M.; Guo, X.; Wenger, O.S. Electron accumulation on naphthalene diimide photosensitized by [Ru (2,2′-Bipyridine)3]2+. Inorg. Chem. 2017, 56, 2432–2439. [Google Scholar] [CrossRef]
  47. Kuss-Petermann, M.; Orazietti, M.; Neuburger, M.; Hamm, P.; Wenger, O.S. Intramolecular light-driven accumulation of reduction equivalents by proton-coupled electron transfer. J. Am. Chem. Soc. 2017, 139, 5225–5232. [Google Scholar] [CrossRef]
  48. Aslan, J.M.; Boston, D.J.; MacDonnell, F.M. Photodriven multi-electron storage in disubstituted RuII dppz analogues. Chem.–Eur. J. 2015, 21, 17314–17323. [Google Scholar] [CrossRef]
  49. Lefebvre, J.F.; Schindler, J.; Traber, P.; Zhang, Y.; Kupfer, S.; Gräfe, S.; Baussanne, I.; Demeunynck, M.; Mouesca, J.M.; Gambarelli, S. An artificial photosynthetic system for photoaccumulation of two electrons on a fused dipyridophenazine (dppz)–pyridoquinolinone ligand. Chem. Sci. 2018, 9, 4152–4159. [Google Scholar] [CrossRef] [Green Version]
  50. Mills, I.N.; Porras, J.A.; Bernhard, S. Judicious design of cationic, cyclometalated Ir (III) complexes for photochemical energy conversion and optoelectronics. Acc. Chem. Res. 2018, 51, 352–364. [Google Scholar] [CrossRef]
  51. Luo, Y.; Maloul, S.; Schönweiz, S.; Wächtler, M.; Streb, C.; Dietzek, B. Yield—not only lifetime—of the photoinduced charge-separated state in iridium complex–polyoxometalate dyads impact their hydrogen evolution reactivity. Chem. Eur. J. 2020, 26, 8045–8052. [Google Scholar] [CrossRef]
  52. Luo, Y.; Maloul, S.; Endres, P.; Schönweiz, S.; Ritchie, C.; Wächtler, M.; Winter, A.; Schubert, U.S.; Streb, C.; Dietzek, B. Organic linkage controls the photophysical properties of covalent photosensitizer–polyoxometalate hydrogen evolution dyads. Sustain. Energy Fuels 2020, 4, 4688–4693. [Google Scholar] [CrossRef]
  53. Shon, J.H.; Teets, T.S. Potent bis-cyclometalated iridium photoreductants with β-diketiminate ancillary ligands. Inorg. Chem. 2017, 56, 15295–15303. [Google Scholar] [CrossRef] [PubMed]
  54. Shon, J.H.; Kim, D.; Rathnayake, M.D.; Sittel, S.; Weaver, J.; Teets, T.S. Photoredox catalysis on unactivated substrates with strongly reducing iridium photosensitizers. Chem. Sci. 2021, 12, 4069–4078. [Google Scholar] [CrossRef] [PubMed]
  55. Shon, J.H.; Kim, D.; Gray, T.G.; Teets, T.S. β-Diketiminate-supported iridium photosensitizers with increased excited-state reducing power. Inorg. Chem. Front. 2021, 8, 3253–3265. [Google Scholar] [CrossRef]
  56. Rohacova, J.; Ishitani, O. Rhenium (I) trinuclear rings as highly efficient redox photosensitizers for photocatalytic CO2 reduction. Chem. Sci. 2016, 7, 6728–6739. [Google Scholar] [CrossRef] [Green Version]
  57. Morimoto, T.; Ishitani, O. Modulation of the photophysical, photochemical, and electrochemical properties of Re (I) diimine complexes by interligand interactions. Acc. Chem. Res. 2017, 50, 2673–2683. [Google Scholar] [CrossRef] [Green Version]
  58. Oppelt, K.; Mosberger, M.; Ruf, J.; Fernández-Terán, R.; Probst, B.; Alberto, R.; Hamm, P. Shedding light on the molecular surface assembly at the nanoscale level: Dynamics of a Re (I) carbonyl photosensitizer with a co-adsorbed cobalt tetrapyridyl water reduction catalyst on ZrO2. J. Phys. Chem. C 2020, 124, 12502–12511. [Google Scholar] [CrossRef]
  59. Fernández-Terán, R.; Sévery, L. Living long and prosperous: Productive intraligand charge-transfer states from a rhenium (I) terpyridine photosensitizer with enhanced light absorption. Inorg. Chem. 2020, 60, 1334–1343. [Google Scholar] [CrossRef]
  60. Wenger, O.S. A bright future for photosensitizers. Nat. Chem. 2020, 12, 323–324. [Google Scholar] [CrossRef]
  61. Kobayashi, A.; Takizawa, S.Y.; Hirahara, M. Photofunctional molecular assembly for artificial photosynthesis: Beyond a simple dye sensitization strategy. Coord. Chem. Rev. 2022, 467, 214624. [Google Scholar] [CrossRef]
  62. Wenger, O.S. Photoactive complexes with earth-abundant metals. J. Am. Chem. Soc. 2018, 140, 13522–13533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Socha, A.L.; Guerinot, M.L. Mn-euvering manganese: The role of transporter gene family members in manganese uptake and mobilization in plants. Front. Plant. Sci. 2014, 5, 106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hou, H.J. Manganese-based materials inspired by photosynthesis for water-splitting. Materials 2011, 4, 1693–1704. [Google Scholar] [CrossRef]
  65. Ye, S.; Ding, C.; Chen, R.; Fan, F.; Fu, P.; Yin, H.; Wang, X.; Wang, Z.; Du, P.; Li, C. Mimicking the key functions of photosystem II in artificial photosynthesis for photoelectrocatalytic water splitting. J. Am. Chem. Soc. 2018, 140, 3250–3256. [Google Scholar] [CrossRef]
  66. Chen, C.; Chen, Y.; Yao, R.; Li, Y.; Zhang, C. Artificial Mn4Ca clusters with exchangeable solvent molecules mimicking the oxygen-evolving center in photosynthesis. Angew. Chem. Int. Ed. 2019, 131, 3979–3982. [Google Scholar] [CrossRef]
  67. Herr, P.; Kerzig, C.; Larsen, C.B.; Häussinger, D.; Wenger, O.S. Manganese (I) complexes with metal-to-ligand charge transfer luminescence and photoreactivity. Nat. Chem. 2021, 13, 956–962. [Google Scholar] [CrossRef]
  68. Wegeberg, C.; Wenger, O.S. Luminescent chromium (0) and manganese (i) complexes. Dalton Trans. 2022, 51, 1297–1302. [Google Scholar] [CrossRef]
  69. Zhang, W.; Alonso-Mori, R.; Bergmann, U.; Bressler, C.; Chollet, M.; Galler, A.; Gawelda, W.; Hadt, R.G.; Hartsock, R.W.; Kroll, T. Tracking excited-state charge and spin dynamics in iron coordination complexes. Nature 2014, 509, 345–348. [Google Scholar] [CrossRef]
  70. Auböck, G.; Chergui, M. Sub-50-fs photoinduced spin crossover in [Fe(bpy)3]2+. Nat. Chem. 2015, 7, 629–633. [Google Scholar] [CrossRef]
  71. Jay, R.M.; Eckert, S.; Vaz da Cruz, V.; Fondell, M.; Mitzner, R.; Föhlisch, A. Covalency-driven preservation of local charge densities in a metal-to-ligand charge-transfer excited iron photosensitizer. Angew. Chem. Int. Ed. 2019, 58, 10742–10746. [Google Scholar] [CrossRef] [Green Version]
  72. Huber-Gedert, M.; Nowakowski, M.; Kertmen, A.; Burkhardt, L.; Lindner, N.; Schoch, R.; Herbst-Irmer, R.; Neuba, A.; Schmitz, L.; Choi, T.K. Fundamental characterization, photophysics and photocatalysis of a base metal iron (II)-cobalt (III) dyad. Chem. Eur. J. 2021, 27, 9905–9918. [Google Scholar] [CrossRef] [PubMed]
  73. Ghobadi, T.G.U.; Ghobadi, A.; Demirtas, M.; Buyuktemiz, M.; Ozvural, K.N.; Yildiz, E.A.; Erdem, E.; Yaglioglu, H.G.; Durgun, E.; Dede, Y. Building an iron chromophore incorporating prussian blue analogue for photoelectrochemical water oxidation. Chem. Eur. J. 2021, 27, 8966–8976. [Google Scholar] [CrossRef] [PubMed]
  74. Jiang, T.; Bai, Y.; Zhang, P.; Han, Q.; Mitzi, D.B.; Therien, M. Electronic structure and photophysics of a supermolecular iron complex having a long MLCT-state lifetime and panchromatic absorption. Proc. Natl. Acad. Sci. USA 2020, 117, 20430–20437. [Google Scholar] [CrossRef] [PubMed]
  75. Zobel, J.P.; Bokareva, O.S.; Zimmer, P.; Wölper, C.; Bauer, M.; González, L. Intersystem crossing and triplet dynamics in an iron(II) N-heterocyclic carbene photosensitizer. Inorg. Chem. 2020, 59, 14666–14678. [Google Scholar] [CrossRef] [PubMed]
  76. Zimmer, P.; Müller, P.; Burkhardt, L.; Schepper, R.; Neuba, A.; Steube, J.; Dietrich, F.; Flörke, U.; Mangold, S.; Gerhards, M. N-heterocyclic carbene complexes of Iron as photosensitizers for light-Induced water reduction. Eur. J. Inorg. Chem. 2017, 2017, 1504–1509. [Google Scholar] [CrossRef]
  77. Duchanois, T.; Etienne, T.; Cebrián, C.; Liu, L.; Monari, A.; Beley, M.; Assfeld, X.; Haacke, S.; Gros, P.C. An iron-based photosensitizer with extended excited-state lifetime: Photophysical and photovoltaic properties. Eur. J. Inorg. Chem. 2015, 2015, 2469–2477. [Google Scholar] [CrossRef]
  78. Zimmer, P.; Burkhardt, L.; Friedrich, A.; Steube, J.; Neuba, A.; Schepper, R.; Müller, P.; Flörke, U.; Huber, M.; Lochbrunner, S. The connection between NHC ligand count and photophysical properties in Fe(II) photosensitizers: An experimental study. Inorg. Chem. 2018, 57, 360–373. [Google Scholar] [CrossRef]
  79. Kunnus, K.; Vacher, M.; Harlang, T.C.; Kjær, K.S.; Haldrup, K.; Biasin, E.; van Driel, T.B.; Pápai, M.; Chabera, P.; Liu, Y. Vibrational wavepacket dynamics in Fe carbene photosensitizer determined with femtosecond X-ray emission and scattering. Nat. Commun. 2020, 11, 634. [Google Scholar] [CrossRef] [Green Version]
  80. Wehry, E.; Sundararajan, S. Intersystem crossing and internal conversion from the lowest charge-transfer singlet excited state of the (2, 9-dimethyl-1, 10-phenanthroline) copper (I) cation. ChemComm 1972, 20, 1135–1136. [Google Scholar]
  81. Ahn, B.T.; McMillin, D.R. Studies of photoinduced electron transfer from bis (2,9-dimethyl-1, 10-phenanthroline) copper (I). Inorg. Chem. 1978, 17, 2253–2258. [Google Scholar] [CrossRef]
  82. Phifer, C.C.; McMillin, D.R. The basis of aryl substituent effects on charge-transfer absorption intensities. Inorg. Chem. 1986, 25, 1329–1333. [Google Scholar] [CrossRef]
  83. Eggleston, M.K.; Fanwick, P.E.; Pallenberg, A.J.; McMillin, D.R. A twist on the copper center in the crystal structure of [Cu(dnpp)2] PF6 and the charge-transfer excited state?(dnpp= 2,9-Dineopentyl-1,10-phenanthroline). Inorg. Chem. 1997, 36, 4007–4010. [Google Scholar] [CrossRef]
  84. Gandhi, B.A.; Green, O.; Burstyn, J.N. Facile oxidation-based synthesis of sterically encumbered four-coordinate bis (2, 9-di-tert-butyl-1,10-phenanthroline) copper (I) and related three-coordinate copper (I) complexes. Inorg. Chem. 2007, 46, 3816–3825. [Google Scholar] [CrossRef] [PubMed]
  85. Sandroni, M.; Kayanuma, M.; Planchat, A.; Szuwarski, N.; Blart, E.; Pellegrin, Y.; Daniel, C.; Boujtita, M.; Odobel, F. First application of the HETPHEN concept to new heteroleptic bis (diimine) copper (I) complexes as sensitizers in dye sensitized solar cells. Dalton Trans. 2013, 42, 10818–10827. [Google Scholar] [CrossRef]
  86. Housecroft, C.E.; Constable, E.C. The emergence of copper (I)-based dye sensitized solar cells. Chem. Soc. Rev. 2015, 44, 8386–8398. [Google Scholar] [CrossRef] [Green Version]
  87. Kohler, L.; Hayes, D.; Hong, J.; Carter, T.J.; Shelby, M.L.; Fransted, K.A.; Chen, L.X.; Mulfort, K.L. Synthesis, structure, ultrafast kinetics, and light-induced dynamics of CuHETPHEN chromophores. Dalton Trans. 2016, 45, 9871–9883. [Google Scholar] [CrossRef]
  88. Kohler, L.; Hadt, R.G.; Hayes, D.; Chen, L.X.; Mulfort, K.L. Synthesis, structure, and excited state kinetics of heteroleptic Cu (I) complexes with a new sterically demanding phenanthroline ligand. Dalton Trans. 2017, 46, 13088–13100. [Google Scholar] [CrossRef] [Green Version]
  89. Zerk, T.J.; Bernhardt, P.V. Redox-coupled structural changes in copper chemistry: Implications for atom transfer catalysis. Coord. Chem. Rev. 2018, 375, 173–190. [Google Scholar] [CrossRef]
  90. Levi, G.; Biasin, E.; Dohn, A.O.; Jónsson, H. On the interplay of solvent and conformational effects in simulated excited-state dynamics of a copper phenanthroline photosensitizer. Phys. Chem. Chem. Phys. 2020, 22, 748–757. [Google Scholar] [CrossRef]
  91. Zhang, Y.; Zedler, L.; Karnahl, M.; Dietzek, B. Excited-state dynamics of heteroleptic copper (I) photosensitizers and their electrochemically reduced forms containing a dipyridophenazine moiety–a spectroelectrochemical transient absorption study. Phys. Chem. Chem. Phys. 2019, 21, 10716–10725. [Google Scholar] [CrossRef]
  92. Velasco, L.; Llanos, L.; Levín, P.; Vega, A.; Yu, J.; Zhang, X.; Lemus, L.; Aravena, D.; Moonshiram, D. Structure and excited-state dynamics of dimeric copper (I) photosensitizers investigated by time-resolved X-ray and optical transient absorption spectroscopy. Phys. Chem. Chem. Phys. 2021, 23, 3656–3667. [Google Scholar] [CrossRef] [PubMed]
  93. Mara, M.W.; Fransted, K.A.; Chen, L.X. Interplays of excited state structures and dynamics in copper (I) diimine complexes: Implications and perspectives. Coord. Chem. Rev. 2015, 282, 2–18. [Google Scholar] [CrossRef]
  94. Stroscio, G.D.; Ribson, R.D.; Hadt, R.G. Quantifying entatic states in photophysical processes: Applications to copper photosensitizers. Inorg. Chem. 2019, 58, 16800–16817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Rosko, M.C.; Wells, K.A.; Hauke, C.E.; Castellano, F.N. Next generation cuprous phenanthroline MLCT photosensitizer featuring cyclohexyl substituents. Inorg. Chem. 2021, 60, 8394–8403. [Google Scholar] [CrossRef]
  96. Garakyaraghi, S.; McCusker, C.E.; Khan, S.; Koutnik, P.; Bui, A.T.; Castellano, F.N. Enhancing the visible-light absorption and excited-state properties of Cu(I) MLCT excited states. Inorg. Chem. 2018, 57, 2296–2307. [Google Scholar] [CrossRef]
  97. Takeda, H.; Kamiyama, H.; Okamoto, K.; Irimajiri, M.; Mizutani, T.; Koike, K.; Sekine, A.; Ishitani, O. Highly efficient and robust photocatalytic systems for CO2 reduction consisting of a Cu (I) photosensitizer and Mn(I) catalysts. J. Am. Chem. Soc. 2018, 140, 17241–17254. [Google Scholar] [CrossRef] [Green Version]
  98. Kim, J.; Whang, D.R.; Park, S.Y. Designing highly efficient CuI photosensitizers for photocatalytic H2 evolution from water. ChemSusChem 2017, 10, 1883–1886. [Google Scholar] [CrossRef]
  99. Heberle, M.; Tschierlei, S.; Rockstroh, N.; Ringenberg, M.; Frey, W.; Junge, H.; Beller, M.; Lochbrunner, S.; Karnahl, M. Heteroleptic Copper Photosensitizers: Why an extended π-system does not automatically lead to enhanced hydrogen production. Chem. Eur. J. 2017, 23, 312–319. [Google Scholar] [CrossRef]
  100. Giereth, R.; Frey, W.; Junge, H.; Tschierlei, S.; Karnahl, M. Copper photosensitizers containing P^ N ligands and their influence on photoactivity and stability. Chem. Eur. J. 2017, 23, 17432–17437. [Google Scholar] [CrossRef]
  101. Takeda, H.; Ohashi, K.; Sekine, A.; Ishitani, O. Photocatalytic CO2 reduction using Cu(I) photosensitizers with a Fe(II) catalyst. J. Am. Chem. Soc. 2016, 138, 4354–4357. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Heberle, M.; Wächtler, M.; Karnahl, M.; Dietzek, B. Determination of side products in the photocatalytic generation of hydrogen with copper photosensitizers by resonance Raman spectroelectrochemistry. RSC Adv. 2016, 6, 105801–105805. [Google Scholar] [CrossRef] [Green Version]
  103. Takeda, H.; Monma, Y.; Sugiyama, H.; Uekusa, H.; Ishitani, O. Development of visible-light driven Cu (I) complex photosensitizers for photocatalytic CO2 reduction. Front. Chem. 2019, 7, 418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Giereth, R.; Reim, I.; Frey, W.; Junge, H.; Tschierlei, S.; Karnahl, M. Remarkably long-lived excited states of copper photosensitizers containing an extended π-system based on an anthracene moiety. Sustain. Energy Fuels 2019, 3, 692–700. [Google Scholar] [CrossRef]
  105. Müller, C.; Schulz, M.; Obst, M.; Zedler, L.; Gräfe, S.; Kupfer, S.; Dietzek, B. Role of MLCT States in the Franck–Condon Region of Neutral, Heteroleptic Cu (I)–4 H-imidazolate Complexes: A Spectroscopic and Theoretical Study. J. Phys. Chem. A 2020, 124, 6607–6616. [Google Scholar] [CrossRef]
  106. Pann, J.; Roithmeyer, H.; Viertl, W.; Pehn, R.; Bendig, M.; Dutzler, J.; Kriesche, B.; Brüggeller, P. Phosphines in artificial photosynthesis: Considering different aspects such as chromophores, water reduction catalysts (WRCs), water oxidation catalysts (WOCs), and dyads. Sustain. Energy Fuels 2019, 3, 2926–2953. [Google Scholar] [CrossRef]
  107. Zhang, Y.; Ren, K.; Wang, L.; Wang, L.; Fan, Z. Porphyrin-based heterogeneous photocatalysts for solar energy conversion. Chin. Chem. Lett. 2021, 33, 33–60. [Google Scholar] [CrossRef]
  108. Al Mogren, M.M.; Ahmed, N.M.; Hasanein, A.A. Molecular modeling and photovoltaic applications of porphyrin-based dyes: A review. J. Saudi Chem. Soc. 2020, 24, 303–320. [Google Scholar] [CrossRef]
  109. Ramasamy, S.; Bhagavathiachari, M.; Suthanthiraraj, S.A.; Pichai, M. Mini review on the molecular engineering of photosensitizer: Current status and prospects of metal-free/porphyrin frameworks at the interface of dye-sensitized solar cells. Dyes Pigm. 2022, 203, 110380. [Google Scholar] [CrossRef]
  110. Zarrabi, N.; Poddutoori, P.K. Aluminum (III) porphyrin: A unique building block for artificial photosynthetic systems. Coord. Chem. Rev. 2021, 429, 213561. [Google Scholar] [CrossRef]
  111. Giannoudis, E.; Benazzi, E.; Karlsson, J.; Copley, G.; Panagiotakis, S.; Landrou, G.; Angaridis, P.; Nikolaou, V.; Matthaiaki, C.; Charalambidis, G. Photosensitizers for H2 evolution based on charged or neutral Zn and Sn porphyrins. Inorg. Chem. 2020, 59, 1611–1621. [Google Scholar] [CrossRef]
  112. Luo, G.F.; Biniuri, Y.; Chen, W.H.; Wang, J.; Neumann, E.; Marjault, H.B.; Nechushtai, R.; Winkler, M.; Happe, T.; Willner, I. Modelling photosynthesis with ZnII-protoporphyrin all-DNA G-quadruplex/aptamer scaffolds. Angew. Chem. Int. Ed. 2020, 132, 9248–9255. [Google Scholar] [CrossRef]
  113. Kunnus, K.; Li, L.; Titus, C.J.; Lee, S.J.; Reinhard, M.E.; Koroidov, S.; Kjær, K.S.; Hong, K.; Ledbetter, K.; Doriese, W.B. Chemical control of competing electron transfer pathways in iron tetracyano-polypyridyl photosensitizers. Chem. Sci. 2020, 11, 4360–4373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Number of publications concerning “photosensitizer” subjects since 2012. Adapted from ISI Web of Science, dated 12 July 2022.
Figure 1. Number of publications concerning “photosensitizer” subjects since 2012. Adapted from ISI Web of Science, dated 12 July 2022.
Catalysts 12 00919 g001
Figure 2. (a) The Z-scheme coupling PSI and PSII in natural photosynthesis; (b) the general scheme of the quenching pathway in photocatalytic reactions. PS: photosensitizer; A: electron acceptor; D: electron donor; Cat: catalyst. *: excited state.
Figure 2. (a) The Z-scheme coupling PSI and PSII in natural photosynthesis; (b) the general scheme of the quenching pathway in photocatalytic reactions. PS: photosensitizer; A: electron acceptor; D: electron donor; Cat: catalyst. *: excited state.
Catalysts 12 00919 g002
Figure 3. Jablonski diagram of a typical metal-based photosensitizer. Adapted with permission from ACS Energy Lett. 2019, 4, 2, 558–566. Copyright 2019 American Chemical Society.
Figure 3. Jablonski diagram of a typical metal-based photosensitizer. Adapted with permission from ACS Energy Lett. 2019, 4, 2, 558–566. Copyright 2019 American Chemical Society.
Catalysts 12 00919 g003
Figure 4. Chemical structures of the (a) [Ru(bpy)3]2+-based, (b) [Ru(phen)3]2+-based, and (c) [Ru(bpy)2(dppz)]2+-based photosensitizers and corresponding ligands in this review.
Figure 4. Chemical structures of the (a) [Ru(bpy)3]2+-based, (b) [Ru(phen)3]2+-based, and (c) [Ru(bpy)2(dppz)]2+-based photosensitizers and corresponding ligands in this review.
Catalysts 12 00919 g004
Figure 5. Chemical structures of the (a) [Ir(ppy)2(bpy)]+-based and (b) β-diketiminate-modified (NacNac) iridium photosensitizers and corresponding ligands in this review.
Figure 5. Chemical structures of the (a) [Ir(ppy)2(bpy)]+-based and (b) β-diketiminate-modified (NacNac) iridium photosensitizers and corresponding ligands in this review.
Catalysts 12 00919 g005
Figure 6. Chemical structures of the (a) Rhenium(I) trinuclear -based, (b) [Re(4,4′-PO3H2-bpy)(CO)3Cl]-based, and (c) Rhenium(I) tricarbonyl photosensitizers in this review.
Figure 6. Chemical structures of the (a) Rhenium(I) trinuclear -based, (b) [Re(4,4′-PO3H2-bpy)(CO)3Cl]-based, and (c) Rhenium(I) tricarbonyl photosensitizers in this review.
Catalysts 12 00919 g006
Figure 7. Chemical structures of the (a) [Mn(Lbi)3]+-based and (b) [Mn(Ltri)2]+-based photosensitizers and corresponding ligands in this review.
Figure 7. Chemical structures of the (a) [Mn(Lbi)3]+-based and (b) [Mn(Ltri)2]+-based photosensitizers and corresponding ligands in this review.
Catalysts 12 00919 g007
Figure 8. Chemical structures of the (a) [Fe(N^N)(CN)4]2−based and (b) [Fe(N^N^N)(R-NHC)4]2−-based photosensitizers and corresponding ligands in this review.
Figure 8. Chemical structures of the (a) [Fe(N^N)(CN)4]2−based and (b) [Fe(N^N^N)(R-NHC)4]2−-based photosensitizers and corresponding ligands in this review.
Catalysts 12 00919 g008
Figure 9. Photochemical process of copper-based photosensitizers (Cu(I): ground state, Cu(II)*: excited state).
Figure 9. Photochemical process of copper-based photosensitizers (Cu(I): ground state, Cu(II)*: excited state).
Catalysts 12 00919 g009
Figure 10. Chemical structures of the (a) Cu-based photosensitizers and (b) corresponding ligands in this review.
Figure 10. Chemical structures of the (a) Cu-based photosensitizers and (b) corresponding ligands in this review.
Catalysts 12 00919 g010
Figure 11. Chemical structures of porphyrin, M-porphyrin (M = Al, Zn, and Sn), protoporphyrin IX(PPIX), and ZnPPIX.
Figure 11. Chemical structures of porphyrin, M-porphyrin (M = Al, Zn, and Sn), protoporphyrin IX(PPIX), and ZnPPIX.
Catalysts 12 00919 g011
Table 1. Summary of the properties of the partial noble-metal-based photosensitizers in this review.
Table 1. Summary of the properties of the partial noble-metal-based photosensitizers in this review.
PhotosensitizerSolvent *E1/2
(V)
λabs
(nm)
λem
(nm)
τ avg ΦTON hTOF hReference
[Ru(bpy)3]2+0.10 M HClO41.26 a454611580 ns0.042//[24]
[Ru(4-CF3-4′-PO3H2-bpy)3]2+0.10 M HClO41.60 a461624667 ns0.0334750.24/s[24]
[Ru(4-CF3-5′-PO3H2-bpy)3]2+0.10 M HClO41.60 a468640222 ns0.0123900.19/s[24]
[Ru(4-PO3H2-bpy)3]2+0.10 M HClO41.34 a458628534 ns0.0379700.21/s[24]
[Ru(5-PO3H2-bpy)3]2+0.10 M HClO41.34 a467640147 ns0.0128550.21/s[24]
[Ru(4-CF3-bpy)3]2+0.10 M HClO41.51 a455617731 ns0.0141850.21/s[24]
[Ru(5-CF3-bpy)3]2+0.10 M HClO41.53 a464630258 ns0.02381400.24/s[24]
[Ru(phen)3]2+CH3CN1.29 b4475950.4 µs/665.5/10 h[45]
[Ru(phen)2(5-pyrenylphen)3]2+CH3CN1.36 b44759532 µs/45237.6/10 h[45]
[Ru(phen)2(3-pyrenylphen)3]2+CH3CN1.36 b39163268.2 µs/112093.3/10 h[45]
[Ru(phen)2(3-pyrenyl ethynylenephen)3]2+CH3CN1.40 b415668118.7 µs/12010/10 h[45]
[Ru(bpy)2(dppz)]2+CH3CN0.84 c,e450/////[49]
[Ru(bpy)2(oxo-dppqp)]2+CH3CN0.86 c,e417/////[49]
[Ir(ppy)2(bpy)]+CH3CN1.25 d/585//275/[50]
[Ir(ppy)2(phen)]+CH3CN1.24 d/579//195/[50]
[Ir(ppy)2(dphphen)]+CH3CN1.23 d/587////[50]
[Ir(ppy)2(NacNacCy)]+CH3CN−0.39 c//////[54]
[Ir(ppy)2(NacNac NMe2)]+CH3CN−0.26 c511 f6340.75 µs0.16//[53]
[Ir(ppy)2(NacNacMe)]+CH3CN−0.07 c460 f5950.2 µs0.053//[53]
[Ir(ppy)2(NacNacOEt)]+CH3CN0.03 c456 f5712 µs0.23//[53]
[Ir(ptz)2(NacNacNMe2)]+CH3CN−0.22 c3805760.13 µs0.025//[55]
[Ir(ptz)2(NacNacCy)]+CH3CN−0.41 c3756420.068 µs0.0032//[55]
[Ir(pmb)2(NacNacNMe2)]+CH3CN−0.26 c4305910.85 µs0.039//[55]
[Ir(pmb)2(NacNaccy)]+CH3CN−0.43 c4396750.12 µs0.0027//[55]
[{Re(4-dmbpy)(CO)22-dppe)}s3]3+DMF−1.73 g3985981.57 µs0.1219.68/[56]
[{Re(5-dmbpy)(CO)22-dppe)}3]3+DMF−1.82 g3905612.32 µs0.3614.2/[56]
[{Re(4,5-dmbpy)(CO)22-dppe)}3]3+DMF−1.91 g3845453.58 µs0.606.31/[56]
[{Re(4-OMebpy)(CO)22-dppe)}3]3+DMF−1.89 g3805437.77 µs0.662.13/[56]
Re(CN-Phtpy)(CO)3ClDMF−1.61 c3896670.582 ns///[59]
Re(CF3-Phtpy)(CO)3ClDMF−1.66 c3876660.79 ns///[59]
Re(Br-Phtpy)(CO)3ClDMF−1.70 c3826631.35 ns///[59]
Re(H-Phtpy)(CO)3ClDMF−1.73 c3806521.53 ns/580~2.5/s[59]
Re(OMe-Phtpy)(CO)3ClDMF−1.77 c3786472.26 ns///[59]
Re(NMe2-Phtpy)(CO)3ClDMF−1.81 c425532380 ns/2130~150/s[59]
* Measurements were performed on the listed solvent unless otherwise noted. a Potential versus NHE in 0.1 M HClO4. b Ferrocene (Fc) was used as an internal reference (E1/2 = + 0.40 V (Fc+/Fc) vs. SCE). c Potential vs. Fc+/0. d Potential vs. SCE. e In DMF. f In THF. g Potential vs. Ag/AgNO3. h 20 μM PSs.
Table 2. Earth abundance (in mass percent) of the metal elements mentioned in this review [62]. Adapted with permission from J. Am. Chem. Soc. 2018, 140, 42, 13522–13533. Copyright 2018 American Chemical Society.
Table 2. Earth abundance (in mass percent) of the metal elements mentioned in this review [62]. Adapted with permission from J. Am. Chem. Soc. 2018, 140, 42, 13522–13533. Copyright 2018 American Chemical Society.
ElementAbundanceElementAbundance
Ru 10 6 Mn0.091
Re 10 7 Fe4.7
Os 5 × 10 7 Ni 7.2 × 10 3
Ir 10 7 Cu0.005
Pt 10 6 Zn0.007
Table 3. Summary of the properties of the partial noble-metal-free photosensitizers in this review.
Table 3. Summary of the properties of the partial noble-metal-free photosensitizers in this review.
PhotosensitizerSolventE1/2 (V)λabs (nm)λem (nm) τ avg ΦReference
[Mn(Lbi)3]+CH3CN1.05 a3854850.74 ns0.05%[68]
[Mn(Ltri)2]+CH3CN1.00 a3955251.73 ns0.03%[68]
[Fe(bpy)(CN)4]2−CH3OH/514 b/0.22 ps/[113]
[Fe(bpyz)(CN)4]2−CH3OH/590 b/0.61 ps/[113]
[Fe(bpym)(CN)4]2−CH3OH/608 b/16.9 ps/[113]
[Fe(tpy)(dipp-NHC)]2−CH3CN0.56 c379, 503, 538///[78]
[Fe(isp-NHC)2]2−CH3CN0.43 c392, 458///[78]
[Fe(bpyz)(dipp-NHC)]2−CH3CN0.46 c376, 409, 466, 506, 538///[78]
[Cu(dsbtmp)2]+CH3CN0.428 c4456491.5 µs2.9%[95]
[Cu(dchtmp)2]+CH3CN0.43 c/6501.5 µs2.6%[95]
[Cu(sbmpep)2]+CH2Cl20.69 c4916871.4 µs2.7%[96]
[Cu(xant)(dppz)]+CH3CN0.815, 1.026 c376, 394/4380 ns/[104]
[Cu(xant)(dmdppz)]+CH3CN0.926, 1.228 c383, 400/1250 ns, 3785 ns/[104]
[Cu(xant)(dmphen)]+CH3CN0.82 c378/64 ns/[104]
[Cu(DPEphos)(bcp)]+CH3CN/3845900.99 µs2.7%[103]
[ZnTMePyP4+]Cl41:1 CH3CN/H2O1.28 d438627/66785 µs/[111]
Sn(OH)2TPyP1:1 CH3CN/H2O1.38 d417597/65078 µs/[111]
Sn(Cl2)TPP-[COOMe]41:1 CH3CN/H2O1.29 d425604/66070 µs/[111]
Sn(Cl2)TPP-[PO(OEt)2]41:1 CH3CN/H2O1.32 d424605/659//[111]
a Data obtained in dichloromethane at 20 °C, potential versus SCE in CH3CN. b Data were calculated from 1240 E v ( e v ) . c Potential versus Fc+/0 standard in CH3CN. d Potential versus Fc+/0 standard in DMF.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, L. Recent Advances in Metal-Based Molecular Photosensitizers for Artificial Photosynthesis. Catalysts 2022, 12, 919. https://doi.org/10.3390/catal12080919

AMA Style

Wang L. Recent Advances in Metal-Based Molecular Photosensitizers for Artificial Photosynthesis. Catalysts. 2022; 12(8):919. https://doi.org/10.3390/catal12080919

Chicago/Turabian Style

Wang, Lei. 2022. "Recent Advances in Metal-Based Molecular Photosensitizers for Artificial Photosynthesis" Catalysts 12, no. 8: 919. https://doi.org/10.3390/catal12080919

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop