Skip to main content

Immobilization of molecular catalysts for artificial photosynthesis

Abstract

Artificial photosynthesis offers a way of producing fuels or high-value chemicals using a limitless energy source of sunlight and abundant resources such as water, CO2, and/or O2. Inspired by the strategies in natural photosynthesis, researchers have developed a number of homogeneous molecular systems for photocatalytic, photoelectrocatalytic, and electrocatalytic artificial photosynthesis. However, their photochemical instability in homogeneous solution are hurdles for scaled application in real life. Immobilization of molecular catalysts in solid supports support provides a fine blueprint to tackle this issue. This review highlights the recent developments in (i) techniques for immobilizing molecular catalysts in solid supports and (ii) catalytic water splitting, CO2 reduction, and O2 reduction with the support-immobilized molecular catalysts. Remaining challenges for molecular catalyst-based devices for artificial photosynthesis are discussed in the end of this review.

1 Introduction

Natural photosynthesis provides over 99% of nutrients to the living organisms on earth. At the beginning of food chains, photosynthetic autotrophs, i.e. plants, algae, and bacteria, convert inorganic substances into energy-intensive food by using sunlight as an energy source. Through the course of 3.4 billion years [1] nature has selected molecular systems to carry out the photochemical processes. Molecular catalysts have abundant advantages: (i) well defined molecular structure and active center (ii) feasible mechanistic study with the help of various spectroscopic techniques, (iii) maximized active sites in homogeneous solutions, (iv) high tunability of their chemical/physical properties by rational molecular design, etc. [2]. These features inspired the development of molecular catalysts for solar energy transduction in artificial photosynthesis.

The concept of storing the solar energy into higher-energy chemicals is the basis of artificial photosynthesis, which is dedicated to producing fuels/high-value chemicals such as H2, methane, methanol, or hydrogen peroxide. A number of molecular catalysts have been developed for water splitting [3,4,5,6,7,8,9,10,11,12], CO2 reduction [13,14,15,16,17,18,19,20,21], and O2 reduction [22, 23]. However, the molecular catalysts in homogeneous solution often suffer from low stability in redox conditions. In addition, debates have been made whether the organometallic catalysts are true homogeneous catalysts or just a precursor of heterogeneous catalysis [24]. One important implication of natural photosynthetic system is that the molecular cofactors are optimally positioned in a protein matrix, so the photochemical processes are carried out in an efficient and robust way [25,26,27,28]. Although replicating the complicated configuration of natural photosystem is not that straightforward, more simple chemistry was addressed to immobilize molecular catalysts in solid matrices, i.e., covalent bonds and non-covalent interactions. This review revisits the techniques for immobilizing molecules in solid supports and summarizes recent advances in artificial photosynthesis using the techniques.

2 Immobilization methods

2.1 Covalent immobilization

Covalent bond between a catalyst and its support can result in stable immobilization. Several synthetic methods have been proposed for covalent attachment of molecular catalysts onto carbon-based surfaces [29,30,31,32]. A widely used strategy to for direct C–C bond formation was pioneered by Pinson and Savéant is based on electrochemical grafting based on reduction of aryl diazonium molecules. The diazonium salts are reduced on the surface of carbon-based electrode to form aryl radicals, which reacts with the carbon surface to form stable C–C bonds [31,32,33]. This method was successfully adopted for the preparation of quinone decorated electrodes showing electrocatalytic activities for catalytic oxygen reduction (Scheme 1a) [34,35,36,37,38,39].

An alternative approach based on electrochemical oxidation of aryl acetates was reported [40,41,42]. The detailed mechanism is more complicated compared to the diazonium pathway and still under debate [42, 43]. González group proposed one possible mechanism, where anodic oxidation of carboxylates results in removal of CO2, thus yields arylmethyl radical. The radical can be further oxidized to form methyl carbocation, which forms C–C bond with carbon surface (Scheme 1b) [42].

Scheme 1
scheme 1

Covalent attachment of a aryl diazonium salts and b aryl acetates onto carbon surface

Pre-functionalization of carbon materials provides synthetic feasibility to incorporate molecular catalysts. Oxidation of carbon surface provides oxygen-containing groups, such as ketone, quinone, phenol, lactone, carboxyl group, ether, acid anhydride, ketene, etc. [44], which can serve as anchoring sites for subsequent functionalization with molecular catalysts (Fig. 1a). More recently, fluorinated graphene (FG) has been shown to be susceptible to nucleophilic substitution to form sulfhydryl [45], amino [46,47,48,49], dichlorocarbene [50], or urea [51] groups for further functionalization (Fig. 1b). Regarding the chemical structure of FG, it has been considered as chemically inert material as an analogue to polytetrafluoroethylene (Teflon®) [52]. However, after Zbořil group reported reductive defluorination of FG [53], a number of methods have been developed for modification of FGs [54].

Fig. 1
figure 1

a Oxidative functionalization of carbon surfaces. Reproduced from [44] with copyright permission from 1993 Elsevier Ltd. b Preparation of graphene derivatives from fluorographene. Reproduced from [54] with copyright permission from 2017 Elsevier Ltd.

Another example is the CuI-catalyzed azide-alkyne cycloaddition (CuAAC) “click” reaction. The click reaction usually uses mild reaction condition and gives high conversion yield. Azide-modification of graphitic carbon surface can be simply carried out by either immersing them in acetonitrile solution of IN3 or exposing them to IN3 gas [55,56,57]. Further modification of carbon surface then can be carried out with CuAAC reaction (Scheme 2).

Scheme 2
scheme 2

Modification of graphitic carbon with azides and coupling to terminal alkynes

Metal-organic frameworks (MOFs) are emerging platforms for immobilization of molecular catalysts. Development of synthetic strategies of metal-organic frameworks (MOFs) provides rational design of molecular catalysts embedded in them. Two major constitutional components of MOF are metal ions/clusters and linking ligands, providing strategies to realize MOF catalysis based on (i) metal ions/clusters as catalytic sites or (ii) modification of organic ligands with molecular catalysts, respectively [58].

2.2 Non-covalent immobilization

Non-covalent immobilization of molecular catalysts allows a facile mix-and-go strategy for catalyst preparation [59, 60]. A figure of merit for non-covalent interaction over covalent bonding is that due to the rather weak electronic interplay, the physical/chemical properties of both anchoring molecules and supporting matrices can be maintained. A number of non-covalent functionalization strategies, using π-π interaction, hydrogen bonding, electrostatic interaction, physical encapsulation, and mechanical interlocking, etc., have been reported so far.

Extended aromatic moieties such as pyrenyl group has been widely explored for immobilization of molecules on graphitic carbons based on the strong π-π interaction between them (Fig. 2a) [61,62,63,64,65]. Especially, a Diels-Alder adduct of pentacene has optimal dimension with the curvature of single-wall carbon nanotubes (CNTs), thus forms tweezer-like binding (Fig. 2b) [66]. The size-specific interaction enabled separation of CNTs mixture in 1–2 nm diameter range. For graphene oxide (GO), additional hydrogen-bond with epoxides, alcohols, ethers, and carboxylic acids can be formed in addition to the π–π interaction. The strong interaction of aromatic moieties with graphitic carbons was further evidenced by microscopic techniques [67,68,69,70,71].

Fig. 2
figure 2

Schematic representation of a pyrenyl moiety sitting on the basal plane of graphene (a) and a Diels-Alder adduct of pentacene tweezing a single-walled carbon nanotube (b) through π–π interaction

Electrostatic interaction can also play a role to immobilize molecular catalysts. Oxidized CNTs or graphenes can afford strong ionic conjugates. Liang et al. demonstrated functionalization of GO and reduced graphene oxide (rGO) with ammonium salts [72]. Electrostatic functionalization of carbon support can be further expanded to end-functional polymers. Kim group reported stable dispersion of rGO in organic solvents by incorporating amine-terminated polystyrene [73]. Hydrophobicity/hydrophilicity of a molecule-support hybrid system can be effectively tuned by this technique, thus opens wider range of solvent selection.

3 Photocatalysis

Photocatalysis is a process where light energy is directly converted into chemical energy. It starts with absorption of photons by photosensitizers, followed by cascade charge/energy transfer to catalysts which convert substrates into products. Artificial photosynthesis schemes are multi-electronic processes, which implies multiple charges should be directed and accumulated at the catalytic center to run the photochemical cycles (Scheme 3a). It is a challenging subject, since photon absorption is generally a monoelectronic process, results in single charge production. In this context, it is required to accumulate electron/holes by multiple rounds of single electronic absorption before exciton decay, charge recombination, or back charge transfer. The best example is photosystem II (PSII) in chlorophyll, which is responsible for O2 evolution in natural photosynthesis. The active units in PSII are optimally positioned in a protein matrix so the charge transfer reactions are kinetically controlled to extract four electrons from a Mn4Ca cluster to run water oxidation (Scheme 3b) [74]. It is not straightforward to directly replicate the complex protein matrix of nature, however, efforts have been made to borrow the strategies in more simple chemical bonds or supramolecular chemistry [75,76,77,78,79].

Scheme 3
scheme 3

a Electron equivalents of various artificial photosynthetic processes. b Arrangement of subunits in PSII. Copyright permission from Elsevier [74]

3.1 MOFs as immobilizing matrices

Many MOFs bearing aromatic ligands have broad UV–vis absorption assigned to metal-to-ligand charge transfer (MLCT), ligand-to-metal charge transfer (LMCT), or ligand-centered (LC) π–π* transition. This feature inspired researchers to use them as immobilizing matrices for molecular photosensitizers/catalyst for various photocatalytic reactions such as degradation [80,81,82,83,84,85] or transformation [86,87,88,89] of organic molecules, and organic photocatalysis.

Ni-based MOFs are reported as visible-light driven photocatalysts for CO2 reduction. Simple ligand modification for PCN-222 and PCN-601 resulted in major products of formate anion and CH4, respectively [90, 91]. The result implies the importance of ligand selection, which affects the charge separation/transfer kinetics and the reaction sphere morphology. Li group introduced amino groups into the bridging ligands, resulting in increased catalytic activities for CO2 reduction compared to the MOFs without amino groups [92, 93]. Systematic studies on Ti-based MOFs, namely MIL-125(Ti) for amine-free MOF and NH2-MIL-125(Ti) for amino-functionalized MOF, showed the amine moiety on the ligands enhances light absorption in the visible range and increases adsorption of CO2 (Fig. 3).

Fig. 3
figure 3

UV–vis absorption (left) and CO2 adsorption isotherms (1 atm, 273 K, right) of MIL-125(Ti) and NH2-MIL-125(Ti). The inset photo shows absorption colors. Copyright permission from Wiley-VCH [93]

Photocatalytic activities can be further enhanced by incorporating photosensitizers and/or catalysts into the MOFs. The UiO MOFs have been extensively studied as immobilizing matrix for molecular catalysts due to their facile ligand tunability and high chemical/photochemical stabilities. The UiO family has a configuration where Zr6(OH)4O412+ clusters are 12-connected by organic dicarboxylate linkers to form a MOF framework with fcc structure. Various organometallic complexes can replace the linking ligands by filling the well-defined cavities while maintaining the structure of parent UiO MOF.

Lin group reported a UiO-67 (Zr6O4(OH)4(bpdc)6, where bpdc is p-biphenyldicarboxylic acid) based MOF containing catalytically active Ir(III), Re(I), or Ru(II) complexes [94]. The MOFs were synthesized through mix-and-match strategy, owing to the matching ligand lengths between bpdc and metal-functionalized ligands (Fig. 4). MOFs 13 containing Cp*Ir(III)(dcppy)Cl (where Cp* is pentamethylcyclopentadienyl and dcppy is 2-phenylpyridine-5,4′-dicarboxylic acid) or Ir(III)(dcppy)2(H2O)2 were effective water oxidation catalyst with tyrnover frequencies up to 4.8 h− 1. The Re(I)(CO)3(dcbpy)Cl (H2L4, where dcbpy is 2,2′-bipyridyl-5,5′-dicarboxylic acid) derivatized MOF 4 served as an active catalyst for photocatalytic CO2 reduction with a total turnover number of 10.9, three times higher than that of the homogeneous complex H2L4.

Fig. 4
figure 4

a Synthesis of organometallic catalysts doped UiO-67. b Structure of L1-doped UiO-67. The inset photo shows absorption colors. Copyright permission from ACS. [94]

Kim et al. selected a bipyridine-embedded UiO-67 MOF, namely BUiO, as a self-healing platform for molecular Pt(II) catalysts and Ir(III) photosensitizers. A series of BUiO-based MOFs, namely, Ptn_Ir_BUiO, comprising 2,2′-bipyridine-5,5′-dicarboxylate (L) as a self-healing site, Pt(II)(L)Cl2 as a H2-evolving catalyst, and Ir(III)(ppy)2(L) as a photosensitizer were synthesized and tested for photocatalytic H2 production [95]. When the metal-diimine bonds were cleaved during photolysis, abundant free diamine ligands promoted re-coordination of the cleaved metal ions to recover the molecular catalytic activity (Fig. 5a). Accordingly, the catalysis with Pt0.1_Ir_BUiO prolonged more than 6.5 days without significant decrease in its activity, while the control MOF without a free diimine ligand lost molecular catalytic activity after 7.5 h by leaching of Pt and forming colloids (Fig. 5b).

Fig. 5
figure 5

a Schematic operation principle of self-healing MOF (Ptn_Ir_BUiO). b H2 evolution rate curve of Pt0.1_Ir_BUiO. Inset: total amount of H2 evolved (left) and HR-TEM image after 6.5 days of photolysis. Scale bar = 200 nm. Copyright permission from ACS [95]

3.2 Dye-sensitized inorganic/organometallic hybrids

In inorganic photocatalysis, metal oxides are typically studied due to their stability and facile synthesis. More importantly, multiple charges can be stored in a nanocrystal, which can be further utilized for multi-electronic photocatalytic cycles. Depending on the size of a nanocrystal, different maximum number of charges can be stored. Mayer and co-workers have shown a ZnO nanocrystal with 3.5 nm radius can take up to 120 electrons per particle [96]. Most of them work well with UV light and only a few visible-light-responsive photocatalytic systems are reported so far [97, 98]. Non-oxide nanocrystals such as CdS or CdSe have suitable bandgaps with strong visible-light absorption, which is suitable for artificial photosynthesis. However, the toxicity of Cd limits their usage in real life. As an analogue to dye-sensitized solar cells, hybridizing inorganic nanocrystals with organic/organometallic dyes offers visible-light driven photocatalysis. Scheme 4 illustrates electronic processes in a Ru(II)–TiO2/Pt system for photocatalytic H2 production: photoexcitation of a Ru(II) photosensitizer (RuIILx), followed by (1) charge transfer from the photoexcited states of the Ru(II) dye (RuII*Lx) to TiO2 nanocrystal, (ii) back electron transfer from TiO2 conduction band to the dye, (iii) electron transfer to Pt catalyst, (iv) water reduction to produce H2, and (v) regeneration of the RuIILx ground state by an electron donor (D).

Scheme 4
scheme 4

Visible-light-driven H2 production with Ru(II) dye-sensitized TiO2. Copyright permission from ACS [99]

Kaneko and co-workers reported sensitization of TiO2/Pt nanoparticles with Ru(II) dyes for visible-light driven H2 evolution from water [100]. The study shows the importance of anchoring groups: Ru(II)(bpym)32+ (where bpym is bipyrimidine) showed much higher H2 evolution rate (80.1 µl h− 1) than Ru(dcbpy)34− (17.8 µl h− 1)or Ru(bpy)32+ (23.3 µl h− 1). Further investigation has been made on Ru(II)-sensitized TiO2/Pt to find structure-catalytic activity relationships for photocatalytic H2 production from water [99, 101,102,103].

Noble metal-free dye-sensitized systems have been also investigated. Yuan et al.. reported a noble metal-free H2 production system with Zn porphyrin sensitized MoS2/ZnO [104]. The 0.50 wt% ZnTCPP–MoS2/ZnO (where TCPP is 5,10,15,20-tetrakis(4-carboxyphenyl)-porphyrin) photocatalyst showed H2 evolution rate of 75 µmol h− 1 g− 1, which is comparable with the Ru(II)–TiO2/Pt photocatalysts. Li and co-workers reported a 5,10,15,20-tetraphenylporphyrin–Cu2O photocatalyst with H2 evolving rate of 1.3 mmol g− 1.

3.3 Modification of carbon materials

Yang and co-workers reported a 5,10,15,20-tetrakis(4-(hydroxyl)phenyl) porphyrin (TPPH) non-covalently adsorbed on rGO with Pt nanoparticles [105]. Strong π–π interaction between TPPH and rGO offers stable photocatalysis (Scheme 5a). The TPPH–RGO/Pt catalyst showed an effective photocatalytic activity with a H2 production rate of 5.29 mmol g− 1. Li and co-workers reported Cr3+ implemented GOs bridged by 5,15-diphenyl-10,20-di(4-pyridyl)porphyrin (DPyP) [106]. Instead of forming π–π stacking between DPyP and GO, DPyP molecules are oriented to the out-of-plane direction of GO sheets by means of electrostatic and coordination interactions (I and II in Scheme 5b). The DPyP–Cr3+–GO showed visible-light-driven H2 production with activities of 928 µmol g− 1, while the control DPyP–GO (III in Scheme 5b) showed that of 686 µmol g− 1.

Scheme 5
scheme 5

a Illustration of non-covalent incorporation of TPPH on the basal plane of rGO surfaces through Strong π-π interaction. b Formation of DPyP–Cr3+–GO (I and II) and DPyP–GO. Copyright permissions from ACS [105, 106]

4 Dye-sensitized photoelectrochemical (DS-PEC) cells

In the early 1970s, Honda and Fujishima reported a photoelectrochemical (PEC) water splitting cell comprising TiO2 photoanode and Pt counter electrode (Fig. 6) [107, 108]. By illumination of the photoanode (λ < 415 nm), O2 and H2 were generated at the photoanode and counter electrode, respectively. Since the pioneering works by Honda and Fujishima, significant progress has been made in pursuit of PEC water splitting systems using TiO2 photoanode or NiO photocathode. However, the large bandgaps of TiO2 and NiO limits the PEC cells operating with visible light. To tackle this issue, visible-light-absorbing semiconductors (VLA-SCs) have been studied, however, the catalytic activities of the VLA-SCs are rather lower than those of TiO2 or NiO [109, 110].

Fig. 6
figure 6

Illustration of a PEC cell reported by Honda and Fujishima: 1 is a TiO2 photoanode, 2 is a Pt counter electrode, 3 is a sintered glass diaphragm, 4 is an external load, and 5 is a voltameter. Copyright permission from The Chemical Society of Japan [107]

One promising strategy is sensitizing the large bandgap inorganic semiconductors with molecular photosensitizers. The concept of dye-sensitized solar cell (DSSC) was first proposed by O’Regan and Grätzel in 1991 [111]. Ru(II) complexes were adsorbed on the surface of TiO2 nanocrystals, which facilitates visible-light absorption by the Ru(II) dyes, followed by electrons from the photoexcited states injected into the TiO2 nanocrystals. The DSSC was first designed to generate electricity; same strategy can be applied to DS-PEC cells for catalytic reactions. Figure 5 depicts the working principle of a water splitting DS-PEC with dye-sensitized photoanode: (i) light absorption by dyes anchored to n-type semiconductor nanoparticles, (ii) injection of photoexcited electrons into the conduction band of semiconductor, (iii) extraction of electrons from water oxidation catalysts (WOCs), (iv) flow of electrons to counter electrode of Pt through outer circuit, and (v) catalysis on the surface of each electordes with supplied holes/electrons (Scheme 6). As an analogue to the dye-sensitized photoanode, dye-sensitized photocathode is also feasible, by incorporating dyes with p-type semiconductor nanoparticles.

Scheme 6
scheme 6

Schematic working principles of a water splitting DS-PEC with dye-sensitized DS-PEC cell. Copyright permission from RSC [112]

Mallouk and co-workers first reported a visible-light-driven water splitting DS-PEC cell [113]. Ru(II) dyes bridge TiO2 nanoparticles and IrO2·nH2O nanoparticles, directing cascade electron/hole transfer between them (Fig. 7a). Low internal quantum yield of 0.9% was reported due to: (i) slow electron transfer between IrO2·nH2O and oxidized Ru(II) dye and (ii) back electron transfer from TiO2 to the dye. To overcome the inefficient charge interplays, an electron mediator was incorporated to the photoanode, resulting in an increased internal quantum efficiency to 2.3%. Inspired by the works by Maooluk group, a number of dye-sensitized photoanodes for water splitting DS-PEC cells have been reported [114,115,116,117,118,119,120,121,122,123,124,125,126,127].

Reductive catalysis such as CO2 reduction or O2 reduction requires development of photocathodes with proper catalysts assembly. Inoue and co-workers reported a reduction of CO2 with dye-sensitized photocathode where a dinuclear Zn(II)–Re(I) complex is anchored to the surface of NiO nanoparticles (Fig. 7b) [128]. Although efficient hole injection efficiency from the dye to the NiO was observed, rather low Faradaic efficiency of 6.2% and turnover number of 10 were recorded due to decomposition of the sensitizing Zn(II) porphyrin part. With the same strategy, Ishitani group reported a dye-sensitized photocathode where Ru(II)–Re(I) dyads are anchored to NiO nanoparticles [129]. Enhanced Faradaic efficiency of 71% and turnover number of 32 were reported with the DS-PEC cell. Further improvement of Faradaic efficiency to 93% was achieved by replacing the NiO nanoparticles to CuGaO2 [130].

Fig. 7
figure 7

DS-PEC cells with a a dye-sensitized photoanode for water splitting and b a dye-sensitized photocathode for CO2 reduction. Copyright permissions from ACS and Elsevier [113, 128]

5 Electrocatalysis

In a broader context, electrocatalytic fuel producing systems wired to external power sources, e.g. photovoltaic cells, is also categorized as artificial photosynthesis. Immobilization of molecular catalysts on electrodes requires few requirements for successful electrocatalysis: (i) charge transfer between electrode surface and immobilizing matrix shoud be assured, (ii) the immobilizing matrix should act as a charge transfer channel, and (iii) substrates and products can freely diffuse in the bulk of immobilizing matrix. By fulfilling these requirements, a number of electrocatalytic systems for water splitting, CO2 reduction, or O2 reduction have been reported.

A number of studies have been made in the development of molecular catalysts for homogeneous water splitting and they are well reviewed previously [21, 131, 132]. The huge library of water splitting catalysts provides facile way to directly immobilize them on the electrode surfaces for electrocatalytic systems. Garrido-Barros et al. demonstrated electrocatalytic water oxidation with Cu(II) complexes strongly bound on the basal plane of graphene sheets (Fig. 8a) [133]. Comparative study with the graphene-immobilized system (G-22−) and a homogeneous system (22−) as a control showed remarkable increase in catalytic activity: G-22− and 22− showed maximum turnover frequencies of 540 s− 1 and 128 s− 1, respectively.

CO2 reduction in protic media inevitably competes with proton production, resulting in production of H2 as a byproduct. Interestingly, immobilization of electrocatalysts in solid matrices often enhances their selectivity to specific products by suppressing proton reduction. Wang and co-workers demonstrated highly selective CO2 electroreduction over H2 evolution with a Cu(I) complex adsorbed on graphene electrode [134]. It was proposed that the mesostructure of graphene favors diffusion of CO2 and limits mass transport of protons. More recently, Zhu and co-workers reported electrocatalytic CO2 reduction by immobilizing pyrrolidinonyl Ni(II) phthalocyanine (PyNiPc) on CNTs (Fig. 8b) [135]. The PyNiPc/CNT catalyst suppressed H2 evolution and selectively promoted CO2 reduction; a high CO/H2 value of 650 was recorded.

Fig. 8
figure 8

Adsorption of molecular catalysts on a the basal plane of graphene and b CNTs. Copyright permissions from Elsevier and ACS [133, 135]

Anthraquinone and its derivatives and its derivatives have been explored as molecular electrocatalysts for O2 reduction, since anthraquinone is low-cost, has suitable chemical functionality, and high selectivity for H2O2 [136,137,138,139]. Attempts to immobilize anthraquinones on electrodes comprise: (i) directly linking to glassy carbon electrodes through covalent bonds [140], (ii) incorporating in insoluble polymer matrix [141], (iii) and covalently linking to CNTs [142]. Wielend et al.. demonstrated a new approach of “mechanically interlocking” anthraquinone-based catalysts around CNTs [143] The rotaxane architecture prevents dissolution problems of physically adsorbed organic molecules upon electrochemical reduction, while retains the electrochemical properties of the pristine molecule (Fig. 9).

Fig. 9
figure 9

Electrocatalytic O2 reduction with a mechanically interlocked CNT with an anthraquinone macrocycle. Copyright permission from ACS [143]

6 Summary and outlook

The purpose of this review has been to understand the techniques to immobilize molecular catalysts in matrix-supports and incorporating them into artificial photosynthesis. Enhancement of catalytic activity and robustness was achieved with a simple chemistry using covalent bonds or non-covalent interactions. In addition, selectivity for specific product can be also enhanced by anchoring them in well-defined mesoporous structure. Synergetic effects were shown by hybridizing complementary molecular catalysts and solid supports.

The results provide a good blueprint for artificial photosynthesis; the researches so far are in a lab-scale and require next step forward for practical application of artificial photosynthesis in real life. Especially, proper device design is required to address the subjects listed below.

6.1 Scalability

The global production of H2 reached 790 Mton in 2018 (https://www.iea.org) and the H2O2 market valued 2.49 billion USD in 2019, (https://www.grandviewresearch.com). H2 and H2O2 are exclusively produced by natural gas reforming and anthraquinone process, respectively. Both of the methods require high energy input; renewable method, such as artificial photosynthesis, is desired to substitute the production. In this regard, scaled artificial photosynthesis with proper device design is required to comply with the large numbers. Atmospheric CO2 concentration of 420 ppm is generally not high enough to directly run the catalytic reactions. The low CO2 solubility in water (ca. 1.5 g L− 1 at the standard condition) also limits the scalability of CO2 reduction. Techniques such as gas-phase electrolysis coupled with flow cells are under development to circumvent the issue.

6.2 Products separation

Artificial photosynthesis comprises two half-reactions, resulting in at least two different products. In terms of monetary expenses, simple and cost-effective separation methods are required to separate a desired product from the admixture of reagents, substrates, and crude products. In water splitting, H2 and O2 are produced in a same phase of gas. Gas separation membranes are widely used to selectively collect H2 from the water splitting products. CO2 reduction is much more complicated due to a variety of possible products, vide supra. The gaseous mixture CO and H2 bubbles out from the reaction solution; it can be further separated or directly used as a syngas. The liquid phase separation requires rather high energy and complicated techniques.

Artificial photosynthesis is a promising, at the same time, a challenging subject. Multidisciplinary collaboration in chemistry, material science, biology, physics is required for successful realization of artificial photosynthesis.

Availability of data and materials

Not applicable.

Abbreviations

CNT:

Carbon nanotube

CuAAC:

CuI-catalyzed azide-alkyne cycloaddition

DS-PEC:

Dye-sensitized photoelectrochemical

DSSC:

Dye-sensitized solar cell

FG:

Fluorinated graphene

GO:

Graphene oxide

LC:

Ligand-centered

LMCT:

Ligand-to-metal charge transfer

MLCT:

Metal-to-ligand charge transfer

MOF:

Metal-organic framework

PEC:

Photoelectrochemical

PSIIPSII:

Photosystem II

rGO:

Reduced graphene oxide

VLA-SC:

Visible-light-absorbing semiconductor

WOC:

Water oxidation catalyst

References

  1. T. Cardona, Heliyon 4, e00548 (2018)

    Article  Google Scholar 

  2. L.H. Gade, P. Hofmann, Molecular Catalysts (Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2014)

    Book  Google Scholar 

  3. T. Liu, B. Zhang, L. Sun, Chem. - An Asian J. 14, 31 (2019)

    Article  CAS  Google Scholar 

  4. P. Du, R. Eisenberg, Energy Environ. Sci. 5, 6012 (2012)

    Article  CAS  Google Scholar 

  5. M.D. Kärkäs, O. Verho, E.V. Johnston, B. Åkermark, Chem. Rev. 114, 11863 (2014)

    Article  CAS  Google Scholar 

  6. K.S. Joya, H.J.M. De Groot, Int. J. Hydrogen Energy 37, 8787 (2012)

    Article  CAS  Google Scholar 

  7. N. Cox, W. Lubitz, in Chem. Energy Storage (DE GRUYTER, 2013)

  8. J.L. Dempsey, J.R. Winkler, H.B. Gray, in Compr. Inorg. Chem. II (Second Ed. From Elem. to Appl. (Elsevier Ltd, 2013), pp. 553–565

  9. T. Stoll, C.E. Castillo, M. Kayanuma, M. Sandroni, C. Daniel, F. Odobel, J. Fortage, M.N. Collomb, Coord. Chem. Rev. 304–305, 20 (2015)

    Article  CAS  Google Scholar 

  10. L. Duan, L. Wang, F. Li, F. Li, L. Sun, Acc. Chem. Res. 48, 2084 (2015)

    Article  CAS  Google Scholar 

  11. P. Garrido-Barros, C. Gimbert-Suriñach, R. Matheu, X. Sala, A. Llobet, Chem. Soc. Rev. 46, 6088 (2017)

    Article  CAS  Google Scholar 

  12. S.M. Lauinger, Q. Yin, Y.V. Geletii, C.L. Hill, in Adv. Inorg. Chem. (Academic Press Inc., 2017), pp. 117–154

  13. K. Elouarzaki, V. Kannan, V. Jose, H.S. Sabharwal, J. Lee, Adv. Energy Mater. 9, 1900090 (2019)

    Article  CAS  Google Scholar 

  14. C.-F. Leung, P.-Y. Ho, Catalysts 9, 760 (2019)

    Article  CAS  Google Scholar 

  15. P. Gotico, Z. Halime, A. Aukauloo, Dalt. Trans. 49, 2381 (2020)

    Article  CAS  Google Scholar 

  16. F. Wang, ChemSusChem 10, 4393 (2017)

    Article  CAS  Google Scholar 

  17. M.R. Dubois, D.L. Dubois, Acc. Chem. Res. 42, 1974 (2009)

    Article  CAS  Google Scholar 

  18. C. Finn, S. Schnittger, L.J. Yellowlees, J.B. Love, Chem. Commun. 48, 1392 (2012)

    Article  CAS  Google Scholar 

  19. N. Elgrishi, M.B. Chambers, X. Wang, M. Fontecave, Chem. Soc. Rev. 46, 761 (2017)

    Article  CAS  Google Scholar 

  20. H. Takeda, C. Cometto, O. Ishitani, M. Robert, ACS Catal. 7, 70 (2017)

    Article  CAS  Google Scholar 

  21. J.W. Wang, W.J. Liu, D.C. Zhong, T.B. Lu, Coord. Chem. Rev. 378, 237 (2019)

    Article  CAS  Google Scholar 

  22. C.W. Machan, ACS Catal. 10, 2640 (2020)

    Article  CAS  Google Scholar 

  23. E. Anxolabéhère-Mallart, F. Banse, Curr. Opin. Electrochem. 15, 118 (2019)

    Article  CAS  Google Scholar 

  24. V. Artero, M. Fontecave, Chem. Soc. Rev. 42, 2338 (2013)

    Article  CAS  Google Scholar 

  25. P. Jordan, P. Fromme, H.T. Witt, O. Klukas, W. Saenger, N. Krauß, Nature 411, 909 (2001)

    Article  CAS  Google Scholar 

  26. A. Amunts, H. Toporik, A. Borovikova, N. Nelson, J. Biol. Chem. 285, 3478 (2010)

    Article  CAS  Google Scholar 

  27. S. Järvi, M. Suorsa, E.M. Aro, Biochim. Biophys. Acta - Bioenerg. 1847, 900 (2015)

    Article  CAS  Google Scholar 

  28. E.J. Boekema, H. Van Roon, F. Calkoen, R. Bassi, J.P. Dekker, Biochemistry 38, 2233 (1999)

    Article  CAS  Google Scholar 

  29. C.P. Andrieux, F. Gonzalez, J.M. Savéant, J. Am. Chem. Soc. 119, 4292 (1997)

    Article  CAS  Google Scholar 

  30. Y.C. Liu, R.L. McCreery, J. Am. Chem. Soc. 117, 11254 (1995)

    Article  CAS  Google Scholar 

  31. P. Allongue, M. Delamar, B. Desbat, O. Fagebaume, R. Hitmi, J. Pinson, J.M. Savéant, J. Am. Chem. Soc. 119, 201 (1997)

    Article  CAS  Google Scholar 

  32. M. Delamar, R. Hitmi, J. Pinson, J. Savéant, J. Am. Chem. Soc. 114, 5883 (1992)

    Article  CAS  Google Scholar 

  33. J. Pinson, F. Podvorica, Chem. Soc. Rev. 34, 429 (2005)

    Article  CAS  Google Scholar 

  34. A. Sarapuu, K. Helstein, D.J. Schififrin, K. Tammeveski, Electrochem. Solid-State Lett. 8, 22 (2005)

    Article  CAS  Google Scholar 

  35. F. Mirkhalaf, K. Tammeveski, D.J. Schiffrin, Phys. Chem. Chem. Phys. 13, 21 (2004)

    Google Scholar 

  36. K. Vaik, A. Sarapuu, K. Tammeveski, F. Mirkhalaf, D.J. Schiffrin, J. Electroanal. Chem. 564, 159 (2004)

    Article  CAS  Google Scholar 

  37. K. Vaik, D.J. Schiffrin, K. Tammeveski, Electrochem. Commun. 6, 1 (2004)

    Article  CAS  Google Scholar 

  38. A. Sarapuu, K. Vaik, D.J. Schiffrin, K. Tammeveski, J. Electroanal. Chem. 541, 23 (2003)

    Article  CAS  Google Scholar 

  39. K. Tammeveski, K. Kontturi, R.J. Nichols, R.J. Potter, D.J. Schiffrin, J. Electroanal. Chem. 515, 101 (2001)

    Article  CAS  Google Scholar 

  40. C.P. Andrieux, F. Gonzalez, J.M. Savéant, J. Electroanal. Chem. 498, 171 (2001)

    Article  CAS  Google Scholar 

  41. E. Coulon, J. Pinson, J.D. Bourzat, A. Commerçon, J.P. Pulicani, Langmuir 17, 7102 (2001)

    Article  CAS  Google Scholar 

  42. M. Galicia, F.J. González, J. Electrochem. Soc. 149, D46 (2002)

    Article  CAS  Google Scholar 

  43. F. Geneste, C. Maël, C. Moinet, G. Jezequel, New J. Chem. 26, 1261 (2002)

    Article  CAS  Google Scholar 

  44. P.E. Fanning, M.A. Vannice, Carbon N. Y. 31, 721 (1993)

    Article  CAS  Google Scholar 

  45. V. Urbanová, K. Holá, A.B. Bourlinos, K. Čépe, A. Ambrosi, A.H. Loo, M. Pumera, F. Karlický, M. Otyepka, R. Zbořil, Adv. Mater. 27, 2305 (2015)

    Article  CAS  Google Scholar 

  46. B. Li, T. He, Z. Wang, Z. Cheng, Y. Liu, T. Chen, W. Lai, X. Wang, X. Liu, Phys. Chem. Chem. Phys. 18, 17495 (2016)

    Article  CAS  Google Scholar 

  47. C. Bosch-Navarro, M. Walker, N.R. Wilson, J.P. Rourke, J. Mater. Chem. C 3, 7627 (2015)

    Article  CAS  Google Scholar 

  48. K.E. Whitener, R. Stine, J.T. Robinson, P.E. Sheehan, J. Phys. Chem. C 119, 10507 (2015)

    Article  CAS  Google Scholar 

  49. R. Stine, J.W. Ciszek, D.E. Barlow, W.K. Lee, J.T. Robinson, P.E. Sheehan, Langmuir 28, 7957 (2012)

    Article  CAS  Google Scholar 

  50. P. Lazar, C.K. Chua, K. Holá, R. Zbořil, M. Otyepka, M. Pumera, Small 11, 3790 (2015)

    Article  CAS  Google Scholar 

  51. X. Ye, L. Ma, Z. Yang, J. Wang, H. Wang, S. Yang, ACS Appl. Mater. Interfaces 8, 7483 (2016)

    Article  CAS  Google Scholar 

  52. R.R. Nair, W. Ren, R. Jalil, I. Riaz, V.G. Kravets, L. Britnell, P. Blake, F. Schedin, A.S. Mayorov, S. Yuan, M.I. Katsnelson, H.M. Cheng, W. Strupinski, L.G. Bulusheva, A.V. Okotrub, I.V. Grigorieva, A.N. Grigorenko, K.S. Novoselov, A.K. Geim, Small 6, 2877 (2010)

    Article  CAS  Google Scholar 

  53. A.B. Bourlinos, K. Safarova, K. Siskova, R. Zbořil, Carbon N. Y. 50, 1425 (2012)

    Article  CAS  Google Scholar 

  54. D.D. Chronopoulos, A. Bakandritsos, M. Pykal, R. Zbořil, M. Otyepka, Appl. Mater. Today 9, 60 (2017)

    Article  Google Scholar 

  55. E.D. Stenehjem, V.R. Ziatdinov, T.D.P. Stack, C.E.D. Chidsey, J. Am. Chem. Soc. 135, 1110 (2013)

    Article  CAS  Google Scholar 

  56. M.A. Pellow, T.D.P. Stack, C.E.D. Chidsey, Langmuir 29, 5383 (2013)

    Article  CAS  Google Scholar 

  57. A. Devadoss, C.E.D. Chidsey, J. Am. Chem. Soc. 129, 5370 (2007)

    Article  CAS  Google Scholar 

  58. C. De Wu, M. Zhao, Adv. Mater. 29, 87 (2017)

    Google Scholar 

  59. C. Janiak, J. Chem. Soc. Dalt. Trans. 38, 85 (2000)

    Google Scholar 

  60. M.R. Axet, J. Durand, M. Gouygou, P. Serp, in Adv. Organomet. Chem. (Academic Press Inc., 2019), pp. 53–174

  61. H. Jaegfeldt, T. Kuwana, G. Johansson, J. Am. Chem. Soc. 105, 1805 (1983)

    Article  CAS  Google Scholar 

  62. L. Zhang, E. Yue, B. Liu, P. Serp, C. Redshaw, W.H. Sun, J. Durand, Catal. Commun. 43, 227 (2014)

    Article  CAS  Google Scholar 

  63. L. Zhang, W. Zhang, P. Serp, W.-H. Sun, J. Durand, ChemCatChem 6, 12 (2014)

    Google Scholar 

  64. R.J. Chen, Y. Zhang, D. Wang, H. Dai, J. Am. Chem. Soc. 123, 3838 (2001)

    Article  CAS  Google Scholar 

  65. V. Georgakilas, J.N. Tiwari, K.C. Kemp, J.A. Perman, A.B. Bourlinos, K.S. Kim, R. Zboril, Chem. Rev. 116, 5464 (2016)

    Article  CAS  Google Scholar 

  66. R.M. Tromp, A. Afzali, M. Freitag, D.B. Mitzi, Z. Chen 26, 18 (2020)

    Google Scholar 

  67. J.S. Foster, J.E. Frommer, Nature 333, 542 (1988)

    Article  Google Scholar 

  68. A.J. Pollard, E.W. Perkins, N.A. Smith, A. Saywell, G. Goretzki, A.G. Phillips, S.P. Argent, H. Sachdev, F. Müller, S. Hüfner, S. Gsell, M. Fischer, M. Schreck, J. Osterwalder, T. Greber, S. Berner, N.R. Champness, P.H. Beton, Angew. Chemie - Int. Ed. 49, 1794 (2010)

    Article  CAS  Google Scholar 

  69. P. Lauffer, K.V. Emtsev, R. Graupner, T. Seyller, L. Ley, Phys. Status Solidi 245, 2064 (2008)

    Article  CAS  Google Scholar 

  70. H. Huang, S. Chen, X. Gao, W. Chen, A.T.S. Wee, ACS Nano 3, 3431 (2009)

    Article  CAS  Google Scholar 

  71. Y. Zhang, C. Liu, W. Shi, Z. Wang, L. Dai, X. Zhang, Langmuir 23, 7911 (2007)

    Article  CAS  Google Scholar 

  72. Y. Liang, D. Wu, X. Feng, K. Müllen, Adv. Mater. 21, 1679 (2009)

    Article  CAS  Google Scholar 

  73. E.Y. Choi, T.H. Han, J. Hong, J.E. Kim, S.H. Lee, H.W. Kim, S.O. Kim, J. Mater. Chem. 20, 1907 (2010)

    Article  CAS  Google Scholar 

  74. H. Dau, M. Haumann, Coord. Chem. Rev. 252, 273 (2008)

    Article  CAS  Google Scholar 

  75. D.R. Whang, S.Y. Park, ChemSusChem 8, 3204 (2015)

    Article  CAS  Google Scholar 

  76. J. Kim, D.R. Whang, S.Y. Park, ChemSusChem 10, 1883 (2017)

    Article  CAS  Google Scholar 

  77. D.R. Whang, D.H. Apaydin, S.Y. Park, N.S. Sariciftci, J. Catal. 363, 191 (2018)

    Article  CAS  Google Scholar 

  78. A. Pannwitz, O.S. Wenger, Chem. Commun. 55, 4004 (2019)

    Article  CAS  Google Scholar 

  79. S. Karlsson, J. Boixel, Y. Pellegrin, E. Blart, H.C. Becker, F. Odobel, L. Hammarström, Faraday Discuss. 155, 233 (2012)

    Article  CAS  Google Scholar 

  80. J.J. Du, Y.P. Yuan, J.X. Sun, F.M. Peng, X. Jiang, L.G. Qiu, A.J. Xie, Y.H. Shen, J.F. Zhu, J. Hazard. Mater. 190, 945 (2011)

    Article  CAS  Google Scholar 

  81. K.G.M. Laurier, F. Vermoortele, R. Ameloot, D.E. De Vos, J. Hofkens, M.B.J. Roeffaers, J. Am. Chem. Soc. 135, 14488 (2013)

    Article  CAS  Google Scholar 

  82. C.G. Silva, I. Luz, F.X. Llabrés, I. Xamena, A. Corma, H. García, Chem. - A Eur. J. 16, 11133 (2010)

    Article  CAS  Google Scholar 

  83. M. Alvaro, E. Carbonell, B. Ferrer, F.X. Llabrés, I. Xamena, H. Garcia, Chem. - A Eur. J. 13, 5106 (2007)

    Article  CAS  Google Scholar 

  84. F.X. Llabrés i Xamena, A. Corma, H. Garcia, J. Phys. Chem. C 111, 80 (2007)

    Article  CAS  Google Scholar 

  85. J. Gao, J. Miao, P.Z. Li, W.Y. Teng, L. Yang, Y. Zhao, B. Liu, Q. Zhang, Chem. Commun. 50, 3786 (2014)

    Article  CAS  Google Scholar 

  86. E.Y. Lee, S.Y. Jang, M.P. Suh, J. Am. Chem. Soc. 127, 6374 (2005)

    Article  CAS  Google Scholar 

  87. H.K. Chae, J. Kim, O.D. Friedrichs, M. O’Keeffe, O.M. Yaghi, Angew. Chemie - Int. Ed. 42, 3907 (2003)

    Article  CAS  Google Scholar 

  88. M.H. Xie, X.L. Yang, C. Zou, C. De Wu, Inorg. Chem. 50, 5318 (2011)

    Article  CAS  Google Scholar 

  89. C. Wang, Z. Xie, K.E. Dekrafft, W. Lin, J. Am. Chem. Soc. 133, 13445 (2011)

    Article  CAS  Google Scholar 

  90. H.Q. Xu, J. Hu, D. Wang, Z. Li, Q. Zhang, Y. Luo, S.H. Yu, H.L. Jiang, J. Am. Chem. Soc. 137, 13440 (2015)

    Article  CAS  Google Scholar 

  91. Z. Bin Fang, T.T. Liu, J. Liu, S. Jin, X.P. Wu, X.Q. Gong, K. Wang, Q. Yin, T.F. Liu, R. Cao, H.C. Zhou, J. Am. Chem. Soc. 142, 12515 (2020)

    Article  CAS  Google Scholar 

  92. D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu, Z. Li, Chem. - A Eur. J. 19, 14279 (2013)

    Article  CAS  Google Scholar 

  93. Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu, Z. Li, Angew. Chemie - Int. Ed. 51, 3364 (2012)

    Article  CAS  Google Scholar 

  94. C. Wang, Z. Xie, K.E. Dekrafft, W. Lin, J. Am. Chem. Soc. 133, 13445 (2011)

    Article  CAS  Google Scholar 

  95. D. Kim, D.R. Whang, S.Y. Park, J. Am. Chem. Soc. 138, 8698 (2016)

    Article  CAS  Google Scholar 

  96. C.N. Valdez, A.M. Schimpf, D.R. Gamelin, J.M. Mayer, J. Am. Chem. Soc. 138, 1377 (2016)

    Article  CAS  Google Scholar 

  97. H. Tong, S. Ouyang, Y. Bi, N. Umezawa, M. Oshikiri, J. Ye, Adv. Mater. 24, 229 (2012)

    Article  CAS  Google Scholar 

  98. A. Kudo, Y. Miseki, Chem. Soc. Rev. 38, 253 (2009)

    Article  CAS  Google Scholar 

  99. E. B. and and W. Choi*, (2006)

  100. K. Hirano, E. Suzuki, A. Ishikawa, T. Moroi, H. Shiroishi, M. Kaneko, J. Photochem. Photobiol. A Chem. 136, 157 (2000)

    Article  CAS  Google Scholar 

  101. H.Q. Zheng, H. Yong, O.Y. Ting, Y.T. Fan, H.W. Hou, in Int. J. Hydrogen Energy (Pergamon, 2013), pp. 12938–12945

  102. K. Maeda, G. Sahara, M. Eguchi, O. Ishitani, ACS Catal. 5, 1700 (2015)

    Article  CAS  Google Scholar 

  103. E. Bae, W. Choi, J. Park, H.S. Shin, S.Bin Kim, J.S. Lee, J. Phys. Chem. B 108, 14093 (2004)

    Article  CAS  Google Scholar 

  104. Y.J. Yuan, J.R. Tu, Z.J. Ye, H.W. Lu, Z.G. Ji, B. Hu, Y.H. Li, D.P. Cao, Z.T. Yu, Z.G. Zou, Dye. Pigment. 123, 285 (2015)

    Article  CAS  Google Scholar 

  105. M. Zhu, Z. Li, B. Xiao, Y. Lu, Y. Du, P. Yang, X. Wang, ACS Appl. Mater. Interfaces 5, 1732 (2013)

    Article  CAS  Google Scholar 

  106. L. Zhang, L. Qin, S.Z. Kang, G.D. Li, X. Li, ACS Sustain. Chem. Eng. 7, 8358 (2019)

    Article  CAS  Google Scholar 

  107. A. Fujishima, K. Honda, Bull. Chem. Soc. Jpn 44, 1148 (1971)

    Article  CAS  Google Scholar 

  108. A. Fujishima, K. Honda, Nature 238, 37 (1972)

    Article  CAS  Google Scholar 

  109. A.G. Tamirat, J. Rick, A.A. Dubale, W.N. Su, B.J. Hwang, Nanoscale Horizons 1, 243 (2016)

    Article  CAS  Google Scholar 

  110. O. Khaselev, J.A. Turner, Science (80-.). 280, 425 (1998)

    Article  CAS  Google Scholar 

  111. B. O’Regan, M. Grätzel, Nature 353, 737 (1991)

    Article  Google Scholar 

  112. Z. Yu, F. Li, L. Sun, Energy Environ. Sci. 8, 760 (2015)

    Article  CAS  Google Scholar 

  113. J.W. Youngblood, S.H.A. Lee, Y. Kobayashi, E.A. Hernandez-Pagan, P.G. Hoertz, T.A. Moore, A.L. Moore, D. Gust, T.E. Mallouk, J. Am. Chem. Soc. 131, 926 (2009)

    Article  CAS  Google Scholar 

  114. J.J. Concepcion, J.W. Jurss, P.G. Hoertz, T.J. Meyer, Angew. Chemie - Int. Ed. 48, 9473 (2009)

    Article  CAS  Google Scholar 

  115. R. Brlmblecombe, A. Koo, G.C. Dismukes, G.F. Swlegers, L. Spiccia, J. Am. Chem. Soc. 132, 2892 (2010)

    Article  CAS  Google Scholar 

  116. D.L. Ashford, A.M. Lapides, A.K. Vannucci, K. Hanson, D.A. Torelli, D.P. Harrison, J.L. Templeton, T.J. Meyer, J. Am. Chem. Soc. 136, 6578 (2014)

    Article  CAS  Google Scholar 

  117. D.L. Ashford, B.D. Sherman, R.A. Binstead, J.L. Templeton, T.J. Meyer, Angew. Chemie Int. Ed. 54, 4778 (2015)

    Article  CAS  Google Scholar 

  118. M.V. Sheridan, B.D. Sherman, R.L. Coppo, D. Wang, S.L. Marquard, K.R. Wee, N.Y. Murakami, Iha, T.J. Meyer, ACS Energy Lett. 1, 231 (2016)

    Article  CAS  Google Scholar 

  119. B.D. Sherman, Y. Xie, M.V. Sheridan, D. Wang, D.W. Shaffer, T.J. Meyer, J.J. Concepcion, ACS Energy Lett. 2, 124 (2017)

    Article  CAS  Google Scholar 

  120. L. Li, L. Duan, Y. Xu, M. Gorlov, A. Hagfeldt, L. Sun, Chem. Commun. 46, 7307 (2010)

    Article  CAS  Google Scholar 

  121. G.F. Moore, J.D. Blakemore, R.L. Milot, J.F. Hull, H.E. Song, L. Cai, C.A. Schmuttenmaer, R.H. Crabtree, G.W. Brudvig, Energy Environ. Sci. 4, 2389 (2011)

    Article  CAS  Google Scholar 

  122. W. Song, C.R.K. Glasson, H. Luo, K. Hanson, M.K. Brennaman, J.J. Concepcion, T.J. Meyer, J. Phys. Chem. Lett. 2, 1808 (2011)

    Article  CAS  Google Scholar 

  123. D.L. Ashford, W. Song, J.J. Concepcion, C.R.K. Glasson, M.K. Brennaman, M.R. Norris, Z. Fang, J.L. Templeton, T.J. Meyer, J. Am. Chem. Soc. 134, 19189 (2012)

    Article  CAS  Google Scholar 

  124. K. Hanson, D.A. Torelli, A.K. Vannucci, M.K. Brennaman, H. Luo, L. Alibabaei, W. Song, D.L. Ashford, M.R. Norris, C.R.K. Glasson, J.J. Concepcion, T.J. Meyer, Angew. Chemie - Int. Ed. 51, 12782 (2012)

    Article  CAS  Google Scholar 

  125. M.R. Norris, J.J. Concepcion, Z. Fang, J.L. Templeton, T.J. Meyer, Angew. Chemie Int. Ed. 52, 13580 (2013)

    Article  CAS  Google Scholar 

  126. A.M. Lapides, D.L. Ashford, K. Hanson, D.A. Torelli, J.L. Templeton, T.J. Meyer, J. Am. Chem. Soc. 135, 15450 (2013)

    Article  CAS  Google Scholar 

  127. X. Ding, Y. Gao, L. Zhang, Z. Yu, J. Liu, L. Sun, ACS Catal. 4, 2347 (2014)

    Article  CAS  Google Scholar 

  128. Y. Kou, S. Nakatani, G. Sunagawa, Y. Tachikawa, D. Masui, T. Shimada, S. Takagi, D.A. Tryk, Y. Nabetani, H. Tachibana, H. Inoue, J. Catal. 310, 57 (2014)

    Article  CAS  Google Scholar 

  129. G. Sahara, R. Abe, M. Higashi, T. Morikawa, K. Maeda, Y. Ueda, O. Ishitani, Chem. Commun. 51, 10722 (2015)

    Article  CAS  Google Scholar 

  130. H. Kumagai, G. Sahara, K. Maeda, M. Higashi, R. Abe, O. Ishitani, Chem. Sci. 8, 4242 (2017)

    Article  CAS  Google Scholar 

  131. J.D. Blakemore, R.H. Crabtree, G.W. Brudvig, Chem. Rev. 115, 12974 (2015)

    Article  CAS  Google Scholar 

  132. M. Yagi, M. Kaneko, Chem. Rev. 101, 21 (2001)

    Article  CAS  Google Scholar 

  133. P. Garrido-Barros, C. Gimbert-Surinifach, D. Moonshiram, A. Picón, P. Monge, V.S. Batista, A. Llobet, J. Am. Chem. Soc. 139, 12907 (2017)

    Article  CAS  Google Scholar 

  134. J. Wang, L. Gan, Q. Zhang, V. Reddu, Y. Peng, Z. Liu, X. Xia, C. Wang, X. Wang, Adv. Energy Mater. 9, 1803151 (2019)

    Article  CAS  Google Scholar 

  135. D.D. Ma, S.G. Han, C. Cao, X. Li, X.T. Wu, Q.L. Zhu, Appl. Catal. B Environ. 264, 118530 (2020)

    Article  CAS  Google Scholar 

  136. G. Goor, J. Glenneberg, S. Jacobi, J. Dadabhoy, E. Candido, Ullmann’s Encycl. Ind. Chem (Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2019), pp. 1–40

    Book  Google Scholar 

  137. F. Mirkhalaf, K. Tammeveski, D.J. Schiffrin, Phys. Chem. Chem. Phys. 6, 1321 (2004)

    Article  CAS  Google Scholar 

  138. K. Tammeveski, K. Kontturi, R.J. Nichols, R.J. Potter, D.J. Schiffrin, J. Electroanal. Chem. 515, 101 (2001)

    Article  CAS  Google Scholar 

  139. A. Sarapuu, K. Helstein, K. Vaik, D.J. Schiffrin, K. Tammeveski, Electrochim. Acta 55, 6376 (2010)

    Article  CAS  Google Scholar 

  140. K. Vaik, U. Mäeorg, F.C. Maschion, G. Maia, D.J. Schiffrin, K. Tammeveski, in Electrochim. Acta (Pergamon, 2005), pp. 5126–5131

  141. P.M. Hoang, S. Holdcroft, B.L. Funt, J. Electrochem. Soc. 132, 2129 (1985)

    Article  CAS  Google Scholar 

  142. M. Mooste, E. Kibena-Põldsepp, L. Matisen, K. Tammeveski, Electroanalysis 29, 548 (2017)

    Article  CAS  Google Scholar 

  143. D. Wielend, M. Vera-Hidalgo, H. Seelajaroen, N.S. Sariciftci, E.M. Pérez, D.R. Whang, ACS Appl. Mater. Interfaces 12, 32615 (2020)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the 2020 Industry-Academia Cooperation Program of Samsung Display and the Asian Office of Aerospace Research and Development (AOARD, FA 2386-17-1-4060, and FA 2386-15-1-4100).

Author information

Authors and Affiliations

Authors

Contributions

DRW wrote the paper. He is the corresponding author of this paper.

Corresponding author

Correspondence to Dong Ryeol Whang.

Ethics declarations

Competing interests

The author declares no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Whang, D.R. Immobilization of molecular catalysts for artificial photosynthesis. Nano Convergence 7, 37 (2020). https://doi.org/10.1186/s40580-020-00248-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40580-020-00248-1

Keywords