Next Article in Journal
Electrocatalysis by Heme Enzymes—Applications in Biosensing
Next Article in Special Issue
Selective and Efficient Olefin Epoxidation by Robust Magnetic Mo Nanocatalysts
Previous Article in Journal
Biocatalytic Transformation of 5-Hydroxymethylfurfural into 2,5-di(hydroxymethyl)furan by a Newly Isolated Fusarium striatum Strain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molybdenum-Containing Metalloenzymes and Synthetic Catalysts for Conversion of Small Molecules

Department of Chemistry, Gwangju Institute of Science and Technology, Gwangju 61005, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(2), 217; https://doi.org/10.3390/catal11020217
Submission received: 2 January 2021 / Revised: 1 February 2021 / Accepted: 3 February 2021 / Published: 6 February 2021
(This article belongs to the Special Issue Molybdenum Catalysis)

Abstract

:
The energy deficiency and environmental problems have motivated researchers to develop energy conversion systems into a sustainable pathway, and the development of catalysts holds the center of the research endeavors. Natural catalysts such as metalloenzymes have maintained energy cycles on Earth, thus proving themselves the optimal catalysts. In the previous research results, the structural and functional analogs of enzymes and nano-sized electrocatalysts have shown promising activities in energy conversion reactions. Mo ion plays essential roles in natural and artificial catalysts, and the unique electrochemical properties render its versatile utilization as an electrocatalyst. In this review paper, we show the current understandings of the Mo-enzyme active sites and the recent advances in the synthesis of Mo-catalysts aiming for high-performing catalysts.

1. Introduction

Fixation of atmospheric nitrogen to ammonia has been promoted by the Haber–Bosch process, and the capacitated mass production of foods accelerated the increase of the world population [1,2]. The rapid growth of fossil fuel-based industry unavoidably increased atmospheric CO2, and resultantly aggravated the global warming and climate crisis. To suggest solutions for those anthropogenic issues and mitigate the current environmental problems, scientific progress to convert fossil fuel-based energy systems toward sustainable ways is keenly desired along with other social and economic efforts [3]. In the line of scientific cognition, we review catalysts for energy conversion reactions including small molecules such as CO2, N2, H2, and O2 as reactants or products. Numerous catalysts have been developed in recent decades owing to the high interest in energy conversion reactions [4,5,6,7,8]. Mo ion is found to be an essential component in biological catalysis [9], as well as a very promising metal candidate among Earth-abundant metals for applications to electrocatalysts and photocatalysts [4,10,11]. Herein, we review versatile utilization of Mo ions in natural and synthetic catalysts.
Mo-containing homogeneous catalysts utilize both low- and high-valent Mo ions, whereas metalloenzymes and their model complexes have high-valent Mo centers. Except for FeMo-nitrogenase [12], Mo-enzyme active sites always use dithiolene ligation of the molybdopterin (MPT) moiety to stabilize a Mo(IV/VI) ion during the redox reaction and to assist electron transfer between metal and ligand [13]. Moreover, the chalcogenide ligand and nearby amino acid play essential roles to determine the selective reactivity of Mo active sites. Although enzyme mechanisms remain unclear despite significant interests, ligand identity and coordination structures seem to be critical factors to determine catalytic activity. We can propose synthetic routes for the design of new Mo-based catalysts from comparing structures, properties, and reactivities of metalloenzymes and artificial catalysts [14]. Although relatively less studied, Mo-clusters have shown promising electrocatalytic activities due to their intermediate behaviors between metal-coordination complexes and nanoparticle catalysts [15]. In the perspective of developing efficient electrocatalysts, Mo-nanomaterials, which have shown remarkable electrocatalytic activities, should be considered together. Preparation methods of Mo-nanomaterials determine their catalytic activities, and proximal atoms and electronic conditions around Mo ions are closely related to reactivities of surficial Mo ions [16,17].

2. CO2 Electrocatalysts

Considering current energy systems, a drastic diminution of fossil fuel consumption seems unrealistic. However, continuous research for recycling anthropogenic CO2 will be able to provide feasible solutions to current energy and climate issues. Among various CO2 transformation methods, electrochemical conversion of CO2 may be an attractive way to be conjugated with solar/electric energy conversion reactions [18]. Mo-enzymes of the formate dehydrogenases [19] (Mo-FDH) and the carbon monoxide dehydrogenases [20] (MoCu-CODH) are the most efficient CO2-catalysts on Earth, and their synthetic models are known, albeit short of the enzyme activity [21]. Moreover, many homogeneous and heterogeneous catalysts have been developed for CO2 conversion reactions.

2.1. Metalloenzymes with CO2 Reactivity

2.1.1. Formate Dehydrogenases

Formate dehydrogenases (FDHs) is a group of metalloenzymes that catalyze a reversible conversion of formate to CO2. The CO2 reactivity of the FDH active site has obtained research interests because the enzyme functions near the potential of −420 mV vs. NHE (Normal hydrogen electrode) scarcely requiring overpotential. It is known that the reaction rate is varied by bacteria. Rhodobacter capsulatus FDH facilitates oxidation of formate to CO2 at a rate of 36.5 s−1 and the reverse direction at a rate of 1.6 s−1 [22]. Desulfovibrio desulfuricans FDH oxidized formate much more quickly at 543 s−1 and reduced CO2 at 46.6 s−1 [23]. The FDH reactivity is directly related to the active site structure, which has a Mo center surrounded by two pyranopterin-conjugated dithiolene ligand, a chalcogenide (hydroxide or sulfide) ligand, and SeCys amino acid as a sixth ligand (Figure 1). In addition, additional roles of adjacent amino acids such as Arg and His seem to be important to assist the FDH reaction by stabilizing reaction intermediates.
CO 2   +   2 e   +   H +     HCOO ,   E °   =   420   mV
In 1997, Boyington and coworkers first reported the crystal structure of E. coli FDH [24]. Later, in 2006, Raaijmakers and coworkers reinterpreted the crystal structure and proposed a reaction pathway of the FDH active site (Scheme 1) [25]. They suggested that SeCys is reversibly bound to the Mo center during the redox reactions so that offers a vacant site to interact with substrates. Once formate bound to Mo6+ site, a proton migrated to His residue, and CO2 was released along with the reduction of Mo6+/4+ site. If two electrons were transferred to [4Fe4S] cluster, SeCys recombined at the Mo site. In the proposed reaction, the sulfide ligand did not participate in the reaction.
Another reaction pathway including the role of sulfide ligand has been proposed (Scheme 2). In 2011, Moura and coworkers suggested that the sulfide ligand assists the reversible dissociation/association of Mo-SeCys bond on the basis of theoretical studies. A probable Mo-S-SeCys intermediate could open a coordination site to interact with formate. Once a proton migrated from formate to SeCys, a thiocarbonate intermediate was formed with the dissociation of the S-SeCys bond, and adjacent Arg+ amino acid possibly stabilized the negatively charged intermediate. The roles of nearby amino acids seemed to be critical for the catalytic reaction at the active site [26,27].
The same group proposed that a hydride transfer occurred in the conversion of Mo(VI) = X and Mo(IV)-XH, and the formate oxidation would likely occur as an indirect pathway [23]. In a substitution experiment of Mo-bound formate for isoelectronic azide, the azide entered between Arg/His and Mo=S, not at the binding site of SeCys (Scheme 3).
Haumann and coworkers reported that the Mo=S bond length of the oxidized form is ~2.16 Å on the basis of X-ray absorption spectroscopy [28].
They suggested that Mo-SH possibly exists based on the Mo-S distance of ~2.4 Å in the X-ray crystal structure and SeCys stays apart from the Mo center during the reaction. They suggested a direct interaction of formate with Mo center based on X-ray absorption spectroscopy. When azide entered in the Mo-SCys site of the oxidized form, the Mo-Cys bond distance decreased. Conversely, when azide was removed, the Mo-Cys distance increased [29].
The reaction mechanism of the FDH active site is a continuous research topic because an in-depth understanding of the active site will lead to the development of efficient electrocatalysts for CO2 conversion.

2.1.2. MoCu-Carbon Monoxide Dehydrogenases

In 2002, MoCu-carbon monoxide dehydrogenase (CODH) from Oligotopha carboxidovorans was found to catalyze the oxidation of CO to CO2 under aerobic conditions. The enzyme exists as a dimer form of heterotrimer [30]. The hetero-trimer is composed of 88.7 kDa L (809 residues), 30.2 kDa M (288 residues), and 17.8 kDa S (166 residues). The L subunit contains molybdopterin-cytosine dinucleotide (MCD), and the M subunit has a flavin adenine dinucleotide (FAD) cofactor. Two [2Fe-2S] clusters stay in the S subunit [31]. As characterized by X-ray crystallography and extended X-ray absorption fine structure (EXAFS), the MoCu-CODH active site is composed of a hetero-bimetallic cluster of (MCD)MoVIO(μ2-S)-CuI(S-CYS), and the bent μ2-S bridge connects the Cu and Mo center (Figure 2a) [32]. The resting state of the active site was characterized as a MoVI(O)2SCuI core, and the catalytic oxidation possibly occurs with CO binding at the Cu center.
In 2002, Dobbek and coworkers reported the X-ray structure of the MoCu-CODH with N-butylisocyanide, where the N-butylisocyanide entered between µ-S–Cu bond by cleaving the bridging sulfide bond [31]. Accordingly, they suggested that CO would interact with the active site in a similar way as the isoelectronic molecule as forming a µ-thiocarbamate intermediate (Figure 2b).
However, Siegbahn and coworker suggested that CO binding to Cu is energetically favored from the computational studies. Next, the Cu-CO likely experiences a nucleophilic attack by the equatorial Mo=O to form a thiocarbamate intermediate, and CO2 can be released with coordination by H2O (or –OH) (Figure 2c) [33].
The formation of a thiocarbamate intermediate was suggested previously, but Hofmann and coworkers claimed that a thiocarbamate intermediate is not relevant to the MoCu-CODH catalytic cycle, because a thiocarbamate moiety is too stable to be involved in the catalytic steps (Figure 2d) [34]. Hille and coworkers also proposed that the Cu-CO-O-Mo bond is formed after CO insertion (Figure 2e). They emphasized the role of nearby Glu763 residue to deprotonate H2O. A nucleophilic attack of –OH to the labile equatorial Mo=O could promote the release of CO2 [30] The aerobic MoCu-CODH catalyzed CO oxidation with kcat 93.3 s−1 at pH 7.2 condition [35].

2.2. Structural Analogs of CO2 Enzymes

2.2.1. FDH Analogs

The high efficiency of the FDH enzyme spurs researchers to synthesize structural analogs of the enzyme active site. Synthetic procedures for various dithiolene complexes enabled modeling studies of the enzyme active site having the Mo-bis(dithiolene) moiety. High-valent Mo(IV-VI)-bis(dithiolene) appeared to be highly reactive. Although the Mo enzyme active site is protected by surrounding proteins, the air-sensitivity of high-valent Mo in solution conditions makes the modeling study difficult. On the other hand, the high reactivity could be a clue to the CO2 reactivity of the Mo site. Previous studies have revealed that the selective reactivity of Mo-bis(dithiolene) complexes is related to auxiliary ligands.
The Mo-FDH and W-FDH active sites have very similar coordination environments except for the metal center, although their reaction conditions are different because of the bacterial habitat. In 2012, Kim and Seo reported CO2 reactivity of the in-situ generated [WIV(OH)(S2C2Ph2)2] (2) [36] as an example of the functional model of the W-FDH active site (Scheme 4) [25]. Complex 2 generated by hydrolysis of [WIV(OPh)(S2C2Ph2)2] (1) [37,38] showed the CO2 reactivity at mild condition, and it was suggested that a nucleophilic attack of –OH to CO2 formed a W-(bi)carbonate intermediate. Although the model complex did not show a catalytic activity, the plausible formation of a W-(bi)carbonate intermediate was similar to the reactivity of the FDH active site [26]. They also reported a stoichiometric reduction of CO2 to formate by [WIVO(S2C2Me2)2]2− (3) at 90 °C [39]. A W-carbonate intermediate was suggested as an intermediate before the CO2 reduction, but the oxidized W complex was dimerized to [WV2O2(μ-S)(μ-S2C2Me2)(S2C2Me2)2]2− (4) at the high temperature as frustrating catalytic reaction (Scheme 4).
Fontecave and coworkers reported catalytic CO2 reduction by Ni-bis(dithiolene) complex using a dithiolene derivative of quinoxaline-pyran-fused dithiolene (qpdt2−), which is similar to the molybdopterin (MPT) structure of the FDH active site (Scheme 4) [40]. In 2018, they used the qpdt2− ligand to prepare Mo complexes, MoIVO(qpdt)2 (5), [MoIVO(H-qpdt)2]2− (6), and [MoVO(2H-qpdt)2] (7), as the Mo-FDH models, and examined the photocatalytic CO2 reduction activity [41]. The reduced forms of H-qpdt and 2H-qpdt had similar oxidation states with the MPT. Using [Ru(bpy)3]2+ photosensitizer, 1,3-dimethyl-2-phenyl-2,3-dihydro-1H-benzoimidazole as a sacrificial electron donor in the solution of triethanolamine and acetonitrile (1:5), and irradiation with a 300 W Xe lamp on the Mo complex solutions afforded CO2-derivated products (CO and HCO2H) and H2. Complex 5 gave the CO2-reduction products with 19% and mostly H2 (81%). However, complex 6 increased the CO2-reduction yield to 47%, and complex 7 provided the CO2-reduction products in the highest yield of 58% with over 100 turnover number (TON) for 68 h.
Structural and functional analogs of the FDH active sites have been reported with rare examples; however, the catalytic activities remain much less than the enzymes. Secondary coordination ligands mimicking amino acids surrounding the enzyme active site would give new catalyst designs to achieve similar activities as the enzyme.

2.2.2. MoCu-CODH Analogs

In 2020, Mougel and coworkers synthesized [(bdt)Mo(O)S2CuCN]2− (8) (bdt = benzenedithiolate) as a MoCu-CODH model complex (Scheme 5) [42]. The Mo and Cu were bridged by μ2-sulfide ligand as the enzyme active site, and a bdt ligand was used to model the MPT. Complex 8 showed a catalytic reduction peak at E1/2 = −2.07 V vs. Fc+/Fc in CO2-saturated acetonitrile. Controlled potential electrolysis (CPE) at −2.62 V vs. Fc+/Fc in the presence of 0.1 M 2,2,2-trifluoroethanol (TFE) produced the reaction products of formate (Faradaic efficiency (FE) = 69%), CO (8%) and H2 (19%). It was a rare example to describe the MoCu-CODH model with the CO2 reduction activity, although the MoCu-CODH enzyme previously showed only the CO oxidation activity [20].

2.3. Synthetic Electrocatalysts for CO2 Reduction

2.3.1. Homogeneous Mo Complexes for CO2 Reduction

In 1978, Reichert and coworkers reported the CO2 reactivity of a dinuclear Mo2(OtBu)6 (9) [43]. Two CO2 molecules were reversibly inserted into the Mo-OtBu bonds to form Mo2(O2COtBu)2(OtBu)4 (Scheme 6) (10). From the X-ray structure, the Mo-Mo triple bond, bridged by two O2COtBu, was measured to be 2.241 Å. The CO2 insertion occurred in both the solid and liquid states of the complex, and sublimation of the CO2 adduct at 100 °C afforded back the complex 9 by releasing CO2.
Metal ligand cooperation (MLC) is a useful method to fixate CO2. The trans-Mo(C2H4)2(PMe3)4 (11) reacted with CO2 under mild conditions [44], where CO2 was inserted into Mo-ethylene bond to form a Mo-acrylate moiety in [Mo(H2CCHCO2H)(C2H4)(PMe3)2]2 (Scheme 7) (12). The acrylate frequency was measured as 1500 cm−1 and the 13C NMR (Nuclear magnetic resonance) chemical shift at 175–180 ppm. Complex 11 reacted with a stoichiometric amount of CO2 under 1 atm to give 12, which could be converted back to 11 by treatment of n-BuLi under H2 with concomitant LiO2CCH2CH3 product. In another MLC type reaction, a PNP-amide-assisted CO2 fixation was shown with Mo(NO)(CO)(PNP) (13) (PNP = N(CH2CH2PiPr2)2) [45]. Complex 13 converted stoichiometric amount of CO2 to formate in the presence of Na[N(SiMe3)2] and additional base of DBU, Et3N (Scheme 7). However, high pressure of CO2 (10 bar) and H2 (70 bar) was also required. The MLC method effectively lowered the binding energy of CO2 to Mo complex, but the formation of a highly stable ML-CO2 adduct rather prevented the catalytic cycle.
Lau and coworkers utilized hetero-bimetallic systems of early/late transition metal ions for electrocatalytic reactions. They synthesized (η5-C5H5)Ru(CO)(μ-dppm)Mo(CO)25-C5H5) (14) (Scheme 8) from a reaction of (η5-C5H5)Ru(dppm)Cl and Na[(η5-C5H5)Mo(CO)3] [46]. Complex 14 promoted catalytic CO2 reduction to formic acid at 120 °C with 43 TON per catalyst for 45 h in benzene containing triethylamine base. Interestingly, the same complex also promoted converse conversion of formic acid to CO2 and H2 at 80 °C. They suggested a hydride-bridged bimetallic species is formed as a reaction intermediate on the basis of a hydride detection at −17.16 ppm by 1H NMR spectroscopy.
Minato and coworkers synthesized a series of Mo-silyl hydrido complexes of [MoH3([Ph2PCH2CH2P(Ph)-C6H4-o]2(R)Si-P,P,P,P,Si)] (R = Ph (15a), C6F5 (15b), 4-Me2NC6H4 (15c), cyclohexyl (15d), and n-C6H13 (15e)) [47]. Silyl ligands with a strong trans effect were employed to control the reactivity of the Mo center (Scheme 9) [48,49]. The RSi-Mo could make an open coordination site at trans position to allow binding of electrophilic molecules such as O2, CO2, and carboxylic acid. Under the reaction conditions of 2.0 M of dimethylamine, 30 atm of CO2, and 20 atm of H2, complexes 15ae catalyzed the CO2 hydration as producing N,N-dimethylformamide (DMF) and H2O. The catalytic conversion of CO2 to DMF could be controlled by the Si-substituent. The complex 15a with a Si–phenyl moiety achieved 92 TON per catalyst, but complex 15b with Si–C6F5 showed only 1 TON. Electron donating substituent increased TON of complexes as seen with 15c (TON = 130) and 15e (110), because the stronger trans effect assisted binding of substrates to increase the catalytic reactivity.
Redox-active ligand could assist central metal ions for reductive CO2 reactions. Trovitch and coworkers have used bis(imino)pyridine ligand for hydrosilylations of ketones [50]. They synthesized (κ6-P,N,N,N,C,P-Ph2PPrPDI)-MoH (17) (PDI = pyridine diamine) by a reduction of ((Ph2PPrPDI)MoI)I (16) with excess K/Hg (Scheme 10) [51]. Complex 17 reacted with CO2 under 0.2 atm of CO2 condition, where CO2 was inserted into Mo-H bond to form η1-formate-bound Mo(OC(O)H) species [52,53]. The Mo-formate 13C NMR resonance appeared at 169.7 ppm and the infrared (IR) frequency was detected at 1625 cm−1. Complex 17 also promoted a catalytic conversion of HBPin to H3COBPin and O(BPin)2 under 1 atm of CO2 as recording 40.4 h−1 turnover frequency (TOF) and 323 TON with consuming 97% of HBPin. Distillation of H3COBPin in water for 24 h generated MeOH, eventually producing 58% of MeOH starting from HBPin. They showed a potential pathway to convert CO2 to methanol. Kubiak and coworkers used bipyridine (bpy) ligands to prepare Mo-carbonyl complexes, Mo(bpy)(CO)4 (18a) and Mo(bpy-tBu)(CO)4 (18b) [54]. Complexes 18a and 18b showed electrocatalytic CO2 reduction activity in acetonitrile solution [55]. Complex 18a had a reversible redox couple at −1.58 V vs. SCE (saturated calomel electrode) and an irreversible wave at −2.14 V vs. SCE, and catalyzed CO2 reduction near the second reduction potential. Complex 18b shifted the reduction potential a bit negatively to −2.20 V vs. SCE. The addition of 1.6 M of TFE into the reaction solutions generated only CO product as recording 1.0 s−1 TOF with 18a and 1.9 s−1 with 18b.
Bernskoetter and coworkers used tertiary amine pincer type ligands to prepare CO2 hydrogenation catalysts [56]. Reduction of (PNMeP)MoCl3 with Na(Hg) under ethylene afforded (PNMeP)Mo(C2H4)2 (19), and the following oxidative C-H addition generated the Mo-H species of (PNCH2P)MoH(C2H4)2 (20) (Scheme 11) [57]. Blowing 4 equiv. of CO2 into complex 20 gave the Mo-O2CH moiety in complex 21. Catalytic production of formate was achieved by using LiOTF as cocatalyst and DBU base under 69 atm of H2 and CO2 (1:1 ratio) at 100 °C as recording 35 TON for 16 h.
The low-valent Mo complexes have shown promising activities for CO2, but the formation of relatively stable adducts or requirement of high-pressure conditions remain issues to solve for efficient catalyst development.

2.3.2. Heterogeneous Mo-Containing CO2 Reduction Electrocatalysts

In 1986, Frese and coworkers reported the reduction of CO2 to methanol using Mo electrode at the potential range of −0.57 to −0.67 V vs. SCE in CO2-saturated aqueous media (0.05 M H2SO4, pH 4.2) [58]. As a result, 18 μmol of methanol was obtained for 40 h with 50% FE together with 0.044 μmol of CO production.
MoS2 surface has shown CO2 reduction ability, but in general, the intrinsic properties such as poor electrical conductivity and few active sites cause low electrocatalytic activity. In 2014, Asadi and coworkers reported the layer-stacked MoS2 nanomaterial [59]. The MoS2 surface conducted selective reduction of CO2 to CO in the presence of 4% 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIM-BF4) with a high reduction current of 65 mA/cm2 at −0.76 V vs. RHE (FE ~98%). The activity at a low overpotential (0.1 V) was measured to be 25 times higher than Au nanoparticles. They reported vertically aligned (VA) MoS2 had higher CO2 reactivity than bulk MoS2. The high catalytic performance was due to the increased MoS2 edge site. In 2017, Salehi-Khojin and coworkers synthesized VA-MoS2 with ~20 nm thickness by chemical vapor diffusion (CVD) method, and they also studied the metal-doping effect for the CO2 reduction activity [60]. The 5% Nb-doped VA-MoS2 began the CO2 reduction with onset overpotential of 31 mV, and it reached 237 mA/cm2 current density at −0.8 V vs. RHE. However, Ta-doped (3~18% Ta) VA-MoS2 showed 98–68 mA/cm2 at −0.8 V vs. RHE, which was rather lower than a pristine VA-MoS2 (121 mA/cm2).
Previously, Bi-deposited glassy carbon has shown a good catalytic activity for reduction of CO2 to CO because CO2 radical species are well stabilized on Bi site [61]. Interestingly, a formation of hetero-bimetallic MoBiSx (Mo/Bi 1:1) nanosheet showed different CO2 reduction product of methanol (FE 71.2%) with generation of side products of methane (7.8%) CO (9.6%) H2 (11.4%) [62]. It was observed that, if the Bi ratio was raised, the CO percentage increased. They explained that the further reduction process is probably promoted by the synergistic effect with MoSx–catalytically active for the proton reduction.
Nano-sized Cu particles have shown CO2-to-hydrocarbon electrocatalytic reduction activity [63]. The Cu-doping effect on the CO2 reactivity of MoS2 was studied by Yu and coworkers [64]. Flower-like MoS2 was prepared by hydrothermal method, and the surface was reacted with Cu(NO3)2·3H2O to give Cu/MoS2 composite with varying the Cu/Mo ratio. Specific Cu ratio (12.76%) exhibited four times higher catalytic current density at −1.7 V vs. SCE than a bare MoS2 as producing CO (FE = 35.19%), CH4 (17.08%), C2H4 (2.93%). The high Faradaic efficiency for CO2 reduction was because the surface-deposited Cu enhanced the electronic conductivity and CO2 adsorption ability of MoS2.
Few active sites have been an issue for the preparation of MoS2 electrodes, thus, synthetic methods are pursued to enlarge the MoS2 edge site area. Wang and coworkers used zeolitic imidazolate frameworks to prepare hollow MoS2 nanostructure, which effectively enlarged the exposed MoS2 edge area. The edge-exposed MoS2 was supported by N-doped carbon to improve electron transfer, which gave two times higher current density (34.31 mA/cm2 at −0.7 V vs. RHE) and 1.4 times higher FE (CO2-to-CO 93% FE) compared to MoS2 [65]. The N-doped carbon served as efficient conductive support, because the more electronegative N atom decreases electron density at nearby carbon sites by improving electron transfer at the electrode surface. An enhanced surface electron transfer to MoS2 edge could lower the energy barrier for the formation of CO2 reduction intermediates on the catalytic site [66]. Zhu and coworkers prepared N-doped MoS2 nanosheets on N-doped carbon nanodots (N-MoS2@NCDs) through a solvothermal method [66]. N-MoS2@NCDs loaded on glassy carbon electrode exhibited high catalytic activity for the CO2 reduction to CO as recording 36 mA/cm2 at −0.9 V vs. RHE (90% FE).
The selection of electrolytes is also a critical factor to improve CO2 reduction efficiency. EMIM-BF4 electrolyte was shown to increase selectivity for CO product from CO2 reduction [59]. Salehi-Khojin and coworkers reported that the use of a hybrid electrolyte of choline chloride/KOH was also effective to improve the CO2 reduction selectivity of the MoS2 electrode [67]. Using the choline chloride/KCl buffer instead of EMIM-BF4 afforded 1.5 times higher current density at −0.25 V vs. RHE. The high catalytic activity was related to an accumulation of small size K+ ion on the MoS2 surface, which allowed exposure of the MoS2 edge site.
Xie and coworkers devised MoS/Se alloy monolayer to induce poor overlapping of frontier orbitals, which was helpful to enhance the surface activity for the CO2 reduction [68]. The MoS/Se alloy monolayer improved CO production efficiency with 45% FE at −1.15 V vs. RHE compared to MoS2 (17%) and MoSe2 (31%) monolayer. On the other hand, in the report by Han and coworkers, the replacement of the chalcogenide with phosphide gave different CO2 reduction products. MoP nanoparticle as supported by In-doped porous carbon (In-PC) converted CO2 to formic acid as recording 43.8 mA/cm2 at −2.2 V vs. Ag/AgNO3 with 97% FE [69].
In a rare example, a Mo-cluster has shown promising photocatalytic CO2 reduction ability under visible light illumination. Hexanuclear [Mo6Br14]2− clusters [70] immobilized on graphene oxide [71] showed photocatalytic CO2 reduction to methanol. High nuclearity transition metal cluster of [Mo6Xi8La6]2− (Xi = inner, La = apical) enabled visible light activities due to delocalized valence electrons overall metal centers [72,73]. The catalytic CO2 reduction activities of the Mo-based electrocatalysts are compared in Table 1.

2.4. FDH-Electrode Biohybrid

Formate dehydrogenase (FDH) is the most efficient electrocatalyst for interconversion between CO2 and formate. There have been research interests in the enzyme activity under electrochemical conditions as adsorbed on an electrode surface, or as free in solution. FDHs are classified as two types of NADH-dependent and metal (Mo or W)-dependent enzymes (NADH = nicotinamide adenine dinucleotide). The bio-electrochemical set-ups of NADH-dependent Candida boidinii (CbFDH) have been reported to show selective CO2 reduction to formate [74,75,76,77,78,79], where the regeneration efficiency of NADH was critical to improving the catalytic system.
Metal-dependent FDH has been studied as attached to an electrode surface. Hirst and coworkers showed reversible interconversion between CO2 and formate by Escherichia coli FDH (EcFDH) as adsorbed on a graphite-epoxy electrode [80]. The reversible redox current for the CO2/formate conversion was observed during the cyclic voltammetry scans, and the formate oxidation became favored at higher pH within 6–8 range. Controlled potential electrolysis at −0.6 V vs. SHE produced only formate product with 101.7 ± 2.0% FE.
Redox-active polymer could enhance electron transfer between FDH and electrode. Milton and coworkers used cobaltocene-functionalized polyallylamine (Cc-PAA) to promote electron transfer between EcFDH and glassy carbon electrodes [81]. The electroenzymatic CO2 reduction was detected at −0.66 V vs. SHE close to the Cc/Cc+ redox couple of −0.576 V vs. SHE. Bulk electrolysis at −0.66 V vs. SHE using 50 mM NaHCO3 as the CO2 source provided formate production with FE 99 ± 5%, but the FE dropped to 65% after continuous electrolysis for 12 h.
Bio-hybrid system gives a synthetic vision to develop electrocatalysts with high selectivity and efficiency, although low current density and limited catalytic sites remain as limitations for scale-up. In addition, a hybrid of molecular catalysts with electrode surface would be an alternative method to obtain well-performing catalysts. Ligand design and synthesis are other difficulties, but the conjugation of coordination complexes by electrode surface can provide chances to modify surface reactivity and efficiency of electrocatalysts.

3. Nitrogen Fixation

Nitrogen is an essential element in the metabolism of living organisms and one of the major elements that make up the body. The only way to take atmospheric N2 into organisms is via the nitrogen fixation process [82,83]. In a symbiotic relation, plants feed nitrogen-fixing bacteria with sugars as a reducing equivalent to perform the reduction of N2 to NH3, which is in return absorbed by plants. Although nature performs the process very efficiently, the direct N2 reduction has been one of the challenges in the research area of synthetic catalysts [84]. The industrial Haber–Bosch process takes charge of global ammonia production, but this process occupies 1% of annual fossil fuel consumption and 3% of annual CO2 emission because of the production conditions demanding high pressure and high temperature [85,86]. Mild reaction conditions are highly craved to decrease the consumption of the limited energy resources as well as anthropogenic CO2 emission [87,88]. There exist many research efforts to elucidate the mechanism of the natural catalytic system, nitrogenase, by investigating the structural features and synthetic model complexes [89,90,91]. In addition, heterogeneous catalysts were developed for nitrogen reductions, and bio-hybrid catalysts were studied in a way of attaching nitrogenase to an electrode surface.

3.1. Nitrogenase

In the 1960s, X-ray structure of the FeMo-nitrogenase active site was obtained, and thereafter the structural features and the relation with the enzyme reactivity have been studied to elucidate the nitrogen reduction process [92,93,94]. The N2 reduction is a series of electrochemical reactions requiring overall 6e/6H+ input. The FeMo-cofactor facilitates the nitrogen reaction, and the co-existing P cluster is in charge of the electron-transfer steps. The FeMo cluster is buried under the enzyme protein to selectively promote the multi-electron reduction process (Figure 3).
N2 + 8H+ + 16MgATP+ 8e → 2NH3 + H2 + 16MgADP + 16Pi
The FeMo-cofactor has a double-cubane type of Mo7Fe7S cluster, where each cubane is bonded by a carbide (C4−) center and a Mo atom is positioned at an end of the cluster [95,96,97]. Although the reaction mechanism remains obscure in the research of N2 reduction catalysts, both the Fe and Mo sites have been considered as the active sites for the nitrogen reactions. In the current review, we focus on the Mo-based model complexes.

3.2. Structural Analogs of Nitrogenase

In 1978, Holm and coworkers synthesized [Mo2Fe6S8] (22) cluster as a model of the FeMo-nitrogenase through a self-assembly method [98,99]. Later, in 1982, the same group reported a cubane-type structure of [MoFe3S4] (23) by removing a Mo ion from [Mo2Fe6S8], [100,101] which showed very similar EXAFS data as the original FeMo-cofactor [102]. In 1993, Coucouvanis and coworkers synthesized a citrate-stabilized [MoFe3S4] (24), similar to the homocitrate-coordinated FeMo-nitrogenase [103]. In addition, [(Tp)MoFe3S4Cl3]1− (25) cluster was synthesized using hydrotrispyrazolylborate(Tp) ligand as a model of homocitrate [104]. Along with the structural modeling of the FeMo-cofactor, Mo-based complexes were studied for the catalytic N2 reduction [105,106,107] (Scheme 12).

3.3. Synthetic Electrocatalysts for N2 Reduction

3.3.1. Homogeneous Mo Complexes for N2 Reduction

In 1988, Hidai reported examples of converting N2 to silylamines using zero-valent Mo coordinated by phosphine ligands, cis-Mo(N2)2(PMe2Ph)4 (26) and Mo(N2)2(dpe)2 (27) (Scheme 13) [108]. Complex 26 produced 23.7% N(SiMe3)3 under optimized conditions of N2 using Me3SiCl and Na (or Li) as reductant, whereas complex 27 showed less activity of 9.7% conversion yield under the same condition. The monodentate phosphine made a more active Mo site than the bidentate phosphine. Even though the conversion yield was low, it gave early examples of Mo-phosphine catalysts for N2 conversion reactions.
In 2003, Yandulov and Schrock used a high-valent Mo(III-IV) ion stabilized by a tetradentate triamidoamine ligand for the reduction of N2 to NH3 [109]. The coordination of a hexaisopropylterphenyltriamido amine (HIPT) ligand to Mo(IV) ion gave [HIPTN3N]MoCl, where the six hexaisopropylterphenyl groups sterically enclosed the Mo active site [110]. The bulky ligand system prevented dimerization of the complex by conserving the single Mo center [106,111]. Unlike the FeMo-cofactor, the Mo complex was coordinated by organic N donors, but the reactivity study of [HIPTN3N]MoIII(N2) (28) species provided implemental results to understand an N2 reduction process (Scheme 13). Using a proton source of [(2,6-lutidinium)(BArF)] and decamethyl chromocene as a reductant, complex 28 generated 7.56 equiv. of NH3 per a Mo ion and 63% yield per a reductant, which was comparable to that of FeMo-nitrogenase (75% NH3, 25% H2) [112]. They proposed an N2 reduction mechanism catalyzed by a single Mo site based on spectral data and X-ray crystal structures of intermediates. X-ray crystal structures and spectral data characterized the nitrogen reaction intermediates such as Mo(N2), Mo-N=NH, [Mo=N-NH2]+, Mo≡N, [Mo=N-NH3]+, [Mo=NH]+, [Mo-NH3]+, and MoNH3, which suggested the N2 reduction pathway including 6e/6H+ processes [113,114,115,116,117]. In the proposed pathway, protons bind to one nitrogen to release the first NH3, and the next protonation occurs on the remaining nitrogen to produce the second NH3 (Scheme 14). Their studies gave a research basis for a Mo-based N2 reduction process [118].
In 2011, Nishibayashi and coworkers tried to improve the N2 reduction activity of Mo complexes by utilizing redox-active ligands. They modified the Hidai complex (26) with ferrocenyl diphosphine ligand to prepare trans-Mo(N2)2(depf)2 (29) (depf = diethylphosphinoferrocene) [119]. Complex 29 performed the silylamine production under 1 atm of N2 at room temperature using Na and Me3SiCl. Me3SiCl formed Me3Si∙ radical in the presence of Na reductant, and a Me3Si∙ radical derived the N2 reaction to produce N(SiMe3)3 (226 equiv. per a Mo atom for 200 h). Acid post-treatment of the produced silylamine afforded NH3. The same group used other redox-active PNP type ligands for Mo catalysts [120]. Reduction of the [MoCl3(PNP)] species with Na/Hg under N2 formed an N2-bridged di-Mo complex [Mo(N2)2(PNP)]2(μ-N2) (30ac) (Scheme 13). Subsequent reaction with 4 equiv. of HBF4∙OEt2 and pyridine generated a Mo-hydrazide species, and further reaction with [LutH]OTf produced NH3 (23.2 equiv. of NH3 per a catalyst). The reaction mechanism was explained similarly as proposed by Schrock, where one nitrogen site receives 3e/3H+ to release NH3 and subsequent 3e/3H+ reaction of the remaining Mo(N) generates (Scheme 15). In the screening of reductants of Cr-Cp (E1/2 = −0.88 V vs. Ag/Ag+ in MeCN), Co-Cp (E1/2 = −1.15 V), Cr-Cp* (E1/2 = −1.35 V) and proton sources such as [LuTH]OTf (pKa 14.4 in MeCN), pyridinium trifluoromethanesulfonate (pKa 12.6), and HOTF (pKa 2.6), the Nishibayashi complex exhibited the N2 reduction activity with Co-Cp and Cr-Cp*, but not with Cr-Cp [121]. The complex showed the best performance with the combinatory use of Co-Cp and [LuTH]OTf.
The electronic property of the pyridine ligand affected the catalytic reaction rate by modulating Mo–N≡N–Mo bond strength [122]. In comparison with the v(NN) frequency (1944 cm−1) of unmodified pyridine (30a), the para-substitution with a phenyl group increased the frequency to 1950 cm−1 (30b), whereas the methoxy group decreased to 1932 cm−1 (30c). The relative bond strength was reflected in the N2 reduction reactivity of catalysts. Complex 30a provided 21–23 equiv. of NH3 production, but 30c with the electron-donating methoxy substituent induced the highly efficient NH3 formation (34 equiv. per catalyst) by accelerating the protonation step. Similar to the P cluster of the nitrogenase, the substitution of a redox-active ferrocene assisted the catalytic reaction (30dg) [123]. Ferrocene-attached Mo complexes showed a similar v(NN) stretching frequency as the complex 30a but increased the NH3 production to 37 equiv. per catalyst. Although unmodified ferrocene did not affect the NN bond strength, modification of the ferrocene with ethyl or phenyl groups caused the slight shift of the v(NN) value indicating electronic connectivity through the chemical bonds. The v(NN) value was measured as 1939 cm−1 with the ethyl-ferrocene and 1951 cm−1 with the phenyl-ferrocene. Consistent with the electron-donating effect, the ethyl-ferrocene resulted in 30 equiv. of NH3 formation per catalyst, which was higher than 10 equiv. of NH3 of the phenyl-ferrocene.
The Nishibayashi type PNP-Mo catalysts began the N2 reduction in the N2-bound Mo state after a halide abstraction step. They recently reported the presence of iodide ligand rather improved the catalytic activity because of the electron-withdrawing property. PNP-MoI3 complexes (31ad) were synthesized with phosphine ligands varied by different substituents such as isopropyl, tert-butyl, adamantyl, and phenyl groups (Scheme 16) [124]. The P(tert-butyl)2 group gave the highest NH3 production. The catalytic activity was further examined by changing the para-position of pyridine with MeO, Me, Ph, Fc, and Rc (32ae) with keeping the P(tert-butyl)2 moiety of PNP ligand. At the time, the para-Ph substituent showed the highest activity of 90 equiv. of NH3 production, which was much higher than the case of para-MeO substituent [125]. The trend was opposite from the case of N2-bridged dinuclear Mo complexes. Complex 31b exhibited a maximum 415 equiv. of NH3 production per catalyst, which was 35 times higher activity than the N2-bridged Mo2 complex. Furthermore, they examined the effect of PCP-type carbene ligands of 1,3-bis((di-tert-butylphosphino)methyl)benzimidazol-2-ylidene (PCP-1), 1,3-bis(2-(di-tert-butylphosphino)ethyl)imidazol-2-ylidene (PCP-2) [126] (Scheme 16). The PCP-1-Mo (33) produced 100 equiv. of NH3 per catalyst, but the PCP-2-Mo (34) showed a low reactivity of 1.6 equiv. of NH3 per catalyst. The identity of the reducing agent and proton source also affected the nitrogen reduction activity of Mo catalysts. Complex 33 showed the high activity when used together with SmI2 and H2O (or ethylene glycol) instead of CoCp* and [LutH]OTf. The combinatory use of SmI2 and ethylene glycol resulted in turnover frequency (TOF) of 7000 h−1 for the NH3 generation, and the condition using H2O as a proton source slightly decreased TOF to 6800 h−1 [127]. Using ethylene glycol as a proton source produced 22% (relative to a reductant) of H2 as a side product, but H2O reduced the ratio to ~2%. The optimized condition using H2O and SmI2 increased the NH3 productivity of complex 33 to 4350 equiv. per catalyst and TOF to 112.9 min−1, which was close to the nitrogenase activity of 40–120 min−1 [128].

3.3.2. Heterogeneous Mo-Containing N2-Reduction Electrocatalysts

Nitrogen reduction catalysts should deliver protons selectively to a bound nitrogen on their active sites, because the thermodynamic potential of a competitive proton reduction is relatively low. For this, the intrinsic property of an active metal center is an important factor to determine the N2 reduction reactivity of catalysts. Novel transition metals have shown nitrogen reactivities, but their limited reserves and high-cost issues are obstacles for their industrial scale-up as catalysts [129,130,131,132,133,134,135]. Earth-abundant metals are promising alternatives for nitrogen catalysts. Mo ions as the forms of MoS2, MoO3, MoN, Mo2N, and Mo2C were proven to be active for the N2 reduction. The specific affinity of Mo ion to nitrogen has been reported in the previous experiments [136,137]. In addition, the utilization of mixed metal elements could increase N2 reactivity and promote electron-transfer by tuning surface energy states.
The N2 reactivity seems to be varied by Mo crystal orientation. Wang and coworkers reported that Mo(110) plane has higher reactivity with N2 than other planes of (200) and (211) [138]. They prepared four types of Mo electrodes of Mo-foil, Mo-A-R, Mo-D-R-1h, and Mo-D-R-5h through electrodeposition, and Mo-D-R-5h showed the highest (110) orientation ratio based on XRD patterns [139,140,141].
Outermost Mo is an active site, but the reactivity and mechanism are determined by supporting elements. Sun and coworkers studied different reactivities of Mo, conjugated by O, S, and N atoms, by comparing MoO3, MoS2, and MoN [142,143,144]. Edge sites of Mo chalcogenides are known to be active for N2 reaction. MoN surface has shown ~0.2 lower overpotential than Mo chalcogenides. It is possibly understood that as a surface Mo-N is reduced to NH3, an empty Mo site facilitates subsequent N2 adsorption and reduction process.
Conductive support was also an important component to improve the electrocatalytic performance of MoS2, because the large conductive surface area supports more MoS2 active area as well as increases surficial electrical conductivity. Tian and coworkers achieved significant improvement of FE (10.94% at –0.3 V vs. RHE) of MoS2 using Ti3C2 MXene [145]. The catalytic efficiencies and reaction rates of the Mo-based electrocatalysts are compared in Table 2.

3.4. Nitrogenase-Electrode Biohybrid

In general, synthetic modeling studies focus on enzyme active sites, but protein structure should be an important factor to control the enzyme reaction. Even a small mutation of a nearby amino acid could distort the entire enzyme reactivity. For example, the substitution of a nearby β-98Tyr amino acid to β-98His deactivated the N2 reduction ability of the mutated FeMo-nitrogenase [146]. However, a reduction of hydrazine (N2H2) to NH3 still occurred in the mutated FeMo-nitrogenase. Interestingly, the mutation by β-98His improved electron-accepting ability from an un-natural reducing agent to promote reductions of azide (N3) or nitrite (NO2) to NH3 [147]. Using electron mediators such as polyaminocarboxylate-ligated Eu [126] and polyallylamine (PAA) polymer-CoCp2 [148] further improved reduction efficiencies. Rather in the presence of Fe-protein, without using an additional electron mediator, electron transfer between Fe-protein and FeMo-cluster became slow. Badalyan and coworkers showed experimentally that electron transfer between FeMo-cluster and Fe protein is the rate-determining step by calculating reaction rate constants using cyclic voltammetry data [149].
King and coworkers made a biohybrid system by attaching FeMo-nitrogenase to CdS nanorods for applications to a photocatalytic reduction of N2 to NH3 [128]. In the biological reaction, electron transfer steps from the P cluster to the FeMo-cluster require 16ATP (Em = −0.42 V) (ATP = adenosine triphosphate). In the bio-hybrid system, irradiation of CdS nanorod with 405 nm wavelength generated excited electrons with −0.8 eV, which was transferred to the FeMo-nitrogenase to proceed with the N2 reduction reaction. The TOF was 75 min−1 and the NH3 production rate was 315 ± 55 nmol(mg MoFe protein−1)min−1, which was 63% less activity compared to the enzyme activity with the presence of Fe protein and ATP. Upon using nitrogenase inhibitors such as acetylene, CO, and H2 [150,151,152], the enzyme activity became silent excluding a possible N2 reduction by CdS only.
Biohybrid of FeMo-nitrogenase with electrode enabled the enzyme to function in the absence of ATP, and thereby provided ways to study the enzyme reaction out of intracellular conditions. In future research, new biohybrid designs are required to decrease overpotential for the N2 reduction as well as to develop methodologies for the transformation of N2 to useful molecules.

4. H2 Evolution

The proton reduction is always a competitive reaction in the electrocatalytic reductions of CO2 and N2, because the 2H+/H2 reduction is thermodynamically favored. However, on the other hand, hydrogen molecules can be used as the simplest energy carrier for solar or electric energy conversion and storage [153]. Due to the importance, various transition metal-based electrocatalysts were developed for H2 evolution [154,155,156].

4.1. Homogeneous Mo Complexes for H2 Evolution

Nature does not use Mo metal for metabolic hydrogen reactions of H2 evolution and splitting, but hydrogenase active site inspired ligand design for Mo complexes. It was shown that anionic cyclopentadienyl (Cp) ligand has a similar electronic property as the Ni ligation sphere of the [NiFe]-hydrogenase active site [157]. Felton and Donovan reported that [(η5-C5H5)Mo(CO)3]2 (Scheme 17) (35) complex promoted the reduction of acetic acid in acetonitrile with 0.9 V overpotential [158]. Fan and Hu utilized polyhapto ligands to prepare Mo-carbonyl complexes of (η3-C3H5)Mo(CO)2(CH3CN)2Br (36), (η3-C3H5)Mo(CO)2(dppe)Br (37), (η5-C5H5)Mo(CO)3I (38), and (η5-C5H5)Mo(CO)2PPh3I (39) (Scheme 17) [159]. Complex 3639 showed similar reactivities of the halide dissociation during the cyclic voltammetry in MeCN solution and also exhibited similar catalytic activity of 16–27 TON for 5 h for the reduction of trifluoroacetic acid as requiring ~1 V overpotentials.
Hydrodesulfurization utilizes Mo-sulfides as a catalytic surface [160]. Similar to the MoS2 active site, Mo-sulfide moiety has been synthesized in the Mo coordination environment. DuBois and coworkers used Cp ligand to obtain a sulfide bridged dinuclear Mo complex, [(η5-C5H5)Mo(µ-S)]2(µ-S2CH2) (Scheme 17) (40). Complex 40 facilitated reduction of p-cyanoanilinium tetrafluoroborate at −0.7 V vs. Fc+/Fc in MeCN solution containing 0.3 M of Et4NBF4 electrolyte [161]. Controlled-potential electrolysis at −0.96 V vs. Fc+/Fc generated 10.6 mol of H2 gas per catalyst with 98 ± 5% FE. The catalytic activity was compared using different proton sources such as triflic acid (pKa = 2.6 in MeCN), p-cyanoanilinium tetrafluoroborate (7.6) and p-cyanoanilinium tetrafluoroborate (9.6). Higher pKa of the proton source shifted the proton reduction potential to the negative direction. Kinetic studies suggested that the elimination of H2 from the catalytic site is the rate-determining step.
Chang and coworkers synthesized the Mo(S2) moiety in a pentadentate polypyridyl Mo complex [(κ5-PY5Me2)MoS2]2+ (42) (PY5Me2 = 2,6-bis(1,1-bis(2-pyridyl)ethyl)pyridine) from a reaction of [(κ5-PY5Me2)Mo((κ1-CF3SO3)](CF3SO3) (41) with S8 (Scheme 18) [162]. Complex 42 generated H2 in acetate buffer (pH 3) at the onset potential of −0.58 V vs. SHE. Controlled potential electrolysis at −0.83 V overpotential produced H2 gas with ~100% FE and 280 mol/s TOF. Albeit less active than 42, Complex 41 also showed catalytic activity to produce H2 gas in 0.6 M phosphate buffer (pH 7) requiring 0.52 V overpotential. Controlled potential electrolysis at 0.64 V overpotential produced H2 gas as 1600 mol/h TOF per catalyst. Complex 41 also produced H2 gas in seawater at the onset potential of −0.81 V vs. SHE with 1200 mol/h TOF at −1.40 V vs. SHE. [163].
Streb and coworkers observed that partial substitution of terminal disulfide in [Mo3S13]2 by H2O to [Mo3S7(H2O)x](2−x)− increased the catalytic activity. Instead, the substitution of terminal disulfide with halides to [Mo3S7×6]2− (X = Cl, Br) decreased significantly the HER activity [164]. They suggested that, if terminal disulfide is protonated, energetically stable and catalytically inactive species are formed. It looks critical the formation of a vacant coordination site on a terminal Mo to generate catalytically active Mo-H.
Eisenberg and coworkers used dithiolene ligands to obtain Mo-sulfides moiety in MoL2(bdt)2 (43ag) (bdt = benzene-1,2-dithiol) (Scheme 19) [11]. Two-electron reduction of complexes 43ae in MeCN/H2O (9:1) solution dissociated isonitrile ligands to generate Mo(bdt)2, which was the active species for the catalytic HER. Complexes 43ag also showed photochemical HER in MeCN/H2O solution under the conditions of using [Ru(bpy)3]Cl2 as a photosensitizer, 0.2 M ascorbic acid (pH 4) as an electron donor. Complexes 43a gave the highest TON of 520 for 24 h than other complexes (<475 TON).
Moly-oxo complexes have shown interesting proton reduction activities. Zhan and coworkers synthesized cis di-oxo Mo complex, [MoVIL(O)2] (Scheme 20) (44) (L = 2-pyridylamino-N,N-bis(2-methylene-4-methoxy-6-tertbutylphenol)ion) [165]. Complex 44 promoted proton reduction at 0.25 M phosphate buffer (pH 7) recording 360 mol/h TOF (907 mV overpotential). Next, fluorine-substitution of [MoL’(O)2] (Scheme 20) (45) (L’ = 2-pyridylamino-N,N-bis(2-methylene-4,6-difluorophenol)ion) gave higher TOF of 756 mol/h (89% FE) at −1.61 V vs. Ag/AgCl in mixed acetonitrile/water (2:3) solution containing phosphate buffer (pH 6) [166].
Kim and coworkers reported the proton reactivity of [MoO(S2C2Ph2)2]2− (46) in MeCN solution. Treatment of protons (p-toluenesulfonic acid) with complex 46 in MeCN processed dehydration of the complex to give [Mo(MeCN)2(S2C2Ph2)2] (47) (Scheme 21) [167]. Complex 47 had poor solubility in polar solvents, and the fast dehydration of 46 prevented the use as an electrocatalyst. However, modification of the dithiolene ligand could make Mo ion function as HER electrocatalyst. Fontecave and coworkers utilized quinoxaline–pyran-fused dithiolene (qpdt2−), analogous to the molybdopterin (MPT), to prepare (Bu4N)2[MoO(qpdt)2] (48) (Scheme 22) [168]. The qpdt2− ligand assisted the Mo-oxo site to catalyze HER. Electrocatalytic proton reduction of trifluoroacetic acid was detected at −0.55 V vs. Ag/AgCl, and TOF of 1030 s−1 was obtained at −1.3 V vs. Ag/AgCl (FE 86% for 3 h). Complex 48 also facilitated photocatalytic HER with [Ru(bpy)3]2+ in 0.1 M ascorbic acid (pH 4) as recording TON 500 for 15 h.

4.2. Heterogeneous Mo-Containing H2-Evolution Electrocatalyst

In 2005, Norskov and coworkers calculated the free energy of atomic hydrogen bonding (∆G°H) to MoS2 surface and compared it with other catalysts such as enzymes (FeMo cofactor and NiFe hydrogenase) and metal surfaces of Au, Pt, Ni, and Mo [169]. Density functional theory (DFT) calculation results suggest that Ni and Mo bind strongly to atomic hydrogen; however, since the next proton/electron transfer steps are thermodynamically uphill, the H2–releasing step becomes slow as making those metals not suitable for HER. Interestingly, the metalloenzymes of FeMo-nitrogenase and NiFe-hydrogenase using Ni and Mo elements had similar hydrogen-binding energy (∆G°H ~ 0) as Pt. Plane MoS2 was inactive for HER, but the edge of MoS2 had a bit positive ∆G°H ~ 0.1 eV in a suitable range for HER. Graphite-supported MoS2 with enlarged edge area showed electrocatalytic HER with 0.1–0.2 V overpotential.
Chorkendorff and coworkers could modify the MoS2 edge area by varying sintering temperature (Figure 4a) [170]. The MoS2 prepared from 400 °C sintering had a longer edge length than another case of 550 °C sintering, and accordingly showed a higher HER current density of 3.1 × 10−7 A/cm2 compared to 1.3 × 10−7 A/cm2 of the latter. Albeit lower than the Pt electrode, the MoS2 edge exhibited higher current density than other common metal elements. They also showed photocatalytic HER activity of mixed MoSx surface as deposited on Ti|n+p-Si photocathode [171]. The MoSx was electrodeposited on Ti-protected n+p-Si electrode by cyclic voltammetry (CV) scans at 0.137–1.247 V vs. RHE. The MoSx|Ti|n+p-Si photocathode showed HER activity above 0.33 V vs. RHE under the illumination of red light.
The electrodeposition method seems to be not suitable to produce a large number of catalysts, and unstable species may not persist during cyclic voltammetry scans. Hu and coworkers reported that unsaturated MoSx provides a more active site for HER [172]. They prepared MoSx film by electro-polymerization method (Figure 4b), which exhibited higher catalytic activity than MoS2 single crystal and MoS2 nanoparticles. They explained that is because amorphous MoSx increases sulfur defects at the surface [173]. The same group investigated the effects of growth mechanism and catalyst mass, but they reported that the catalyst mass dominantly affected the catalytic efficiency (Figure 4c) [174].
Mo sulfur clusters have shown similar active sites as MoS2. Chorkendorff and coworkers adsorbed cubane type [Mo3S4]4+ cluster to highly oriented pyrolytic graphite [175]. [Mo3S4]4+ cluster showed similar overpotential as MoS2 nanoparticle. Even higher TOF of 0.07 s−1 was measured than 0.02 s−1 of MoS2 nanoparticle, but it was unstable during successive potential scanning.
Bensenbacher and coworkers synthesized [Mo3S13]2− nanocluster by a wet chemical method of (NH4)6Mo7O24 and ammonium polysulfide [176]. [Mo3S13]2− showed HER activity at 0.1 V onset potential and 0.18 V overpotential at 10 mA/cm2 (TOF 3 s−1). Li and coworkers synthesized Mo3S13 film by electrodeposition, which showed a higher catalytic rate and lower overpotential by ~40 mV than Mo3S13 film prepared by drop-casting of (NH4)2Mo3S13·2H2O [177]. The increased electrical conductivity enhanced catalytic efficiency. Xu and coworkers improved HER efficiency by attaching [Mo3S13]2− to highly conductive reduced graphene oxide-carbon nanotube (rGO-CNTs) aerogels, which recorded a lower overpotential by 74 mV than pure [Mo3S13]2− [178]. Wu and coworkers reported that dimeric [Mo2S12]2− cluster slightly improved the HER activity by recording 0.16 V overpotential at 10 mA/cm2 (TOF ~3 s−1) compared to [Mo3S13]2− [179] They reported hydrogen adsorption free energy △Gads(H) of [Mo2S12]2− (−0.05 eV) is closer to zero than other surfaces such as [Mo3S13]2− (−0.08 eV), MoS2 (0.08 eV), and Pt(111)(−0.09 eV).
Min and coworkers reported HER activity of [Mo3S13]2− using [Ru(bpy)3]Cl2 and ascorbic acid under visible light (≥420 nm). Initial TOF of 335 h−1 was measured, but it had a problem of gradual decomposition [180]. The photocatalytic stability of [Mo3S13]2− was improved by Zang and coworkers [181]. They encapsulated [Mo3S13]2− in ethidium bromide covalent organic frameworks (EB-COFs). The COF-encapsulation did not decrease significantly the catalytic activity relative to a free [Mo3S13]2− cluster. Mo3S13@EB-COF maintained photocatalytic HER for 8 h by recording 21,465 μmol g−1 h−1 rate, furthermore which showed the stable catalytic performance during the recycling test of 4 times of 5 h experiment.
Artero and coworkers studied the structure and reaction mechanism of amorphous MoSx (α-MoSx), which had polymer-based on [Mo3S13]2− cluster sharing disulfide ligand [182]. They suggested that the reaction of terminal disulfide with proton and electron releases HS to generate unsaturated MoIV active site.
Octahedral molybdenum clusters are red near-IR phosphorescent emitters and showed high photocatalytic HER activity. Feliz and coworkers studied the catalytic performance of the (TBA)2[Mo6Bri8Fa6] (TBA = tetra-n-butylammonium) cluster in aqueous solution in the presence of triethylamine (TEA) [183]. The catalytic activity of the {Mo6Bri8}4+ cluster unit was enhanced by the in situ generation of the [Mo6Bri8Fa5(OH)a]2− and [Mo6Bri8Fa3(OH)a3]2− and [Mo6Bri8(OH)a6]2− species. In the same work, the cluster unit was coordinatively immobilized onto graphene oxide (GO) surfaces. The resulting material, (TBA)2Mo6Bri8@GO, enhanced the cluster stability in a water/MeOH mixture and under photoirradiation. Its catalytic performance was superior to that of GO, it decreased with respect to that of the molecular cluster complex. The TOF values with respect to atomic molybdenum were 5 × 10−6 s−1 and 3 × 10−4 s−1 for the heterogeneous and the homogeneous materials, respectively. Recently, the (TBA)2[Mo6Ii8(O2CCH3)a6] was also coordinatively immobilized onto GO to give (TBA)2Mo6Ii8@GO [184]. To assure the cluster stability of the cluster units under photocatalytic conditions, both materials were tested in vapor water photoreduction. Even after longer radiation exposure times, the catalysts remained stable and recyclability of both catalysts was demonstrated. The TOF of (TBA)2Mo6Ii8@GO is three times higher than that of the microcrystalline (TBA)2[Mo6Ii8(O2CCH3)a6], in agreement with the better accessibility of catalytic cluster sites for water molecules in the gas phase. In the framework of the study of the photocatalytic properties of the {Mo6I8}4+ cluster units, one of the most emissive octahedral metal clusters, [Mo6Ii8(OCOC2F5)a6]2−, was immobilized onto graphene sheets through pyrene-containing organic cations as supramolecular linkers [185]. These non-covalent interactions enhanced the photocatalytic activity of the nanocomposite by 280% with respect to the molecular cluster and graphene counterparts. The improvement of the H2 production activity is attributed to the synergetic effect between graphene and the hybrid cluster complex because graphene facilitates the charge-separation activity and enhances electron transfer of the cluster photocatalyst.
Lana-Villarreal and coworkers reported photocatalytic HER activity of Mo3S7 cluster as immobilized on TiO2 surface [186]. Functionalized bipyridyl ligand of Mo3S7Br4(diimino) cluster enabled adsorption of the cluster on TiO2 surface. The immobilized Mo3S7 cluster decreased the HER overpotential by 0.3 V as recording 1.4 s−1 TOF. Despite the moderate catalytic activity, it gave an example to immobilize molecular cluster on electrode surface.
Fabrication of hetero-metal chalcogenides was an effective method to improve catalytic activities of Mo-based electrode materials. Doping of metal promoters possibly improves the intrinsic activity of unsaturated Mo sites. The electrocatalytic effect of Cu/Mo hetero-metals was examined by Tran and coworkers [187]. Cu2MoS4 was synthesized by the solvothermal reaction of [Cu(CH3CN)](BF4) and (NH4)2[MoS4], and the obtained Cu2MoS4 crystals were stable under air for weeks.
Both MoS2 and CoSe2 were active for HER, and the synergistic effect of combined MoS2 and CoSe2 was examined by Gao and coworkers [188]. MoS2/CoSe2 showed the HER activity close to commercial Pt/C. Interaction of the first-row transition metal Co with S could form S2− and S22− states, which possibly assisted MoS2 growth.
MoSe2 was relatively less studied compared to MoS2. Since the atomic hydrogen adsorption energy was measured to be lower with MoSe2 than that of MoS2, MoSe2 also showed an efficient HER activity [189]. Sasaki and coworkers improved corrosion stability of Ni/Mo-alloy by NiMoNx nanosheet, where metal stabilizing effect of nitride possibly improved the stability [190]. Mo phosphide (MoP) is a known hydrodesulfurization catalyst. Jaramillo and coworkers compared catalytic HER activity of crystalline MoP and molybdenum phosphosulfide (MoP/S) [191]. Both the MoP and MoP/S were active for HER, but MoP/S showed the higher electrocatalytic activity than MoP.
Electrocatalytic HER activities of commercial MoB and Mo2C have been reported [192]. Two electrode materials showed similar catalytic activity and were usable under both acidic and basic conditions, exhibiting similar activity and durability in continuous electrolysis for 48 h.
Leonard and coworkers reported the synthesis of various Mo-carbide materials such as α-MoC1−x, β-Mo2C, η-MoC, and γ-MoC with different stacking sequences, and compared their HER activities [193]. The β-Mo2C, η-MoC, and γ-MoC had similar hexagonal crystal structure, but the stacking sequence of β-Mo2C was ABAB, η-MoC was ABCABC, and γ-MoC was AAAA packing. α-MoC1-x had a cubic structure as ABCABC stacking sequence. The β-Mo2C showed the highest HER activity, and the catalytic activities of the others were obtained in the order of: γ-MoC > η-MoC > α-MoC1−x. The γ-MoC showed stable catalytic activity with −1.95 mA/cm2 at −340 mV vs. RHE in the continuous electrolysis for 18 h.
Nakanishi and coworkers reported the synthesis of Mo carbonitride (MoCN) nanomaterial [194]. The reaction of Na2MoO4 with diaminopyridine (DAP) in acidic condition gave DAP-bound Mo oxide, which was polymerized by NaHCO3/(NH4)2S2O8 as emitting CO2 gas. The CO2 emission step formed nano-sized (PDAP)-2H+/MoO42− (PDAP = polydiaminopyridine) complex causing a high density of the catalytically active site. Finally, the polymer was pyrolyzed at 800 °C to become MoCN nanomaterial (Figure 4f). MoCN nanomaterial showed the HER activity above −0.05 V vs. RHE and reached 10 mA/cm2 at −0.14 V vs. RHE in sulfuric acid (pH 1) solution. MoCN nanomaterial had electrocatalytic stability during 1000 CV scans between −0.35 and 0.2 V vs. RHE.
Yu and coworkers used hybrid phosphorous-doped nanoporous carbon (PC) and RGO to support MoO2 [195]. MoO2@PC-RGO had a carbon skeleton, which prevented aggregation of catalytically active MoO2 nanoparticles. Additionally, the synergistic combination of PC and RGO assisted to enhance the catalytic activity.
The use of adhesives for power-deposition on electrode generally lowers catalytic efficiency, so nickel foam (NF) has been used to prepare self-supported electrode. Huang and coworkers used NF to support Mo2N/CeO2 hetero-nanoparticles (Figure 4d) [196]. CeO2 was coated on NF by hydrothermal method, and the CeO2-fabricated NF was covered by Mo2N with varying amounts of (NH4)4Mo7O24·4H2O (0.02, 0.04, 0.05, 0.06 mmol) precursor by solvothermal method. Post-annealing at 500 °C for 5 h under NH3 atmosphere afforded Mo2N/CeO2@NF. The synthetic condition using 0.05 mmol Mo source gave the best HER activity. Mo2N/CeO2@NF-0.05 sample showed smaller overpotential than commercial 20 wt% Pt/C. The high catalytic activity was probably because electronic interaction between Mo2N and CeO2 lowers the energy barrier for hydrogen intermediate formation (Figure 4e). The catalytic efficiencies of the above Mo-based HER electrocatalysts are compared in Table 3.

5. Heterogeneous Mo-Containing O2-Evolution Electrocatalysts

Water splitting is an ideal method for the conversion of electric and solar energy to the chemical bond energy of dihydrogen. However, oxidation of water to O2, the counterpart reaction of H2 evolution, is kinetically slow, which decreases the overall efficiency of the water-splitting reaction. Since oxygen evolution reaction (OER) is a multi-electron process demanding large overpotential, efficient electrocatalysts are highly desired. RuO2 is a benchmark catalyst for the OER, but the high cost and easy decomposition to RuO4 are drawbacks of using Ru. Research efforts have been made to develop Earth-abundant metal-based electrocatalysts with high-performance and durability, and, among those, Mo-containing electrocatalysts have shown promising OER activities.
MoS2 edge site is also known to be active for the OER, and, thus, various synthetic methods were developed to enlarge the edge site area. Mohanty and coworkers investigated the OER activity of MoS2 quantum dots [197]. The MoS2 quantum dot compounds (MSQDs) of 2–5 nm size were synthesized without aggregation through single-step hydrothermal reaction using (NH4)2MoS4) precursor (Figure 5a,b). Repeating linear sweep voltammetry (LSV) scans up to 50 cycles increased the OER current density, because the electrochemically active surface area of MSQDs was enlarged with the potential scans.
Direct growing methods, useful to increase contact, conductivity, and electron transfer, are generally applied to prepare OER electrocatalysts. In 2016, Yan and Lu prepared a porous MoS2 microsphere on nickel foam (NF) [198]. The MoS2 on NF showed a higher catalytic rate (105 mV/dec Tafel slope) than commercial RuO2 on NF (127 mV/dec) and 20% Pt/C (150 mV dec−1). Cui and coworkers synthesized mesoporous MoO2 nanosheets on NF [199]. The mesoporosity of MoO2 enlarged the active surface area and slightly improved the catalytic activity than that of compact MoO2.
Boride increases a reverse electron transfer to the metal site, which accordingly increases electron density on the catalytic site. Gupta and coworkers synthesized ternary Co-Mo-B hetero-metal nanocomposite, which gave enhanced OER activity [200].
Li and coworkers used a well-defined Cu nanowire to deposit Ni/Mo-alloy catalyst [201]. Alkaline anodization and subsequent cathodic reduction process modified the surface of Cu foam. A mixture of Ni(SO3NH2)2 and Na2MoO4 were electrodeposited on the Cu nano-scaffold to obtain Ni-Mo/Cu nanowire. The Ni-Mo/Cu nanowire exhibited comparable OER activity with commercial Pt/C(RuO2) on Cu foam.
Yang and coworkers examined the synergistic effect of trimetallic FeCoMo nanocomposite for the OER test [202]. Amorphous FeCoMo nanocomposite was synthesized as a rod-like shape by hydrolysis of a mixture of CoCl2∙6H2O, FeCl3∙6H2O and (NH4)6Mo7O24∙4H2O, hexamethylene tetramine at 90 °C. The existence of high-valent Mo6+ was characterized by Raman spectroscopy. In situ X-ray absorption near edge structure (XANES) showed that Mo6+ tends to attract electrons from 3d transition metals, which possibly promoted the OER process.
Gao and coworkers prepared a bifunctional catalyst using MoS2 HER catalyst and Ni3S2 OER catalyst (Figure 5c–e) [203]. MoS2−Ni3S2 heteronanorods on NF afforded higher OER activity than Ni3S2/NF and MoS2/NF.
The bimetallic system enables manipulation of electronic structure and gives abundant active sites, but a synthesis of pure phase is difficult. Lan and coworkers could make pure-phase bimetallic Co/Mo carbides using mesoporous carbon substrate (Figure 5f) [204]. Transition metal carbide also gave the high catalytic activity because of the good electrical conductivity, metallic property, and chemical stability.
Ni oxyhydroxide is an active site for OER, but the low stability decreases the catalytic efficiency. Shen and coworkers increased the stability by fabricating MoFe:Ni(OH)2/NiOOH nanosheet and studied synergistic effect [205].
Ma and coworkers synthesized one-dimensional MoO2-Co2Mo3O8@C nanorods with the enlarged catalytic surface area [206]. The ZIF-67 MOF was attached to MoO3 nanorods to form MoO3@ZIF-67, and subsequent hydrothermal reaction at 700 °C gave MoO2-Co2Mo3O8@C nanorods (containing Co/Mo in 9.1/10.5% ratio). The organic ligand of ZIF-67 was carbonized into carbon, which served as a reductant to reduce excess MoO3 to MoO2. The MoO2-Co2Mo3O8@C nanorods showed similar activity as the benchmark RuO2.
Yang and coworkers studied the OER activity of Ni2Mo3N hetero-metal nitride [207]. They reported that an amorphous surface oxygen-rich activation layer (SOAL) was generated on Ni2Mo3N nanoparticle surface by applying high oxidative potential in alkaline conditions. The computational study suggested that SOAL could increase the OER activity, because the conjugation of Ni-nitrides active sites and the secondary Mo-electron pump promoted the catalytic reaction.
Jiang and coworkers reported the synergistic effect of Co3Mo alloy nanoparticles as attached to nanoporous Cu skeleton [208]. They synthesized Mo-doped Co3O4 nanoflakes on CuO/Cu skeleton by electro-oxidation (EO) at 1.57 V vs. RHE. The EO Co3Mo/Cu showed lower OER onset overpotential than Ir/C supported by nanoporous Cu. Since Co3Mo/Cu also acted as a highly active HER catalyst, they composed a water electrolysis set-up by Co3Mo/Cu cathode and EO Co3Mo/Cu anode. The Co3Mo/Cu-based electrolyzer showed overall water splitting delivering a current density of 100 mA/cm2 at 1.62 V vs. RHE. The water-splitting system acted more efficiently than Pt/C/Cu-Ir/C/Cu set-up (Figure 5g,h). The catalytic efficiency parameters of the Mo-based OER electrocatalysts are compared in Table 4.

6. Conclusions

This review summarizes Mo-containing metalloenzymes and their model complexes, homogeneous and heterogeneous catalysts under categories of reactivities with CO2, N2, H2, and O2. Compared to the Mo enzymes, catalytic activities of synthetic systems remain at a low-efficiency level. Thus, continuous research efforts in synthetic and theoretical perspectives are requested to achieve high-performing catalysts. The revealed Mo-enzyme structures enabled synthetic research on the enzyme reactivities, and concomitant theoretical studies gave an in-depth understanding of the enzyme mechanism. However, previously proposed mechanisms need further supports by synthetic experiments, and spectroscopic research on living cells still has limitations; thus, more active modeling studies are required.
Along with the modeling studies, new catalysts by ligand design should be developed to find an optimal condition to prepare active Mo ion. Ligand design for a low-valent Mo ion to function under lower gas pressure is pursued. Examples of homogeneous catalysts using high-valent Mo ions are relatively rare. Mimicking the enzyme active site partially or conceptually could provide synthetic ideas to develop both low- and high-valent Mo-based catalysts. In addition, catalytic stability of the Mo complex is a prerequisite for commercialization. Utilization of hetero-metallic surface is a method of preventing deactivation of catalytic sites as well as providing synergistic effects. Another method would be to use polymer or organic scaffold for regulating access of reagents into a catalytic site, which is similar to the enzymatic strategy to protect the active site by surrounding it with polypeptide chains.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2020R1C1C1007106) and GIST Research Institute (GRI) grant funded by the GIST in 2020.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. van der Ham, C.J.M.; Koper, M.T.M.; Hetterscheid, D.G.H. Challenges in reduction of dinitrogen by proton and electron transfer. Chem. Soc. Rev. 2014, 43, 5183–5191. [Google Scholar] [CrossRef] [PubMed]
  2. John, J.; Lee, D.-K.; Sim, U. Photocatalytic and electrocatalytic approaches towards atmospheric nitrogen reduction to ammonia under ambient conditions. Nano Converg. 2019, 6, 15. [Google Scholar] [CrossRef] [Green Version]
  3. He, M.; Sun, Y.; Han, B. Green Carbon Science: Scientific Basis for Integrating Carbon Resource Processing, Utilization, and Recycling. Angew. Chem. Int. Ed. 2013, 52, 9620–9633. [Google Scholar] [CrossRef]
  4. Zou, X.; Zhang, Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chem. Soc. Rev. 2015, 44, 5148–5180. [Google Scholar] [CrossRef] [PubMed]
  5. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar Water Splitting Cells. Chem. Rev. 2010. [Google Scholar] [CrossRef] [PubMed]
  6. Rosca, V.; Duca, M.; de Groot, M.T.; Koper, M.T.M. Nitrogen Cycle Electrocatalysis. Chem. Rev. 2009, 109, 2209–2244. [Google Scholar] [CrossRef] [PubMed]
  7. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef]
  8. Janani, G.; Choi, H.; Surendran, S.; Sim, U. Recent advances in rational design of efficient electrocatalyst for full water splitting across all pH conditions. MRS Bull. 2020, 45, 539–547. [Google Scholar] [CrossRef]
  9. Andreini, C.; Bertini, I.; Cavallaro, G.; Holliday, G.L.; Thornton, J.M. Metal ions in biological catalysis: From enzyme databases to general principles. J. Biol. Inorg. Chem. 2008, 13, 1205–1218. [Google Scholar] [CrossRef]
  10. Bullock, R.M.; Chen, J.G.; Gagliardi, L.; Chirik, P.J.; Farha, O.K.; Hendon, C.H.; Jones, C.W.; Keith, J.A.; Klosin, J.; Minteer, S.D.; et al. Using nature’s blueprint to expand catalysis with Earth-abundant metals. Science 2020, 369. [Google Scholar] [CrossRef]
  11. Eckenhoff, W.T.; Brennessel, W.W.; Eisenberg, R. Light-Driven Hydrogen Production from Aqueous Protons using Molybdenum Catalysts. Inorg. Chem. 2014, 53, 9860–9869. [Google Scholar] [CrossRef]
  12. Dos Santos, P.C.; Dean, D.R.; Hu, Y.; Ribbe, M.W. Formation and Insertion of the Nitrogenase Iron−Molybdenum Cofactor. Chem. Rev. 2004, 104, 1159–1174. [Google Scholar] [CrossRef]
  13. Dobbek, H. Structural aspects of mononuclear Mo/W-enzymes. Coord. Chem. Rev. 2011, 255, 1104–1116. [Google Scholar] [CrossRef]
  14. Cordas, C.M.; Moura, J.J.G. Molybdenum and tungsten enzymes redox properties—A brief overview. Coord. Chem. Rev. 2019, 394, 53–64. [Google Scholar] [CrossRef]
  15. Lin, Y.-W. Rational Design of Artificial Metalloproteins and Metalloenzymes with Metal Clusters. Molecules 2019, 24, 2743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Zhu, J.; Hu, L.; Zhao, P.; Lee, L.Y.S.; Wong, K.-Y. Recent Advances in Electrocatalytic Hydrogen Evolution Using Nanoparticles. Chem. Rev. 2020. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, L.; Liu, P.; Li, J.; Xiang, B. Two-Dimensional Material Molybdenum Disulfides as Electrocatalysts for Hydrogen Evolution. Catalysts 2017, 7, 285. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, S.; Fan, Q.; Xia, R.; Meyer, T.J. CO2 Reduction: From Homogeneous to Heterogeneous Electrocatalysis. Acc. Chem. Res. 2020, 53, 255–264. [Google Scholar] [CrossRef]
  19. Axley, M.J.; Grahame, D.A. Kinetics for formate dehydrogenase of Escherichia coli formate-hydrogenlyase. J. Biol. Chem. 1991, 266, 13731–13736. [Google Scholar] [CrossRef]
  20. Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.; Hille, R.; Kenis, P.J.A.; et al. Frontiers, Opportunities, and Challenges in Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 6621–6658. [Google Scholar] [CrossRef] [Green Version]
  21. Hille, R.; Hall, J.; Basu, P. The Mononuclear Molybdenum Enzymes. Chem. Rev. 2014, 114, 3963–4038. [Google Scholar] [CrossRef] [Green Version]
  22. Hartmann, T.; Leimkühler, S. The oxygen-tolerant and NAD+-dependent formate dehydrogenase from Rhodobacter capsulatus is able to catalyze the reduction of CO2 to formate. FEBS J. 2013, 280, 6083–6096. [Google Scholar] [CrossRef]
  23. Maia, L.B.; Fonseca, L.; Moura, I.; Moura, J.J.G. Reduction of Carbon Dioxide by a Molybdenum-Containing Formate Dehydrogenase: A Kinetic and Mechanistic Study. J. Am. Chem. Soc. 2016, 138, 8834–8846. [Google Scholar] [CrossRef] [PubMed]
  24. Boyington, J.C.; Gladyshev, V.N.; Khangulov, S.V.; Stadtman, T.C.; Sun, P.D. Crystal structure of formate dehydrogenase H: Catalysis involving Mo, molybdopterin, selenocysteine, and an Fe4S4 cluster. Science 1997, 275, 1305–1308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Raaijmakers, H.C.A.; Romao, M.J. Formate-reduced E-coli formate dehydrogenase H: The reinterpretation of the crystal structure suggests a new reaction mechanism. J. Biol. Inorg. Chem. 2006, 11, 849–854. [Google Scholar] [CrossRef]
  26. Mota, C.S.; Rivas, M.G.; Brondino, C.D.; Moura, I.; Moura, J.J.G.; González, P.J.; Cerqueira, N.M.F.S.A. The mechanism of formate oxidation by metal-dependent formate dehydrogenases. J. Biol. Inorg. Chem. 2011, 16, 1255–1268. [Google Scholar] [CrossRef] [Green Version]
  27. Cerqueira, N.M.F.S.A.; Fernandes, P.A.; Gonzalez, P.J.; Moura, J.J.G.; Ramos, M.J. The Sulfur Shift: An Activation Mechanism for Periplasmic Nitrate Reductase and Formate Dehydrogenase. Inorg. Chem. 2013, 52, 10766–10772. [Google Scholar] [CrossRef] [PubMed]
  28. Schrapers, P.; Hartmann, T.; Kositzki, R.; Dau, H.; Reschke, S.; Schulzke, C.; Leimkuhler, S.; Haumann, M. Sulfido and Cysteine Ligation Changes at the Molybdenum Cofactor during Substrate Conversion by Formate Dehydrogenase (FDH) from Rhodobacter capsulatus. Inorg. Chem. 2015, 54, 3260–3271. [Google Scholar] [CrossRef]
  29. Duffus, B.R.; Schrapers, P.; Schuth, N.; Mebs, S.; Dau, H.; Leimkuhler, S.; Haumann, M. Anion Binding and Oxidative Modification at the Molybdenum Cofactor of Formate Dehydrogenase from Rhodobacter capsulatus Studied by X-ray Absorption Spectroscopy. Inorg. Chem. 2020, 59, 214–225. [Google Scholar] [CrossRef]
  30. Hille, R.; Dingwall, S.; Wilcoxen, J. The aerobic CO dehydrogenase from Oligotropha carboxidovorans. J. Biol. Inorg. Chem. 2015, 20, 243–251. [Google Scholar] [CrossRef]
  31. Dobbek, H.; Gremer, L.; Kiefersauer, R.; Huber, R.; Meyer, O. Catalysis at a dinuclear [CuSMo(=O)OH] cluster in a CO dehydrogenase resolved at 1.1-A resolution. Proc. Natl. Acad. Sci. USA 2002, 99, 15971–15976. [Google Scholar] [CrossRef] [Green Version]
  32. Gnida, M.; Ferner, R.; Gremer, L.; Meyer, O.; Meyer-Klaucke, W. A Novel Binuclear [CuSMo] Cluster at the Active Site of Carbon Monoxide Dehydrogenase:  Characterization by X-ray Absorption Spectroscopy. Biochemistry 2003, 42, 222–230. [Google Scholar] [CrossRef] [PubMed]
  33. Siegbahn, P.E.M.; Shestakov, A.F. Quantum chemical modeling of CO oxidation by the active site of molybdenum CO dehydrogenase. J. Comput. Chem. 2005, 26, 888–898. [Google Scholar] [CrossRef]
  34. Hofmann, M.; Kassube, J.K.; Graf, T. The mechanism of Mo-/Cu-dependent CO dehydrogenase. J. Biol. Inorg.Chem. 2005, 10, 490–495. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, B.; Hemann, C.F.; Hille, R. Kinetic and spectroscopic studies of the molybdenum-copper CO dehydrogenase from Oligotropha carboxidovorans. J. Biol. Chem. 2010, 285, 12571–12578. [Google Scholar] [CrossRef] [Green Version]
  36. Seo, J.; Kim, E. O-Atom Exchange between H2O and CO2 Mediated by a Bis(dithiolene)tungsten Complex. Inorg. Chem. 2012, 51, 7951–7953. [Google Scholar] [CrossRef] [PubMed]
  37. Lim, B.S.; Holm, R.H. Bis(Dithiolene)molybdenum Analogues Relevant to the DMSO Reductase Enzyme Family:  Synthesis, Structures, and Oxygen Atom Transfer Reactions and Kinetics. J. Am. Chem. Soc. 2001, 123, 1920–1930. [Google Scholar] [CrossRef]
  38. Sung, K.-M.; Holm, R.H. Oxo Transfer Reactions Mediated by Bis(dithiolene)tungsten Analogues of the Active Sites of Molybdoenzymes in the DMSO Reductase Family:  Comparative Reactivity of Tungsten and Molybdenum. J. Am. Chem. Soc. 2001, 123, 1931–1943. [Google Scholar] [CrossRef] [PubMed]
  39. Seo, J.; Shearer, J.; Williard, P.G.; Kim, E. Reactivity of a biomimetic W(iv) bis-dithiolene complex with CO2 leading to formate production and structural rearrangement. Dalton Trans. 2019, 48, 17441–17444. [Google Scholar] [CrossRef] [PubMed]
  40. Fogeron, T.; Todorova, T.K.; Porcher, J.-P.; Gomez-Mingot, M.; Chamoreau, L.-M.; Mellot-Draznieks, C.; Li, Y.; Fontecave, M. A Bioinspired Nickel(bis-dithiolene) Complex as a Homogeneous Catalyst for Carbon Dioxide Electroreduction. ACS Catal. 2018, 8, 2030–2038. [Google Scholar] [CrossRef] [Green Version]
  41. Fogeron, T.; Retailleau, P.; Chamoreau, L.-M.; Li, Y.; Fontecave, M. Pyranopterin Related Dithiolene Molybdenum Complexes as Homogeneous Catalysts for CO2 Photoreduction. Angew. Chem. Int. 2018, 57, 17033–17037. [Google Scholar] [CrossRef] [PubMed]
  42. Mouchfiq, A.; Todorova, T.K.; Dey, S.; Fontecave, M.; Mougel, V. A bioinspired molybdenum–copper molecular catalyst for CO2 electroreduction. Chem. Sci. 2020, 11, 5503–5510. [Google Scholar] [CrossRef] [PubMed]
  43. Chisholm, M.H.; Cotton, F.A.; Extine, M.W.; Reichert, W.W. The molybdenum-molybdenum triple bond Insertion reactions of hexakis(alkoxy)dimolybdenum compounds with carbon dioxide and single-crystal x-ray structural characterization of bis(tert-butylcarbonato)tetrakis(tert-butoxy)dimolybdenum. J. Am. Chem. Soc. 1978, 100, 1727–1734. [Google Scholar] [CrossRef]
  44. Alvarez, R.; Carmona, E.; Cole-Hamilton, D.J.; Galindo, A.; Gutierrez-Puebla, E.; Monge, A.; Poveda, M.L.; Ruiz, C. Formation of acrylic acid derivatives from the reaction of carbon dioxide with ethylene complexes of molybdenum and tungsten. J. Am. Chem. Soc. 1985, 107, 5529–5531. [Google Scholar] [CrossRef]
  45. Chakraborty, S.; Blacque, O.; Berke, H. Ligand assisted carbon dioxide activation and hydrogenation using molybdenum and tungsten amides. Dalton Trans. 2015, 44, 6560–6570. [Google Scholar] [CrossRef] [Green Version]
  46. Man, M.L.; Zhou, Z.; Ng, S.M.; Lau, C.P. Synthesis, characterization and reactivity of heterobimetallic complexes (η5-C5R5)Ru(CO)(μ-dppm)M(CO)25-C5H5) (R = H, CH3; M = Mo, W). Interconversion of hydrogen/carbon dioxide and formic acid by these complexes. Dalton Trans. 2003, 3727–3735. [Google Scholar] [CrossRef]
  47. Minato, M.; Zhou, D.-Y.; Sumiura, K.-I.; Oshima, Y.; Mine, S.; Ito, T.; Kakeya, M.; Hoshino, K.; Asaeda, T.; Nakada, T.; et al. Reactivity Patterns of O2, CO2, Carboxylic Acids, and Triflic Acid with Molybdenum Silyl Hydrido Complexes Bearing Polydentate Phosphinoalkyl–Silyl Ligands: Pronounced Effects of Silyl Ligands on Reactions. Organometallics 2012, 31, 4941–4949. [Google Scholar] [CrossRef]
  48. McWeeny, R.; Mason, R.; Towl, A.D.C. The geometries of and bonding in certain transition metal complexes. Discuss. Faraday Soc. 1969, 47, 20–26. [Google Scholar] [CrossRef]
  49. Haszeldine, R.N.; Parish, R.V.; Setchfield, J.H. Organosilicon chemistry: XI. The stereochemistry of Ir(H)Cl(SiR3)CO(PPh3)2, and the trans-influence of substituted silyl, germyl, and stannyl groups. J. Organomet. Chem. 1973, 57, 279–285. [Google Scholar] [CrossRef]
  50. Mukhopadhyay, T.K.; Flores, M.; Groy, T.L.; Trovitch, R.J. A Highly Active Manganese Precatalyst for the Hydrosilylation of Ketones and Esters. J. Am. Chem. Soc. 2014, 136, 882–885. [Google Scholar] [CrossRef]
  51. Pal, R.; Groy, T.L.; Trovitch, R.J. Conversion of Carbon Dioxide to Methanol Using a C–H Activated Bis(imino)pyridine Molybdenum Hydroboration Catalyst. Inorg. Chem. 2015, 54, 7506–7515. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, Y.; Hanna, B.S.; Dineen, A.; Williard, P.G.; Bernskoetter, W.H. Functionalization of Carbon Dioxide with Ethylene at Molybdenum Hydride Complexes. Organometallics 2013, 32, 3969–3979. [Google Scholar] [CrossRef]
  53. Fong, L.K.; Fox, J.R.; Cooper, N.J. Reactions of carbon dioxide with the electron-rich polyhydride complex [Mo(dmpe)2H4]. Organometallics 1987, 6, 223–231. [Google Scholar] [CrossRef]
  54. Hor, T.S.A.; Chee, S.-M. Substituted metal carbonyls: III. Chromium, molybdenum and tungsten tricarbonyl complexes containing bipyridyl and a unidentate diphosphine: Facile synthesis via trimethylamine N-oxide-induced decarbonylations. J. Organomet. Chem. 1987, 331, 23–28. [Google Scholar] [CrossRef]
  55. Clark, M.L.; Grice, K.A.; Moore, C.E.; Rheingold, A.L.; Kubiak, C.P. Electrocatalytic CO2 reduction by M(bpy-R)(CO)4 (M = Mo, W, R = H, tBu) complexes. Electrochemical, spectroscopic, and computational studies and comparison with group 7 catalysts. Chem. Sci. 2014, 5, 1894–1900. [Google Scholar] [CrossRef] [Green Version]
  56. Zhang, Y.; MacIntosh, A.D.; Wong, J.L.; Bielinski, E.A.; Williard, P.G.; Mercado, B.Q.; Hazari, N.; Bernskoetter, W.H. Iron catalyzed CO2 hydrogenation to formate enhanced by Lewis acid co-catalysts. Chem. Sci. 2015, 6, 4291–4299. [Google Scholar] [CrossRef] [Green Version]
  57. Zhang, Y.; Williard, P.G.; Bernskoetter, W.H. Synthesis and Characterization of Pincer-Molybdenum Precatalysts for CO2 Hydrogenation. Organometallics 2016, 35, 860–865. [Google Scholar] [CrossRef]
  58. Summers, D.P.; Leach, S.; Frese, K.W. The electrochemical reduction of aqueous carbon dioxide to methanol at molybdenum electrodes with low overpotentials. J. Electroanal. Chem. 1986, 205, 219–232. [Google Scholar] [CrossRef]
  59. Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B.A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R.; et al. Robust carbon dioxide reduction on molybdenum disulphide edges. Nat. Commun. 2014, 5, 4470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Abbasi, P.; Asadi, M.; Liu, C.; Sharifi-Asl, S.; Sayahpour, B.; Behranginia, A.; Zapol, P.; Shahbazian-Yassar, R.; Curtiss, L.A.; Salehi-Khojin, A. Tailoring the Edge Structure of Molybdenum Disulfide toward Electrocatalytic Reduction of Carbon Dioxide. ACS Nano 2017, 11, 453–460. [Google Scholar] [CrossRef]
  61. DiMeglio, J.L.; Rosenthal, J. Selective Conversion of CO2 to CO with High Efficiency Using an Inexpensive Bismuth-Based Electrocatalyst. J. Am. Chem. Soc. 2013, 135, 8798–8801. [Google Scholar] [CrossRef] [Green Version]
  62. Sun, X.; Zhu, Q.; Kang, X.; Liu, H.; Qian, Q.; Zhang, Z.; Han, B. Molybdenum–Bismuth Bimetallic Chalcogenide Nanosheets for Highly Efficient Electrocatalytic Reduction of Carbon Dioxide to Methanol. Angew. Chem. Int. Ed. 2016, 55, 6771–6775. [Google Scholar] [CrossRef]
  63. Gattrell, M.; Gupta, N.; Co, A. A review of the aqueous electrochemical reduction of CO2 to hydrocarbons at copper. J. Electroanal. Chem 2006, 594, 1–19. [Google Scholar] [CrossRef]
  64. Shi, G.; Yu, L.; Ba, X.; Zhang, X.; Zhou, J.; Yu, Y. Copper nanoparticle interspersed MoS2 nanoflowers with enhanced efficiency for CO2 electrochemical reduction to fuel. Dalton Trans. 2017, 46, 10569–10577. [Google Scholar] [CrossRef] [PubMed]
  65. Li, H.; Liu, X.; Chen, S.; Yang, D.; Zhang, Q.; Song, L.; Xiao, H.; Zhang, Q.; Gu, L.; Wang, X. Edge-Exposed Molybdenum Disulfide with N-Doped Carbon Hybridization: A Hierarchical Hollow Electrocatalyst for Carbon Dioxide Reduction. Adv. Energy Mater. 2019, 9, 1900072. [Google Scholar] [CrossRef]
  66. Lv, K.; Suo, W.; Shao, M.; Zhu, Y.; Wang, X.; Feng, J.; Fang, M.; Zhu, Y. Nitrogen doped MoS2 and nitrogen doped carbon dots composite catalyst for electroreduction CO2 to CO with high Faradaic efficiency. Nano Energy 2019, 63, 103834. [Google Scholar] [CrossRef]
  67. Asadi, M.; Motevaselian, M.H.; Moradzadeh, A.; Majidi, L.; Esmaeilirad, M.; Sun, T.V.; Liu, C.; Bose, R.; Abbasi, P.; Zapol, P.; et al. Highly Efficient Solar-Driven Carbon Dioxide Reduction on Molybdenum Disulfide Catalyst Using Choline Chloride-Based Electrolyte. Adv. Energy Mater. 2019, 9, 1803536. [Google Scholar] [CrossRef]
  68. Xu, J.; Li, X.; Liu, W.; Sun, Y.; Ju, Z.; Yao, T.; Wang, C.; Ju, H.; Zhu, J.; Wei, S.; et al. Carbon Dioxide Electroreduction into Syngas Boosted by a Partially Delocalized Charge in Molybdenum Sulfide Selenide Alloy Monolayers. Angew. Chem. Int. Ed. 2017, 56, 9121–9125. [Google Scholar] [CrossRef] [PubMed]
  69. Sun, X.; Lu, L.; Zhu, Q.; Wu, C.; Yang, D.; Chen, C.; Han, B. MoP Nanoparticles Supported on Indium-Doped Porous Carbon: Outstanding Catalysts for Highly Efficient CO2 Electroreduction. Angew. Chem. Int. Ed. 2018, 57, 2427–2431. [Google Scholar] [CrossRef] [PubMed]
  70. Kumar, P.; Kumar, S.; Cordier, S.; Paofai, S.; Boukherroub, R.; Jain, S.L. Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2− cluster units under visible light irradiation. RSC Adv. 2014, 4, 10420–10423. [Google Scholar] [CrossRef]
  71. Kumar, P.; Mungse, H.P.; Cordier, S.; Boukherroub, R.; Khatri, O.P.; Jain, S.L. Hexamolybdenum clusters supported on graphene oxide: Visible-light induced photocatalytic reduction of carbon dioxide into methanol. Carbon 2015, 94, 91–100. [Google Scholar] [CrossRef]
  72. Jackson, J.A.; Turro, C.; Newsham, M.D.; Nocera, D.G. Oxygen quenching of electronically excited hexanuclear molybdenum and tungsten halide clusters. J. Phys. Chem. A 1990, 94, 4500–4507. [Google Scholar] [CrossRef]
  73. Cordier, S.; Kirakci, K.; Méry, D.; Perrin, C.; Astruc, D. Mo6X8i Nanocluster cores (X=Br, I): From inorganic solid state compounds to hybrids. Inorganica Chim. Acta 2006, 359, 1705–1709. [Google Scholar] [CrossRef]
  74. Srikanth, S.; Maesen, M.; Dominguez-Benetton, X.; Vanbroekhoven, K.; Pant, D. Enzymatic electrosynthesis of formate through CO2 sequestration/reduction in a bioelectrochemical system (BES). Bioresour. Technol. 2014, 165, 350–354. [Google Scholar] [CrossRef] [PubMed]
  75. Kim, S.; Kim, M.K.; Lee, S.H.; Yoon, S.; Jung, K.-D. Conversion of CO2 to formate in an electroenzymatic cell using Candida boidinii formate dehydrogenase. J. Mol. Catal. B Enzym. 2014, 102, 9–15. [Google Scholar] [CrossRef]
  76. Kim, S.-H.; Chung, G.-Y.; Kim, S.-H.; Vinothkumar, G.; Yoon, S.-H.; Jung, K.-D. Electrochemical NADH regeneration and electroenzymatic CO2 reduction on Cu nanorods/glassy carbon electrode prepared by cyclic deposition. Electrochim. Acta 2016, 210, 837–845. [Google Scholar] [CrossRef]
  77. Amao, Y.; Shuto, N. Formate dehydrogenase–viologen-immobilized electrode for CO2 conversion, for development of an artificial photosynthesis system. Res. Chem. Intermed. 2014, 40, 3267–3276. [Google Scholar] [CrossRef]
  78. Srikanth, S.; Alvarez-Gallego, Y.; Vanbroekhoven, K.; Pant, D. Enzymatic Electrosynthesis of Formic Acid through Carbon Dioxide Reduction in a Bioelectrochemical System: Effect of Immobilization and Carbonic Anhydrase Addition. ChemPhysChem 2017, 18, 3174–3181. [Google Scholar] [CrossRef]
  79. Zhang, L.; Liu, J.; Ong, J.; Li, S.F.Y. Specific and sustainable bioelectro-reduction of carbon dioxide to formate on a novel enzymatic cathode. Chemosphere 2016, 162, 228–234. [Google Scholar] [CrossRef]
  80. Bassegoda, A.; Madden, C.; Wakerley, D.W.; Reisner, E.; Hirst, J. Reversible Interconversion of CO2 and Formate by a Molybdenum-Containing Formate Dehydrogenase. J. Am. Chem. Soc. 2014, 136, 15473–15476. [Google Scholar] [CrossRef] [Green Version]
  81. Yuan, M.; Sahin, S.; Cai, R.; Abdellaoui, S.; Hickey, D.P.; Minteer, S.D.; Milton, R.D. Creating a Low-Potential Redox Polymer for Efficient Electroenzymatic CO2 Reduction. Angew. Chem. Int. Ed. 2018, 57, 6582–6586. [Google Scholar] [CrossRef]
  82. Ferguson, S.J. Nitrogen cycle enzymology. Curr. Opin. Chem. Biol. 1998, 2, 182–193. [Google Scholar] [CrossRef]
  83. Raymond, J.; Siefert, J.L.; Staples, C.R.; Blankenship, R.E. The Natural History of Nitrogen Fixation. Mol. Biol. Evol. 2004, 21, 541–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Shilov, A.E. Catalytic reduction of molecular nitrogen in solutions. Russ. Chem. Bull. 2003, 52, 2555–2562. [Google Scholar] [CrossRef]
  85. Haber, F.J.N. The production of ammonia from nitrogen and hydrogen. Naturwissenschaften 1922, 10, 1041. [Google Scholar] [CrossRef]
  86. Service, R.F. New recipe produces ammonia from air, water, and sunlight. Science 2014. [Google Scholar] [CrossRef]
  87. Eady, R.R. Structure−Function Relationships of Alternative Nitrogenases. Chem. Rev. 1996, 96, 3013–3030. [Google Scholar] [CrossRef]
  88. Garagounis, I.; Kyriakou, V.; Skodra, A.; Vasileiou, E.; Stoukides, M. Electrochemical Synthesis of Ammonia in Solid Electrolyte Cells. Front. Energy Res. 2014, 2. [Google Scholar] [CrossRef] [Green Version]
  89. Jia, H.-P.; Quadrelli, E.A. Mechanistic aspects of dinitrogen cleavage and hydrogenation to produce ammonia in catalysis and organometallic chemistry: Relevance of metal hydride bonds and dihydrogen. Chem. Soc. Rev. 2014, 43, 547–564. [Google Scholar] [CrossRef]
  90. MacKay, B.A.; Fryzuk, M.D. Dinitrogen Coordination Chemistry:  On the Biomimetic Borderlands. Chem. Rev. 2004, 104, 385–402. [Google Scholar] [CrossRef]
  91. Deng, H.; Hoffmann, R. How N2 Might Be Activated by the FeMo-Cofactor in Nitrogenase. Angew. Chem. Int. Ed. 1993, 32, 1062–1065. [Google Scholar] [CrossRef]
  92. Burgess, B.K.; Lowe, D.J. Mechanism of Molybdenum Nitrogenase. Chem. Rev. 1996, 96, 2983–3012. [Google Scholar] [CrossRef]
  93. Howard, J.B.; Rees, D.C. Structural Basis of Biological Nitrogen Fixation. Chem. Rev. 1996, 96, 2965–2982. [Google Scholar] [CrossRef]
  94. Hoffman, B.M.; Lukoyanov, D.; Yang, Z.-Y.; Dean, D.R.; Seefeldt, L.C. Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage. Chem. Rev. 2014, 114, 4041–4062. [Google Scholar] [CrossRef]
  95. Spatzal, T.; Aksoyoglu, M.; Zhang, L.; Andrade, S.L.A.; Schleicher, E.; Weber, S.; Rees, D.C.; Einsle, O. Evidence for Interstitial Carbon in Nitrogenase FeMo Cofactor. Science 2011, 334, 940. [Google Scholar] [CrossRef] [Green Version]
  96. Lancaster, K.M.; Roemelt, M.; Ettenhuber, P.; Hu, Y.; Ribbe, M.W.; Neese, F.; Bergmann, U.; DeBeer, S. X-ray Emission Spectroscopy Evidences a Central Carbon in the Nitrogenase Iron-Molybdenum Cofactor. Science 2011, 334, 974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wiig, J.A.; Hu, Y.; Lee, C.C.; Ribbe, M.W. Radical SAM-Dependent Carbon Insertion into the Nitrogenase M-Cluster. Science 2012, 337, 1672. [Google Scholar] [CrossRef] [Green Version]
  98. Wolff, T.E.; Berg, J.M.; Warrick, C.; Hodgson, K.O.; Holm, R.H.; Frankel, R.B. The molybdenum-iron-sulfur cluster complex [Mo2Fe6S9(SC2H5)8]3−. A synthetic approach to the molybdenum site in nitrogenase. J. Am. Chem. Soc. 1978, 100, 4630–4632. [Google Scholar] [CrossRef]
  99. Wolff, T.E.; Berg, J.M.; Hodgson, K.O.; Frankel, R.B.; Holm, R.H. Synthetic approaches to the molybdenum site in nitrogenase. Preparation and structural properties of the molybdenum-iron-sulfur “double-cubane” cluster complexes [Mo2Fe6S8(SC2H5)9]3− and [Mo2Fe6S9(SC2H5)8]3. J. Am. Chem. Soc. 1979, 101, 4140–4150. [Google Scholar] [CrossRef]
  100. Armstrong, W.H.; Holm, R.H. Synthesis and structure of a new type of molybdenum-iron-sulfur double-cubane cluster and evidence for formation of magnetically uncoupled S = 3/2 MoFe3S4 subclusters. J. Am. Chem. Soc. 1981, 103, 6246–6248. [Google Scholar] [CrossRef]
  101. Armstrong, W.H.; Mascharak, P.K.; Holm, R.H. Doubly bridged double cubanes containing MFe3S4 clusters (M = Mo, W). Synthesis, structure, and conversion to spin-quartet single clusters in solution. J. Am. Chem. Soc. 1982, 104, 4373–4383. [Google Scholar] [CrossRef]
  102. Cramer, S.P.; Gillum, W.O.; Hodgson, K.O.; Mortenson, L.E.; Stiefel, E.I.; Chisnell, J.R.; Brill, W.J.; Shah, V.K. The molybdenum site of nitrogenase. 2. A comparative study of molybdenum-iron proteins and the iron-molybdenum cofactor by x-ray absorption spectroscopy. J. Am. Chem. Soc. 1978, 100, 3814–3819. [Google Scholar] [CrossRef]
  103. Coucouvanis, D.; Demadis, K.D.; Kim, C.G.; Dunham, R.W.; Kampf, J.W. Single and double MoFe3S4 cubanes with molybdenum-coordinated polycarboxylate ligands. Syntheses and structural characterization of (Et4N)4{[MoFe3S4Cl4]2(μ-C2O4)} and (Et4N)3([MoFe3S4Cl4(C2O4)] clusters. J. Am. Chem. Soc. 1993, 115, 3344–3345. [Google Scholar] [CrossRef]
  104. Fomitchev, D.V.; McLauchlan, C.C.; Holm, R.H. Heterometal Cubane-Type MFe3S4 Clusters (M = Mo, V) Trigonally Symmetrized with Hydrotris(pyrazolyl)borate(1−) and Tris(pyrazolyl)methanesulfonate(1−) Capping Ligands. Inorg. Chem. 2002, 41, 958–966. [Google Scholar] [CrossRef]
  105. Bjornsson, R.; Neese, F.; Schrock, R.R.; Einsle, O.; DeBeer, S. The discovery of Mo(III) in FeMoco: Reuniting enzyme and model chemistry. J. Biol. Inorg. Chem. 2015, 20, 447–460. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, J. Progress in Synthesizing Analogues of Nitrogenase Metalloclusters for Catalytic Reduction of Nitrogen to Ammonia. Catalysts 2019, 9, 939. [Google Scholar] [CrossRef] [Green Version]
  107. Tanifuji, K.; Ohki, Y. Metal–Sulfur Compounds in N2 Reduction and Nitrogenase-Related Chemistry. Chem. Rev. 2020, 120, 5194–5251. [Google Scholar] [CrossRef]
  108. Komori, K.; Oshita, H.; Mizobe, Y.; Hidai, M. Preparation and properties of molybdenum and tungsten dinitrogen complexes. J. Am. Chem. Soc. 1989, 111, 1939–1940. [Google Scholar] [CrossRef]
  109. Yandulov, D.V.; Schrock, R.R. Catalytic Reduction of Dinitrogen to Ammonia at a Single Molybdenum Center. Science 2003, 301, 76. [Google Scholar] [CrossRef] [Green Version]
  110. Yandulov, D.V.; Schrock, R.R. Reduction of Dinitrogen to Ammonia at a Well-Protected Reaction Site in a Molybdenum Triamidoamine Complex. J. Am. Chem. Soc. 2002, 124, 6252–6253. [Google Scholar] [CrossRef]
  111. Foster, S.L.; Bakovic, S.I.P.; Duda, R.D.; Maheshwari, S.; Milton, R.D.; Minteer, S.D.; Janik, M.J.; Renner, J.N.; Greenlee, L.F. Catalysts for nitrogen reduction to ammonia. Nat. Catal. 2018, 1, 490–500. [Google Scholar] [CrossRef]
  112. Schrock, R.R. Catalytic reduction of dinitrogen under mild conditions. Chem. Commun. 2003, 19, 2389–2391. [Google Scholar] [CrossRef]
  113. Yandulov, D.V.; Schrock, R.R.; Rheingold, A.L.; Ceccarelli, C.; Davis, W.M. Synthesis and Reactions of Molybdenum Triamidoamine Complexes Containing Hexaisopropylterphenyl Substituents. Inorg. Chem. 2003, 42, 796–813. [Google Scholar] [CrossRef] [PubMed]
  114. Ritleng, V.; Yandulov, D.V.; Weare, W.W.; Schrock, R.R.; Hock, A.S.; Davis, W.M. Molybdenum Triamidoamine Complexes that Contain Hexa-tert-butylterphenyl, Hexamethylterphenyl, or p-Bromohexaisopropylterphenyl Substituents. An Examination of Some Catalyst Variations for the Catalytic Reduction of Dinitrogen. J. Am. Chem. Soc. 2004, 126, 6150–6163. [Google Scholar] [CrossRef]
  115. Schrock, R.R. Catalytic Reduction of Dinitrogen to Ammonia at a Single Molybdenum Center. Acc. Chem. Res. 2005, 38, 955–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yandulov, D.V.; Schrock, R.R. Studies Relevant to Catalytic Reduction of Dinitrogen to Ammonia by Molybdenum Triamidoamine Complexes. Inorg. Chem. 2005, 44, 1103–1117. [Google Scholar] [CrossRef]
  117. Schrock, R.R. Reduction of dinitrogen. Proc. Natl. Acad. Sci. USA 2006, 103, 17087. [Google Scholar] [CrossRef] [Green Version]
  118. Schrock, R.R. Catalytic Reduction of Dinitrogen to Ammonia by Molybdenum: Theory versus Experiment. Angew. Chem. Int. Ed. 2008, 47, 5512–5522. [Google Scholar] [CrossRef]
  119. Tanaka, H.; Sasada, A.; Kouno, T.; Yuki, M.; Miyake, Y.; Nakanishi, H.; Nishibayashi, Y.; Yoshizawa, K. Molybdenum-Catalyzed Transformation of Molecular Dinitrogen into Silylamine: Experimental and DFT Study on the Remarkable Role of Ferrocenyldiphosphine Ligands. J. Am. Chem. Soc. 2011, 133, 3498–3506. [Google Scholar] [CrossRef]
  120. Arashiba, K.; Miyake, Y.; Nishibayashi, Y. A molybdenum complex bearing PNP-type pincer ligands leads to the catalytic reduction of dinitrogen into ammonia. Nat. Chem. 2011, 3, 120–125. [Google Scholar] [CrossRef]
  121. Nishibayashi, Y. Molybdenum-catalyzed reduction of molecular dinitrogen under mild reaction conditions. Dalton Trans. 2012, 41, 7447–7453. [Google Scholar] [CrossRef]
  122. Kuriyama, S.; Arashiba, K.; Nakajima, K.; Tanaka, H.; Kamaru, N.; Yoshizawa, K.; Nishibayashi, Y. Catalytic Formation of Ammonia from Molecular Dinitrogen by Use of Dinitrogen-Bridged Dimolybdenum–Dinitrogen Complexes Bearing PNP-Pincer Ligands: Remarkable Effect of Substituent at PNP-Pincer Ligand. J. Am. Chem. Soc. 2014, 136, 9719–9731. [Google Scholar] [CrossRef]
  123. Kuriyama, S.; Arashiba, K.; Nakajima, K.; Tanaka, H.; Yoshizawa, K.; Nishibayashi, Y. Nitrogen fixation catalyzed by ferrocene-substituted dinitrogen-bridged dimolybdenum–dinitrogen complexes: Unique behavior of ferrocene moiety as redox active site. Chem. Sci. 2015, 6, 3940–3951. [Google Scholar] [CrossRef] [Green Version]
  124. Itabashi, T.; Mori, I.; Arashiba, K.; Eizawa, A.; Nakajima, K.; Nishibayashi, Y. Effect of substituents on molybdenum triiodide complexes bearing PNP-type pincer ligands toward catalytic nitrogen fixation. Dalton Trans. 2019, 48, 3182–3186. [Google Scholar] [CrossRef] [PubMed]
  125. Arashiba, K.; Eizawa, A.; Tanaka, H.; Nakajima, K.; Yoshizawa, K.; Nishibayashi, Y. Catalytic Nitrogen Fixation via Direct Cleavage of Nitrogen–Nitrogen Triple Bond of Molecular Dinitrogen under Ambient Reaction Conditions. Bull. Chem. Soc. Jpn. 2017, 90, 1111–1118. [Google Scholar] [CrossRef]
  126. Eizawa, A.; Arashiba, K.; Egi, A.; Tanaka, H.; Nakajima, K.; Yoshizawa, K.; Nishibayashi, Y. Catalytic Reactivity of Molybdenum–Trihalide Complexes Bearing PCP-Type Pincer Ligands. Chem. Asian J. 2019, 14, 2091–2096. [Google Scholar] [CrossRef] [PubMed]
  127. Ashida, Y.; Arashiba, K.; Nakajima, K.; Nishibayashi, Y. Molybdenum-catalysed ammonia production with samarium diiodide and alcohols or water. Nature 2019, 568, 536–540. [Google Scholar] [CrossRef]
  128. Brown, K.A.; Harris, D.F.; Wilker, M.B.; Rasmussen, A.; Khadka, N.; Hamby, H.; Keable, S.; Dukovic, G.; Peters, J.W.; Seefeldt, L.C.J.S. Light-driven dinitrogen reduction catalyzed by a CdS: Nitrogenase MoFe protein biohybrid. Science 2016, 352, 448–450. [Google Scholar] [CrossRef]
  129. Nazemi, M.; Panikkanvalappil, S.R.; El-Sayed, M.A. Enhancing the rate of electrochemical nitrogen reduction reaction for ammonia synthesis under ambient conditions using hollow gold nanocages. Nano Energy 2018, 49, 316–323. [Google Scholar] [CrossRef]
  130. Manjunatha, R.; Schechter, A. Electrochemical synthesis of ammonia using ruthenium–platinum alloy at ambient pressure and low temperature. Electrochem. Commun. 2018, 90, 96–100. [Google Scholar] [CrossRef]
  131. Tao, H.; Choi, C.; Ding, L.-X.; Jiang, Z.; Han, Z.; Jia, M.; Fan, Q.; Gao, Y.; Wang, H.; Robertson, A.W.; et al. Nitrogen Fixation by Ru Single-Atom Electrocatalytic Reduction. Chem 2019, 5, 204–214. [Google Scholar] [CrossRef]
  132. Liu, H.-M.; Han, S.-H.; Zhao, Y.; Zhu, Y.-Y.; Tian, X.-L.; Zeng, J.-H.; Jiang, J.-X.; Xia, B.Y.; Chen, Y. Surfactant-free atomically ultrathin rhodium nanosheet nanoassemblies for efficient nitrogen electroreduction. J. Mater. Chem. A 2018, 6, 3211–3217. [Google Scholar] [CrossRef]
  133. Huang, H.; Xia, L.; Shi, X.; Asiri, A.M.; Sun, X. Ag nanosheets for efficient electrocatalytic N2 fixation to NH3 under ambient conditions. Chem. Commun. 2018, 54, 11427–11430. [Google Scholar] [CrossRef]
  134. Oshikiri, T.; Ueno, K.; Misawa, H. Selective Dinitrogen Conversion to Ammonia Using Water and Visible Light through Plasmon-induced Charge Separation. Angew. Chem. Int. Ed. 2016, 55, 3942–3946. [Google Scholar] [CrossRef]
  135. Wang, H.; Li, Y.; Li, C.; Deng, K.; Wang, Z.; Xu, Y.; Li, X.; Xue, H.; Wang, L. One-pot synthesis of bi-metallic PdRu tripods as an efficient catalyst for electrocatalytic nitrogen reduction to ammonia. J. Mater. Chem. A 2019, 7, 801–805. [Google Scholar] [CrossRef]
  136. Montoya, J.H.; Tsai, C.; Vojvodic, A.; Nørskov, J.K. The Challenge of Electrochemical Ammonia Synthesis: A New Perspective on the Role of Nitrogen Scaling Relations. ChemSusChem 2015, 8, 2180–2186. [Google Scholar] [CrossRef]
  137. Skúlason, E.; Bligaard, T.; Gudmundsdóttir, S.; Studt, F.; Rossmeisl, J.; Abild-Pedersen, F.; Vegge, T.; Jónsson, H.; Nørskov, J.K. A theoretical evaluation of possible transition metal electro-catalysts for N2 reduction. Phys. Chem. Chem. Phys. 2012, 14, 1235–1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Yang, D.; Chen, T.; Wang, Z. Electrochemical reduction of aqueous nitrogen (N2) at a low overpotential on (110)-oriented Mo nanofilm. J. Mater. Chem. A 2017, 5, 18967–18971. [Google Scholar] [CrossRef]
  139. Zhu, D.; Zhang, L.; Ruther, R.E.; Hamers, R.J. Photo-illuminated diamond as a solid-state source of solvated electrons in water for nitrogen reduction. Nat. Mater. 2013, 12, 836–841. [Google Scholar] [CrossRef]
  140. Kordali, V.; Kyriacou, G.; Lambrou, C. Electrochemical synthesis of ammonia at atmospheric pressure and low temperature in a solid polymer electrolyte cell. Chem. Commun. 2000, 1673–1674. [Google Scholar] [CrossRef]
  141. Chen, S.; Perathoner, S.; Ampelli, C.; Mebrahtu, C.; Su, D.; Centi, G. Electrocatalytic Synthesis of Ammonia at Room Temperature and Atmospheric Pressure from Water and Nitrogen on a Carbon-Nanotube-Based Electrocatalyst. Angew. Chem. Int. Ed. 2017, 56, 2699–2703. [Google Scholar] [CrossRef] [PubMed]
  142. Han, J.; Ji, X.; Ren, X.; Cui, G.; Li, L.; Xie, F.; Wang, H.; Li, B.; Sun, X. MoO3 nanosheets for efficient electrocatalytic N2 fixation to NH3. J. Mater. Chem. A 2018, 6, 12974–12977. [Google Scholar] [CrossRef]
  143. Zhang, L.; Ji, X.; Ren, X.; Luo, Y.; Shi, X.; Asiri, A.M.; Zheng, B.; Sun, X. Efficient Electrochemical N2 Reduction to NH3 on MoN Nanosheets Array under Ambient Conditions. ACS Sustain. Chem. Eng. 2018, 6, 9550–9554. [Google Scholar] [CrossRef]
  144. Zhang, L.; Ji, X.; Ren, X.; Ma, Y.; Shi, X.; Tian, Z.; Asiri, A.M.; Chen, L.; Tang, B.; Sun, X. Electrochemical Ammonia Synthesis via Nitrogen Reduction Reaction on a MoS2 Catalyst: Theoretical and Experimental Studies. Adv. Mat. 2018, 30, 1800191. [Google Scholar] [CrossRef] [PubMed]
  145. Xu, X.; Sun, B.; Liang, Z.; Cui, H.; Tian, J. High-Performance Electrocatalytic Conversion of N2 to NH3 Using 1T-MoS2 Anchored on Ti3C2 MXene under Ambient Conditions. ACS Appl. Mater. Interfaces 2020, 12, 26060–26067. [Google Scholar] [CrossRef] [PubMed]
  146. Danyal, K.; Inglet, B.S.; Vincent, K.A.; Barney, B.M.; Hoffman, B.M.; Armstrong, F.A.; Dean, D.R.; Seefeldt, L.C. Uncoupling Nitrogenase: Catalytic Reduction of Hydrazine to Ammonia by a MoFe Protein in the Absence of Fe Protein-ATP. J. Am. Chem. Soc. 2010, 132, 13197–13199. [Google Scholar] [CrossRef] [Green Version]
  147. Milton, R.D.; Abdellaoui, S.; Khadka, N.; Dean, D.R.; Leech, D.; Seefeldt, L.C.; Minteer, S.D. Nitrogenase bioelectrocatalysis: Heterogeneous ammonia and hydrogen production by MoFe protein. Energy Environ. Sci. 2016, 9, 2550–2554. [Google Scholar] [CrossRef] [Green Version]
  148. Lee, Y.S.; Yuan, M.; Cai, R.; Lim, K.; Minteer, S.D. Nitrogenase Bioelectrocatalysis: ATP-Independent Ammonia Production Using a Redox Polymer/MoFe Protein System. ACS Catal. 2020, 10, 6854–6861. [Google Scholar] [CrossRef]
  149. Badalyan, A.; Yang, Z.-Y.; Seefeldt, L.C. A Voltammetric Study of Nitrogenase Catalysis Using Electron Transfer Mediators. ACS Catal. 2019, 9, 1366–1372. [Google Scholar] [CrossRef]
  150. Jensen, B.B.; Burris, R.H. Nitrous oxide as a substrate and as a competitive inhibitor of nitrogenase. Biochemistry 1986, 25, 1083–1088. [Google Scholar] [CrossRef]
  151. Davis, L.C. Hydrazine as a substrate and inhibitor of Azotobacter vinelandii nitrogenase. Arch. Biochem. Biophys. 1980, 204, 270–276. [Google Scholar] [CrossRef]
  152. Vaughn, S.A.; Burgess, B.K. Nitrite, a new substrate for nitrogenase. Biochemistry 1989, 28, 419–424. [Google Scholar] [CrossRef] [PubMed]
  153. Hosseini, S.E.; Wahid, M.A. Hydrogen production from renewable and sustainable energy resources: Promising green energy carrier for clean development. Renew. Sustain. Energy Rev. 2016, 57, 850–866. [Google Scholar] [CrossRef]
  154. Vesborg, P.C.K.; Seger, B.; Chorkendorff, I. Recent Development in Hydrogen Evolution Reaction Catalysts and Their Practical Implementation. J. Phys. Chem. Lett. 2015, 6, 951–957. [Google Scholar] [CrossRef]
  155. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef]
  156. Kong, D.; Cha, J.J.; Wang, H.; Lee, H.R.; Cui, Y. First-row transition metal dichalcogenide catalysts for hydrogen evolution reaction. Energy Environ. Sci. 2013, 6, 3553–3558. [Google Scholar] [CrossRef]
  157. Darensbourg, D.J.; Reibenspies, J.H.; Lai, C.-H.; Lee, W.-Z.; Darensbourg, M.Y. Analysis of an Organometallic Iron Site Model for the Heterodimetallic Unit of [NiFe]Hydrogenase. J. Am. Chem. Soc. 1997, 119, 7903–7904. [Google Scholar] [CrossRef]
  158. Donovan, E.S.; Felton, G.A.N. Electrochemical analysis of cyclopentadienylmetal carbonyl dimer complexes: Insight into the design of hydrogen-producing electrocatalysts. J. Organomet. Chem. 2012, 711, 25–34. [Google Scholar] [CrossRef]
  159. Hu, C.; Fan, W.Y. Molybdenum carbonyl complexes as HER electrocatalysts. Mol. Catal. 2019, 479, 110615. [Google Scholar] [CrossRef]
  160. DuBois, M.R.; VanDerveer, M.C.; DuBois, D.L.; Haltiwanger, R.C.; Miller, W.K. Characterization of reactions of hydrogen with coordinated sulfido ligands. J. Am. Chem. Soc. 1980, 102, 7456–7461. [Google Scholar] [CrossRef]
  161. Appel, A.M.; DuBois, D.L.; Rakowski DuBois, M. Molybdenum−Sulfur Dimers as Electrocatalysts for the Production of Hydrogen at Low Overpotentials. J. Am. Chem. Soc. 2005, 127, 12717–12726. [Google Scholar] [CrossRef]
  162. Karunadasa, H.I.; Montalvo, E.; Sun, Y.; Majda, M.; Long, J.R.; Chang, C.J. A Molecular MoS2 Edge Site Mimic for Catalytic Hydrogen Generation. Science 2012, 335, 698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Karunadasa, H.I.; Chang, C.J.; Long, J.R. A molecular molybdenum-oxo catalyst for generating hydrogen from water. Nature 2010, 464, 1329–1333. [Google Scholar] [CrossRef] [PubMed]
  164. Dave, M.; Rajagopal, A.; Damm-Ruttensperger, M.; Schwarz, B.; Nägele, F.; Daccache, L.; Fantauzzi, D.; Jacob, T.; Streb, C. Understanding homogeneous hydrogen evolution reactivity and deactivation pathways of molecular molybdenum sulfide catalysts. Sustain. Energ. Fuels 2018, 2, 1020–1026. [Google Scholar] [CrossRef]
  165. Cao, J.-P.; Zhou, L.-L.; Fu, L.-Z.; Zhao, J.-X.; Lu, H.-X.; Zhan, S.-Z. A molybdenum–Schiff base complex, a new molecular electro-catalyst for generating hydrogen from acetic acid or water. Catal. Commun. 2014, 57, 1–4. [Google Scholar] [CrossRef]
  166. Cao, J.-P.; Zhou, L.-L.; Fu, L.-Z.; Zhan, S. A molecular molybdenum electrocatalyst for generating hydrogen from acetic acid or water. J. Power Sources 2014, 272, 169–175. [Google Scholar] [CrossRef]
  167. Seo, J.; Williard, P.G.; Kim, E. Deoxygenation of Mono-oxo Bis(dithiolene) Mo and W Complexes by Protonation. Inorg. Chem. 2013, 52, 8706–8712. [Google Scholar] [CrossRef]
  168. Porcher, J.-P.; Fogeron, T.; Gomez-Mingot, M.; Derat, E.; Chamoreau, L.-M.; Li, Y.; Fontecave, M. A Bioinspired Molybdenum Complex as a Catalyst for the Photo- and Electroreduction of Protons. Angew. Chem. Int. Ed. 2015, 54, 14090–14093. [Google Scholar] [CrossRef]
  169. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jørgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Nørskov, J.K. Biomimetic Hydrogen Evolution:  MoS2 Nanoparticles as Catalyst for Hydrogen Evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef]
  170. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100. [Google Scholar] [CrossRef] [Green Version]
  171. Seger, B.; Laursen, A.B.; Vesborg, P.C.K.; Pedersen, T.; Hansen, O.; Dahl, S.; Chorkendorff, I. Hydrogen Production Using a Molybdenum Sulfide Catalyst on a Titanium-Protected n+p-Silicon Photocathode. Angew. Chem. Int. Ed. 2012, 51, 9128–9131. [Google Scholar] [CrossRef]
  172. Merki, D.; Fierro, S.; Vrubel, H.; Hu, X. Amorphous molybdenum sulfide films as catalysts for electrochemical hydrogen production in water. Chem. Sci. 2011, 2, 1262–1267. [Google Scholar] [CrossRef] [Green Version]
  173. Vrubel, H.; Merki, D.; Hu, X. Hydrogen evolution catalyzed by MoS3 and MoS2 particles. Energy Environ. Sci. 2012, 5, 6136–6144. [Google Scholar] [CrossRef] [Green Version]
  174. Vrubel, H.; Hu, X. Growth and Activation of an Amorphous Molybdenum Sulfide Hydrogen Evolving Catalyst. ACS Catal. 2013, 3, 2002–2011. [Google Scholar] [CrossRef] [Green Version]
  175. Jaramillo, T.F.; Bonde, J.; Zhang, J.; Ooi, B.-L.; Andersson, K.; Ulstrup, J.; Chorkendorff, I. Hydrogen Evolution on Supported Incomplete Cubane-type [Mo3S4]4+ Electrocatalysts. J. Phys. Chem. C 2008, 112, 17492–17498. [Google Scholar] [CrossRef]
  176. Kibsgaard, J.; Jaramillo, T.F.; Besenbacher, F. Building an appropriate active-site motif into a hydrogen-evolution catalyst with thiomolybdate [Mo3S13]2− clusters. Nat. Chem. 2014, 6, 248–253. [Google Scholar] [CrossRef] [Green Version]
  177. Du, K.; Zheng, L.; Wang, T.; Zhuo, J.; Zhu, Z.; Shao, Y.; Li, M. Electrodeposited Mo3S13 Films from (NH4)2Mo3S13·2H2O for Electrocatalysis of Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2017, 9, 18675–18681. [Google Scholar] [CrossRef] [PubMed]
  178. Shang, Y.; Xu, X.; Gao, B.; Ren, Z. Thiomolybdate [Mo3S13]2– Nanoclusters Anchored on Reduced Graphene Oxide-Carbon Nanotube Aerogels for Efficient Electrocatalytic Hydrogen Evolution. ACS Sustain. Chem. Eng. 2017, 5, 8908–8917. [Google Scholar] [CrossRef]
  179. Huang, Z.; Luo, W.; Ma, L.; Yu, M.; Ren, X.; He, M.; Polen, S.; Click, K.; Garrett, B.; Lu, J.; et al. Dimeric [Mo2S12]2− Cluster: A Molecular Analogue of MoS2 Edges for Superior Hydrogen-Evolution Electrocatalysis. Angew. Chem. Int. Ed. 2015, 54, 15181–15185. [Google Scholar] [CrossRef]
  180. Lei, Y.; Yang, M.; Hou, J.; Wang, F.; Cui, E.; Kong, C.; Min, S. Thiomolybdate [Mo3S13]2− nanocluster: A molecular mimic of MoS2 active sites for highly efficient photocatalytic hydrogen evolution. Chem. Commun. 2018, 54, 603–606. [Google Scholar] [CrossRef]
  181. Cheng, Y.-J.; Wang, R.; Wang, S.; Xi, X.-J.; Ma, L.-F.; Zang, S.-Q. Encapsulating [Mo3S13]2− clusters in cationic covalent organic frameworks: Enhancing stability and recyclability by converting a homogeneous photocatalyst to a heterogeneous photocatalyst. Chem. Commun. 2018, 54, 13563–13566. [Google Scholar] [CrossRef] [PubMed]
  182. Tran, P.D.; Tran, T.V.; Orio, M.; Torelli, S.; Truong, Q.D.; Nayuki, K.; Sasaki, Y.; Chiam, S.Y.; Yi, R.; Honma, I.; et al. Coordination polymer structure and revisited hydrogen evolution catalytic mechanism for amorphous molybdenum sulfide. Nat. Mat. 2016, 15, 640–646. [Google Scholar] [CrossRef] [PubMed]
  183. Feliz, M.; Puche, M.; Atienzar, P.; Concepción, P.; Cordier, S.; Molard, Y. In Situ Generation of Active Molybdenum Octahedral Clusters for Photocatalytic Hydrogen Production from Water. ChemSusChem 2016, 9, 1963–1971. [Google Scholar] [CrossRef] [PubMed]
  184. Puche, M.; García-Aboal, R.; Mikhaylov, M.A.; Sokolov, M.N.; Atienzar, P.; Feliz, M.J.N. Enhanced Photocatalytic Activity and Stability in Hydrogen Evolution of Mo6 Iodide Clusters Supported on Graphene Oxide. Nanomaterials 2020, 10, 1259. [Google Scholar] [CrossRef]
  185. Feliz, M.; Atienzar, P.; Amela-Cortés, M.; Dumait, N.; Lemoine, P.; Molard, Y.; Cordier, S. Supramolecular Anchoring of Octahedral Molybdenum Clusters onto Graphene and Their Synergies in Photocatalytic Water Reduction. Inorg. Chem. 2019, 58, 15443–15454. [Google Scholar] [CrossRef] [Green Version]
  186. Recatalá, D.; Llusar, R.; Gushchin, A.L.; Kozlova, E.A.; Laricheva, Y.A.; Abramov, P.A.; Sokolov, M.N.; Gómez, R.; Lana-Villarreal, T. Photogeneration of Hydrogen from Water by Hybrid Molybdenum Sulfide Clusters Immobilized on Titania. ChemSusChem 2015, 8, 148–157. [Google Scholar] [CrossRef]
  187. Tran, P.D.; Nguyen, M.; Pramana, S.S.; Bhattacharjee, A.; Chiam, S.Y.; Fize, J.; Field, M.J.; Artero, V.; Wong, L.H.; Loo, J.; et al. Copper molybdenum sulfide: A new efficient electrocatalyst for hydrogen production from water. Energy Environ. Sci. 2012, 5, 8912–8916. [Google Scholar] [CrossRef]
  188. Gao, M.-R.; Liang, J.-X.; Zheng, Y.-R.; Xu, Y.-F.; Jiang, J.; Gao, Q.; Li, J.; Yu, S.-H. An efficient molybdenum disulfide/cobalt diselenide hybrid catalyst for electrochemical hydrogen generation. Nat. Commun. 2015, 6, 5982. [Google Scholar] [CrossRef]
  189. Tang, H.; Dou, K.; Kaun, C.-C.; Kuang, Q.; Yang, S. MoSe2 nanosheets and their graphene hybrids: Synthesis, characterization and hydrogen evolution reaction studies. J. Mater. Chem. A 2014, 2, 360–364. [Google Scholar] [CrossRef]
  190. Chen, W.-F.; Sasaki, K.; Ma, C.; Frenkel, A.I.; Marinkovic, N.; Muckerman, J.T.; Zhu, Y.; Adzic, R.R. Hydrogen-Evolution Catalysts Based on Non-Noble Metal Nickel–Molybdenum Nitride Nanosheets. Angew. Chem. Int. Ed. 2012, 51, 6131–6135. [Google Scholar] [CrossRef]
  191. Kibsgaard, J.; Jaramillo, T.F. Molybdenum Phosphosulfide: An Active, Acid-Stable, Earth-Abundant Catalyst for the Hydrogen Evolution Reaction. Angew. Chem. Int. Ed. 2014, 53, 14433–14437. [Google Scholar] [CrossRef]
  192. Vrubel, H.; Hu, X. Molybdenum Boride and Carbide Catalyze Hydrogen Evolution in both Acidic and Basic Solutions. Angew. Chem. Int. Ed. 2012, 51, 12703–12706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Wan, C.; Regmi, Y.N.; Leonard, B.M. Multiple Phases of Molybdenum Carbide as Electrocatalysts for the Hydrogen Evolution Reaction. Angew. Chem. Int. Ed. 2014, 53, 6407–6410. [Google Scholar] [CrossRef] [PubMed]
  194. Zhao, Y.; Kamiya, K.; Hashimoto, K.; Nakanishi, S. In Situ CO2-Emission Assisted Synthesis of Molybdenum Carbonitride Nanomaterial as Hydrogen Evolution Electrocatalyst. J. Am. Chem. Soc. 2015, 137, 110–113. [Google Scholar] [CrossRef] [PubMed]
  195. Tang, Y.-J.; Gao, M.-R.; Liu, C.-H.; Li, S.-L.; Jiang, H.-L.; Lan, Y.-Q.; Han, M.; Yu, S.-H. Porous Molybdenum-Based Hybrid Catalysts for Highly Efficient Hydrogen Evolution. Angew. Chem. Int. Ed. 2015, 54, 12928–12932. [Google Scholar] [CrossRef] [PubMed]
  196. Wang, C.; Lv, X.; Zhou, P.; Liang, X.; Wang, Z.; Liu, Y.; Wang, P.; Zheng, Z.; Dai, Y.; Li, Y.; et al. Molybdenum Nitride Electrocatalysts for Hydrogen Evolution More Efficient than Platinum/Carbon: Mo2N/CeO2@Nickel Foam. ACS Appl. Mater. Interfaces 2020, 12, 29153–29161. [Google Scholar] [CrossRef]
  197. Mohanty, B.; Ghorbani-Asl, M.; Kretschmer, S.; Ghosh, A.; Guha, P.; Panda, S.K.; Jena, B.; Krasheninnikov, A.V.; Jena, B.K. MoS2 Quantum Dots as Efficient Catalyst Materials for the Oxygen Evolution Reaction. ACS Catal. 2018, 8, 1683–1689. [Google Scholar] [CrossRef]
  198. Hartley, C.L.; DiRisio, R.J.; Screen, M.E.; Mayer, K.J.; McNamara, W.R. Iron Polypyridyl Complexes for Photocatalytic Hydrogen Generation. Inorg. Chem. 2016, 55, 8865–8870. [Google Scholar] [CrossRef]
  199. Jin, Y.; Wang, H.; Li, J.; Yue, X.; Han, Y.; Shen, P.K.; Cui, Y. Porous MoO2 Nanosheets as Non-noble Bifunctional Electrocatalysts for Overall Water Splitting. Adv. Mat. 2016, 28, 3785–3790. [Google Scholar] [CrossRef]
  200. Gupta, S.; Patel, N.; Fernandes, R.; Hanchate, S.; Miotello, A.; Kothari, D.C. Co-Mo-B Nanoparticles as a non-precious and efficient Bifunctional Electrocatalyst for Hydrogen and Oxygen Evolution. Electrochim. Acta 2017, 232, 64–71. [Google Scholar] [CrossRef]
  201. Zhao, S.; Huang, J.; Liu, Y.; Shen, J.; Wang, H.; Yang, X.; Zhu, Y.; Li, C. Multimetallic Ni–Mo/Cu nanowires as nonprecious and efficient full water splitting catalyst. J. Mater. Chem. A 2017, 5, 4207–4214. [Google Scholar] [CrossRef]
  202. Liu, P.F.; Yang, S.; Zheng, L.R.; Zhang, B.; Yang, H.G. Mo6+ activated multimetal oxygen-evolving catalysts. Chemi. Sci. 2017, 8, 3484–3488. [Google Scholar] [CrossRef] [Green Version]
  203. Yang, Y.; Zhang, K.; Lin, H.; Li, X.; Chan, H.C.; Yang, L.; Gao, Q. MoS2–Ni3S2 Heteronanorods as Efficient and Stable Bifunctional Electrocatalysts for Overall Water Splitting. ACS Catal. 2017, 7, 2357–2366. [Google Scholar] [CrossRef]
  204. Tang, Y.-J.; Liu, C.-H.; Huang, W.; Wang, X.-L.; Dong, L.-Z.; Li, S.-L.; Lan, Y.-Q. Bimetallic Carbides-Based Nanocomposite as Superior Electrocatalyst for Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2017, 9, 16977–16985. [Google Scholar] [CrossRef]
  205. Jin, Y.; Huang, S.; Yue, X.; Du, H.; Shen, P.K. Mo- and Fe-Modified Ni(OH)2/NiOOH Nanosheets as Highly Active and Stable Electrocatalysts for Oxygen Evolution Reaction. ACS Catal. 2018, 8, 2359–2363. [Google Scholar] [CrossRef]
  206. Li, Y.; Xu, H.; Huang, H.; Wang, C.; Gao, L.; Ma, T. One-dimensional MoO2–Co2Mo3O8@C nanorods: A novel and highly efficient oxygen evolution reaction catalyst derived from metal–organic framework composites. Chem. Commun. 2018, 54, 2739–2742. [Google Scholar] [CrossRef] [PubMed]
  207. Yuan, Y.; Adimi, S.; Guo, X.; Thomas, T.; Zhu, Y.; Guo, H.; Priyanga, G.S.; Yoo, P.; Wang, J.; Chen, J.; et al. A Surface-Oxide-Rich Activation Layer (SOAL) on Ni2Mo3N for a Rapid and Durable Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2020, 59, 18036–18041. [Google Scholar] [CrossRef]
  208. Shi, H.; Zhou, Y.-T.; Yao, R.-Q.; Wan, W.-B.; Ge, X.; Zhang, W.; Wen, Z.; Lang, X.-Y.; Zheng, W.-T.; Jiang, Q. Spontaneously separated intermetallic Co3Mo from nanoporous copper as versatile electrocatalysts for highly efficient water splitting. Nat. Commun. 2020, 11, 2940. [Google Scholar] [CrossRef]
Figure 1. Oxidized form of the FDH active site from Escherichia coli (PDB 2IV2).
Figure 1. Oxidized form of the FDH active site from Escherichia coli (PDB 2IV2).
Catalysts 11 00217 g001
Scheme 1. A reaction pathway proposed by Raaijmakers.
Scheme 1. A reaction pathway proposed by Raaijmakers.
Catalysts 11 00217 sch001
Scheme 2. A reaction pathway including the role of sulfide ligand.
Scheme 2. A reaction pathway including the role of sulfide ligand.
Catalysts 11 00217 sch002
Scheme 3. Indirect reaction pathway through Mo-SH and Arg interaction.
Scheme 3. Indirect reaction pathway through Mo-SH and Arg interaction.
Catalysts 11 00217 sch003
Figure 2. (a) MoCu-CODH active site (O. carboxidovorans CODH, PDB code 1N5W) and reaction pathways proposed by (b) Dobbek, (c) Siegbahn, (d) Hofmann, and (e) Hille.
Figure 2. (a) MoCu-CODH active site (O. carboxidovorans CODH, PDB code 1N5W) and reaction pathways proposed by (b) Dobbek, (c) Siegbahn, (d) Hofmann, and (e) Hille.
Catalysts 11 00217 g002aCatalysts 11 00217 g002b
Scheme 4. FDH model complexes.
Scheme 4. FDH model complexes.
Catalysts 11 00217 sch004
Scheme 5. MoCu-CODH active site and the model complex.
Scheme 5. MoCu-CODH active site and the model complex.
Catalysts 11 00217 sch005
Scheme 6. CO2 reactivity of Mo2(OtBu)6 complex.
Scheme 6. CO2 reactivity of Mo2(OtBu)6 complex.
Catalysts 11 00217 sch006
Scheme 7. Metal ligand cooperation (MLC) complexes with CO2 reactivity.
Scheme 7. Metal ligand cooperation (MLC) complexes with CO2 reactivity.
Catalysts 11 00217 sch007
Scheme 8. Heterobimetallic (η5-C5H5)Ru(CO)(μ-dppm)Mo(CO)25-C5H5) complex.
Scheme 8. Heterobimetallic (η5-C5H5)Ru(CO)(μ-dppm)Mo(CO)25-C5H5) complex.
Catalysts 11 00217 sch008
Scheme 9. Mo-silyl hydrido complexes.
Scheme 9. Mo-silyl hydrido complexes.
Catalysts 11 00217 sch009
Scheme 10. Redox active ligand assisted Mo complexes.
Scheme 10. Redox active ligand assisted Mo complexes.
Catalysts 11 00217 sch010
Scheme 11. CO2 reactivity of (PNMeP)Mo(C2H4)2 complex.
Scheme 11. CO2 reactivity of (PNMeP)Mo(C2H4)2 complex.
Catalysts 11 00217 sch011
Figure 3. Structure of molybdenum nitrogenase, 4Fe4S cluster, 8Fe7S cluster, FeMo-cofactor, PDB code 2AFK.
Figure 3. Structure of molybdenum nitrogenase, 4Fe4S cluster, 8Fe7S cluster, FeMo-cofactor, PDB code 2AFK.
Catalysts 11 00217 g003
Scheme 12. Nitrogenase model complexes.
Scheme 12. Nitrogenase model complexes.
Catalysts 11 00217 sch012
Scheme 13. Homogeneous Mo complexes.
Scheme 13. Homogeneous Mo complexes.
Catalysts 11 00217 sch013
Scheme 14. N2 reduction pathways by Schrock complex.
Scheme 14. N2 reduction pathways by Schrock complex.
Catalysts 11 00217 sch014
Scheme 15. N2 reduction pathway by Nishibayashi complex.
Scheme 15. N2 reduction pathway by Nishibayashi complex.
Catalysts 11 00217 sch015
Scheme 16. Homogeneous PNP-Mo complexes.
Scheme 16. Homogeneous PNP-Mo complexes.
Catalysts 11 00217 sch016
Scheme 17. Homogeneous Mo complexes for H2 evolution.
Scheme 17. Homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch017
Scheme 18. Structure of homogeneous Mo complexes for H2 evolution.
Scheme 18. Structure of homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch018
Scheme 19. Structure of homogeneous Mo complexes for H2 evolution.
Scheme 19. Structure of homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch019
Scheme 20. Structure of homogeneous Mo complexes for H2 evolution.
Scheme 20. Structure of homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch020
Scheme 21. Structure of homogeneous Mo complexes for H2 evolution.
Scheme 21. Structure of homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch021
Scheme 22. Structure of homogeneous Mo complexes for H2 evolution.
Scheme 22. Structure of homogeneous Mo complexes for H2 evolution.
Catalysts 11 00217 sch022
Figure 4. (a) STM image of atomically re-dissolved MoS2 particle on Au(111) [170]; (b) electron micrograph of MoS3-CV film on ITO [172]; (c) CV scan dependency of the mass of MoS2+x film [174]; (d) TEM images (TEM, Ori) and EDS mapping of Mo carbonitride [194]; (e) synthetic process of Mo2N/CeO2@NF [196]; (f) polarization curves of Mo2N/CeO2@NF-0.05, 20 wt% Pt/C, CeO2@NF and Mo2N@NF [196]. Ref. [170] with permission from copyright 2007 American Association for the Advancement of Science, Ref. [172] with permission from the Royal Society of Chemistry, and Refs. [174,194,196] with permission from copyright American Chemical Society.
Figure 4. (a) STM image of atomically re-dissolved MoS2 particle on Au(111) [170]; (b) electron micrograph of MoS3-CV film on ITO [172]; (c) CV scan dependency of the mass of MoS2+x film [174]; (d) TEM images (TEM, Ori) and EDS mapping of Mo carbonitride [194]; (e) synthetic process of Mo2N/CeO2@NF [196]; (f) polarization curves of Mo2N/CeO2@NF-0.05, 20 wt% Pt/C, CeO2@NF and Mo2N@NF [196]. Ref. [170] with permission from copyright 2007 American Association for the Advancement of Science, Ref. [172] with permission from the Royal Society of Chemistry, and Refs. [174,194,196] with permission from copyright American Chemical Society.
Catalysts 11 00217 g004
Figure 5. (a,b) HRTEM images of MSQCs [197]; (c) EDS and elemental mapping of MoS2-Ni3S2 heteronanorods [203]; (d) polarization curves and (e) Tafel slopes of NF, Ni3S2/NF, MoS2-Ni3S2 HNRs/NF, and Pt/C on NF and MoS2/NF [203]; (f) overpotentials of IrO2 and Co/Mo carbides (Co6Mo6C2/NCRGO with different Co/Mo ratios 1:1, 2:1, 3:1) at 10 mA/cm2 [204]; (g) polarization curves of EO Co3Mo/Cu, EO Co/Cu, EO Cu, and Ir/C/Cu (Inset: SEM image of EO Co3Mo/Cu) [208]; and (h) polarization curves for electrocatalytic water splitting of nanoporous Co3Mo/Cu and EO Co3Mo/Cu electrodes, nanoporous NiFe/Cu and EO NiFe/Cu electrodes and Pt/C/Cu and Ir/C/Cu, Inset: Comparison of current density at 1.65 V [208]. Reproduced from Refs. [197,203,204] with permission from copyright American Chemical Society and Ref. [208] with permission from copyright 2020 Springer Nature.
Figure 5. (a,b) HRTEM images of MSQCs [197]; (c) EDS and elemental mapping of MoS2-Ni3S2 heteronanorods [203]; (d) polarization curves and (e) Tafel slopes of NF, Ni3S2/NF, MoS2-Ni3S2 HNRs/NF, and Pt/C on NF and MoS2/NF [203]; (f) overpotentials of IrO2 and Co/Mo carbides (Co6Mo6C2/NCRGO with different Co/Mo ratios 1:1, 2:1, 3:1) at 10 mA/cm2 [204]; (g) polarization curves of EO Co3Mo/Cu, EO Co/Cu, EO Cu, and Ir/C/Cu (Inset: SEM image of EO Co3Mo/Cu) [208]; and (h) polarization curves for electrocatalytic water splitting of nanoporous Co3Mo/Cu and EO Co3Mo/Cu electrodes, nanoporous NiFe/Cu and EO NiFe/Cu electrodes and Pt/C/Cu and Ir/C/Cu, Inset: Comparison of current density at 1.65 V [208]. Reproduced from Refs. [197,203,204] with permission from copyright American Chemical Society and Ref. [208] with permission from copyright 2020 Springer Nature.
Catalysts 11 00217 g005
Table 1. Compared CO2 reduction catalytic efficiencies of Mo-based electrocatalysts.
Table 1. Compared CO2 reduction catalytic efficiencies of Mo-based electrocatalysts.
CatalystFECurrent DensityCO2 Reduction Product (Major)Ref.
Mo electrode50%-Methanol[58]
layer-stacked MoS2~98%−65 mA/cm2 at −0.76 V vs. RHECO[59]
5% Nb-doped VA-MoS282%−237 mA/cm2 at −0.8 V vs. RHECO[60]
Ta-doped VA-MoS2-−98~68 mA/cm2 at −0.8 V vs. RHECO[60]
MoBiSx nanosheets71%−12.1 mA/cm2 at −0.7 V vs. SHEMethanol[62]
MoS/Se monolayer45%−43 mA/cm2 at −1.15 V vs. RHECO[68]
Cu-doped MoS285%−17 mA/cm2 at −1.7 V vs. SCECO[64]
NCMSH93%−34.31 mA/cm2 at −0.7 V vs. RHECO[65]
MoS2 nanoflake
(choline chloride)
93%−315 mA/cm2 at −0.8 V vs. RHECO[67]
N-MoS2@NCDs90%−36 mA/cm2 at −0.9 V vs. RHECO[66]
MoP@In-PC97%−43.8 mA/cm2 at −2.2 V vs. Ag/AgNO3Formic acid[69]
Table 2. Faradaic efficiencies, NH3 formation rates of different Mo-based catalysts.
Table 2. Faradaic efficiencies, NH3 formation rates of different Mo-based catalysts.
CatalystFENH3 Formation RateRef.
(110)-oriented Mo nanofilm0.72%3.09 × 10−11 mols−1cm−2 at −0.49 V vs. RHE[138]
MoO3 nanosheets1.9%4.80 × 10−10 mols−1cm−2 at −0.5 V vs. RHE[142]
MoN nanosheets1.15%3.01 × 10−10 mols−1cm−2 at −0.3 V vs. RHE[143]
MoS21.17%8.08 × 10−11 mols−1cm−2 at −0.5 V vs. RHE[144]
1T-MoS2@Ti3C210.94%30.33 μg h−1mg−1cat. at −0.3 V vs. RHE[145]
Table 3. Compared catalytic efficiencies of Mo-based HER electrocatalysts.
Table 3. Compared catalytic efficiencies of Mo-based HER electrocatalysts.
CatalystOnset PotentialOverpotentialTafel Slope (mV/dec)Exchanged Current Density (A/cm2)Ref.
MoS2 nanoparticles--55~601.3 × 10−7[170]
[Mo3S4]4+ cluster−0.2 V vs. NHE-1202.2 × 10−7[175]
MoSx|Ti|n+p-Si photocathode0.33 V vs. RHE-39-[171]
MoS3–CV-200 mV at −15 mA/cm2401.3 × 10−7[172]
amorphous MoS3-200 mV at −4.8 mA/cm242-[173]
MoS2+x film-170 mV at −20 mA/cm2--[174]
MoSe2 nanosheets−0.15 V vs. RHE290 mV at −10 mA/cm2101-[189]
MoSe2|RGO−0.05 V vs. RHE115 mV at −10 mA/cm269-[189]
Cu2MoS4 135 mV (onset)954.0 × 10−5[187]
MoS2/CoSe2−11 mV vs. RHE68 mV at −10 mA/cm2367.3 × 10−5[188]
NiMoNx nanosheet−78 mV vs. RHE-35.92.4 × 10−4[190]
commercial MoB 100 mV (onset)551.4 × 10−6[192]
commercial Mo2C 100 mV (onset)561.3 × 10−6[192]
MoP/S-90 mV at −10 mA/cm2502.0 × 10−4[191]
MoP-117 mV at −10 mA/cm2505.0 × 10−5[191]
γ-MoC−0.24 V vs. RHE
(at 0.18 mA/cm2)
-121.63.2 × 10−6[193]
Mo carbonitride−0.05 V vs. RHE-46~51-[194]
MoO2@PC-RGO0 V vs. RHE64 mV at −10 mA/cm2414.8 × 10−4[195]
Mo2N/CeO2@NF-0.05-26 mV at −10 mA/cm237.8-[196]
[Mo3S13]2 nanocluster 180 mV at −10 mA/cm240 [176]
Mo3S13 film−130 mV vs. RHE200 mV at −10 mA/cm237 [177]
[Mo3S13]2 attached at (rGO-CNTs) aerogels−110 mV vs. RHE179 mV at −10 mA/cm260.2 [178]
dimeric [Mo2S12]2 cluster 161mV at −10 mA/cm239 [179]
Table 4. Onset potentials, overpotentials, Tafel slopes Mo-based OER catalysts.
Table 4. Onset potentials, overpotentials, Tafel slopes Mo-based OER catalysts.
CatalystCurrent Density
(mA/cm2)
Overpotential Tafel Slope (mV/dec)Ref.
MoS2 on NF20310 mV 105[198]
Mesoporous MoO2 nanosheets on NF10260 mV 54[199]
Co-Mo-B10320 mV 56[200]
Ni-Mo/Cu nanowire20280 mV 66[201]
FeCoMo nanocomposite10277 mV 27.74[202]
MoS2-Ni3S2 heteronanorods10249 mV 57[203]
bimetallic Co/Mo carbides10260 mV 50[204]
MoFe:Ni(OH)2/NiOOH nanosheet100280 mV 47[205]
MSQDs-AC10370 mV 39[197]
MoO2-Co2Mo3O8@C nanorods10320 mV 88[206]
Ni2Mo3N hetero-metal nitride10270 mV 59[207]
EO Co3Mo alloy nanoparticles164350 mV 82[208]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, D.; Lee, J.; Seo, J. Molybdenum-Containing Metalloenzymes and Synthetic Catalysts for Conversion of Small Molecules. Catalysts 2021, 11, 217. https://doi.org/10.3390/catal11020217

AMA Style

Kim D, Lee J, Seo J. Molybdenum-Containing Metalloenzymes and Synthetic Catalysts for Conversion of Small Molecules. Catalysts. 2021; 11(2):217. https://doi.org/10.3390/catal11020217

Chicago/Turabian Style

Kim, Donghyeon, Jaeheon Lee, and Junhyeok Seo. 2021. "Molybdenum-Containing Metalloenzymes and Synthetic Catalysts for Conversion of Small Molecules" Catalysts 11, no. 2: 217. https://doi.org/10.3390/catal11020217

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