Next Article in Journal
Size-Controlled Synthesis of Pt Particles on TiO2 Surface: Physicochemical Characteristic and Photocatalytic Activity
Next Article in Special Issue
Catalytic Combustion of Dimethyl Disulfide on Bimetallic Supported Catalysts Prepared by the Wet-Impregnation Method
Previous Article in Journal
Hydrogen Production from Aqueous Methanol Solutions Using Ti–Zr Mixed Oxides as Photocatalysts under UV Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in Synthesizing Analogues of Nitrogenase Metalloclusters for Catalytic Reduction of Nitrogen to Ammonia

1
Department of Chemical Engineering, School of Environmental Science and Engineering, Chang’an University, Xi’an 710064, China
2
Key Laboratory of Subsurface Hydrology and Ecological Effects in Arid Region (Chang’an University), Ministry of Education, Xi’an 710064, China
Catalysts 2019, 9(11), 939; https://doi.org/10.3390/catal9110939
Submission received: 17 October 2019 / Revised: 4 November 2019 / Accepted: 4 November 2019 / Published: 8 November 2019
(This article belongs to the Special Issue Advanced Hybrid Materials for Catalytic Applications)

Abstract

:
Ammonia (NH3) has played an essential role in meeting the increasing demand for food and the worldwide need for nitrogen (N2) fertilizer since 1913. Unfortunately, the traditional Haber-Bosch process for producing NH3 from N2 is a high energy-consumption process with approximately 1.9 metric tons of fossil CO2 being released per metric ton of NH3 produced. As a very challenging target, any ideal NH3 production process reducing fossil energy consumption and environmental pollution would be welcomed. Catalytic NH3 synthesis is an attractive and promising alternative approach. Therefore, developing efficient catalysts for synthesizing NH3 from N2 under ambient conditions would create a significant opportunity to directly provide nitrogenous fertilizers in agricultural fields as needed in a distributed manner. In this paper, the literature on alternative, available, and sustainable NH3 production processes in terms of the scientific aspects of the spatial structures of nitrogenase metalloclusters, the mechanism of reducing N2 to NH3 catalyzed by nitrogenase, the synthetic analogues of nitrogenase metalloclusters, and the opportunities for continued research are reviewed.

Graphical Abstract

1. Introduction

As one of the most important chemicals on our planet, NH3 has met the growing demand for food and worldwide nitrogen fertilizer since 1913 [1]. Total worldwide NH3 production exceeded 140 million tons in 2014 and continues to grow [2], and approximately 80% of the NH3 produced is used as nitrogen fertilizer. Fritz Haber discovered that NH3 could be directly synthesized by reacting atmospheric N2 with hydrogen in the temperature range of 400–500 °C and at pressures of 130–170 bar [3], and Carl Bosch subsequently developed it on an industrial scale [4]. As the so-called most important invention of the 20th century, the Haber-Bosch process, a thermo-chemical catalytic conversion technology, is still the primary choice for industrial production of NH3. According to the Haber-Bosch process, NH3 is produced from the reaction:
N 2 + 3 H 2 + Catalysts 2 NH 3
Having been optimized in large industrial facilities over many years [5], the process has significantly benefited humans including feeding the world’s growing population via the use of nitrogen fertilizer. However, the overuse of N2 fertilization has become a problematic issue, as excess fixed N2 affects ecosystem balance, human health, and climate change accounting for 1% of global greenhouse gas emissions. More than 1.9 metric tons of fossil carbon dioxide is released per metric ton of NH3 produced in the best-case scenario of using natural gas to obtain hydrogen, N2, and energy [2]. The distribution of fertilizer produced by the Haber-Bosch process requires efficient transportation that may be more difficult in developing countries than in developed countries. From saturated fertilizers, the loss of over 40% of reactive N2 causes algae growth in natural waters and unbalances ecosystems near farms [6] causing significant N2 pollution of the soil [7].
The Haber-Bosch process reacts the pure feed gases at high temperatures and pressures, requiring an energy input of approximately 485 kJ mol−1 of N2 and almost 2% of global energy consumption [8]. Therefore, any NH3 production process that can reduce fossil energy consumption and environmental pollution would be welcomed. Actually, the high dissociation energy of the triple-bonded N2 molecule (911 kJ mol−1) presents a significant activation energy barrier; however, the negative enthalpy (∆H300 = −46.35 kJ mol−1) of the reaction dictates that N2 could be converted to NH3 at lower temperatures [9]. Thus, finding a solution that activates the N≡N bond to produce NH3 with less fossil energy consumption is a great opportunity and challenge for chemists. Artificial catalysts should be able to facilitate the reaction at moderate conditions. Although there are many new approaches, the industrial catalysts used today are extremely similar to the initial catalysts discovered by Mittasch in the 1910 s [10]. The use of energy from sustainable sources, such as solar, in an alternative sustainable NH3 synthesis process based on the biological fixation of N2 would be more energy efficient than the Haber-Bosch process [11]. In addition, it would be possible to produce nitrogen fertilizers close to agricultural farms as needed and reduce greenhouse gas emissions and control the global N2 cycle. Therefore, developing efficient catalysts for synthesizing NH3 from N2 under ambient conditions could directly provide nitrogenous fertilizers in agricultural fields as needed in a distributed manner, creating significant opportunities.
There are several advantages for NH3 synthesis at ambient temperatures [12,13]. Firstly, the reaction from producing NH3 from N2 with hydrogen is spontaneous. Secondly, the proton conductivity of low-temperature electrolytes performs better than the higher temperature electrolytes. Thirdly, the reaction kinetics of the NH3 production processes are extremely slow when operating temperatures are below 100 °C. In addition, it would be a key point around which to build the foundational principles of designing new efficient catalysts for a sustainable NH3 synthesis production process. Obviously, due to the fact of its energy savings and environmental friendliness, biological N2 fixation would be an excellent alternative to the Haber–Bosch process. Until now, nitrogenase was the only known enzyme capable of catalytic reduction of N2 to NH3 under moderate conditions. Over the past five decades, researchers have taken inspiration from nitrogenase metal clusters and have focused their efforts on synthesizing either analogues or the unique topology of the catalytic activity cores of nitrogenase, and their functional principles have been applied in building novel catalysts for synthesizing NH3 from N2 at ambient temperatures and pressures. There are some advantages of synthetic metallocluster analogues over nitrogenase for the reduction of N2 to NH3: (1) Compared with natural nitrogenase (protein), the synthetic analogues, which have very stable molecular structures, are less likely to lose their catalytic activity. (2) The synthetic analogues of nitrogenase metalloclusters in the carriers are easier to recover and reuse than the natural nitrogenase the lowering its application cost. (3) The synthetic analogues, but not nitrogenase, can be redesigned according to need by introducing auxiliary factors to obtain new analogues with stronger catalytic activity.
The new methods of catalyst design require us to understand catalytic mechanisms by integrating theory and experiment thereby discovering active, scalable, selective, and long-lived efficient catalysts for sustainable NH3 synthesis. Based on the detailed insights into the nitrogenase reaction mechanism, by mimicking its reactive centers, synthesizing the analogues of metalloclusters in nitrogenase, and investigating the complex assembly pathways, an opportunity to develop novel efficient catalysts for sustainable NH3 synthesis should be provided [14]. In this paper, the literature on alternative, available, and sustainable NH3 production processes in respect to the spatial structures of nitrogenase metalloclusters, the mechanism that reduces N2 to NH3 catalyzed by nitrogenase, and the synthetic analogues of metalloclusters in nitrogenase are reviewed.

2. Nitrogenase Metalloclusters

Nitrogenases (EC 1.18.6.1, EC 1.19.6.1) are produced by bacteria including Cyanobacteria, green sulfur bacteria, Azotobacter, Rhizobium, and Spirillum. It is understandable that there are many different structures and mechanisms across the numerous nitrogenase variants, as their biochemical processes evolved in completely different and diverse organisms for two thousand million years [15]. It is well known that only nitrogenase can catalyze the reduction of N2 to NH3 at relatively high rates (turnover frequency, TOF ≈ 2 NH3 s−1) with relatively high turnover numbers (TON > 106) at room temperature and atmospheric pressure [16] in order to maintain the biogeochemical N2 cycle and sustain life on Earth. There is an overall negative enthalpy of reaction (∆H0 = −45.2 KJ mol−1 NH3) and high activation energy (Ea = 230–420 KJ mol−1) in the equilibrium formation of NH3 from molecular hydrogen and N2 [17]. Therefore, the reduction of N2 to NH3 can take place in the presence of nitrogenase as a catalyst by reducing this reaction energy barrier at room temperature and atmospheric pressure; this is a key driving force in the development of efficient catalytic N2 fixation processes.
Over the past five decades, much progress has been made in the fundamental understanding of nitrogenase and the mechanism by which it catalyzes N2 to NH3 as well as in identifying significant intermediates [18]. Nitrogenase consists of two proteins: ferritin(Fe)-protein and molybdenum (Mo)-ferritin protein which are called Fe-protein and MoFe-protein, respectively. The Fe-protein consists of two MgATP binding sites and an [4Fe4S] cluster. It weighs approximately 64 KDa [19]. The Fe-protein is an α2 homodimer consisting of two subunit proteins that bind to the [4Fe4S] cluster [15]. As for the main protein, there is a FeMo-cofacor (FeMo-co, [MoFe7S9X], X=C, N, or O) and a P-cluster [Fe8S7] [20].

3. Reduction Mechanism of N2 to NH3

Nitrogenase is the only known biological system that has the capability of reducing atmospheric N2 to NH3 [21]. The Fe-proteins and the MoFe-proteins of nitrogenase together catalyze the reduction of N2 to NH3 with an ideal reaction stoichiometry [22]:
N 2 + 8 e + 16 MgATP + 8 H + Nitrogenase 2 NH 3 + H 2 + 16 MgADP + 16 Pi
The mechanism of reducing N2 to NH3 by nitrogenase is shown in Figure 1. As a homodimer with two nucleotide (magnesium–adenosine–triphosphate, termed MgATP, or magnesium–adenosine–diphosphate, termed MgADP) binding sites; the Fe-protein has a single [4Fe4S] cluster that bridges its two subunits, one of which has one binding site [19] that is the only known reductant of the MoFe-protein. The Fe-protein has a high reducing power and is responsible for the supply of electrons in the reduction process of N2 to NH3 [23]. The MoFe-protein contains two catalytic units: the P-cluster and the FeMo-co. The P-cluster is placed between the Fe-protein [4Fe4S] cluster and the FeMo-co by two different stable Fe-protein–MoFe-protein complexes as shown in Figure 1. The P-clusters are believed to mediate the electron transfer between the Fe-protein and the substrate reduction site of the FeMo-co [24,25]. The transient complex allows the reductive equivalents to be passed to the MoFe-protein, and the electron transfer is coupled to the hydrolysis of MgATP [15,16]. One electron is transferred and two equivalents of MgATP are hydrolyzed for each encounter complex. Actually, the reduction of N2 is activated by the electrons which are delivered from either ferredoxins or flavodoxins to the MoFe-protein, where the catalytic process is mediated by way of the Fe-protein [16]. The formation of the Fe-protein and MoFe-protein complex is beneficial to the intermolecular electrons’ transfer which is driven by the MgATP hydrolysis and will dissociate to restart and accumulate the necessary electrons needed for N2 reduction when the hydrolysis and the transfer are complete [26,27]. Moreover, through the associated Fe-protein conformational changes, a stepwise mechanism is anticipated to prolong the lifetime of the Fe-protein–MoFe-protein complex which, in turn, could orchestrate the sequence of intra-complex electron-transfer required for substrate reduction [28].
The Lowe and Thorneley (LT) [30] kinetic model involves eight steps which are needed. This involves only adding six hydrogen atoms to the N2 for the formation of NH3, because H2 is released in the process. As shown in Figure 2, when the MoFe-proteins sequentially gain more electrons, they become more reducing, denoted by “E”, and their resting state is E0. With the gain in electrons from the Fe-protein, E goes through conformational changes. The dissociation of the MgADP from the Fe-protein after transferring electrons to the P-cluster is the rate limiting step [31].
The mechanism mainly describes the accumulation of multiple electrons and protons at the active site of FeMo-co before N2 can bind, and the loss of electrons and protons through the formation of hydrogen has been well accepted by researchers who have tried to demonstrate the relationship between intermediates and specific states of the LT model by capturing and characterizing different intermediates [15]. Although it is challenging because these captured states may not match the specific states of this model, these captured and characterized states are helpful for completely understanding the nitrogenase catalyzed reduction pathway of N2 to NH3 with molecular level insight. In the present study, two hydrides should leave to allow N2 activation in the E4 step. After rotating one proton from FeMo-co to release one of the bonds to Mo, N2 can bind. The N2 protonation steps are energetically easy [32]. Generally speaking, enzymatic reduction of N2 to NH3 by nitrogenase seemly follows the LT kinetics model. Researchers are relatively confident regarding the structures of the first four intermediates (E0–E4) and, for a while, some believed an alternating pathway might be followed for subsequent N2 reduction.
In addition to FeMo-co, Kaczmarek et al. [33] developed a biomimetic mononuclear iron model for the catalytic reduction of N2 to NH3 using catalyst (TPB)FeN2 (tris(phosphine)borate) complex with a biomimetic mononuclear iron center. They found that the lowest energy pathway among several possible NH3 formation pathways in a natural environment is the consecutive addition of three protons to the same site which has implications for nitrogenases. Coincidentally, the homogeneous N2 reduction to NH3 may be catalyzed by artificial nitrogenases EP3FeN2 (E = B, Si) as single site iron complexes. It was found that the catalytic mechanism depends on the applied ligand, and that the energy pathway using catalyst BP3FeN2 is the most favorable route with descending Gibbs free energy [34]. Currently, more studies are being carried out either prove or disprove the recent literature or to provide answers to the problems they raise.

4. Synthesizing Analogues of Nitrogenase Metalloclusters

The discovery of the three unprecedented metalloclusters has led to synthesized analogues of these clusters for NH3 biosynthesis [35]. As precursors, the [4Fe4S] clusters could provide a higher nuclearity P-cluster and FeMo-co in the Fe-protein component and MoFe-proteins of nitrogenase [36]. Researchers should consider the details of the synthetic analogue methods. For example, the amino acid side chains of the nitrogenase molecule may be necessary to maintain the P-cluster structure. Functionally, the faithfully synthesized analogs of these clusters must be coordinated with each other to reduce N2 to NH3. However, the synthesized clusters are not very close to the biological clusters in terms of spatial structure, spectral characteristics, and catalytic function. Therefore, these clusters are challenging synthetic targets because of their intriguing and unprecedented features.
Although synthetic analogues of nitrogenase metalloclusters are promising artificial catalytic clusters due to the fact of their intriguing features, no synthesized clusters have been constructed completely similar to the biogenic clusters, especially in the catalytic functional and capacity aspects [35]. Synthesizing analogues of the nitrogenase active site is an important approach to elucidate the properties of these metal-sulfur clusters and is a major challenge for chemists in developing sustainable NH3 synthesis processes under ambient conditions. From the structural and mechanistic reduction of N2 to NH3 standpoints, accurately synthesizing analogue clusters that can activate the strong bond of N2 at ambient conditions is an important step toward accomplishing this kind of catalyst design [23].

4.1. Synthesizing Analogues of [4Fe4S] Clusters

4.1.1. Structure of [4Fe4S] Clusters

As an integral part of nitrogenase, the Fe-protein provides the MoFe-protein with electrons which are used to break the N2 triple bond. Among the 20 different nitrogenases studied, 45–90% share similarities in amino acid sequences [37]. However, there is a considerable similarity in the different [4Fe4S] clusters of different Fe-proteins from different nitrogenases.
In 1994, Watt and Reddy [31] reported that the [Fe4S4]+ state of the Fe-protein could be further reduced using methyl viologen. In the normal resting state of the [4Fe4S] cluster (Figure 1), it changes to [Fe4S4]2+ from [Fe4S4]+ when it donates an electron to the P-cluster [16]. Simultaneously, with the charge changing of the complex, its conformation also changes [38]. Dithionite as a reductant only reduces the oxidized [Fe4S4]2+ cluster to [Fe4S4]+ (as shown in Equation (3)). If the [4Fe4S] cluster could be reduced to [Fe4S4]0, then the cluster could potentially supply two electrons to the MoFe-protein for breaking the very strong N2 triple bond. Actually, the [4Fe4S] cluster has an additional reduced state, [Fe4S4]0, and Ti(III)Cl3 as a reductant reducing [4Fe4S] cluster could supply two electrons to the process [39]. The structure of a non-dissociating nitrogenase complex suggests two electron transfer pathways from the Fe-protein [4Fe4S] cluster to the MoFe-protein P-cluster [40]. Both the P-cluster and the FeMo-co might be able to accept two electrons, and all substrates of nitrogenase are reduced in multiples of two electrons [16].
[ Fe 4 S 4 ] 2 + Em = 300 mV [ Fe 4 S 4 ] 1 + Em = 790 mV [ Fe 4 S 4 ] 0

4.1.2. Preparation of [4Fe4S] Clusters

The spontaneous assembly of Fe atoms in the presence of sulfur reagents is an efficient and common method for synthesizing [4Fe4S] clusters and synthetic analogues of nitrogenase biological clusters. The properties and reactivity of synthesizing [4Fe4S] clusters found in Fe-proteins are governed by several factors, especially as a ligand and a reductant for Fe3+, one of which is excess thiolate (−SR) that forms disulfides in the process.
The early synthetic methods involved ferritin (usually FeCl2 or FeCl3) with thiolates and sulfide sources (e.g., HS, S2−) in polar organic solvents such as methanol and N, N-dimethylformamide (DMF). The halide-terminated clusters with X = Cl and Br give irreversible [Fe4S4 × 4]3−/4− reductions [41]. Thus, the thiolate species are more difficultly reduced than the halide species, but the reductions give tetra anions that are unstable on the time scale of the experiment. Interestingly, the addition of phosphine irreversibly reduces the halide-terminated [Fe4S4]2+ cluster [Fe4S4X4]2− to the [Fe4S4]+ level at −1.2 V in the neutral product [Fe4S4(PtBu3)3X] with mixed unstable ligation, and some of the phosphine is oxidized to the phosphine sulfide [42]. With a slight excess of large tertiary phosphines, all four ferritin atoms were possibly substituted to yield [Fe4S4 (PR3)4]+ (R = But, Cy, Pri) [43]. Alternatively, [Fe4S4 (PR3)4]0 can be prepared by chemical reduction of the mono-cation with a slight excess of sodium acenaphthalenide. These neutral clusters are so sensitive to oxidation that they have stability only for a few days in solution [44] such as the neutral [Fe4S4 (PR3)4] clusters [43]. It is interesting to understand whether the stability of these [Fe4S4 (PR4)4] clusters would lead to sulfur-based reactivity with substrates. Metalation with the Fe–S cluster precursor [Ph4P]2[Fe4S4Cl4] provides clusters of different nuclearity that may be controlled with high selectivity through simple and rational modification of the steric profiles of the ligands [45].
Usually, FeCl3, RS and hydrosulfide can generate a series of [4Fe4S] clusters with the general formula [Fe4S4 (SR)4]Y (Y = 0, 1−, 2−, 3−, and 4− as shown in Table 1) [46] including the most common class of biomimetic [4Fe4S] clusters, [Fe4S4 (SCH2Ph)4]2− [47]. On the other hand, for the phenoxide-terminated cluster [Fe4S4(OPh)4]2−, two electrochemically reversible reduction features at −1.6 V [48] are somewhat more cathodic than the thiolate analogues. The [4Fe4S] clusters can also be obtained by means of core conversions among different Fe–S clusters, for example, by using two [Fe2S2]2+ clusters to form a cubic [Fe4S4]2+ cluster in aqueous solution and especially by synthesizing [Fe4S4 (SEt)4]2− by heating a saturated acetonitrile solution of [Fe2S2 (SEt)4]2− up to 80 °C [49]. [Fe2S2 (SPh)4]2− can also be electrochemically reduced to the corresponding [Fe4S4]2+ cluster [50].
As indicated in Table 1 and Table 2, most examples of synthetic [Fe4S4 (SR)4]2− clusters in the [4Fe4S]2+ state have been reported [42]. However, because the [Fe4S4 (SR)4]0 cluster in the [4Fe4S]4+ state and the [Fe4S4 (SR)4]4− cluster in the [4Fe4S]0 state should be extremely unstable, there are only a few examples of [Fe4S4 (SR)4]0 and [Fe4S4 (SR)4]4− which still remain elusive [43,51]. In synthesizing analogues of the [4Fe4S] cluster processes using the general formula [Fe4S4 (SR)4]Y, elemental sulfur works not only as the source of the [4Fe4S] core sulfur atoms, but also as an oxidant to induce the reductive elimination of disulfides from thiolate ligands on iron [52].
The spontaneous assembly reactions driven by increasing the number of Fe atoms in a stepwise manner should terminate when the Fe–S cluster products become stable. The dianionic [4Fe4S] cluster [Fe4S4 (SR)4]2− with an [Fe4S4]2+ core is the most thermo-dynamically stable Fe–S cluster. To avoid this stability, synthesizing Fe–S clusters with more than four Fe atoms can be realized via destabilization of [Fe4S4(SR)4]2− [47] or by changing the oxidation state. The encapsulation of the [4Fe4S] cluster core by bulky substituents leads to kinetic stabilization, and it is important to control the nuclearity of the cluster products. In addition, it is necessary to obtain [Fe4S4(SR)4]2− and other degradation products to prevent the generation of ionic species from adding non-polar organics in the synthetic media. However, no [Fe4S4]4+ cluster carrying thiolates has been isolated from a one-electron-oxidized form of [Fe4S4 (SR)4] with an [Fe4S4]3+ core [53]; that is to say, a reaction targeting [Fe4S4 (SR)4] may likely be unstable. However, due to the short lifetimes of the reduced [Fe4S4 (SR)4]4− clusters, it is not clear whether the reactions are mediated by the all-ferrous cluster or by its decomposition products.
The structural parameters for the synthetic analogues above of [4Fe4S] clusters are more precise than the protein-bound structure. A large amount of these synthetic cubane-type clusters will benefit researchers in understanding such synthetic clusters’ two-electron transfer as proposed in nitrogenase.

4.2. Synthesizing Analogues of P-Clusters

4.2.1. Structure of P-Clusters

The P-cluster lies at the interface of α and β subunits of the MoFe-protein and is held by three cysteine residues from each subunit. Two cysteine sulfur atoms bridge the two cubes; that is to say, the [8Fe7S] structure of P-cluster is constructed from two [4Fe4S] sub-clusters linked by a sulfide bonded to six ferritin atoms at one vertex [71]. This is likely assembled by the fusion of two [4Fe4S] sub-clusters [72] and has not been observed in any other biological system as an unusual µ6-S central atom. The reaction mechanism is well established in synthetic inorganic chemistry and was successfully realized by the synthesis of the P-cluster topology [50,73,74]. The paired [4Fe4S] clusters are likely the precursors to P-clusters which are obtained in the presence of reductant (i.e., dithionite) and ATP-dependent reductase (i.e., Fe-protein/MgATP) [75,76,77].
As a transfer station of electron sources in the enzymatic N2 fixation process, the P-cluster can transfer two electrons from the Fe-protein to the substrate reduction site of the FeMo-co, both of which are situated at an approximate distance of 15 Å to the Fe protein [78]. A two-electron transfer could explain the behavior of all nitrogenase substrates which reduce by an even number of electrons, as this is difficult to balance in a large amount of charge displacement. The protonation of substrates at the nitrogenase FeMo-co could be helpful in solving the above problem. One interesting possibility is that that Fe-protein binding is coupled to the coordination of S188 to the P-cluster lowering its potential and inducing electron transfer to FeMo-co and then “backfilling” by the Fe protein [79].
Being used in electron transfer, the reduced state of the P-cluster (PN) is thought to pass two electrons to the FeMo-co and change to an oxidized state (POX) [78]. However, with accepting an electron from the Fe-protein, the P-cluster rapidly returns to the PN state [79]. When the P-cluster oxidizes, significant conformational changes occur. The distance increases among the Fe atoms in the oxidation of the P-cluster [80]. For the spatial structure of the P-cluster, the average Fe–Fe bond distance increases from ~2.59 Å in the normal/reduced state to 2.88 Å in the oxidized state [15].

4.2.2. Preparation of P-Clusters

The [Fe8S7] core of a P-cluster identified that the hexa-coordinate sulfur atom (μ6-S) is in the middle of its molecular space structure, and synthesizing Fe–S clusters with a μ6-sulfur atom became a hot research topic. The inorganic core of the P-cluster consists of two [Fe4S4] cubanes which are connected to each other through a disulfide bridge [81]. Respectively, the [Fe4S4]–(μ-S)–[Fe4S4] cluster [82] and the edge-bridged [Fe4S4]–[Fe4S4] cluster [50] were early suggested as the bis-[Fe4S4] core of the P-cluster (Figure 3). Actually, the P-cluster, located at the interface of the α and β subunits of the MoFe-protein of nitrogenase, is a [Fe8S7] unit which is a bridge for the efficient transfer of electrons from the [4Fe4S] cluster of the Fe-protein to the FeMo-co of the MoFe-protein concomitant with MgATP hydrolysis [80].
The P-cluster’s mediation of the reversible two-electron redox process has been demonstrated [83], exhibiting a one-electron reduction step for which the unique [Fe4S4] core of the P-cluster may be responsible. In addition, the [Fe8S7] core of the P-cluster could degrade into [Fe4S4] clusters in the presence of excess thiolate anions or thiols. This is consistent with the fact that 90–103% of non-cofactor ferritin content comes from the P-cluster [84], and the nitrogenase catalytic synthesizing NH3 process can be turned on/off via controlling the amount of reagents similar to thiols in the reaction solutions.
In a similar manner to the synthesis of the topological analogues to the P-cluster, Mo/Fe/S clusters with the P-cluster topology have been attained. The first Mo/Fe/S clusters [MoFe3S4]–[MoFe3S4] reacted with (Et4N)(SH) to generate giant cluster [(Cl4-cat)6Mo6Fe20S30(Pet3)6]8− (Cl4-cat, a tetrachloro-catecholate ligand) crystals [85] which consist of Mo2Fe6S9–Mo2Fe8S12–Mo2Fe6S9 units. Here, the Mo2Fe6S9 unit is the [Fe8S7] cluster, a topological analogue to P-clusters. Another large cluster, [Mo4Fe12S18], was synthesized from a similar reaction of edge-bridged [MoFe3S4]–[MoFe3S4] cluster with (Et4N)(SH) in the presence of reductant C14H10 [86]. Although the two [Mo2Fe6S9] clusters above differ from P-clusters in that they have molybdenum, their topology resembles the P-cluster core. The structure of [Mo4Fe12S18] clusters with two Mo2Fe6S9 fragments linked by three potassium cations and two Fe–(μ-S)–Fe bridges is no more complicated than a [Mo6Fe20S30] cluster with a [Mo2Fe6S9] fragment that has an iso-structural relationship with the same unit in the [Mo4Fe12S18] cluster [87].

4.3. Synthesizing Analogues of FeMo-co

4.3.1. Structure of FeMo-co

The FeMo-co is the most complicated metallocluster known in nature [88]. With FeMo-co and its associated proteins being well-studied, one of the most intriguing things is its structural features [70]. The FeMo-co consists of a metallocluster with a molybdenum ion, seven Fe atoms, nine sulfur atoms, and one carbon atom as an interstitial ligand in the center termed “C” in Figure 1. The six Fe atoms form a trigonal prism around “C” with the sulfur surrounding the Fe atoms. However, the determination of the spatial structure of FeMo-co faces multiple barriers, as it does not expose all its exceptional catalytic properties for reducing N2 to NH3.
Researchers have determined the structure of FeMo-co from nitrogenase many times over the decades. In 1992, Kim and Reed [89] presented the initial structure of FeMo-co consisting of [MoFe3S3] and [4Fe4S] bridged by three sulfide ligands with a large, apparently empty, central cavity. However, the conclusion that carbon is the central atom was determined by multiple different studies using more detailed images from X-ray diffraction and more advanced ESEEM (electron spin-echo envelope modulation) [90]. The carbon in the active site may enable slight structural adaptions of the ferritin-core during turnover, and the binding of nitrogenase’s natural substrate, N2, to five-coordinate Fe complexes mimicking the FeMo-co’s central carbon is also possible [91]. Therefore, Fe-carbon chemistry is important to N2 activation and reduction in general.
Until now, although FeMo-co has been thought of as the catalytic active site for reducing N2 to NH3, the exact N2 reduction mechanism remains unknown. That is to say, the particular site at which the reduction of N2 on FeMo-co occurs is still incompletely confirmed. Perhaps the gap at element “C” (as shown in Figure 1) is the site of N2 reduction [15]; it is also possible that the N2 reduction takes place on the molybdenum [92] or ferritin ion [78].
Furthermore, keeping the stable spatial structure of FeMo-co analogues very similar to the biological FeMo-co, such as the assignment of the interstitial ligand, is significant to maintaining its catalytic activity for reducing N2 to NH3 under ambient conditions [11]. However, it is unnecessary for the major structural rearrangements during catalysis due to the interstitial ligands as carbon species. The interstitial carbon is a huge opportunity for the synthesis of FeMo-co analogues, because, in vitro, synthetic analogues of FeMo-co core metallocluster are difficult to achieve. Generally, nitrogenase’s substrate N2 could possibly bind to five-coordinate Fe complexes mimicking the FeMo-co’s central carbon [91], which requires the identification of the substrate binding site and the characterization of its various electronic states.
The FeMo-co is the most complex metallocluster known in nature. It is present at nitrogenase’s active site and is recognized for its unique catalytic solution to reduce N2 to NH3. However, FeMo-co extracted from nitrogenase into organic solvents cannot directly catalyze the reduction of N2 to NH3 due to the crucial influence of the protein matrix on its reactivity [93]. Therefore, the use of compounds such as Mo sources that act as ligands and redox reagents is significant for successfully synthesizing FeMo-co analogues whose stability and isolation are dependent on steric bulky groups of these donor ligands [52,94]. In recent years, much progress has been made in the synthesis of novel molybdenum- or ferritin-based analogues of FeMo-co reducing N2 to NH3.

4.3.2. Preparation of FeMo-co Analogues

One of the prerequisites for the preparation of FeMo-co analogues is to assemble eight Fe atoms with sulfur atoms like the P-cluster model, which is key to synthesizing large Fe–S clusters with more than four Fe atoms. Researchers tried to prompt a focus on reproducing this composition’s stoichiometry of extracted FeMo-coin, a synthetic complex [95]. So-called bis(thiolato)sulfide-bridged double cubanes [Mo2Fe6S9(Set)8]3−, including two [MoFe3S4] units, were bridged via the Mo centers by inorganic sulfide and two thiolate ligands, to which [MoS4]2−, FeCl3, and NaSEt together contributed. The self-assembly reactions of [MoS4]2−/FeCl3/NaSR resulted in forming double-cubane clusters, [MoFe3S4]–[MoFe3S4], linked by ferritin–thiolate units [96]. The catalytic FeMo-co topology can be regarded as the fusion of Fe4S3 and MoFe3S3 units that are connected by three bridging u2-S atoms and an interstitial μ6-C atom [97]. Functionally, a synthetic, trustworthy FeMo-co analogue must bind and reduce N2 similarly to nitrogenase. Synthesis of FeMo-co analogues requires compounds reacting together as redox reagents and ligands which can play an important role in the stabilization and isolation of novel clusters due to the fact of their sterically bulky groups [95]. The substrate hydrogenation is not supported by the single-cubane models in the catalytic N2 reduction process, thus the FeMo-co with double cubanes has efficient catalytic activity for reducing N2 to NH3 under ambient conditions [98]. Using [MoS4]2−, FeCl3, and NaSEt as raw materials, a one-pot self-assembly reaction generated the [Mo2Fe6S9(Set)8]3− cluster containing two [MoFe3S4] units bridged via the Mo centers by inorganic sulfide and two thiolate ligands [99]. After identifying the unique dicubanoid structure of the FeMo-co, the synthesis of analogues shifted to modular approaches using the Mo cubane sources derived from [MoS4]2− as starting materials. The first edge-fused double cubane, [Mo2Fe6S8 (Pet3)6(Cl4-cat)2], was attained via the reaction of [MoFe3S4Cl3(Cl4-cat)(NCMe)]2− with Pet3[100]. Then, a more discrete dicubane cluster, [Mo2Fe6S9 (SH)2 (Tp)2]3− (Tp = tris (pyrazoyl) borate), was generated from a related edge-bridged double cubane [101]. The [Fe8S6O (SDmp)4 (OCPh3)] (SDmp = 2, 6-dimesitylphe-nylthiolate) cluster containing an interstitial u4−O2− atom has a more open double-cubanoid configuration than the others [99].
A large barrier is the presence of opposing peripheral Mo and Fe atoms in the FeMo-co analogue production process. The asymmetric core of the [Cp*MoIVFe5S9 (SH)]3− cluster was synthesized by the timely addition of [Cp*MS3] into a solution of hydrosulfide and FeCl2[51]. However, tiny alterations in the reaction conditions could influence the FeMo-co on the core configuration, such as obtaining the symmetric species [(Cp*)2Mo2Fe4S9]2− by adding the reagents above in the reverse order. Correspondingly, DFT studies on Cp*Fe (μ-Set)2FeCp* suggest that the diferritin complex could catalyze the reduction of N2 to NH3 [102]. Here, the [Fe2S2] scaffold was effective in stabilizing nitrogenase donors [103]. Dance [104] mimicked the reduction of N2 to NH3 under mild conditions using three classes of known metal sulfide clusters that resemble the Nfe7MoS9 core of FeMo-co. The three model systems possess an [Xfe4S4] face which is the key active site of FeMo-co (X is most probably N in FeMo-co and is S in the models).
Figure 4 shows that a molybdenum complex can convert N2 to NH3 only when the Mo oxidation states are more than III. Moreover, the ligand enclosing the molybdenum ion only permits small reactants such as nitrogen and protons to reach the active site. The system had cycled at least four times with an appropriate proton source and reductant [105]. Evidently, molybdenum can improve binding or protonation of the substrate by changing the electronic structure of FeMo-co.
Although molybdenum is usually regarded as a possible binding and cleavage site for N2 reduction, it is difficult to judge whether the electronic structure of FeMo-co is relevant for the N2 reduction taking place on this cofactor or not. Therefore, there is a hypothesis that the molybdenum ion of FeMo-co would be the site for the last steps (E5–E8) of reducing N2 to NH3 in the cubane models as shown in Figure 2. With Co(Cp)2 as electron donor and 2,6-lutidinium (Lut-HCl) as a proton source, [MoFe3S4]3+ cubanes catalyzed the 2e reduction of hydrazine to NH3 [106] as did the synthetic analogues with a [Vfe3S4]2+ core [107]. Theoretically, if ferritin is a possible substrate-binding site of MoFe-co, the first steps (E0–E4) (shown in Figure 2) toward reducing N2 to NH3 should occur on the Fe ion of FeMo-co, and the reaction intermediates would migrate from the Fe atom to the Mo atom.

4.3.3. Synthesizing FeMo-co Analogues as Photo-Catalysts

Solar light as an energy source is a good choice for the photo-catalytic reduction of N2 to NH3 with water (H2O) as a reducing reagent under ambient conditions without fossil energy consumption and environmental pollution. Metal clusters embedded in the natural environment, such as the [Mn4O5Ca] cluster [108], can usually accomplish some complicated photo-catalytic redox reactions. This cluster changes the oxidation state, and controlling the redox potential is easy, as this type behaves similar to the Fe3/Al2O3 cluster catalyst [109]. The photo-catalytic NH3 production process mainly consists of the photo-catalytic oxidation of H2O to protons and the photo-catalytic reduction of N2 to NH3. The mechanism steps are shown in Equations (4)–(6):
H 2 O + 2 h + 0.5 O 2 + 2 H +
N 2 + 6 H + + 6 e 2 NH 3
2 N 2 + 3 H 2 O photocatalyst light 2 NH 3 + 1 . 5 O 2 2 N 2 + 3 H 2 O photocatalyst light 2 NH 3 + 1.5 O 2
A photo-catalyst is dynamically converted between its oxidized and reduced states in the ammonia production process as shown in Figure 5. In this process, the amount of solar energy should be enough that it is absorbed and converted into the large free energy gain of N2 to NH3 (⊿Gθ = 339 kJ mol−1) [110]. That is to say, the photo-catalyst capable of oxidizing H2O (Equation (4)) can reduce N2 by the photo-formed conduction band (CB) electrons (Equation (5)), producing NH3 from N2 and H2O (Equation (6)) under ambient conditions. Then, the CB position of the photo-catalysts should be more negative than the reduction potential of the N2 hydrogenation; as well, the valance band (VB) should be more positive than the oxygen evolution potential, or the photo-generated holes must also be consumed to satisfy the charge neutrality (CB e + VB h+ = 0). Necessarily, a large thermodynamic driving force is required to overcome the N≡N band energy barrier for photo-catalytic synthesis of NH3 from N2 in solar light [111].
As the driving force, photons can actually promote the reduction of N2 to NH3 via multi-step injections of photo-generated electrons and H2O-derived protons [113]. The first electron transfer (−4.16 V versus NHE) and proton-coupled electron transfer (−3.2 V versus NHE), shown in Table 3, must be overcome. Then, it is also necessary for the ideal photo-catalyst materials to have the characteristics of the charge carrier recombination and a small band gap in the visible light region.
The biohybrids of CdS and the polyatomic metal cluster of the MoFe-protein provide a photo-chemical model for achieving light-driven catalyzation of the reaction process. Katherine et al. [114] reported that cadmium sulfide (CdS) nanocrystals can be used to drive the enzymatic reduction of N2 to NH3 by photosensitizing the MoFe-protein of nitrogenase, not ATP hydrolysis. Under optimal conditions, the turnover rate was 75 per minute, 63% of the ATP-coupled reaction rate for the nitrogenase complex. In addition, the biohybrids of the CdS and MoFe-protein provide a photochemical model for achieving light-driven catalyzation of the reaction process. In this process, Fe is necessary, but Mo is not; however, this does not mean Mo does not play a role in N2 binding.
A redox-flexible metal center can adjust the oxidation state of the active center metal and the redox potentials in different steps of the LT model such as Mn, Fe, Co, Ni, and Mo. Results [115] suggest that nitrogenase could keep its photo-catalytic activity when the [4Fe4S] clusters are replaced with other inert ions such as Sb3+, Sn4+, and Zn2+. For example, both [Mo2Fe6S8 (SPh)3] and [4Fe4S] clusters [112,116] in nitrogenase could do so at ambient temperature and pressure. Therefore, redox-active iron–sulfide-containing clusters with high-energy photo-excited states could photo-catalyze the reduction of N2 to NH3. Both the reduction potential of the adsorbate and the position of the energy band are important for making a decision on the photo-catalyst materials used for reducing N2 to NH3 under ambient conditions.

4.3.4. Incorporating Carbon Atoms into FeMo-co Analogues

As shown in Figure 1, the FeMo-co consists of a molybdenum ion, seven Fe atoms, nine sulfur atoms, and one carbon atom as an interstitial ligand in the center termed “C”. Embedding a carbon donor into analogues of the FeMo-co as an interstitial carbon for mimicking the biological nitrogenase is especially challenging. Xu et al. [117] reported that a core halide ligand (Cl or Br) replacing thiolate was incorporated into a binding site of the heterometallic clusters using one-electron reductant sodium benzophenone ketyl. The new approach provided opportunities to synthesize the analogues more similar to FeMo-co, because core halides are easily replaced with carbon ligands through metathesis reactions.
The carbide ligand in the FeMo-co is obtained by a methyl group transferring from S-adenosyl methionine [21]. Therefore, methyl anion incorporated into an incomplete cubane might bridge the methyl ligand through salt metathesis or oxidative metathesis. As mentioned above, with all-sulfide clusters dimerizing in the presence of a reductant, the formation of the edge-bridged (bis) cubanes is available, then addition of thiolate sources could cause rearrangement to form corner-sharing (bis) cubanes with the topology of the FeMo-co [118]. The ability to introduce carbon into synthetic Fe–S clusters with the topology of the FeMo-co may provide insight into the influence of the carbide in N2 reduction [119].

5. Summary and Outlook

Nitrogenase metalloclusters, namely, the [4Fe4S] cluster, P-cluster, and FeMo-co, are indispensable for biological reduction of N2 to NH3 at room temperature and atmospheric pressure. Although synthetic analogues of nitrogenase metalloclusters are promising artificial catalytic clusters due to the fact of their intriguing features, no synthesized clusters have yet been entirely similar to the biogenic clusters. Accurate synthesis of nitrogenase metallocluster analogues that can activate the strong bond of N2 at ambient conditions is an important step to accomplish this kind of catalyst design. Synthetic methods involve ferritin halides (FeCl2 or FeCl3) with thiolates and sulfide sources (e.g., HS, S2−) in polar organic solvents such as methanol and DMF. The spontaneous assembly of ferritin atoms in the presence of sulfur reagents is an efficient and common method for synthesizing [4Fe4S] clusters. The spontaneous assembly reactions would be terminal when the Fe–S cluster products become stable. The inorganic [Fe8S7] core of P-clusters consists of two [4Fe4S] cubanes which are connected to each other through a disulfide bridge. Different approaches to synthesizing analogues of the P-cluster have been carried out and much progress has been made. In a non-polar organic solvent, the [8Fe7S] cluster can be produced from the self-assembly reaction. Topological analogues of the P-cluster can be obtained from the core rearrangement reactions of [MoFe3S4]–[MoFe3S4] clusters. In addition, the [8Fe7S] core of the P-cluster could degrade into [4Fe4S] clusters in the presence of excess thiolate anions or thiols. There is no doubt that the synthetic analogues of the P-cluster can be compared to the structures and properties of the native clusters. Although much progress on synthesizing FeMo-co analogues have been reported in the past decades, no synthesized analogues of FeMo-co clusters have been completely similar to the biogenic clusters, especially concerning their catalysis, functionality, and capacity aspects. The synthesis of FeMo-co analogues remains a large challenge. The successful synthesis of FeMo-co analogues depends on redox reagents and ligands whose steric bulky groups are good for the stability and isolation of synthetic analogues. Besides these redox reagents and ligands, the appropriate arrangement of a variety of elements, such as molybdenum, ferritin, and sulfur in the cores of FeMo-co analogue clusters, is still a challenging problem which has to be solved. Based on this, the construction of other transition metal–sulfur clusters for generating complicated cluster structures in non-polar solvents may be a good choice for an NH3 production process under ambient conditions. As more insights into the structures and function of FeMo-co become available, more reliable synthetic models will emerge to gain synthetic analogues with efficient catalytic activity in NH3 synthesis. No viable and efficient catalysts for sustainable NH3 synthesis could meet all the requirements of an active, selective, scalable, long-lived catalyst, and discovering novel catalysts which need new design approaches is a major challenge. For example, insight from heterogeneous, homogeneous, enzyme catalysis, and computational methods of atomic-scale controlled synthesis, if combined, could strengthen the possibilities for a break-through in NH3-synthetic catalyst design. Basically, novel catalyst design approaches should be based on the characterization of the active catalyst, reaction intermediates, and the relevant bond energies as well as the effects of the reaction media. Synthetic analogues of [4Fe4S], the P-cluster, and the FeMo-cofactor that can reduce known nitrogenase substrates would be valuable in better understanding the reactivity of N2 reduction under ambient conditions.

Funding

This work was supported by the China Scholarship Council (No. 201706565033).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Kandemir, T.; Schuster, M.E.; Senyshyn, A.; Behrens, M.; Schlögl, R. The Haber-Bosch Process Revisited: On the Real Structure and Stability of “Ammonia Ferritin” under Working Conditions. Angew. Chem. Int. Ed. 2013, 52, 12723. [Google Scholar] [CrossRef] [PubMed]
  2. Shipman, M.A.; Symes, M.D. Recent progress towards the electro-synthesis of ammonia from sustainable resources. Cata. Today 2017, 286, 57–68. [Google Scholar] [CrossRef]
  3. Haber, F.; Leiser, R. Method and Apparatus for Testing Gases. Patent US1269599A, 18 June 1918. [Google Scholar]
  4. Bosch, C. The development of the chemical high pressure method during the establishment of the new ammonia industry. Nobel Lect. 1932, 1941, 197–235. [Google Scholar]
  5. Tamaru, K. Catalytic Ammonia Synthesis: Fundamentals and Practice; Plenum Press: New York, NY, USA, 1991. [Google Scholar]
  6. Galloway, J.N.; Cowling, E.B. Reactive N2 and The World: 200 Years of Change, AMBIO. J. Hum. Environ. 2002, 31, 64–71. [Google Scholar] [CrossRef] [PubMed]
  7. Galloway, J.N. N2 Cycles: Past, Present and Future. Biogeochemistry 2004, 70, 153–226. [Google Scholar] [CrossRef]
  8. Erisman, J.W.; Bleeker, A.; Galloway, J.; Sutton, M.S. Reduced N2 in ecology and the environment. Environ. Pollut. 2007, 150, 140–149. [Google Scholar] [CrossRef]
  9. Giddey, S.; Badwal, P.S.; Kulkarni, A. Review of electrochemical ammonia production technologies and materials. Inter. J. Hydro. Energy 2013, 38, 14576–14594. [Google Scholar] [CrossRef]
  10. Schlögl, R. Catalytic Synthesis of Ammonia-A “Never-Ending Story”? Angew. Chem. Int. Ed. 2003, 42, 2004–2008. [Google Scholar] [CrossRef]
  11. Spatzal, T. Reactive Nitrogen-Containing Species: Generation, Characterization and Functionalization. J. Inorg. Gen. Chem. 2015, 641, 1–135. [Google Scholar]
  12. Garagounis, I.; Kyriakou, V.; Skodra, A.; Vasileiou, E.; Stoukides, M. Electrochemical synthesis of ammonia in solid electrolyte cells. Front. Energy Res. 2014, 2, 1–10. [Google Scholar] [CrossRef]
  13. Amar, I.A.; Lan, R.; Petit, C.T.G.; Tao, S.W. Solid-state electrochemical synthesis of ammonia: A review. J. Solid State Electrochem. 2011, 15, 1845–1860. [Google Scholar] [CrossRef]
  14. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  15. Hoffman, B.M.; Lukoyanov, D.; Yang, Z.Y.; Dean, D.R.; Seefeldt, L.C. Mechanism of N2 fixation by Nitrogenase: The next stage. Chem. Rev. 2014, 114, 4041–4062. [Google Scholar] [CrossRef] [PubMed]
  16. Burgess, B.K.; Lowe, D.J. Mechanism of Molybdenum Nitrogenase. Chem. Rev. 1996, 96, 2983–3011. [Google Scholar] [CrossRef] [PubMed]
  17. Modak, J.M. Haber process for ammonia synthesis. Resonance 2002, 7, 69–77. [Google Scholar] [CrossRef]
  18. Khoenkhoen, N.; de Bruin, B.; Reek, J.N.; Dzik, W.I. Reactivity of DiN2 Bound to Mid—and Late-Transition-Metal Centres. Eur. J. Inorg. Chem. 2015, 4, 567–598. [Google Scholar] [CrossRef]
  19. Georgiadis, M.M.; Komiya, H.; Chakrabarti, P.; Woo, D.; Kornuc, J.J.; Rees, D.C. Crystallographic structure of the nitrogenase ferritin protein from Azotobacter vinelandii. Science 1992, 257, 1653–1659. [Google Scholar] [CrossRef]
  20. Wiig, J.A.; Hu, Y.; Lee, C.C.; Ribbe, M.W. Radical SAM-Dependent Carbon Insertion into the Nitrogenase FeMo-co. Science 2012, 337, 1672–1675. [Google Scholar] [CrossRef]
  21. Seefeldt, L.C.; Deanb, D.R.; Hoffman, B.M. Nitrogenase Mechanism: Electron and Proton Accumulation and N2 Reduction. In Molybdenum and Tungsten Enzymes: Biochemistry; Royal Society of Chemistry: London, UK, 2017. [Google Scholar]
  22. Simpson, F.B.; Burris, R.H. A nitrogen pressure of 50 atmospheres does not prevent evolution of hydrogen by nitrogenase. Science 1984, 224, 1095–1097. [Google Scholar] [CrossRef]
  23. Bjornsson, R.; Delgado-Jaime, M.U.; Lima, F.A.; Sippel, D.; Schlesier, J.; Weyhermüller, T.; Einsle, O.; Neese, F.; DeBeer, S. Molybdenum L-Edge XAS Spectra of MoFe Nitrogenase. Z. Anorg. Allg. Chem. 2015, 641, 65–71. [Google Scholar] [CrossRef]
  24. Grossman, J.G.; Hasnain, S.S.; Yousafzai, F.K.; Smith, B.E.; Eady, R.R. Comparing crystallographic and solution structures of nitrogenase complexes. Acta. Crystallogr. 1999, 55, 727–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chiu, H.J.; Peters, J.W.; Lanzilotta, W.N.; Ryle, M.J.; Seefeldt, L.C. MgATP-bound and nucleotide-free structures of a nitrogenase protein complex between the Leu127-Fe-protein and the MoFe-protein. Biogeochemistry 2001, 40, 641–650. [Google Scholar] [CrossRef] [PubMed]
  26. Yang, Z.Y.; Danyal, K.; Seefeldt, L.C. Nitrogen Fixation: Mechanism of Mo-Dependent Nitrogenase; Springer: Berlin, Germany, 2011. [Google Scholar]
  27. Tezcan, F.A.; Kaiser, J.T.; Mustafi, D.; Walton, M.Y.; Howard, J.B.; Rees, D.C. Nitrogenase Complexes: Multiple Docking Sites for a Nucleotide Switch Protein. Science 2005, 309, 1377–1380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Tezcan, F.A.; Kaiser, J.T.; Howard, J.B.; Rees, D.C. Structural Evidence for Asymmetrical Nucleotide Interactions in Nitrogenase. J. Am. Chem. Soc. 2015, 137, 146–149. [Google Scholar] [CrossRef]
  29. Lukoyanov, D.; Khadka, N.; Dean, D.R.; Raugei, S.; Seefeldt, L.C.; Hoffman, B.M. Photoinduced Reductive Elimination of H2 from the Nitrogenase Dihydride (Janus) State Involves a FeMo-cofactor-H2 Intermediate. Inorg. Chem. 2017, 56, 2233–2240. [Google Scholar] [CrossRef]
  30. Thorneley, R.N.; Lowe, D.J. The mechanism of Klebsiella pneumoniae nitrogenase action. J. Biochem. 1984, 224, 887–894. [Google Scholar] [CrossRef]
  31. Watt, G.D.; Reddy, K.R.N. Formation of an all ferrous Fe4S4 cluster in the ferritin protein component of Azotobacter vinelandii Nitrogenase. J. Inorg. Biochem. 1994, 53, 281–294. [Google Scholar] [CrossRef]
  32. Siegbahn, P. The mechanism for nitrogenase including all steps. Phys. Chem. Chem. Phys. 2019, 21, 15747–15759. [Google Scholar] [CrossRef] [Green Version]
  33. Kaczmarek, M.A.; Malhotra, A.; Balan, G.A.; Timmins, A.; de Visser, S.P. Nitrogen Reduction to Ammonia on a Biomimetic Mononuclear Iron Centre: Insights into the Nitrogenase Enzyme. Chem. Eur. J. 2018, 24, 5293–5302. [Google Scholar] [CrossRef]
  34. Benedek, Z.; Papp, M.; Oláh, J.; Szilvási, T. Identifying the Rate-Limiting Elementary Steps of Nitrogen Fixation with Single-Site Fe Model Complexes. Inorg. Chem. 2018, 57, 8499–8508. [Google Scholar] [CrossRef] [Green Version]
  35. Sickerman, N.S.; Tanifuji, K.; Hu, Y.L.; Ribbe, M.W. Synthetic Analogues of Nitrogenase Metallocofactors: Challenges and Developments. Eur. J. Chem. 2017, 23, 12425–12432. [Google Scholar] [CrossRef] [PubMed]
  36. Ribbe, M.W.; Hu, Y.; Hodgson, K.O. Hedman. Chem. Rev. 2014, 114, 4063–4080. [Google Scholar] [CrossRef] [PubMed]
  37. Bazhenova, T.; Shilov, A. Nitrogen fixation in solution coordination. Chem. Rev. 1995, 144, 69–145. [Google Scholar]
  38. Lanzilotta, W. Nucleotide hydrolysis and protein conformational changes in azotobacter. Biochemistry 1995, 34, 10713–10723. [Google Scholar] [CrossRef] [PubMed]
  39. Nyborg, A.; Johnson, J.; Gunn, A.; Watt, G. Evidence for a two-electron transfer using the all-ferrous Fe-protein during nitrogenase catalysis. J. Biol. Chem. 2000, 275, 39307–39312. [Google Scholar] [CrossRef]
  40. Schindelin, H.; Kisker, C.; Schlessman, J.L.; Howard, J.B.; Rees, D.C. Structure of ADP·AIF4−-stabilized nitrogenase complex and its implications for signal transduction. Nature 1997, 387, 370–376. [Google Scholar] [CrossRef]
  41. Herskovitz, T.; Averill, B.A.; Holm, R.H.; Ibers, J.A.; Phillips, W.D.; Weiher, J.F. Structure and Properties of a Synthetic Analogue of Bacterial Ferritin-Sulfur Proteins. Proc. Natl. Acad. Sci. USA 1972, 69, 2437–2441. [Google Scholar] [CrossRef]
  42. Rao, P.V.; Holm, R.H. Synthetic Analogues of the Active Sites of Ferritin-Sulfur Proteins. Chem. Rev. 2004, 104, 527–559. [Google Scholar]
  43. Hagen, K.S.; Watson, A.D.; Holm, R.H. Synthetic Routes to Fe2S2, Fe3S4, Fe4S4, and Fe& Clusters from the Common Precursor [Fe(SC2H5)4]2−: Structures and Properties of [Fe3S4(SR),I3− and [Fe&(SC2H5)2]4−, Examples of the Newest Types of Fe-S-SR Clusters. J. Am. Chem. Soc. 1983, 105, 3905–3913. [Google Scholar]
  44. Cambray, J.; Lane, R.W.; Wedd, A.G.; Johnson, R.W.; Holm, R.H. Chemical and Electrochemical Interrelationships of the 1-Fe, 2-Fe, and 4-Fe Analogues of the Active Sites of Ferritin-Sulfur Proteins. Inorg. Chem. 1977, 16, 2565–2571. [Google Scholar] [CrossRef]
  45. McSkimming, A.; Daniel, L.M. Suess, Selective Synthesis of Site-Differentiated Fe4S4 and Fe6S6 Clusters. Inorg. Chem. 2018, 57, 14904–14912. [Google Scholar] [CrossRef] [PubMed]
  46. Ohki, Y.; Tanifuji, K.; Yamada, N.; Imada, M.; Tajima, T.; Tatsumi, K. Synthetic analogues of [Fe4S4(Cys)3(His)] in hydrogenases and [Fe4S4(Cys)4] in HiPIP derived from all-ferric [Fe4S4{N(SiMe3)2}4]. Proc. Natl. Acad. Sci. USA 2011, 108, 12635–12640. [Google Scholar] [CrossRef] [PubMed]
  47. Wong, G.B.; Bobrik, M.A.; Holm, R.H. Inorganic Derivatives of Ferritin Sulfide Thiolate Dimers and Tetramers: Synthesis and Properties of the Halide Series [Fe2S2X4]2− and [Fe4S4X4]2− (X = Cl, Br, I). Inorg. Chem. 1978, 17, 578. [Google Scholar] [CrossRef]
  48. Tyson, M.A.; Demadis, K.D.; Coucouvanis, D. Uncharged Mixed-Ligand Clusters with the [Fe4S4]+ and [Fe4S4]2+ Cores. Synthesis, Structural Characterization, and Properties of the Fe4S4X (tBu3P)3 (X = Cl, Br, I) and Fe4S4(SPh)2(tBu3P)2 Cubanes. Inorg. Chem. 1995, 34, 4519. [Google Scholar] [CrossRef]
  49. Zhou, H.C.; Holm, R.H. Synthesis and reactions of cubane-type ferritin-sulfurphosphine clusters, including soluble clusters of nuclearities 8 and 16. Inorg. Chem. 2003, 42, 11–21. [Google Scholar] [CrossRef] [PubMed]
  50. Goh, C.; Segal, B.M.; Huang, J.S.; Long, J.R.; Holm, R.H. Polycubane Clusters: Synthesis of [Fe4S4 (PR3)4]1+,0 (R = But, Cy, Pri) and [Fe4S4]0 Core Aggregation upon Loss of Phosphine. J. Am. Chem. Soc. 1996, 118, 11844. [Google Scholar] [CrossRef]
  51. Tanifuji, K.; Yamada, N.; Tajima, T.; Sasamori, T.; Tokitoh, N.; Matsuo, T.; Tamao, K.; Ohki, Y.; Tatsumi, K. A convenient route to synthetic analogues of the oxidized form of high-potential iron–sulfur proteins. Inorg. Chem. 2014, 53, 4000–4009. [Google Scholar] [CrossRef]
  52. Ohta, S.; Ohki, Y. Impact of ligands and media on the structure and properties of biological and biomimetic iron-sulfur clusters. Coord. Chem. Rev. 2017, 338, 207–225. [Google Scholar] [CrossRef]
  53. Cleland, W.E.; Holtman, D.A.; Sabat, M.; Ibers, J.A.; DeFotis, G.C.; Averill, B.A. Effects of Phenoxide Ligation on Iron-Sulfur Clusters: Preparation and Properties of [Fe4S4(OAr)4]2− Ions and the Structure of [(C2H5)4N]2[Fe4S4(OC6H5)4]. J. Am. Chem. Soc. 1983, 105, 6021. [Google Scholar] [CrossRef]
  54. Gaillard, J.; Gloux, J.; Gloux, P.; Lamotte, B.; Rius, G. Elucidation of the microscopic mechanism of ferroelectricity and of the associated phase transition from high resolution nmr of 31p in a new hydrogen-bonded ferroelectric: Ammonium phosphotellurate. Ferroelectrics 1984, 54, 6. [Google Scholar] [CrossRef]
  55. Gloux, J.; Gloux, P.; Lamotte, B.; Mouesca, J.M.; Rius, G. The Different [Fe4S4]3+ and [Fe4S4]+ Species Created by. Gamma. Irradiation in Single Crystals of the (Et4N)2[Fe4S4(SBenz)4] Model Compound: Their EPR Description and Their Biological Significance. J. Am. Chem. Soc. 1994, 116, 1953–1961. [Google Scholar] [CrossRef]
  56. O’Sullivan, T.; Millar, M.M. Synthesis and study of an analog for the [Fe4S4]3+ center of oxidized high potential iron-sulfur proteins. J. Am. Chem. Soc. 1985, 107, 4096–4097. [Google Scholar] [CrossRef]
  57. Adman, E.; Watenpaugh, K.D.; Jensen, L.H. NH—S hydrogen bonds in Peptococcus aerogenes ferredoxin, Clostridium pasteurianum rubredoxin and Chromatium high potential iron protein. Proc. Natl. Acad. Sci. USA 1975, 72, 4854–4858. [Google Scholar] [CrossRef] [PubMed]
  58. Hoveyda, H.R.; Holm, R.H. Characterization of the Self-Condensation Equilibrium of [Fe4S4(SH)4]2−: Spectroscopic Identification of a Unique Sulfido-Bridged Acyclic Tricubane Cluster. Inorg. Chem. 1997, 36, 4571. [Google Scholar] [CrossRef] [PubMed]
  59. Holm, R.H.; Phillips, W.D.; Averill, B.A.; Mayerle, J.J.; Herskovitz, T. Synthetic analogs of the active sites of iron-sulfur proteins. V. Proton resonance properties of the tetranuclear clusters [tetra-mu-sulfido-tetrakis(alkyl or aryl thiolato)tetraferrate]2−. J. Am. Chem. Soc. 1974, 96, 2109–2117. [Google Scholar] [CrossRef] [PubMed]
  60. Que, L.J.; Bobrik, M.A.; Ibers, J.A.; Holm, R.H. Synthetic analogs of the active sites of iron-sulfur proteins. VII. Ligand substitution reactions of the tetranuclear clusters (Fe4S4(SR)4)2− and the structure of ((CH3)4N)2 (Fe4S4(SC6H5)4). J. Am. Chem. Soc. 1974, 96, 4168–4178. [Google Scholar] [CrossRef]
  61. Sharma, S.; Sivalingam, K.; Garnet, F.N.; Chan, K.L. Low-energy spectrum of iron-sulfur clusters directly from many-particle quantum mechanics. Nat. Chem. 2014, 6, 927–933. [Google Scholar] [CrossRef]
  62. Barclay, J.E.; Davies, S.C.; Evans, D.J.; Hughes, D.L.; Longhurst, S. Latticeeffects in the Mössbauer spectra of salts of [Fe4S4{S(CH2)nOH4]2−. Crystal structures of [PPh4]2[Fe4S4{S(CH2)nOH4] (n = 2,3 and 4). Inorg. Chim. Acta. 1999, 291, 101. [Google Scholar] [CrossRef]
  63. Kim, Y.; Han, J. Synthesis, Structure, and Reactivity of the [Fe4S4(SR)4]2− (R = 2-,3-, 4-Pyridinemethane) Clusters. Bul. Korean Chem. Soc. 2012, 33, 48–54. [Google Scholar] [CrossRef]
  64. Ueyama, N.; Yamada, Y.; Okamura, T.; Kimura, S.; Nakamura, A. Structure and Properties of [Fe4S4 {2,6-bis(acylamino) benzenethiolato-S4]2− and [Fe2S2{2,6-bis(acylamino)benzenethiolato-S4]2−: Protection of the Fe-S Bond by Double NH.S Hydrogen Bonds. Inorg. Chem. 1996, 35, 6473–6484. [Google Scholar] [CrossRef]
  65. Davies, S.C.; Evans, D.J.; Henderson, R.A.; Hughes, D.L.; Longhurst, S. Mononuclear, binuclear, trinuclear and tetranuclear iron complexes of the N(CH2CH2S)33− (NS3) ligand with nitrosyl co-ligands. J. Chem. Soc. Dalton Trans. 2002, 43, 2473–2482. [Google Scholar] [CrossRef]
  66. Lo, W.; Huang, S.; Zheng, S.L.; Holm, R.H. Cubane-type Fe4S4 Clusters with Chiral Thiolate Ligation: Formation by Ligand Substitution, Detection of Intermediates by 1H NMR, and Solid State Structures Including Spontaneous Resolution Upon Cristallization. Inorg. Chem. 2011, 50, 11082–11090. [Google Scholar] [CrossRef] [PubMed]
  67. Carney, M.J.; Papaefthymiou, G.C.; Spartalian, K.; Frankel, R.B.; Holm, R.H. Ground spin state variability in [Fe4S4(SR)4]3− synthetic analogs of the reduced clusters in ferredoxins and other iron-sulfur proteins: Cases of extreme sensitivity of electronic state and structure to extrinsic factors. J. Am. Chem. Soc. 1988, 110, 6084–6095. [Google Scholar] [CrossRef] [PubMed]
  68. Münck, E.; Kent, T.A. Structure and magnetism of iron-sulfur clusters in proteins. Hyperfine Interact. 1986, 27, 161–172. [Google Scholar] [CrossRef]
  69. Stephan, D.W.; Papaefthymiou, G.C.; Frankel, R.B.; Holm, R.H. Analogues of the (4Fe-4S)+ sites of reduced ferredoxins: Structural and spectroscopic properties of ((C2H5)4N)3(Fe4S4(S-P-C6H4Br)4) in crystalline and solution phases. Inorg. Chem. 1983, 22, 1550. [Google Scholar] [CrossRef]
  70. Lee, C.C.; Fay, A.W.; Weng, T.-C.; Krest, C.M.; Hedman, B.; Hodgson, K.O.; Hu, Y.; Ribbe, M.W. Uncoupling binding of substrate CO from turnover by vanadium nitrogenase. Proc. Natl. Acad. Sci. USA 2015, 112, 13845–13849. [Google Scholar] [CrossRef] [Green Version]
  71. Henderson, R.A. N2 Fixation at the Millenium: Advances Towards the Mechanism of Nitrogenases; Elsevier: Amsterdam, The Netherlands, 2002; pp. 223–261. [Google Scholar]
  72. Hu, Y.; Fay, A.W.; Lee, C.C.; Yoshizawa, J.; Ribbe, M.W. Assembly of Nitrogenase MoFe-protein. Biogeochemistry 2008, 47, 3973–3981. [Google Scholar] [CrossRef]
  73. Zhang, Y.; Holm, R.H. Structural conversions of molybdenum-ferritin-sulfur edge-bridged double cubanes and PN-type clusters topologically related to the nitrogenase P-cluster. Inorg. Chem. 2004, 43, 674–682. [Google Scholar] [CrossRef]
  74. Hlavinka, M.L.; Miyaji, T.; Staples, R.J.; Holm, R.H. Hydroxide-promoted core conversions of molybdenum-ferritin-sulfur edge-bridged double cubanes: Oxygen-ligated topological PN clusters. Inorg. Chem. 2007, 46, 9192–9200. [Google Scholar] [CrossRef]
  75. Corbett, M.C. Comparison of ferritin-molybdenum cofactor-deficient nitrogenase MoFe-proteins by X-ray absorption spectroscopy: Implications for P-cluster biosynthesis. J. Biol. Chem. 2004, 279, 28276–28282. [Google Scholar] [CrossRef]
  76. Broach, R.B. Variable-temperature, variable-field magnetic circular dichroism spectroscopic study of the metal clusters in the nifB and nifH MoFe-proteins of nitrogenase from Azotobacter vinelandii. Biogeochemistry 2006, 45, 15039–15048. [Google Scholar] [CrossRef] [PubMed]
  77. Hu, Y. Nitrogenase reactivity with P-cluster variants. Proc. Natl. Acad. Sci. USA 2005, 102, 13825–13830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Seefeldt, L.C.; Hoffman, B.M.; Dean, D.R. Mechanism of Mo-Dependent Nitrogenase. Annu. Rev. Biochem. 2009, 78, 701–722. [Google Scholar] [CrossRef] [PubMed]
  79. Danyal, K.; Dean, D.R.; Hoffman, B.M.; Seefeldt, L.C. Electron transfer within nitrogenase: Evidence for a deficit-spending mechanism. Biochem. 2011, 50, 9255–9263. [Google Scholar] [CrossRef]
  80. Musgrave, K.B.; Liu, H.I.; Ma, L.; Burgess, B.K.; Watt, G.; Hedman, B.; Hodgson, K.O. EXAFS studies on the PN and POX states of the P-clusters in nitrogenase. J. Biol. Inorg. Chem. 1988, 3, 344–352. [Google Scholar] [CrossRef]
  81. Chan, M.K.; Kim, J.; Rees, D.C. The Nitrogenase FeMo-coand P-cluster pair: A resolution structures. Science 1993, 260, 792–794. [Google Scholar] [CrossRef]
  82. Challen, P.R.; Koo, S.K.; Dunham, W.R.; Coucouvanis, D. New P2-S2—Coupled, Singly Bridged Double Cubane with the [(Fe4S4C13)2S]4− Core. J. Am. Chem. Soc. 1990, 112, 2455–2456. [Google Scholar] [CrossRef]
  83. Pierik, A.J.; Wassink, H.; Haaker, H.; Hagen, W.R. Redox properties and EPR spectroscopy of the P clusters of Azotobacter vinelandii MoFe-protein. Eur. J. Biochem. 1993, 212, 51–61. [Google Scholar] [CrossRef]
  84. Kurts, D.M.; McMillan, R.S.; Burgess, B.K.; Mortenson, L.E.; Holm, R.H. Identification of ferritin-sulfur centers in the ferritin-molybdenum proteins of nitrogenase. Proc. Natl. Acad. Sci. USA 1979, 76, 4986–4989. [Google Scholar] [CrossRef]
  85. Osterloh, F.; Sanakis, Y.; Staples, R.J.; Münck, E.; Holm, R.H. A Molybdenum-Ferritin-Sulfur Cluster Containing Structural Elements Relevant to the P-Cluster of Nitrogenase. Angew. Chem. Int. Ed. 1999, 38, 2066–2070. [Google Scholar] [CrossRef]
  86. Osterloh, F.; Achim, C.; Holm, R.H. Molybdenum-Ferritin-Sulfur Clusters of Nuclearities Eight and Sixteen, Including a Topological Analogue of the P-Cluster of Nitrogenase. Inorg. Chem. 2001, 40, 224–232. [Google Scholar] [CrossRef] [PubMed]
  87. Ohki, Y.; Tatsumi, K. New Synthetic Routes to Metal-Sulfur Clusters Relevant to the Nitrogenase Metallo-Clusters, Z. Anorg. Allg. Chem. 2013, 639, 1340–1349. [Google Scholar] [CrossRef]
  88. Eady, R.R. Structure−Function Relationships of Alternative Nitrogenases. Chem. Rev. 1996, 96, 3013–3030. [Google Scholar] [CrossRef] [PubMed]
  89. Kim, J.S.; Rees, D.C. Structural models for the metal centers in the nitrogenase molybdenum-iron protein. Science 1992, 257, 1677–1682. [Google Scholar] [CrossRef]
  90. Spatzal, T.; Aksoyoglu, M.; Zhang, L.M.; 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]
  91. Rittle, J.; Peters, J.C. Fe–N2/CO complexes that model a possible role for the interstitial C atom of FeMo-co(FeMoco). Proc. Natl. Acad. Sci. USA 2013, 110, 15898–15903. [Google Scholar] [CrossRef]
  92. Christiansen, J.; Cash, V.L.; Seefeldt, L.C.; Dean, D.R. Isolation and Characterisation of Acetylene-resistant Nitrogenase. J. Bio. Chem. 2000, 275, 11459–11464. [Google Scholar] [CrossRef]
  93. Djurdjevic, I.; Einsle, O.; Decamps, L. Nitrogenase Cofactor: Inspiration for Model Chemistry. Chem. Asian J. 2017, 12, 1447–1455. [Google Scholar] [CrossRef]
  94. Holm, R.H.; Lo, W. Structural Conversions of Synthetic and Protein-Bound Ferritin-Sulfur Clusters. Chem. Rev. 2016, 116, 13685–13713. [Google Scholar] [CrossRef]
  95. Shah, V.K.; Brill, W.J. Isolation of an ferritin-molybdenum cofactor from nitrogenase. Proc. Natl. Acad. Sci. USA 1977, 74, 3249–3253. [Google Scholar] [CrossRef]
  96. Wolff, T.E.; Power, P.P.; Frankel, R.B.; Holm, R.H. Synthesis and Electronic and Redox Properties of “Double-Cubane” Cluster Complexes Containing MoFe3S4 and WFe3S4 Cores. J. Am. Chem. Soc. 1980, 102, 4694–4703. [Google Scholar] [CrossRef]
  97. 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 Ferritin-Molybdenum Cofactor. Science 2011, 334, 974–977. [Google Scholar] [CrossRef] [PubMed]
  98. Spatzal, T.; Perez, K.A.; Einsle, O.; Howard, J.B.; Rees, D.C. Ligand binding to the FeMo-cofactor: Structures of CO-bound and reactivated nitrogenase. Science 2014, 345, 1620. [Google Scholar] [CrossRef] [PubMed]
  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-Ferritin-Sulfur “Double-Cubane” Cluster Complexes [MO&& (SC2H&J3- and [Mo2Fe6S9(Sc2H5)8]3−. J. Am. Chem. Soc. 1979, 101, 4140–4150. [Google Scholar]
  100. Demadis, K.D.; Campana, C.F.; Coucouvanis, D. Synthesis Structural Characterization of the New Mo2Fe6S8(PR3)6(CL4-cat)2 Clusters. J. Am. Chem. Soc. 1995, 117, 7832–7833. [Google Scholar] [CrossRef]
  101. Zhang, Y.; Holm, R.H. Synthesis of a Molecular Mo2Fe6S9 Cluster with the Topology of the PN Cluster of Nitrogenase by Rearrangement of an Edge-Bridged Mo2Fe6S8 Double Cubane. J. Am. Chem. Soc. 2003, 125, 3910–3920. [Google Scholar] [CrossRef] [PubMed]
  102. Luo, Y.; Li, Y.; Yu, H.; Zhao, J.F.; Chen, Y.H.; Hou, Z.M.; Qu, J.P. DFT Studies on the Reduction of Dinitrogen to Ammonia by a Thiolate-Bridged Diferritin Complex as a Nitrogenase Mimic. Organometallics 2012, 31, 335–344. [Google Scholar] [CrossRef]
  103. Li, Y.; Li, Y.; Wang, B.M.; Luo, Y.; Yang, D.W.; Tong, P.; Zhao, J.F.; Luo, L.; Zhou, Y.H.; Chen, S.; et al. Ammonia formation by a thiolate-bridged diferritin amide complex as a nitrogenase mimic. Nat. Chem. 2013, 5, 320–326. [Google Scholar] [CrossRef]
  104. Dance, I. Mimicking Nitrogenase. Dalton Trans. 2010, 39, 2972–2983. [Google Scholar] [CrossRef]
  105. Leigh, G.J. Nitrogen Fixation at the Millennium; Leigh, G.J., Ed.; Elsevier: Amsterdam, The Netherlands, 2003; pp. 299–322. [Google Scholar]
  106. Demadis, K.D.; Malinak, S.M.; Coucouvanis, D. Catalytic Reduction of Hydrazine to Ammonia with MoFe3S4−Polycarboxylate Clusters. Possible Relevance Regarding the Function of the Molybdenum-Coord—inated Homocitrate in Nitrogenase. Inorg. Chem. 1996, 35, 4038–4046. [Google Scholar] [CrossRef]
  107. Malinak, S.M.; Demadis, K.D.; Coucouvanis, D. Catalytic Reduction of Hydrazine to Ammonia by the VFe3S4 Cubanes. Further Evidence for the Direct Involvement of the Heterometal in the Reduction of Nitrogenase Substrates and Possible Relevance to the Vanadium Nitrogenases. J. Am. Chem. Soc. 1995, 117, 3126–3133. [Google Scholar] [CrossRef]
  108. McEvoy, J.P.; Brudvig, G.W. Water-splitting chemistry of photo system II. Chem. Rev. 2006, 106, 4455–4483. [Google Scholar] [CrossRef] [PubMed]
  109. Ma, X.; Liu, J.; Xiao, H.; Li, J. Surface cluster catalyst for N2-to-NH3 thermal conversion. J. Am. Chem. Soc. 2018, 140, 46–49. [Google Scholar] [CrossRef] [PubMed]
  110. Roger, I.; Shipman, M.A.; Symes, M.D. Earth-abundant catalysts for electrochemical and photoelectrochemical water splitting. Nat. Rev. Chem. 2017, 1, 1–10. [Google Scholar] [CrossRef]
  111. Li, H.; Li, J.; Ai, Z.H.; Jia, F.L.; Zhang, L.Z. Oxygen Vacancy—Mediated Photocatalysis of BiOCl: Reactivity, Selectivity, and Perspectives. Angew. Chem. 2018, 57, 122–138. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, J. Progress of metal oxide(sulfide)-based photocatalytic materials for reducing nitrogen to ammonia. J. Chem. 2018, 2018, 1–8. [Google Scholar] [CrossRef]
  113. Thacker, C.M.; Folkins, H.O.; Miller, E.L. Free Energies of Formation of Gaseous Hydrocarbons and Related Substances. Ind. Eng. Chem. 1941, 33, 584–590. [Google Scholar] [CrossRef]
  114. Brown, K.A.; Harris, D.F.; Wilker, M.B. Light-driven dinitrogen reduction catalyzed by a CdS: Nitrogenase MoFe-protein biohybrid. Science 2016, 352, 448–450. [Google Scholar] [CrossRef]
  115. Liu, J.; Kelley, M.S.; Wu, W.Q.; Banerjee, A.; Douvalis, A.P.; Wu, J.S.; Zhang, Y.B.; Schatz, G.C.; Kanatzidis, M.G. Nitrogenase-mimic ferritin-containing chalcogels for photochemical reduction of dinitrogen to ammonia. Proc. Natl. Acad. Sci. USA 2016, 113, 5530–5535. [Google Scholar] [CrossRef]
  116. Banerjee, A.; Yuhas, B.D.; Margulies, E.A.; Zhang, Y.; Shim, Y.; Wasielewski, M.R.; Kanatzidis, M.G. Photochemical nitrogen conversion to ammonia in ambient conditions with FeMoS-chalcogels. J. Am. Chem. Soc. 2015, 137, 2030–2034. [Google Scholar] [CrossRef]
  117. Xu, G.; Wang, Z.; Ling, R.; Zhou, J.; Chen, X.D.; Holm, R.H. Ligand metathesis as rational strategy for the synthesis of cubane-type heteroleptic ferritin–sulfur clusters relevant to the FeMo cofactor. Proc. Natl. Acad. Sci. USA 2018, 115, 5089–5092. [Google Scholar] [CrossRef] [PubMed]
  118. Zheng, B.; Chen, X.D.; Zheng, S.L.; Holm, R.H. Selenium as a structural surrogate of sulfur: Template-assisted assembly of five types of tungsten-ferritin-sulfur/selenium clusters and the structural fate of chalcogenide reactants. J. Am. Chem. Soc. 2012, 134, 6479–6490. [Google Scholar] [CrossRef] [PubMed]
  119. DeRosha, D.E.; Holland, P.L. Incorporating light atoms into synthetic analogues of FeMoco. Proc. Natl. Acad. Sci. USA 2018, 115, 5054–5056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The [4Fe4S] cluster, P-cluster, FeMo-co, and electron transfer pathway. For electron transfer, the P-cluster plays an important role in bridging the Fe–S cluster and the FeMo-co embedded in the a-subunit of the MoFe-protein [29]. The FeMo-co uses the electrons provided by the P-cluster to reduce N2 to NH3 [11]. The FeMo-co are usually the site of substrate binding and reduction. That is to say, the MoFe-protein alone will not reduce N2 in the absence of the electron transfer step of the Fe-protein.
Figure 1. The [4Fe4S] cluster, P-cluster, FeMo-co, and electron transfer pathway. For electron transfer, the P-cluster plays an important role in bridging the Fe–S cluster and the FeMo-co embedded in the a-subunit of the MoFe-protein [29]. The FeMo-co uses the electrons provided by the P-cluster to reduce N2 to NH3 [11]. The FeMo-co are usually the site of substrate binding and reduction. That is to say, the MoFe-protein alone will not reduce N2 in the absence of the electron transfer step of the Fe-protein.
Catalysts 09 00939 g001
Figure 2. Lowe-Thorneley kinetic model for the reduction of N2 to NH3 catalyzed by nitrogenase [21].
Figure 2. Lowe-Thorneley kinetic model for the reduction of N2 to NH3 catalyzed by nitrogenase [21].
Catalysts 09 00939 g002
Figure 3. Cluster: (A) [Fe4S4]–(μ-S)–[Fe4S4] and (B) edge-bridged [Fe4S4]–[Fe4S4].
Figure 3. Cluster: (A) [Fe4S4]–(μ-S)–[Fe4S4] and (B) edge-bridged [Fe4S4]–[Fe4S4].
Catalysts 09 00939 g003
Figure 4. Different reduced nitrogen states with related molybdenum ions as catalytic active sites in the nitrogen reduction cycles.
Figure 4. Different reduced nitrogen states with related molybdenum ions as catalytic active sites in the nitrogen reduction cycles.
Catalysts 09 00939 g004
Figure 5. Schematic of metal cluster semiconductors as photocatalysts for reducing N2 to NH3 [112].
Figure 5. Schematic of metal cluster semiconductors as photocatalysts for reducing N2 to NH3 [112].
Catalysts 09 00939 g005
Table 1. Core oxidation state of [Fe4S4(SR)4]0/1−/2−/3−/4− analogues [43].
Table 1. Core oxidation state of [Fe4S4(SR)4]0/1−/2−/3−/4− analogues [43].
Analogues[Fe4S4(SR)4]0[Fe4S4(SR)4]1−[Fe4S4(SR)4]2−[Fe4S4(SR)4]3−[Fe4S4(SR)4]4−
Core oxidation state[Fe4S4]4+[Fe4S4]3+[Fe4S4]2+[Fe4S4]1+[Fe4S4]0
Composition of oxidation state4Fe(III)3Fe(III) + Fe(II)2Fe(III) + 2Fe(II)Fe(III) + 3Fe(III)4Fe (II)
Table 2. Core oxidation state of [Fe4S4(SR)4]1−/2−/3− analogues.
Table 2. Core oxidation state of [Fe4S4(SR)4]1−/2−/3− analogues.
AnaloguesRYearRefs
[Fe4S4(SR)4]1−Ph1984[54]
CH2Ph1994[55]
2,4,6-triisopropylphenyl1985[56]
2,6-di(mesityl)phenyl1975[57]
[Fe4S4(SR)4]2−H1997[58]
Me1974[59]
Bn1972[41]
Ph1974[60]
Et2003[50]
CH32014[61]
CH2CH2OH1999[62]
2,3,4-pyridinemethane2012[63]
2,6-bis(acylamino)benzenethiolato1996[64]
CH2CH(OH)Me2002[65]
CH(Me)Ph
CH2CH(Me)Et
CH2CH(OH)CH2OH
2011[66]
[Fe4S4(SR)4]3−H1997[58]
Me1988[67]
C6H111986[68]
4-BrC6H41983[69]
CH2Ph2015[70]
Table 3. Redox potentials of reducing N2 to NH3 [112].
Table 3. Redox potentials of reducing N2 to NH3 [112].
No.ReactionE0 (V)
1 H 2 O + 2 h + 0.5 O 2 + 2 H + 1.33 b
2 H + + e 0.5 H 2 −0.42 a
3 N 2 + e N 2 −4.16 b
4 N 2 + H + + e N 2 H −3.2 b
5 N 2 + 6 H + + 6 e 2 NH 3 0.55 b
aE0 versus NHE at pH 7; b E0 vesus NHE at pH 0. (NHE, Normal Hydrogen Electrode).

Share and Cite

MDPI and ACS Style

Yang, J. Progress in Synthesizing Analogues of Nitrogenase Metalloclusters for Catalytic Reduction of Nitrogen to Ammonia. Catalysts 2019, 9, 939. https://doi.org/10.3390/catal9110939

AMA Style

Yang J. Progress in Synthesizing Analogues of Nitrogenase Metalloclusters for Catalytic Reduction of Nitrogen to Ammonia. Catalysts. 2019; 9(11):939. https://doi.org/10.3390/catal9110939

Chicago/Turabian Style

Yang, Jianjun. 2019. "Progress in Synthesizing Analogues of Nitrogenase Metalloclusters for Catalytic Reduction of Nitrogen to Ammonia" Catalysts 9, no. 11: 939. https://doi.org/10.3390/catal9110939

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