Sei sulla pagina 1di 11

Review Article

https://doi.org/10.1038/s41929-018-0092-7

Catalysts for nitrogen reduction to ammonia


Shelby L. Foster1,5, Sergio I. Perez Bakovic1,5, Royce D. Duda2, Sharad Maheshwari3, Ross D. Milton4,
Shelley D. Minteer   4*, Michael J. Janik3*, Julie N. Renner2* and Lauren F. Greenlee1*

The production of synthetic ammonia remains dependent on the energy- and capital-intensive Haber–Bosch process. Extensive
research in molecular catalysis has demonstrated ammonia production from dinitrogen, albeit at low production rates.
Mechanistic understanding of dinitrogen reduction to ammonia continues to be delineated through study of molecular catalyst
structure, as well as through understanding the naturally occurring nitrogenase enzyme. The transition to Haber–Bosch alter-
natives through robust, heterogeneous catalyst surfaces remains an unsolved research challenge. Catalysts for electrochemi-
cal reduction of dinitrogen to ammonia are a specific focus of research, due to the potential to compete with the Haber–Bosch
process and reduce associated carbon dioxide emissions. However, limited progress has been made to date, as most electro-
catalyst surfaces lack specificity towards nitrogen fixation. In this Review, we discuss the progress of the field in develop-
ing a mechanistic understanding of nitrogenase-promoted and molecular catalyst-promoted ammonia synthesis and provide a
review of the state of the art and scientific needs for heterogeneous electrocatalysts.

A
mmonia is essential to the global economy as a fertilizer 97% overall conversion is achieved by the utilization of these
feedstock, industrial and household chemical, and chemi- unreacted gases. Even with a high conversion efficiency, the HBP
cal precursor in addition to also being considered a future remains energy intensive, and the entire process scheme needed to
fuel alternative and hydrogen storage molecule. Despite drawbacks, achieve reaction conversions of 97% is complex. Overall, the HBP
such as energy use, process complexity, greenhouse gas emissions accounts for ~1% of the world’s yearly natural gas consumption and
and limiting economies of scale in production, there remain no is responsible for ~1% of global energy consumption2. Further, the
viable alternatives to the incumbent Haber–Bosch process (HBP). large plant infrastructure is only economically viable at large econo-
In this Review, we discuss progress in the fields of biocatalysis, mies of scale, where natural gas feedstock is responsible for 50% of
homogeneous catalysis and heterogeneous catalysis to understand the production cost.
the reaction of dinitrogen (N2) reduction to ammonia (NH3) and to Today, there is a need for alternative technologies, and there is a
enable a more sustainable path to NH3 production. growing interest as to whether an electrochemically based system
NH3 is predominately used as an agricultural feedstock in the pro- may possibly succeed as a replacement to the HBP. Electrochemical
duction of synthetic fertilizers. High-yield nutritious crops are depen- reduction of N2 to NH3 is thermodynamically predicted to be more
dent on the addition of these fertilizers and the associated nitrogen energy efficient than the HBP by about 20% (ref. 3). In addition, an
supply due to the degradation of agriculturally usable soil. The need electrochemical process could provide the advantage of eliminating
for NH3 production continues to increase in order to support a grow- fossil fuels as the source of H2 and energy via the use of water mol-
ing global population, and abundant and low-cost NH3 production is ecules as the H2 source and integration with renewable energy tech-
ultimately necessary to provide a stable and affordable food supply. nology. In this scenario, ammonia would be synthesized directly
Demand for NH3 is also likely to increase with additional future uses from humidified air in a carbon-neutral manner. Electrochemical
as a carbon-neutral fuel and hydrogen storage molecule, as a result of systems offer additional advantages, including modularity, scalabil-
its high energy density and easy handling and storage. ity and on-site, on-demand NH3 generation. The HBP is restricted
The current industrial production method, the HBP, produces in economies of scale due to its natural gas feedstock dependence.
500 million tons of NH3 per year1. In the HBP, a mixture of hydro- A smaller plant infrastructure, electrochemical or otherwise, that
gen gas (H2) and nitrogen gas, termed synthetic gas, is passed over utilizes renewable resources would result in the ability to decentral-
an iron-based catalyst commonly promoted with K2O and Al2O3. ize ammonia production and provide availability in remote areas.
Steam reforming of coal and/or natural gas to produce the inlet The electrochemical catalysis community has made some
H2 stream leads to 1.87 tons of the greenhouse gas carbon dioxide progress towards enabling efficient electrochemical N2 reduction
(CO2) released per 1 ton of NH3 (ref. 2). The exothermic N2 reduc- although efforts are plagued by low Faradaic efficiencies due to
tion reaction, shown in reaction (1), requires temperatures of 300– the competing hydrogen evolution reaction, which dominates all
500 °C to improve the kinetics and pressures of 150–200 atm to shift metal-based catalyst surfaces. Few catalysts have resulted in effi-
the reaction equilibrium to be in favour of the products. ciencies greater than 1%, with efficiencies of >​30% achieved only
at high temperature and for limited time and/or current produced4,
N2 + 3H 2 ⇌ 2NH 3 Δ f H0 = −45.9 kJ mol−1 (1) or when used in combination with ionic liquids5. In this Review, we
discuss the key research efforts to delineate the mechanistic aspects
Recycling of unreacted synthetic gas is required, because each of natural biocatalysts (nitrogenases) that catalyse N2 reduction,
pass results in a conversion efficiency of only 15%. However, a as well as homogeneous (molecular) catalyst systems and current

1
University of Arkansas, Ralph E. Martin Department of Chemical Engineering, Fayetteville, AR, USA. 2Case Western Reserve University, Department of
Chemical and Biomolecular Engineering, Cleveland, OH, USA. 3Department of Chemical Engineering, Penn State University, University Park, PA, USA.
4
Department of Chemistry, University of Utah, Salt Lake City, UT, USA. 5These authors contributed equally: Shelby L. Foster, Sergio I. Perez Bakovic.
*e-mail: minteer@chem.utah.edu; mjj13@psu.edu; jxr484@case.edu; greenlee@uark.edu

490 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal


NaTuRE CaTalysIs Review Article
computational theory. The findings of recent and ongoing nitro- a MoFe p
rotein
genase research continue to direct current homogeneous catalyst
research. The synergy of these two fields suggests that a funda- rote
in
Fe p
mental understanding of the roles of specific species (ligands) and
the local catalyst environment in the adsorption and desorption
of nitrogen species will ultimately lead to the development of a
ATP
better heterogeneous catalyst. Therefore, the last section of this 2ATP
Review offers potential insights for selectivity in heterogeneous [4Fe4S] P cluster
catalyst design. 2ADP –
+ 2Pi e
ATP N2
Biological catalysts e–
Biological nitrogen fixation occurs naturally in diazotrophic micro-
organisms through the enzyme nitrogenase. Notably, nitrogenase FeMo-co
operates at mild conditions (<​40 °C, atmospheric pressure) com-
2NH3
pared with catalysts operating in the traditional HBP (300–500 °C,
>​150 bar). The study of this enzyme is thus of great interest in
meeting the grand challenge of sustainable and efficient ammonia
synthesis. The synthesis of ammonia from dinitrogen by nitroge-
nase follows reaction (2) under optimal conditions (where ATP is
adenosine triphosphate, ADP is adenosine diphosphate and Pi is Homocitrate
inorganic phosphate):

N2 + 8H + + 16MgATP + 8e − → 2NH 3 + H 2 + 16MgADP + 16Pi (2)


e– e–
Second event First event
The reaction involves the obligatory hydrolysis of ATP to release
stored chemical energy and thermodynamic or kinetic barriers of
nitrogen reduction. For every molecule of nitrogen that is reduced, [4Fe4S] P cluster FeMo-co
two molecules of ammonia are generated, and protons are also
reduced to form one molecule of H2. Therefore, 25% of the energy b
consumed results in hydrogen production. In the absence of nitro-
gen or other substrates, nitrogenase reverts to the reduction of pro- α-96Arg
tons. α-70Val
Three distinct nitrogenase enzymes sharing similar character-
istics have been found; they are distinguished by their metal-cen-
tred catalytic cofactors (co): FeMo-co, FeFe-co and VFe-co. The
Homocitrate S3B
most widely studied and understood nitrogenase enzyme contains
S2B S2A α-195His
the FeMo-co, and is known as MoFe nitrogenase. Therefore, this Fe6

Review provides a brief overview on the structure and functions of Fe2


Fe7
the MoFe nitrogenase. Fe3
X-ray crystallography has given insight into the crystal struc- Fe1
ture of the MoFe nitrogenase enzyme, and a representation of the
structure is shown in Fig. 16. Nitrogenase consists of two multi-sub- α-442His α-275Cys
unit proteins, both of which are oxygen sensitive, with one protein
serving as a catalytic domain and the other serving as a reducing
domain. The MoFe protein is the catalytic protein to which nitrogen Fig. 1 | Nitrogenase enzyme structure and functions. a, Diagram of one
binds and is reduced to NH3, whereas the second protein, or the half of the nitrogenase complex and electron transfer. b, Detailed diagram
Fe protein, hydrolyses MgATP molecules and transfers electrons to of the FeMo cofactor and surrounding environment. Figure adapted from
the MoFe protein (more specifically, to the FeMo-co) for catalysis. ref. 6, Elsevier.
Table 1 outlines the basic components of these proteins and their
functions.
The Fe protein is a homodimer and each subunit in the protein it is then transferred to the FeMo-co. This model would require the
contains a nucleotide binding site for a MgATP molecule and two P cluster in the MoFe protein to reach a super-reduced state, which
cysteine residues to which the bridging [4Fe4S] cluster binds. The has not been observed. The direct electron transfer model involves
MoFe protein is a larger α​2β​2 tetramer containing two Fe8S7 ‘P clus- the FeMo-co being reduced directly by the Fe protein, meaning that
ters’ (one bridging cluster between each α​ and β​subunit dimer) and the P cluster is not involved. As the P cluster has been shown to
two FeMo cofactors (Fe7MoS9C, located in the α​subunit). The P change oxidation states during the reaction7, this model is unlikely.
cluster is thought to play an exclusive role in transferring electrons The deficit-spending model proposes that an electron is first trans-
originating from coupled ATP hydrolysis in the Fe protein to the ferred from the P cluster to the FeMo-co within the MoFe protein,
FeMo-co of the MoFe protein. The electron transfer to each compo- and only after this is the P cluster reduced again by the F cluster of
nent is depicted in Fig. 1a. the Fe protein. This mechanism for electron transfer is consistent
with recent stopped-flow kinetic experiments8.
Electron transfer. Three models exist to describe the electron Dissociation of the Fe protein from the MoFe protein after elec-
transfer process between the Fe protein and the MoFe protein: the tron transfer has been considered the rate-limiting step9. More
‘sequential’, ‘direct’ and ‘deficit spending’ models. In the sequential recent studies, however, indicate that the hydrolysis of ATP at the
model, for each Fe protein that complexes with the MoFe protein, a Fe protein, and more specifically, the release of phosphate from the
single electron is transferred from the F cluster to the P cluster, where protein, is the rate-limiting step10.

Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal 491


Review Article NaTuRE CaTalysIs

H+/e– H+/e– H+/e– H+/e–


Table 1 | Basic components and functions of the nitrogenase E0 E1H1 E2H2 E3H3 E4H4
enzyme
N2
Protein Domain Function
Fe protein Fe4S4, F cluster Facilitates hydrolysis of MgATP
homodimer and electron transfer to the MoFe H2
H+/e– H+/e– H+/e– H+/e–
(~66 kDa) protein
E8 E7 E6 E5 E4H2N2
Nucleotide binding Facilitates binding of MgATP
sites
Fig. 2 | Single electron–proton transfer model for nitrogenase-mediated
MoFe FeMo-co clusters Catalyses reduction of nitrogen
protein α​2β​2 to ammonia, buried to prevent nitrogen fixation. Simplified representation of the Lowe–Thorneley kinetic
tetramer access to H2O and improve scheme for nitrogen reduction21. The full mechanistic model also features
(~240 kDa) nitrogen selectivity over hydrogen N2 binding at E3.
evolution
Fe8S7, P clusters Responsible for transferring
electrons to the FeMo-co from the uses density functional theory (DFT) calculations to explain how
Fe4S4 cluster of the Fe protein the dissociation of belt-sulfur atoms as H2S (S2B sulfur, Fig. 1b)
from the FeMo-co unveils the reactive Fe sites. The work was based
FeMo-co structure. Researchers have recently made great strides on structural experiments supporting the displacement of a belt-
in determining the structure of nitrogenase, specifically FeMo-co. sulfur atom by CO (ref. 16). According to the model, this step is
The complete molecular structure of the FeMo-co has been deter- critical to the initiation of the N2 reduction process. The subsequent
mined with the identification of an interstitial carbon atom at the re-sealing of the active Fe site by H2S re-absorption, required to free
centre (Fig. 1b)11 . Despite this progress, the understanding of the the second NH3 product, releases H2, which can account for the
electronic structure of the FeMo-co is still debated. Electronic stud- requisite H2 produced per reduced N2 by nitrogenase in the Lowe–
ies suggest that FeMo-co only cycles through one redox couple, with Thorneley kinetic model16. Mechanisms determined in additional
one resting stage MN and a one electron-reduced stage MR (ref. 12). studies discuss the vital role played by the interstitial carbon and
Although studies have often accepted Moiv to be the oxidation state protein environment in initiating the reduction process, which was
of the molybdenum, recent studies propose the reassignment of the not considered by the previous work17.
oxidation state to Moiii (ref. 13). This finding has prompted studies
that clearly assign the charges of the iron and molybdenum atoms in Role of protein environment and interstitial carbon. Polypeptides
the FeMo-co so that substrate reactivity with the Fe and Mo atoms and amino acids located near the FeMo-co have been found to affect
in the FeMo-co can be better understood13,14. Studies have shown the reactivity of the MoFe protein. Using mutant enzymes and mod-
that three of the seven iron atoms in the FeMo-co (iron atoms elling, several important residues have been determined, as sum-
labelled one, three and seven in Fig. 1b) are relatively reduced com- marized in Table 2. Many key amino acids are near the active centre,
pared with the remaining four iron atoms. Given this information, a but studies have also determined that the activity of the enzyme
FeMo-co with three iron atoms in the 2+​oxidation state, four iron can be modified by making certain amino acid substitutions in the
atoms in the 4+​oxidation state and a molybdenum atom with a 3+​ β​-chain, away from the FeMo-cofactor. Specifically, substituting his-
oxidation state would agree with these observations of the FeMo-co. tidine for tyrosine at position β​-98 improved reactivity of the MoFe
protein compared with the wild-type for the conversion of hydra-
Reaction mechanism. With the existence of three types of metal- zine to ammonia while using unnatural reducing agents18. Given the
centred catalytic cofactors (FeMo-co, FeFe-co and VFe-co), the position of the substitution, near where the Fe protein would trans-
precise substrate binding site is still debated. With respect to the fer an electron to the MoFe protein, it is believed that this amino
FeMo-co, the involvement of amino acids in the catalysis and the acid substitution modifies the structure of the MoFe protein in a
spatial positioning (Fig. 1b) suggests the active site to be the Fe–S way that mimics the conformational change that occurs when the
face, but determining the location and binding mode of dinitrogen Fe protein binds to the MoFe protein.
within the nitrogenase remains a challenge. The peptide environment and the presence of interstitial carbon
Despite knowledge of the nitrogenase structure and active site, may play an even greater role as theorized by recent computational
the mechanism of nitrogenase-mediated N2 reduction to NH3 studies17. Current proposals suggest that the first hydrogen binding
remains unsolved. Amino acid substitutions and freeze-quench takes place at the interstitial carbon atom rather than at the S2B sul-
trapping, however, has isolated intermediates in support of draft fur, as originally predicted by Dance19.
mechanisms. Lowe and Thorneley developed an eight-step kinetic In addition, high-resolution crystal structure analysis has
model for reduction of nitrogen to ammonia by nitrogenase (Fig. 2). revealed that polypeptide chains around the FeMo-co are likely
More recently, Hoffman et al. proposed a nitrogenase mecha- blocking water from the active site20. This analysis suggests that the
nism of N2 activation and reduction that unifies with the Lowe– α​-helix containing α​-70Val is arranged such that the hydrophobic
Thorneley kinetic model. In the Lowe–Thorneley kinetic model, amino acids face the active site, and the more hydrophilic amino
one proton and one electron bind to the cofactor during each stage acids, and thus water molecules, are arranged on the opposite side.
from one to eight, designated from E0 to E8. This draft mechanism The chain containing α​-195His is also is thought to be arranged where
proposes H2 generation to proceed through a reductive elimination hydrophobic residues are facing the active site20. This same study
of hydrides producing a highly reduced FeMo-co intermediate stage also postulates an interesting feature of the nitrogenase enzyme:
(E4). At stage E4, four protons are bound to the FeMo-co including a chain of water molecules from the surface of the enzyme to the
two protons that are bound to two iron atoms each (Fe–H–Fe). At FeMo-co. This water chain consists of a ‘proton bay’ and a chain
this stage, N2 is able to bind to the FeMo-co and the dinitrogen triple of eight water molecules that acts as a pathway for transmitting
bond can be split. protons from the outer layers of the enzyme to the FeMo-co. This
Multiple computational studies have also proposed possible mech- structural feature may also be interpreted as a conduit for ammonia
anisms for dinitrogen reduction. A model proposed by Varley et al.15 product to escape, as H2O and NH4+ are hard to distinguish.

492 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal


NaTuRE CaTalysIs Review Article
catalyst, the first non-enzymatic system was developed to reduce
Table 2 | Important primary and secondary structural features
dinitrogen to ammonia at ambient conditions through a distal type
in the MoFe protein with their function
mechanism25. The only other reported Mo-based catalytic system
Amino acid/helix Function for dinitrogen reduction at ambient conditions is a di-Mo-based
α​-70 Val
Controls access of substrates to the active sites catalyst with PNP-type pincer ligands. Using DFT calculations,
on FeMo-co30. a distal type pathway was proposed and it also explained the key
role played by the dinuclear complex for catalytic activity by allow-
α​-Helix containing Blocks water from the active site through
ing stable bridging of nitrogen ligand, which acts as a medium to
α​-70Val orientation of hydrophobic amino acids20.
transfer e– from one Mo core to the active site of the other core26.
α​-96Arg Contributes to the control of substrates to the Although work has continued for Mo-based molecular catalysts,
FeMo-co. Hydrogen bonds to S5A sulfur atom in including recent advancements with PNP-type (with the highest
the FeMo-co17.
turnover number for molecular catalysts of 415 per Mo atom), PCP-
α​-195His Facilitates the delivery of protons to the type and NNN-type pincer ligands27–29, more bioinspired mimic
active site. Stabilizes the protonated nitrogen catalysts have become an emerging area. Key molecular catalyst
intermediate96. complexes are summarized in Table 3.
α​-Helix containing Presents hydrophobic residues to the active site20.
α​-195His Bioinspiration and mechanistic investigation. Recent biochemi-
α​-191Gln Impacts reactivity of the FeMo-co97. cal and spectroscopic studies on the nitrogenase enzyme provide
convincing evidence that dinitrogen reduction occurs at the Fe atom
β​-98Tyr Facilitates electron transfer between the Fe
protein and the MoFe protein18.
in the FeMo-co30–35. Similarly, the industrial production of ammo-
nia utilizes an Fe-based catalyst. In the field of Fe-based complexes,
research has focused on nitrogenase enzyme mimics36–39, HBP mim-
Homocitrate, represented in (Fig. 1b), is a bidentate ligand that ics39–41 and the environmental influence of the ligand structure42–44.
binds to the molybdenum atom21. The homocitrate located next to In a single Mo-atom system, binding of N2 occurs through the
the FeMo-co also has an effect on the catalytic activity of the MoFe reduction of Moiv to Moiii (ref. 45), the reduction of Moi to Mo0 (ref. 26),
nitrogenase. Studies have shown that the lysine residue at the α​-426 or Moii or Moiv to Moi (refs 27,28). In Fe, Ni and Co systems, binding
position in the MoFe protein may hydrogen bond with the homoci- of N2 is apt to occur through the reduction of the metal from the
trate located with the FeMo-co22. It is believed that the hydrogen Mii to the Mi oxidation state, where further reduction to the M0 oxi-
bonding between the α​-146Lys and the homocitrate orients the dation state will yield a different binding alignment. Experiments
homocitrate in the direction that best allows N2 to bind with the performed on synthesized complexes with intermediate species
FeMo-co, thus optimizing the catalysis of nitrogen to ammonia. already attached showed that a single Fe atom stabilizes significant
intermediates44. However, DFT and experimental studies show that
Nitrogenase enzymes in electrochemistry. Recent work has stud- additional metal centres increase N—N activation through bond
ied the MoFe protein in electrochemical applications, with multiple lengthening41,46,47. More specifically, low-coordinate complexes (that
substrates and reducing agents. In one key study, the MoFe protein is, electronically unsaturated)39 containing three Fe atoms demon-
was immobilized on a glassy carbon electrode surface and demon- strated cooperation between the centres to make N–N cleavage
strated electrochemical activity23. The cobaltocene/cobaltocenium thermodynamically feasible46,47.
redox couple was used to mediate electrons between the electrode Different mechanisms have been presented over time, includ-
surface and the MoFe protein where it was found that under an ing the Chatt type (including either the alternating (or symmetric)
applied voltage, the immobilized nitrogenase could reduce azide or reaction pathways or the distal (or asymmetric) pathway; Fig. 3)36,48.
nitrite to ammonia. While this system was unable to reduce dinitro- The symmetric pathway has been favoured as the mechanism for
gen into ammonia, the results demonstrate that mediated electron the nitrogenase enzyme21,31, as well as catalysts producing hydra-
transport to MoFe nitrogenase immobilized on an electrode sur- zine. However, recent findings27,49 demonstrate the production of
face is possible and represents a promising step towards being able N3– as the N–N bond is cleaved and reduced to NH3 in a multi-iron
to produce ammonia from nitrogen in electrochemical systems. complex and that iron nitrides have been observed for single iron
Further, this method provided a novel approach to study substrate complexes50, resulting in growing support for the asymmetric path-
turnover by the MoFe protein free of the rate-limiting steps associ- way. Therefore, much of the recent research in the field of molecular
ated with Fe protein-coupled catalysis. catalysts is focused towards designing molecular catalysts for the
Nitrogenase has also been incorporated in an enzymatic fuel competing symmetric pathway or asymmetric pathway.
cell. Using a proton exchange membrane as a separator, a NH3-
producing nitrogenase cathodic compartment was coupled with Nitrogenase mimics. Although there have been a variety of bioin-
a hydrogenase-based anodic compartment, where hydrogen was spired molecular catalysts for nitrogen reduction, there has also been
employed as the terminal electron donor24. In this configuration, a wealth of research attempting to mimic the cofactor and ligand
the enzyme was not immobilized at the electrode surface and the Fe binding of the natural nitrogenases. Recently, it has been demon-
protein of nitrogenase was also included. When using methylviolo- strated by Peters and co-workers that a single Fe atom in a molecular
gen as an electron mediator between the cathode and nitrogenase complex is capable of binding44 and reducing51 dinitrogen to pro-
(via the Fe protein), it was found that a current could be generated duce ammonia. These [(EPR3)Fe–N2]− (R =​Ph and iPr) complexes
while small quantities of ammonia were produced; the hydrolysis of are tetradentate ligands containing E =​Si (ref. 44) B or C (ref. 51),
ATP was required for nitrogenase turnover. and three phosphine ligands. P- or N-based ligands are electron
rich, allowing the coordination of dinitrogen and increasing π​ back-
Molecular catalysts bonding, which weakens the triple N2 bond40,44,51,52. These ligands
Synthetic molecular complexes provide an arguably less complex are continuing to be advanced, including recent work on carbazole-
model to study the mechanism of N2 fixation, compared with bio- based PNP-type pincer ligands53.
logical or heterogeneous surface systems. Molybdenum was long Phosphorus-based ligand systems are generally a popular elec-
thought to be the essential transition metal for nitrogen fixation tron donor system used to stabilize various metals for ammonia
by the nitrogenase enzyme. Using a single Mo-based molecular synthesis. PNP ligands have been used to stabilize Co in various

Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal 493


Review Article NaTuRE CaTalysIs

Table 3 | Summary of key molecular catalyst complexes


Metal core Complex Temperature (°C) Pressure (atm) NH3 equivalent
Single Fe (TPB)Fe(N2[Na(12-crown-4)2] −​78 1 59.0–88.1 (refs 98,99)
(CAAC)2Fe −​95 1 3.3±​1.1 (ref. 52)
(CAAC)2Fe[BArF4] −​95 1 3.4±​1.0 (ref. 52)
C-atom anchor −​78 1 36 (ref. 51)
Boron-substituted complexes −​78 1 3.8 (ref. 98)
Anionic boron-substituted complexes −​78 1 84 (ref. 100)
PNP −​78 1 14.3–22.7 (refs 101,102)
PPP −​78 1 66.7 (ref. 99)
[Fe(N2)(depe)2] −​78 1 0.95 (ref. 103)
Single Co PNP −​78 1 15.9 (ref. 54)
[(TPB)Co(N2)][Na(12-crown-4)2] −​78 1 2.4 (ref. 55)
Single Ru/Os Tris(phosphine)silyl ligands −​78 1 4.3 Ru/120 Os (ref. 63)
Single Mo [HIPTN3N]Mo(N2) 25 1 7.56±​0.11 (ref. 25)
PPP 25 1 63 (ref. 56)
Pyridine-based diamide 22 1 10.8 (ref. 28)
PCP 25 1 115 (ref. 29)
Two Mo PNP 25 1 26 (ref. 104)
Metallocene-substituted PNP 25 1 22 (ref. 105)
PNP 25 1 415 (ref. 27)
Ammonia production amounts are normalized to one metal used for the core reactive centre. TPB, tris(phosphino)borane; depe, Et2PCH2CH2PEt2; HIPT, hexa-iso-propyl-terphenyl; BArF4, tetrakis[3,5-
bis(trifluoromethyl)phenyl]borate.

NH3 NH3
*N2 *NNH *NHNH *NHNH2 *NH2NH2 *NH3
H H

N N

H H H H H H H H H H
N N

N N N N N H

Alternating
N N N H N H H N H H N H

*N
+
NH3 NH3
*N2 *NNH *NNH2 *NH *NH2 *NH3
H H

N N
H H H H H H H
N N

N N N H H H H

Distal
N N N N N N H N H

Fig. 3 | Associative nitrogen reduction pathways. Chatt-type pathway for nitrogen fixation for symmetric or alternating addition of hydrogen versus
asymmetric or distal addition of hydrogen48,106.

oxidation states, producing 4.2 ±​1 NH3 equivalents54. Large equivalents, indicating PNP is important to the production of NH3
amounts of both reducing agent and proton source increased the (ref. 54). It was found that shorter Co–N2 distances resulted in the
NH3 production to 15.9 ±​0.2 NH3 equivalents, along with 1.0 ±​ 0.4 best catalyst, and the ligand hardly affected N≡​N bond stretching54.
NH2NH2 intermediate54. Using a boron-containing pincer (PBP)- A tris(phosphine)boron complex resulted in a 2.4 equivalents of
type ligand with similar molecular structure only catalysed 0.4 NH3 ammonia per Co (ref. 55). A triphosphine-based ligand system was

494 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal


NaTuRE CaTalysIs Review Article
shown to produce up to 63 equivalents of NH3/Mo over a 20 hour breaking to occur as long as there are three iron atoms in the com-
experiment56. This result is the highest reported conversion using a plex42. The size of the alkali metal can influence the shape of trime-
transition metal catalyst and is believed to occur because of the ease tallic clusters, but the primary influence of the alkali metal on the
of reduction for these complexes56. overall complex is to provide stability to the supporting ligands65. It
Ligand chemistry has been shown to change the coordination, was shown that oxygen bridges between Fe and potassium, which
charge and spin state of an atom43. Seeking to mimic the nitro- have been observed on the HBP catalyst, hold these components in
genase structure, sulfur and carbon donors have been incorpo- place to provide N2 binding locations47.
rated into the ligand inner sphere, as they are similarly found in Computational studies suggest alkali metals such as potassium
the FeMo-co structure. Recently, the first Fe complex containing enable dinitrogen reduction through arrangement of the Fe atoms47,
FeMo-co analogue ligands (sulfur and carbon) was synthesized and stabilization of the N3– species46 and/or increases in the π back-
demonstrated to bind N2 (ref. 57). The studies undertaken with this bonding of the Fe39,65. Thermodynamically, the positively charged
complex demonstrated that, on reduction, a Fe–S bond is broken alkali metal helps stabilize highly reduced complexes, pulls electron
and a Fe–N2 complex with a weakened triple N2 bond is formed38,57. density away from the N2 and produces stable reduced products64.
Electropositive aryl compounds were also developed in a ligand sys- Kinetically, the alkali metal cation π bonding arranges multiple Fe
tem to offset negative charges formed on a carbon atom to produce atoms close together and with sufficient flexibility to allow multi-
a stable compound51. While the complex was shown to be catalyti- step and multi-electron reactions64. It has been suggested that the
cally active, over time, the reaction produced a significant amount environment provided by the alkali metal cations is analogous to
of catalytically inactive complexes, decreasing the overall viability of that provided by the positively charged histidine and arginine resi-
the complex51. It is believed that the flexibility of the Fe–C interac- dues found near the Fe–S face in the nitrogenase64.
tion greatly influences N2 binding and reduction at a single Fe site51.
A cyclic alkyl(amino) carbene (CAAC)-type ligand was shown to Heterogeneous catalysts
produce ammonia in small amounts through an end-on binding of Over the years, the HBP has been improved and optimized, but the
N2 at a flexible two/three-coordinate system at low temperatures52. industrial catalyst used today is surprisingly similar to the original
However, attempts to utilize this complex at room temperature catalyst developed in the early to mid-1900s66. Studies on single-
proved ineffective producing only 0.4 ±​0.2 NH3 equivalents52. crystal iron catalysts have shown that ammonia synthesis is a struc-
In a similar manner, hydrides have been incorporated into turally sensitive reaction with the Fe(111) crystal face proving to
the ligand sphere to mimic the bridging hydrides of FeMo-co33. be the most reactive67. The reactivity is theorized to result from the
Bridging hydrides are strongly implicated as the key N2 binding spe- highly coordinated Fe sites (C7 within the Fe(111)), which experi-
cies, where low or intermediate spin Fe ions are formed. Hydrides ence the largest electronic fluctuations68. The presence of promoter
have been demonstrated to act as a base or nucleophile in a di-iron species has been shown to greatly impact the catalyst structure and
sulfide complex, where reduction of the hydride led to N2 binding37. performance in the high-temperature, high-pressure environment
Late transition metal hydride complexes display weakly activated of the HBP67,69. Promoter oxides (for example, Al2O3 and K2O) in
dinitrogen binding, whereas early transition metal hydride com- operational industrial HBP catalysts are not fully reduced67,69. The
plexes result in complete reduction of the triple bond of N2 on coor- addition of non-reducible oxides prevents the sintering of the iron,
dination58. This trend shows the importance of the central metal for increases the reaction rate67, and increases the Brunauer–Emmett–
initial reduction, while later reactions are shown to be sensitive to Teller surface area69.
steric and electronic characteristics provided by ligands58. Non-ferrous catalysts did not receive notoriety until Ozaki et al.
In a different approach, molecular complexes that differentiate proposed that the energetics of chemical adsorption and desorption
from the FeMo-co have been developed. The ease of conversion of of nitrogen species on the surface could be associated, leading to
tris(trimethylsilyl)amine (N(SiMe3)3) into NH3 on hydrolysis pres- so-called volcano plots70. Iron, osmium and ruthenium were at the
ents an alternative approach to nitrogen fixation59. Nishibayashi optimum energy, or peak, of these volcano graphs, which shifted
and co-workers produced NH3 based on tris(trimethylsilyl)amine catalysis research toward ruthenium-based catalysts70. Ruthenium
(N(SiMe3)3) through the use of an iron pentacarbonyl (Fe(CO)5) catalysts promoted with Al2O3 (ref. 71), MgO (refs 71–73), MgAl2O4
complex under ambient conditions59. This concept has also been (ref. 74) and a mixture of barium with cesium75 proved to be effective
applied to vanadium, chromium, molybdenum and cobalt com- for ammonia synthesis, leading to commercialization in the Kellog
plexes. Brown and Root (KBR) advanced ammonia process (KAAP)76 . The
Recently, functionalization of dinitrogen to afford nitrogen-con- KAAP was implemented in Canada in 1992 with a Ba/Cs-promoted
taining compounds has been surveyed for almost all of the transi- Ru catalyst supported on graphitized carbon, which is ten times
tion metal complexes, including hafnium and rhenium60–62, as well more active than the HBP iron catalyst at low pressure76. As of
as ruthenium and osmium complexes63. Although these are not con- 2010, only 16 ammonia plants used Ru catalysts77. In addition to the
sidered to be bioinspired, they have started inspiring new classes of higher cost of Ru over Fe, the catalyst suffers from loss of support
molecular catalysts for nitrogen reduction. The future of molecular and shorter catalyst life77,78. Therefore, research for Ru catalysts and
catalysts will focus on improved performance, but also immobiliza- potential catalyst supports, such as boron nitride79, non-thermal
tion of the molecular catalyst on the electrode surface. There are a plasma80, BaCeO3 nanocrystals81, lanthanide oxides (MgO (refs 82,83),
variety of strategies being investigated, including encapsulation in CeO2 (refs 82,83), Sm2O3 (ref. 82)), graphitic nanofilaments84 and zeo-
polymeric coatings (that is, Nafion), covalent binding to electrode lites85 continues. In recent work, Hara et al.86 demonstrated that a
surfaces, and non-covalent immobilization with pyrene tethers that calcium aluminum oxide-based electride ([Ca24Al28O64]4+ (e−)4)
π​–π​stack to carbon electrodes. acts as an electron-donating support for an Ru catalyst, resulting in
ammonia production rates as high as 2,120 μ​mol g–1 h–1 (gas inlet
Heterogeneous catalyst mimics. In general, alkali metals have of H2/N2 (3:1), 1 atm, 673 K), compared with traditional alumina-
been used as promoters in the iron-catalysed HBP, but a mecha- or calcium oxide-supported Ru catalysts with ammonia production
nistic understanding of these promoters has evolved from work on rates of 50–160 μ​mol g–1 h–1.
molecular catalysts. Holland and co-workers developed an Fe potas-
sium complex capable of cleaving N2 and producing NH3 through Sabatier principle limits heterogeneous catalysis of N2 reduction.
the formation of two nitrides (N3–)40. Alkali metals are shown to The N2 reduction reaction, whether the reductant is H2 gas or pro-
provide easier reduction for the overall complex64, allowing N–N ton/electron pairs, must break the N≡​N triple bond and form three

Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal 495


Review Article NaTuRE CaTalysIs

log10[TOF (s–1)]
a b –20 –16 –12 –8 – 4 0 4 12
c
8 0 0
0 3 Rh
‘CoMo’ *N Mo Rh

N2 transition state energy (EN-N) (eV)


Ru W Ru
Fe Ir
Fe Rh Ru

U versus SHE (V)


Mo 2 Zr Re Co
–5 Co –1 Ni –1
Rh
Ni Ta Nb Cr
log10[TOF (s–1)]

Ti
Se Y Zr V Fe

–ΔG (eV)
Re Rh Ru
Re
1 Pd
–10 Y Sc –ΔG, *NH → *NH2 (flat)
Ti
–ΔG, N2(g) → *NH (flat) Pd
–2 *H –ΔG, N2(g) → *2N (flat) Pt Pt –2
0 Dissociative (flat) (step) Pt
Pt
–15 Associative (flat) (step)
–ΔG, *NH2 → NH3(g) (step)
–ΔG, N2(g) → *N2H (step) Pd
Pd –1 Pd
–ΔG, N2(g) → *2N (step)
–20 –3 –3
–2.0 –1.5 –1.0 – 0.5 0.0 – 0.5 –1.0 –2.0 –1.5 –1.0 –0.5 0.0 – 0.5 –1.0 –3 –2 –1 0 1
Nitrogen adsorption energy (EN) (eV) Nitrogen adsorption energy (EN) (eV) ΔEN* (eV)

Fig. 4 | Computational predictions and theory-based limitations for heterogenous (electro)catalysts. a, Volcano plot for ammonia synthesis on late
transition metals107. b, Scaling relationship for N2 dissociation transition state intermediate on late transition state metals107. TOF, turnover frequency.
c, Proposed volcano plot for electrochemical nitrogen reduction on late transition metals87. Panels reproduced from: a,b, ref. 107, Elsevier; c, ref. 87, RSC.

N–H bonds for ammonia formation. On a heterogeneous catalyst, dissociative paths87. As the hydrogen evolution reaction will have
N2 will initially bind to the surface and may then dissociate to two lower overpotentials, N2 electrochemical reduction will also suffer
adsorbed N atoms, or first partially reduce by forming an N2–H from a selectivity challenge. This result is consistent with the experi-
bond and subsequently dissociating into two molecules. According mental observation that no late transition metal is highly active or
to the Sabatier principle, the optimal catalyst will have an interme- selective to N2 electrochemical reduction at low temperature.
diate binding energy of a key reaction intermediate, which in the The above analysis provides guidance in considering the design
case of N2 reduction would be either NHx* or N2Hx* species (where of electrocatalysts for electrochemical N2 reduction. The corre-
x =​0–2 and * denotes a surface-bound species). lations used to establish the plots in Fig. 4 do have noise, and the
DFT calculations and microkinetic analysis have helped to illus- existence of BEP relationships for elementary electrochemical reac-
trate how this trade-off results in Fe and Ru catalysts being the tions across late transition metals is yet to be established (Fig. 4c
optimal materials for the HBP. For dissociative N2 reduction, it was considered only elementary reaction energies with an assumption
shown that the binding energies of NHx (x =​0–2) species corre- that BEP relationships would hold). Though the noise in scaling
late with each other on late transition metals87. This correlation thus or BEP relationships provides some hope for late transition metal
relates the binding energy of all reaction intermediates on various catalysis, other materials or mechanisms may be needed to provide
surfaces using a single descriptor — the binding energy of N* to significant N2 electroreduction. Catalytic systems/materials that
the surface. Bronsted–Evans–Polanyi (BEP) relationships have been either lower activation barriers relative to the late transition metal
shown to hold for elementary N–N dissociation and N–H bond for- BEP relationships, or break scaling or BEP relationships altogether,
mation steps, linearly correlating the elementary activation barri- are needed.
ers with the reaction energies88. This leaves the metal–N* binding Though these limitations lead to difficulty in discovering active
energy as a single descriptor that dictates all elementary reaction and selective N2 reduction catalysts, the nitrogenase enzyme sys-
energies and activation barriers, and therefore, the overall rate of tem demonstrates significant turnover rates, for what is effectively
the ammonia synthesis reaction on late transition metal surfaces. an electrochemical reduction process at atmospheric conditions.
The consequences of the Sabatier principle on the rate of N2 Homogeneous catalysts demonstrate the possibility to reduce N2
reduction results in a ‘volcano’ curve when activity is plotted against through different elementary reaction mechanisms, and both
N* binding energy (Fig. 4a). Metals to the left (strong N* binding) enzymes and homogeneous systems demonstrate the potential to
result in lower ammonia synthesis rates due to slow N–H forma- use complex active sites to further impact catalyst performance.
tion and metals to the right (weak N* binding) are limited by N2
activation. Figure 4b shows the BEP linear correlation between the Heterogeneous electrocatalysts. Electrochemical systems pro-
N2 dissociation transition state stability and the N* binding energy. vide an easily scalable system, alleviate dependence on fossil fuels
No metal exists with a combination of intermediate N* binding and decrease energy consumption by utilizing renewable electric
and a low energy N2 dissociation transition state (bottom middle of energy, thereby increasing accessibility. However, existing cata-
plot) As the optimal material at the top of the ‘volcano’ still displays lysts are plagued by low Faradaic efficiencies due to the competing
a significant activation barrier for both N2 dissociation and N–H hydrogen evolution reaction (HER). Components such as the cat-
bond formation, the Haber–Bosch reaction must be performed at alyst material and electrolyte type (solid polymer, liquid or mol-
elevated temperature (400–500 °C) to reach an acceptable rate. As ten salt)4,89–93, have been modified to increase the electrocatalytic
higher temperature limits equilibrium conversion for the overall performance. The wide variety of different experimental param-
reaction, high pressure must be used to reach reasonable conver- eters tested thus far lead to challenges in attempting to directly
sions, as described earlier. compare results. However, it is clear that the primary challenge
Though N2 electrochemical reduction may first form a N2H* that must be solved to enable electrocatalytic ammonia synthesis
species before breaking the N–N bond, Nørskov et al. suggested is the design of the electrocatalyst itself to suppress HER while
that similar scaling and BEP relationships will lead to an equivalent supporting an optimized nitrogen reduction reaction (NRR) to
‘volcano’ relationship that would limit the ability of late transition ammonia. Further, experimental4,93–95 and theoretical1,87 results on
metals to reduce N2 electrochemically (Fig. 4c). Analysis of their traditional transition metal catalysts and catalysts designed with
DFT data suggests that all late transition metals will have a signifi- well-known strategies (for example, monometallic or bimetal-
cant overpotential for N2 reduction regardless of the associative or lic composition, well-known transition metal catalyst materials

496 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal


NaTuRE CaTalysIs Review Article
such as platinum or iron, morphological control of size or shape) Outlook
suggest that these traditional strategies, alone, will not achieve The extensive research performed on delineating nitrogenase struc-
high Faradaic efficiencies for the NRR. Alternative approaches ture and developing molecular catalysts has resulted in an increas-
must thus be explored for electrocatalyst design that strategically ingly comprehensive view of the necessary structural environment
combine optimization of dinitrogen adsorption and proton addi- for efficient N2 reduction to NH3. In particular, nitrogenase research
tion thermodynamics with successful suppression of HER at the has used amino acid substitutions to control the enzyme catalytic
electrocatalyst surface. The successful experimental demonstra- environment, allowing the study of substrate and intermediates
tion of potential strategies necessitates continued development availability and ammonia production. Homogeneous catalysis
of the theoretical understanding of the NRR reaction mechanism research has developed successful ligand systems for N2 reduction,
thermodynamics and kinetics, as well as detailed and careful as well as provided catalyst molecular structures that demonstrate
experimental work to identify, evaluate and understand synthetic necessary components for the N2 reduction mechanism. Simple
heterogeneous electrocatalysts. metal-based heterogeneous catalyst surfaces are highly unlikely
For the majority of the electrochemical systems, despite cata- to promote N2 reduction to NH3 at high Faradaic efficiency, and
lyst type, Faradaic efficiencies are still typically below 1%. One mechanistic insights from the nitrogenase and molecular catalyst
exception, a nano-Fe2O3 catalyst dispersed in a molten hydrox- research communities must be used to innovatively design success-
ide electrolyte solution, has far exceeded other catalysts with a ful heterogeneous catalysts.
Faradaic efficiency of 35% (2.4 ×​ 10–9 mol NH3 cm–2 s–1 at 200 °C DFT reports coupled with in situ electrochemical observation of
and atmospheric pressure)4. Interestingly, this study found that adsorption and desorption of nitrogen species will lead to a better
changing the electrolyte molar ratio and composition did not understanding of surface interactions to reduce competing reactions
affect the ammonia evolution potential but did contribute to on heterogeneous catalysts. DFT studies point researchers in an
changes in production rate results4. It is likely that the combi- educated direction for selective catalyst design, but do not address
nation of higher temperature with a controlled amount of water the overall design of an electrochemical nitrogen fixation system.
molecules within the electrocatalytic system both contributed to There is currently a lack of polymers to promote the transport of
the high reported Faradaic efficiencies. However, the complexity OH– ions for a stable membrane separator, and proper experimental
of system operations and restriction to higher temperatures for controls. Ammonia contamination in current research setups leads
a molten hydroxide electrolyte decrease the economic feasibility. to the uncertainty of results and affects reproducibility of results.
In contrast, polymer electrolyte membranes such as Nafion have Ultimately, the problems facing the electrochemical ammonia syn-
promising prospects to be utilized in low-temperature ammonia thesis industry require efforts in both theory and experimental
synthesis due to the high conductivity of the polymer at low tem- work from a broad array of researchers to achieve a low-tempera-
perature ranges. Recently, several advances have suggested poten- ture, cost-effective and efficient system.
tial strategies in catalyst design that may overcome the challenge With future efforts identifying novel NH3-evolving catalysts or
of low Faradaic efficiency. An Au–TiO2 nanocomposite catalyst novel approaches to further study known (bio)catalysts, the impor-
resulted in Faradaic efficiencies as high as 8.11% (3.5 ×​ 10–10 mol tance of product analysis must not be overlooked. This is important
NH3 cm–2 s–1 at 20 °C and atmospheric pressure, inlet gas as wet for determining substrate/product selectivity as well as confirm-
N2)91, suggesting a bimetallic catalyst comprised of elements from ing the production of NH3. While NH3 can be easily quantified
either side of the ‘volcano’ plot (Fig. 4c)87 may provide a more opti- by ultraviolet-visible or fluorescence spectroscopy, such probes
mized surface. However, an increase in current density (that is, are susceptible to false-positive interference from other N species
increase in cell potential) leads to increased ammonia production (such as amino acids or creatine, a product of ATP regeneration
along with a decrease in Faradaic efficiency91, which is a general in nitrogenase enzymatic assays) and an accompanying technique
trend seen with other electrochemical systems94,95. The decrease should also be employed, such as NMR spectroscopy of 1H or 14N;
in Faradaic efficiency occurs because the competing electro- the use of NMR also introduces the ability to easily confirm NH3
chemical reaction, hydrogen evolution, dominates at the catalyst production from 15N-labelled nitrogen gas, which is important in
surface, leaving no vacant sites for nitrogen reduction to occur. eliminating contaminant sources of NH3. However, please note
These results suggest that while the Ti–Au bimetallic combina- that sources of labelled nitrogen gas frequently have a contami-
tion seems to enhance the NRR, the catalyst does not sufficiently nation of labelled ammonia, which needs to be considered when
suppress HER as the current density increases. Further research designing the analytical method. Overall, there has been a wealth of
is needed to understand if these, and other, catalyst design strate- materials and mechanistic research in nitrogen reduction catalysts
gies, such as the electrode-supported catalyst materials of Hara for ammonia production over the past five to seven years that is
et al.86, will prove beneficial in electrochemical systems. Inherent showing the ability to translate nitrogenase and molecular catalyst
in the need for further research on electrocatalyst design is the knowledge to heterogeneous catalysts is dramatically improving
need for fundamental experimental studies to understand the catalyst performance.
NRR reaction mechanism and reaction dynamics at the catalyst
surface, as well as to understand how material properties of cata- Received: 8 October 2017; Accepted: 9 May 2018;
lysts directly impact reaction dynamics. The recent results from Published online: 12 July 2018
novel catalyst design suggest that, even though optimization of
the overall electrochemical system is important, without a better References
understanding of how to control the adsorption and desorption of 1. Montoya, J. H., Tsai, C., Vojvodic, A. & Nørskov, J. K. The challenge of
species on the catalyst surface, the Faradaic efficiencies of hetero- electrochemical ammonia synthesis: a new perspective on the role of
nitrogen scaling relations. ChemSusChem 8, 2180–2186 (2015). Density
geneous electrocatalyst will not improve. Further, due to ambient functional theory results for close-packed and stepped metal surfaces
ammonia contamination, any evaluation of NH3 electrocatalysts suggest that linear scaling between N-surface binding and other N2Hxor
should report, at a minimum, a set of control experiments with NHx intermediates limits activity and selectivity for electrochemical
argon to have an accurate measurement of Faradaic efficiency. ammonia synthesis, and successful catalyst design must break or
Ideally, multiple approaches to experimentally verify ammonia circumvent these linear scaling relationships.
2. Strait, R. & Nagveka, M. Carbon dioxide capture and storage in the
production should be used, including ion chromatography, 15N nitrogen and syngas industries. Nitrogen Syngas 303, 16–19 (2010).
isotope studies and the common indophenol blue spectropho- 3. Lipman, T. & Shah, N. Ammonia as an alternative energy storage medium
tometry method. for hydrogen fuel cells: scientific and technical review for near-term

Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal 497


Review Article NaTuRE CaTalysIs
stationary power demonstration project, final report. Univ. California, 29. Eizawa, A. et al. Remarkable catalytic activity of dinitrogen-bridged
Research Report UCB-ITS-RR-2007-5 (eScholarship Repository, Berkeley, dimolybdenum complexes bearing NHC-based PCP-pincer ligands toward
2007); http://www.escholarship.org/uc/item/7z69v4wp nitrogen fixation. Nat. Commun. 8, 12 (2017).
4. Licht, S. et al. Ammonia synthesis by N2 and steam electrolysis in molten 30. Barney, B. M., Igarashi, R. Y., Dos Santos, P. C., Dean, D. R. & Seefeldt, L.
hydroxide suspensions of nanoscale Fe2O3. Science 345, 637–640 (2014). C. Substrate interaction at an iron–sulfur face of the FeMo-cofactor during
Nanoscale iron oxide, combined with steam and air in a molten nitrogenase catalysis. J. Biol. Chem. 279, 53621–53624 (2004).
hydroxide electrolyte, enable Faradaic efficiencies of up to 35% for 31. Seefeldt, L. C., Hoffman, B. M. & Dean, D. R. Mechanism of Mo-dependent
electrochemical nitrogen reduction to ammonia. nitrogenase. Annu. Rev. Biochem. 78, 701–722 (2009).
5. Zhou, F. et al. Electro-synthesis of ammonia from nitrogen at ambient 32. Kim, C. H., Newton, W. E. & Dean, D. R. Role of the MoFe protein α​
temperature and pressure in ionic liquids. Energy Environ. Sci. 10, -subunit histidine-195 residue in FeMo-cofactor binding and nitrogenase
2516–2520 (2017). catalysis. Biochemistry 34, 2798–2808 (1995).
6. Seefeldt, L. C., Hoffman, B. M. & Dean, D. R. Electron transfer in 33. Igarashi, R. Y. et al. Trapping H– bound to the nitrogenase FeMo-cofactor
nitrogenase catalysis. Curr. Opin. Chem. Biol. 16, 19–25 (2012). active site during H2 evolution: characterization by ENDOR spectroscopy.
7. Chan, J. M., Christiansen, J., Dean, D. R. & Seefeldt, L. C. Spectroscopic J. Am. Chem. Soc. 127, 6231–6241 (2005).
evidence for changes in the redox state of the nitrogenase P-cluster during 34. Dos Santos, P. C., Mayer, S. M., Barney, B. M., Seefeldt, L. C. & Dean, D. R.
turnover. Biochemistry 38, 5779–5785 (1999). Alkyne substrate interaction within the nitrogenase MoFe protein. J. Inorg.
8. Danyal, K., Dean, D. R., Hoffman, B. M. & Seefeldt, L. C. Electron transfer Biochem. 101, 1642–1648 (2007).
within nitrogenase: evidence for a deficit-spending mechanism. 35. Scott, D. J., May, H. D., Newton, W. E., Brigle, K. E. & Dean, D. R. Role for
Biochemistry 50, 9255–9263 (2011). the nitrogenase MoFe protein α​-subunit in FeMo-cofactor binding and
9. Thorneley, R. N. F. & Lowe, D. J. Nitrogenase of Klebsiella pneumoniae. catalysis. Nature 343, 188–190 (1990).
Kinetics of the dissociation of oxidized iron protein from molybdenum-iron 36. Hendrich, M. P. et al. On the feasibility of N2 fixation via a single-site Fei/
protein: identification of the rate-limiting step for substrate reduction. Feiv cycle: spectroscopic studies of Fei (N2) Fei, Feiv ≡​N, and related species.
Biochem. J. 215, 393–403 (1983). Proc. Natl. Acad. Sci. USA 103, 17107–17112 (2006).
10. Yang, Z.-Y. et al. Evidence that the Pi release event is the rate-limiting step 37. Arnet, N. A. et al. Synthesis, characterization, and nitrogenase-relevant
in the nitrogenase catalytic cycle. Biochemistry 55, 3625–3635 (2016).
reactions of an iron sulfide complex with a bridging hydride. J. Am. Chem.
11. Spatzal, T. et al. Evidence for interstitial carbon in nitrogenase FeMo
Soc. 137, 13220–13223 (2015).
cofactor. Science 334, 940–940 (2011).
38. Coric, I. & Holland, P. L. Insight into the iron-molybdenum cofactor of
12. Doan, P. E. et al. 57Fe ENDOR spectroscopy and ‘electron inventory’
nitrogenase from synthetic iron complexes with sulfur, carbon, and hydride
analysis of the nitrogenase E4 intermediate suggest the metal-ion core of
ligands. J. Am. Chem. Soc. 138, 7200–7211 (2016).
FeMo-cofactor cycles through only one redox couple. J. Am. Chem. Soc.
39. Smith, J. M. et al. Studies of low-coordinate iron dinitrogen complexes.
133, 17329–17340 (2011).
J. Am. Chem. Soc. 128, 756–769 (2006).
13. Bjornsson, R. et al. Identification of a spin-coupled Mo(III) in the
40. Rodriguez, M. M., Bill, E., Brennessel, W. W. & Holland, P. L. N2 reduction
nitrogenase iron–molybdenum cofactor. Chem. Sci. 5, 3096–3103 (2014).
and hydrogenation to ammonia by a molecular iron-potassium complex.
14. Bjornsson, R., Neese, F., Schrock, R. R., Einsle, O. & DeBeer, S. The
discovery of Mo(III) in FeMoco: reuniting enzyme and model chemistry. J. Science 334, 780–783 (2011).
Biol. Inorg. Chem. 20, 447–460 (2015). 41. MacLeod, K. C., Vinyard, D. J. & Holland, P. L. A multi-iron system capable
15. Varley, J. B., Wang, Y., Chan, K., Studt, F. & Nørskov, J. K. Mechanistic of rapid N2 formation and N2 cleavage. J. Am. Chem. Soc. 136,
insights into nitrogen fixation by nitrogenase enzymes. Phys. Chem. Chem. 10226–10229 (2014).
Phys. 17, 29541–29547 (2015). 42. Grubel, K., Brennessel, W. W., Mercado, B. Q. & Holland, P. L. Alkali metal
16. Spatzal, T., Perez, K. A., Einsle, O., Howard, J. B. & Rees, D. C. Ligand control over N–N cleavage in iron complexes. J. Am. Chem. Soc. 136,
binding to the FeMo-cofactor: structures of CO-bound and reactivated 16807–16816 (2014).
nitrogenase. Science 345, 1620–1623 (2014). 43. MacLeod, K. C. & Holland, P. L. Recent developments in the homogeneous
17. Rao, L., Xu, X. & Adamo, C. Theoretical investigation on the role of the reduction of dinitrogen by molybdenum and iron. Nat. Chem. 5,
central carbon atom and close protein environment on the nitrogen 559–565 (2013). A review of molecular catalyst advances provides recent
reduction in Mo nitrogenase. ACS Catal. 6, 1567–1577 (2016). developments and recommended next steps, including the need for
18. Danyal, K. et al. Uncoupling nitrogenase: catalytic reduction of hydrazine to redox active ligands, ligand-driven steric control, and structural
ammonia by a MoFe protein in the absence of Fe protein-ATP. J. Am. correlations between molecular catalysts and the nitrogenase enzyme.
Chem. Soc. 132, 13197–13199 (2010). 44. Lee, Y., Mankad, N. P. & Peters, J. C. Triggering N2 uptake via redox-
19. Dance, I. Calculated details of a mechanism for conversion of N2 to NH3 at induced expulsion of coordinated NH3 and N2 silylation at trigonal
the FeMo cluster of nitrogenase. Chem. Commun., 165–166 (1997). bipyramidal iron. Nat. Chem. 2, 558–565 (2010).
20. Dance, I. The controlled relay of multiple protons required at the active site 45. Schrock, R. R. Catalytic reduction of dinitrogen to ammonia by
of nitrogenase. Dalton Trans. 41, 7647–7659 (2012). molybdenum: theory versus experiment. Angew. Chem. Int. Ed. 47,
21. Hoffman, B. M., Lukoyanov, D., Yang, Z.-Y., Dean, D. R. & Seefeldt, L. C. 5512–5522 (2008).
Mechanism of nitrogen fixation by nitrogenase: the next stage. Chem. Rev. 46. Figg, T. M., Holland, P. L. & Cundari, T. R. Cooperativity between
114, 4041–4062 (2014). A thorough review provides recently low-valent iron and potassium promoters in dinitrogen fixation. Inorg.
discovered details about the nitrogenase structures and functions, Chem. 51, 7546–7550 (2012).
including definitive determination of the central carbon atom of the 47. Chiang, K. P., Bellows, S. M., Brennessel, W. W. & Holland, P. L.
FeMo cofactor. Multimetallic cooperativity in activation of dinitrogen at iron–potassium
22. Durrant, M. C., Francis, A., Lowe, D. J., Newton, W. E. & Fisher, K. sites. Chem. Sci. 5, 267–274 (2014).
Evidence for a dynamic role for homocitrate during nitrogen fixation: the 48. Hinrichsen, S., Broda, H., Gradert, C., Söncksen, L. & Tuczek, F. Recent
effect of substitution at the α​-Lys426 position in MoFe-protein of Azotobacter developments in synthetic nitrogen fixation. Inorg. Chem. 108,
vinelandii. Biochem. J. 397, 261–270 (2006). 17–47 (2012).
23. Milton, R. D. et al. Nitrogenase bioelectrocatalysis: heterogeneous ammonia 49. Ritleng, V. et al. Molybdenum triamidoamine complexes that contain
and hydrogen production by MoFe protein. Energy Environ. Sci. 9, hexa-tertbutylterphenyl, hexamethylterphenyl, or p-
2550–2554 (2016). bromohexaisopropylterphenyl substituents. An examination of some catalyst
24. Milton, R. D. et al. Bioelectrochemical Haber–Bosch process: an ammonia- variations for the catalytic reduction of dinitrogen. J. Am. Chem. Soc. 126,
producing H2/N2 fuel cell. Angew. Chem. Int. Ed. 56, 2680–2683 (2017). 6150–6163 (2004).
25. Yandulov, D. V. & Schrock, R. R. Catalytic reduction of dinitrogen to 50. Thompson, N. B., Green, M. T. & Peters, J. C. Nitrogen fixation via a
ammonia at a single molybdenum center. Science 301, 76–78 (2003). terminal Fe(iv) nitride. J. Am. Chem. Soc. 139, 15312–15315 (2017).
26. Tanaka, H. et al. Unique behaviour of dinitrogen-bridged dimolybdenum 51. Creutz, S. E. & Peters, J. C. Catalytic reduction of N2 to NH3 by an Fe–N2
complexes bearing pincer ligand towards catalytic formation of ammonia. complex featuring a C-atom anchor. J. Am. Chem. Soc. 136,
Nat. Commun. 5, 3737 (2014). 1105–1115 (2014).
27. Arashiba, K. et al. Catalytic nitrogen fixation via direct cleavage of 52. Ung, G. & Peters, J. C. Low-temperature N2 binding to two-coordinate L2Fe0
nitrogen–nitrogen triple bond of molecular dinitrogen under ambient enables reductive trapping of L2FeN2– and NH3 generation. Angew. Chem.
reaction conditions. Bull. Chem. Soc. Jpn 90, 1111–1118 (2017). Int. Ed. 54, 532–535 (2015).
28. Wickramasinghe, L. A., Ogawa, T., Schrock, R. R. & Müller, P. Reduction of 53. Higuchi, J. et al. Preparation and reactivity of iron complexes bearing
dinitrogen to ammonia catalyzed by molybdenum diamido complexes. J. anionic carbazole-based PNP-type pincer ligands toward catalytic nitrogen
Am. Chem. Soc. 139, 9132–9135 (2017). fixation. Dalton Trans. 47, 1117–1121 (2018).

498 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal


NaTuRE CaTalysIs Review Article
54. Kuriyama, S. et al. Direct transformation of molecular dinitrogen into 82. Niwa, Y. & Aika, K. I. The effect of lanthanide oxides as a support for
ammonia catalyzed by cobalt dinitrogen complexes bearing anionic PNP ruthenium catalysts in ammonia synthesis. J. Catal. 162,
pincer ligands. Angew. Chem. Int. Ed. 55, 14291–14295 (2016). 138–142 (1996).
55. Del Castillo, T. J., Thompson, N. B., Suess, D. L. M., Ung, G. & Peters, J. C. 83. Saito, M., Itoh, M., Iwamoto, J., Li, C. Y. & Machida, K. I. Synergistic effect
Evaluating molecular cobalt complexes for the conversion of N2 to NH3. of MgO and CeO2 as a support for ruthenium catalysts in ammonia
Inorg. Chem. 54, 9256–9262 (2015). synthesis. Catal. Lett. 106, 107–110 (2006).
56. Arashiba, K. et al. Catalytic reduction of dinitrogen to ammonia by use of 84. Liang, C., Li, Z., Qiu, J. & Li, C. Graphitic nanofilaments as novel support
molybdenum-nitride complexes bearing a tridentate triphosphine as of Ru–Ba catalysts for ammonia synthesis. J. Catal. 211,
catalysts. J. Am. Chem. Soc. 137, 5666–5669 (2015). 278–282 (2002).
57. Coric, I., Mercado, B. Q., Bill, E., Vinyard, D. J. & Holland, P. L. Binding of 85. Fishel, C. T., Davis, R. J. & Garces, J. M. Ammonia synthesis catalyzed by
dinitrogen to an iron–sulfur–carbon site. Nature 526, 96–99 (2015). ruthenium supported on basic zeolites. J. Catal. 163, 148–157 (1996).
58. Ballmann, J., Munha, R. F. & Fryzuk, M. D. The hydride route to the 86. Hara, M., Kitano, M. & Hosono, H. Ru-loaded C12A7:e– electride as a
preparation of dinitrogen complexes. Chem. Commun. 46, catalyst for ammonia synthesis. ACS Catal. 7, 2313–2324 (2017). New
1013–1025 (2010). results on an electride-based catalyst suggest that an electron-donating
59. Yuki, M. et al. Iron-catalysed transformation of molecular dinitrogen into character accelerates N2 cleavage and opens a potential pathway for
silylamine under ambient conditions. Nat. Commun. 3, 1254 (2012). understanding how to better design heterogeneous catalysts for nitrogen
60. MacKay, B. A., Johnson, S. A., Patrick, B. O. & Fryzuk, M. D. fixation.
Functionalization and cleavage of coordinated dinitrogen via hydroboration 87. Skulason, E. et al. A theoretical evaluation of possible transition metal
using primary and secondary boranes. Can. J. Chem. 83, electro-catalysts for N2 reduction. Phys. Chem. Chem. Phys. 14, 1235–1245
315–323 (2005). (2012). One of the first extensive theoretical efforts to evaluate transition
61. Hidai, M. & Mizobe, Y. Research inspired by the chemistry of nitrogenase metals for electrochemical nitrogen reduction predicts optimal metal
— novel metal complexes and their reactivity toward dinitrogen, nitriles, surfaces and demonstrates most metal surfaces will be dominated by *H
and alkynes. Can. J. Chem. 83, 358–374 (2005). rather than *N species on the surface.
62. Ohki, Y. & Fryzuk, M. D. Dinitrogen activation by group 4 metal 88. Wang, S. et al. Universal transition state scaling relations for (de)
complexes. Angew. Chem. Int. Ed. 46, 3180–3183 (2007). hydrogenation over transition metals. Phys. Chem. Chem. Phys. 13,
63. Fajardo, J. & Peters, J. C. Catalytic nitrogen-to-ammonia conversion by 20760–20765 (2011).
osmium and ruthenium complexes. J. Am. Chem. Soc. 139, 89. Köleli, F., Röpke, D., Aydin, R. & Röpke, T. Investigation of N2-fixation on
16105–16108 (2017). polyaniline electrodes in methanol by electrochemical impedance
64. McWilliams, S. F. & Holland, P. L. Dinitrogen binding and cleavage by spectroscopy. J. Appl. Electrochem. 41, 405–413 (2010).
multinuclear iron complexes. Acc. Chem. Res. 48, 2059–2065 (2015). 90. Pappenfus, T. M., Lee, K. M., Thoma, L. M. & Dukart, C. R. Wind to
65. McWilliams, S. F. et al. Alkali metal variation and twisting of the FeNNFe ammonia: electrochemical processes in room temperature ionic liquids.
core in bridging diiron dinitrogen complexes. Inorg. Chem. 55, ECS Trans. 16, 89–93 (2009).
2960–2968 (2016). 91. Shi, M.-M. et al. Au sub-nanoclusters on TiO2 toward highly efficient and
66. Vojvodic, A. et al. Exploring the limits: a low-pressure, low-temperature selective electrocatalyst for N2 conversion to NH3 at ambient conditions.
Haber–Bosch process. Chem. Phys. Lett. 598, 108–112 (2014). A theoretical Adv. Mater. 29, 1606550 (2017). One of the highest Faradaic efficiencies
study that suggests that a low-pressure, low-temperature process is reported to date for ambient temperature and pressure electrochemical
energetically possible for nitrogen reduction to ammonia. nitrogen reduction suggests that this reported catalyst structure and
67. Spencer, M. S. On the rate-determining step and the role of potassium in composition may be promising for greater N2 selectivity.
the catalytic synthesis of ammonia. Catal. Lett. 13, 45–53 (1992). 92. Murakami, T., Nishikiori, T., Nohira, T. & Ito, Y. Electrolytic synthesis of
68. Strongin, D. R., Carrazza, J., Bare, S. R. & Somorjai, G. A. The importance ammonia in molten salts under atmospheric pressure. J. Am. Chem. Soc.
of C7 sites and surface roughness in the ammonia synthesis reaction over 125, 334–335 (2003).
iron. J. Catal. 103, 213–215 (1987). 93. Nash, J. et al. Electrochemical nitrogen reduction reaction on noble metal
69. Ertl, G., Prigge, D., Schloegl, R. & Weiss, M. Surface characterization of catalysts in proton and hydroxide exchange membrane electrolyzers. J.
ammonia synthesis catalysts. J. Catal. 79, 359–377 (1983). Electrochem. Soc. 164, F1712–F1716 (2017).
70. Ozaki, A., Taylor, H. & Boudrart, M. Kinetics and mechanism of the 94. Chen, S. et al. Room-temperature electrocatalytic synthesis of NH3 from
ammonia synthesis. Proc. R. Soc. Lond. Ser. A 258, 47–62 (1960). H2O and N2 in a gas–liquid–solid three-phase reactor. ACS Sustain. Chem.
71. Aika, K. et al. Support and promoter effect of ruthenium catalyst: I. Eng. 5, 7393–7400 (2017). A nanoparticulate iron oxide catalyst on a
Characterization of alkali-promoted ruthenium/alumina catalysts for carbon nanotube support is tested in a gas-phase electrochemical cell,
ammonia synthesis. J. Catal. 92, 296–304 (1985). and the data show that electrolyte and nanoparticle loading can impact
72. Rosowski, F. et al. Ruthenium catalysts for ammonia synthesis at high ammonia production and are important design considerations.
pressures: preparation, characterization, and power-law kinetics. Appl. 95. Kong, J. et al. Electrochemical synthesis of NH3 at low temperature and
Catal. A 151, 443–460 (1997). atmospheric pressure using a γ​-Fe2O3 catalyst. ACS Sustain. Chem. Eng. 5,
73. Bossi, A., Garbassi, F., Petrini, G. & Zanderighi, L. Support effects on the 10986–10995 (2017). Experimental results are presented for an iron oxide
catalytic activity and selectivity of ruthenium in CO and N2 activation. nanoparticle catalyst in both aqueous and gas-phase electrochemical test
J. Chem. Soc. Faraday Trans. 1 78, 1029–1038 (1982). cells, and results of lower Faradaic efficiency and ammonia production
74. Fastrup, B. On the interaction of N2 and H2 with Ru catalyst surfaces. Catal. rate in a membrane electrode assembly demonstrate the need for
Lett. 48, 111–119 (1997). optimized membrane electrode assembly engineering.
75. Szmigiel, D. et al. The kinetics of ammonia synthesis over ruthenium-based 96. Igarashi, R. Y. et al. Localization of a catalytic intermediate bound to the
catalysts: the role of barium and cesium. J. Catal. 205, 205–212 (2002). FeMo-cofactor of nitrogenase. J. Biol. Chem. 279,
76. Brown, D. E., Edmonds, T., Joyner, R. W., McCarroll, J. J. & Tennison, S. R. 34770–34775 (2004).
The genesis and development of the commercial BP doubly promoted 97. Fisher, K., Dilworth, M. J. & Newton, W. E. Differential effects on N2
catalyst for ammonia synthesis. Catal. Lett. 144, 545–552 (2014). The only binding and reduction, HD formation, and azide reduction with α​
other commercialized catalyst for the Haber–Bosch process is based on a -195His-and α​-191Gln-substituted MoFe proteins of Azotobacter vinelandii
highly active ruthenium–alkali metal–carbon catalyst, with an activity nitrogenase. Biochemistry 39, 15570–15577 (2000).
that is 20 times greater than the iron-based commercial Haber–Bosch 98. Del Castillo, T. J., Thompson, N. B. & Peters, J. C. A synthetic single-site Fe
catalyst. nitrogenase: high turnover, freeze-quench 57Fe Mössbauer data, and a
77. Liu, H. Ammonia synthesis catalyst 100 years: practice, enlightenment and hydride resting state. J. Am. Chem. Soc. 138, 5341–5350 (2016).
challenge. Chin. J. Catal. 35, 1619–1640 (2014). 99. Buscagan, T. M., Oyala, P. H. & Peters, J. C. N2‐to‐NH3 conversion by a
78. Hrbek, J. Coadsorption of oxygen and hydrogen on ruthenium (001): triphos–iron catalyst and enhanced turnover under photolysis. Angew.
blocking and electronic effects of preadsorbed oxygen. J. Phys. Chem. 90, Chem. Int. Ed. 56, 6921–6926 (2017).
6217–6222 (1986). 100. Chalkley, M. J., Del Castillo, T. J., Matson, B. D., Roddy, J. P. & Peters, J. C.
79. Jacobsen, C. J. Boron nitride: a novel support for ruthenium-based Catalytic N2-to-NH3 conversion by Fe at lower driving force: a proposed
ammonia synthesis catalysts. J. Catal. 200, 1–3 (2001). role for metallocene-mediated PCET. ACS Cent. Sci. 3,
80. Akay, G. & Zhang, K. Process intensification in ammonia synthesis using 217–223 (2017).
novel coassembled supported micro-porous catalysts promoted by 101. Kuriyama, S. et al. Catalytic transformation of dinitrogen into ammonia
nonthermal plasma. Ind. Eng. Chem. Res. 56, 457–468 (2016). and hydrazine by iron-dinitrogen complexes bearing pincer ligand. Nat.
81. Yang, X. L. et al. Low temperature ruthenium catalyst for ammonia Commun. 7, 12181 (2016).
synthesis supported on BaCeO3 nanocrystals. Catal. Commun. 11, 102. Sekiguchi, Y. et al. Synthesis and reactivity of iron-dinitrogen complexes
867–870 (2010). bearing anionic methyl- and phenyl-substituted pyrrole-based PNP-type

Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal 499


Review Article NaTuRE CaTalysIs
pincer ligands toward catalytic nitrogen fixation. Chem. Commun. 53, Acknowledgements
12040–12043 (2017). S.L.F. and L.F.G. gratefully acknowledge support from the US Department of Agriculture
103. Hill, P. J., Doyle, L. R., Crawford, A. D., Myers, W. K. & Ashley, A. E. Small Business Innovation Research Program, under award # EC-015996-02. S.I.P.B., R.D.,
Selective catalytic reduction of N2 to N2H4 by a simple Fe complex. J. Am. S.M., M.J.J., J.N.R. and L.F.G. gratefully acknowledge support from the US Department of
Chem. Soc. 138, 13521–13524 (2016). Energy, Office of Science, Basic Energy Sciences, Catalysis Science Program, under award
104. Kuriyama, S. et al. Catalytic formation of ammonia from molecular # DE-SC0016529. R.D.M. and S.D.M. thank the Army Research Office Multidisciplinary
dinitrogen by use of dinitrogen-bridged dimolybdenum–dinitrogen University Research Initiative (MURI) (W911NF-14-1-0263) for funding.
complexes bearing PNP-pincer ligands: remarkable effect of substituent at
PNP-pincer ligand. J. Am. Chem. Soc. 136, 9719–9731 (2014).
105. Kuriyama, S. et al. Nitrogen fixation catalyzed by ferrocene-substituted Competing interests
dinitrogen-bridged dimolybdenum–dinitrogen complexes: unique behavior The authors declare no competing interests.
of ferrocene moiety as redox active site. Chem. Sci. 6,
3940–3951 (2015). Additional information
106. van der Ham, C. J. M., Koper, M. T. M. & Hetterscheid, D. G. H. Reprints and permissions information is available at www.nature.com/reprints.
Challenges in reduction of dinitrogen by proton and electron transfer.
Chem. Soc. Rev. 43, 5183–5191 (2014). Correspondence should be addressed to S.D.M. or M.J.J. or J.N.R. or L.F.G.
107. Medford, A. J. et al. From the Sabatier principle to a predictive theory of Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
transition-metal heterogeneous catalysis. J. Catal. 328, 36–42 (2015). published maps and institutional affiliations.

500 Nature Catalysis | VOL 1 | JULY 2018 | 490–500 | www.nature.com/natcatal

Potrebbero piacerti anche