Catalytic ammonia oxidation mediated by high-spin Fe(III) complex: combined experimental and DFT study

Mohammad Usman a, Mohamed A.T. Hussein b, Tarek A. Kandiel ab, Zain H. Yamani ac and M. Nasiruzzaman Shaikh *ad
aInterdisciplinary Research Center for Hydrogen Technologies and Carbon Management (IRC-HTCM), King Fahd University of Petroleum and Minerals (KFUPM), Dhahran-31261, Saudi Arabia. E-mail: mnshaikh@kfupm.edu.sa
bDepartment of Chemistry, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran-31261, Saudi Arabia
cDepartment of Physics, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran-31261, Saudi Arabia
dMaterial Science and Engineering Department (MSE), King Fahd University of Petroleum and Minerals (KFUPM), Dhahran-31261, Saudi Arabia

Received 11th September 2025 , Accepted 4th November 2025

First published on 5th November 2025


Abstract

Ammonia offers high hydrogen density and favorable transport properties, making it an appealing hydrogen carrier; yet conventional cracking methods for hydrogen release are energy-intensive. Molecular iron complexes offer a sustainable route for the homogeneous conversion of NH3 to N2 under mild conditions. Here, we describe a high-spin FeIII-complex bearing a tetradentate N,N,O,O-donor trans-1,2-bis(2-hydroxy-3-methoxyphenyl-methaniminyl)cyclohexane ligand framework that catalyzes ammonia oxidation at room temperature. In combination with a triarylaminium oxidant and 2,4,6-collidine base, the catalyst produces up to 2.20 equivalents of N2 per Fe center. Comprehensive characterization of the FeIII-complex by FTIR, UV-vis, XPS, and X-ray diffraction, with Mössbauer and DFT analysis, confirmed its high-spin state. Moreover, DFT studies revealed that N–N bond formation in ammonia oxidation proceeds through nucleophilic attack followed by sequential proton- and electron-transfer steps. Together, these findings underscore the potential of high-spin FeIII-complexes in ammonia oxidation catalysis and provide crucial mechanistic understanding of N–N bond formation.


Introduction

Global energy demand is rapidly increasing, intensifying concerns about carbon dioxide emissions associated with fossil fuel combustion.1 Addressing this critical issue requires a transition toward renewable and sustainable energy systems.2 Among various strategies, chemical energy storage stands out as highly promising, particularly if the stored energy can be efficiently utilized. In this regard, ammonia represents an attractive clean energy carrier, combining high energy density and carbon-free use with ease of liquefaction and a well-established worldwide infrastructure for production, storage, and distribution.3–6 Despite these benefits, significant technical challenges remain, notably the issue of nitrogen oxide (NOx) emissions when ammonia is combusted directly.7

Nevertheless, the intrinsic chemical energy stored in ammonia, particularly within its N–H bonds, can be efficiently harnessed to produce electricity using gas turbines, analogous to the energy contained in C–H and C–C bonds in conventional hydrocarbon fuels.2–6 This utilization of ammonia as a sustainable fuel aligns closely with natural biological processes, such as anaerobic ammonia oxidation, which produces nitrite (NO2) and nitrate (NO3).8–10 Through ammonia oxidation (AO), energy can be directly extracted as two ammonia molecules are oxidized to dinitrogen, generating six electrons and six protons. This transformation is hindered by the requirement for multiple proton- and electron-transfer steps to overcome the strong N–H bonds and facilitate N–N bond formation. Specifically, overcoming the high dissociation energy of ammonia's initial N–H bond (99.4 kcal mol−1) is critical for advancing AO catalysis.11,12

Historically, ammonia conversion to dinitrogen has primarily involved heterogeneous catalysis approaches, including electrochemical oxidation and thermal decomposition.13–15 In contrast, homogeneous catalytic ammonia oxidation through coordination chemistry has been comparatively less explored, with only limited examples in the literature.15b,16 Early studies in the 1980s by Meyer17a and Collman17b,c provided initial insights into homogeneous electrochemical AO; however, significant recent advancements by researchers such as Smith18 and Nakajima,19 employing ruthenium polypyridyl complexes, have renewed interest in molecular AO catalysis. Further expanding homogeneous AO catalysis, significant contributions from Bullock et al.,16 Chirik et al.,20,21 and others22 have demonstrated catalytic homolytic N–H bond cleavage using organic radicals and metal complexes based on molybdenum,20,21,23–25 manganese,26 and ruthenium.27,28 Nevertheless, it is essential to investigate catalysts made of metals that are plentiful on Earth to improve sustainability and practical feasibility.

Motivated by biological nitrification in comammox bacteria29 and the nitrogen-fixation role of the iron–molybdenum cofactor (FeMoco) in nitrogenase,30,31 Peters and co-workers32 have recently reported iron-based molecular catalysts for AO. The results point to high-valent iron–amido species as precursors to N–N bond formation, emphasizing iron's potential as an accessible catalyst for ammonia oxidation.

Motivated by these developments, our research investigates explicitly iron complex featuring salen ligand, renowned for their robust catalytic performance in various oxidation reactions, including olefin epoxidation and benzylic oxidation.33 Leveraging these established catalytic properties, we hypothesize that iron–salen complexes can efficiently catalyze ammonia oxidation under chemical oxidation conditions. Herein, we present our recent findings demonstrating successful ammonia oxidation catalysis using a high-spin iron complex, marking significant progress toward realizing full potential of ammonia as a sustainable and carbon-free energy carrier. In this study, ammonium triflate (NH4OTf) was employed as a highly soluble and convenient ammonia source for homogeneous oxidation in non-aqueous media. To examine the effect of ammonia source on catalytic performance, ammonium nitrate and gaseous NH3 (1 M MeCN solution) were also tested under comparable conditions, providing consistent results that confirm the intrinsic catalytic activity of the system.

Experimental section

Details of materials, characterization techniques, and DFT methods are provided in the SI. The Schiff base ligand was synthesized according to a reported procedure.34

Synthesis of FeIII-complex, [FeIII(L)Cl] (1)

Schiff base ligand H2L (1.93 g, 5 mmol) was dissolved in 30 mL of ethanol in a round bottom flask. A solution of FeCl3·6H2O (1.24 g, 5 mmol) in 15 mL of ethanol was added to the previous solution and stirred for 1–2 hour at room temperature. The colour of the solution changed from yellow to dark brown colour. The reaction mixture was filtered, and the filtrate was dried under vacuum. The crude product was dissolved in a small amount of dichloromethane and slowly layered with n-hexane. After 1 week in the air at room temperature, complex 1 was obtained as a needle shape dark brown crystalline solid (yield 2.13 g, 87.29%). Anal. calc. for [C22H24N2O4FeCl] (%) C, 56.01; H, 5.13; N, 5.94; found: C, 59.12; H, 5.67; N, 6.27. UV-vis (MeCN) [λmax, nm]: 269, 305, 358, 518. FTIR (cm−1): 480, 549, 1391, 1376, 1473, 1550, 1617, 2935, 2853, 3062.

Catalytic ammonia oxidation reactions

The catalytic oxidation of ammonia was evaluated using an FeIII-complex under oxidative conditions, following the benchmark procedures reported by Nakajima et al.19 All manipulations were carried out under an inert atmosphere. Catalytic reactions were performed in a 50 mL borosilicate glass batch reactor equipped with gas inlet/outlet ports and a liquid injection port (Fig. S7). The gas inlet was connected to a mass flow controller (Bronkhorst) to regulate helium flow, while the outlet was directly connected to a gas chromatograph (GC) for on-line monitoring of gaseous products.

In a typical experiment, tris(4-bromophenyl)aminium hexachloroantimonate, [(p-BrC6H4)3N˙]SbCl6 (735 mg, 0.90 mmol), ammonium triflate (NH4OTf, 501 mg, 3.0 mmol) and anhydrous acetonitrile (15 mL), previously degassed under helium were introduced into the reactor. In the case of gaseous ammonia, 1 M NH3 solution in MeCN was injected after purging. The reactor was purged with helium using the mass flow controller until residual gases were completely removed. The reaction was initiated by injecting a degassed MeCN solution of the FeIII-complex (0.01, 0.05, 0.25, 0.5 mmol), followed by 2,4,6-collidine (0.40 mL, 3.0 mmol), through the liquid injection port using a gas-tight syringe (Fig. S7). The mixture was stirred magnetically at room temperature for 1 h under helium atmosphere. The evolved N2 was continuously monitored and quantified using a Shimadzu Nexis GC-2030 equipped with a BID detector and Micropacked ST column (2.0 m × 1.0 mm ID) with helium as the carrier gas. The N2 yield was calculated by comparing the integrated peak area with that of a calibrated standard under identical conditions.

Results and discussion

Synthesis and structural analysis

The Schiff base ligand H2L, trans-1,2-bis(2-hydroxy-3-methoxyphenyl-methaniminyl)cycrohexane, was synthesized following a previously reported method.34 The ethanolic solution of the free base ligand was slowly mixed with FeCl3·6H2O solution in ethanol over 30 minutes. The resulting mixture was stirred for 1–2 hours at 60 °C, changing the solution color from yellow to a dark brown. The solvent was then evaporated under vacuum, yielding the isolated penta-coordinated FeIII-complex (1).

The IR spectra of the free Schiff base ligand H2L, and FeIII-complex (1) were recorded in the solid state (Fig. S1). The spectra reveal bands at 3011 cm−1, and 3062 cm−1, respectively, corresponding to sp2 C–H stretching vibrations.34 Another set of bands appeared at 2939 cm−1 (H2L), and 2935 cm−1 (1) attributed to sp3 C–H stretching vibrations. The characteristic imine stretching vibration, ν(C[double bond, length as m-dash]N), was observed at 1624 cm−1 in the free ligand (H2L). Upon coordination with FeIII, this band shifted to 1617 cm−1 in FeIII-complex (1), indicating coordination through the imine nitrogen (–C[double bond, length as m-dash]N).34 The ν(C[double bond, length as m-dash]C) stretching vibrations appeared in the range of 1550–1474 cm−1. Additionally, the B3LYP/DFT approach was used to calculate the IR spectra and identify vibrational bands (Table S2). The calculated vibrational frequencies closely match the experimental data, with slight differences due to DFT being performed on isolated gas-phase molecules while the experimental data reflect solid-state material. Moreover, the spectrum of FeIII-complex (1), in comparison to that of the free ligand, displays the emergence of two bands at 540 and 480 cm−1, corresponding to Fe–N and Fe–O stretching vibrations, respectively, thereby confirming coordination of the ligand to the metal center.

The electronic absorption properties of both the Schiff base ligand precursor (H2L) and its FeIII-complex (1) were examined using UV–Visible spectroscopy. Measurements were conducted in acetonitrile (3.0 μM) at 298 K, with absorbance recorded over the 200–800 nm range (Fig. 1). The spectra obtained for both H2L and FeIII-complex (1) are consistent with literature-reported data for related systems.35 The UV-vis spectrum of FeIII-complex (1) exhibits two strong absorption bands at 269 nm and 305 nm, which are also observed in the free ligand spectrum at 272 nm and 334 nm. These bands are attributed to π → π* and intra-ligand (IL) transitions within the aromatic rings and azomethine (C[double bond, length as m-dash]N) units of the Schiff base framework. These transitions, arising from the delocalized π-electron system, are only slightly shifted upon metal coordination, indicating a minor electronic perturbation due to complexation. Importantly, two additional low-intensity bands appear at 358 nm and 518 nm in the spectrum of FeIII-complex (1), which are absent in the free ligand. These are assigned to ligand-to-metal charge transfer (LMCT) transitions, specifically from the non-bonding orbitals of the phenolate oxygen atoms to the partially vacant d-orbitals of the high-spin Fe(III) center.35 Time-dependent density functional theory (TDDFT) calculations further support the spectral assignments by providing a frontier molecular orbital (FMO) interpretation consistent with the experimental data. The computed spectrum closely matches the experimental transitions in both energy and intensity. Notably, the LMCT bands at 358 nm and 518 nm are attributed to HOMO−1 → LUMO and HOMO → LUMO transitions, respectively, involving Fe-based donor orbitals (dz2 and dx2y2) and ligand-centered π* acceptor orbitals. The higher-energy bands at 269 nm and 305 nm are similarly attributed to π → π* and IL transitions, also corresponding to HOMO−1 → LUMO excitations within the ligand framework. Thus, the combined UV-vis spectroscopic and TDDFT analyses confirm successful metal coordination and are consistent with a five-coordinate Fe(III) center in a square pyramidal geometry.


image file: d5cy01107h-f1.tif
Fig. 1 Electronic absorption spectra of the ligand H2L (red line) and the FeIII-complex (1) (blue line) in acetonitrile at 298 K, with TDDFT-calculated oscillator strengths for FeIII-complex (1) shown as vertical blue lines (right axis).

The XPS spectrum of the FeIII-complex (1) shows distinct peaks corresponding to Fe 2p orbitals, characteristic of the Fe(III) oxidation state (Fig. 2A). The spectrum reveals Fe 2p3/2 and Fe 2p1/2 peaks at binding energies consistent with Fe(III), typically found at 711.91 eV for Fe 2p3/2 and 723.47 eV for Fe 2p1/2. Additionally, intense satellite peaks at 715.40 and 729.10 eV accompany the main Fe 2p signals, arising from shake-up transitions and ligand to metal charge transfer (LMCT), indicative of strong Fe–N/O covalency and a high-spin electronic configuration. These findings are consistent with literature-reported XPS features for high-spin FeIII in nitrogen- and oxygen-donor ligand environments.36 The spectral data thus validate the successful formation and oxidation state integrity of the FeIII center in the synthesized complex.


image file: d5cy01107h-f2.tif
Fig. 2 Spectroscopic characterization of FeIII-complex (1): (A) XPS spectrum highlighting the Fe 2p core-level region; (B) Mössbauer spectrum recorded at 295 K.

Mössbauer spectroscopy provides direct evidence for the electronic configuration of the Fe center. The spectrum of FeIII-complex (1) (Fig. 2B) exhibits a well-defined quadrupole-split doublet with parameters δ = 0.35 mm s−1 and ΔEq = 0.80 mm s−1. These values are characteristic of a high-spin Fe(III) (S = 5/2) center, consistent with the presence of five unpaired electrons occupying the majority-spin d orbitals. The relatively large isomer shift reflects the higher electron density at the iron nucleus typical of Fe(III), while the moderate quadrupole splitting indicates an asymmetric ligand field around the metal center. The clear doublet without additional hyperfine features confirms a single electronic environment, supporting the assignment of a uniform high-spin state. These results are in line with other reported high-spin Fe(III) Schiff-base complexes, further validating the oxidation state and spin configuration of FeIII-complex (1).37

Crystallographic characterization

Single crystals of the FeIII-complex (1) suitable for X-ray crystallographic analysis were successfully grown by the slow diffusion of n-hexane vapor into a dichloromethane solution of the complex at ambient temperature under aerobic conditions. The crystal structure was unambiguously determined by single-crystal X-ray diffraction analysis. The FeIII-complex (1) crystallizes in the triclinic space group P[1 with combining macron], possessing unit cell parameters a = 12.1450(8) Å, b = 15.8095(10) Å, c = 16.2008(10) Å, α = 87.8740(10)°, β = 82.5000(10)°, γ = 68.3490(10)°, a unit cell volume (V) of 2164.19(11) Å3 with Z = 4 molecules per unit cell, measured at T = 100 K.

The asymmetric unit of the FeIII-complex (1) comprises a central FeIII ion coordinated by a planar tetradentate di-anionic Schiff base ligand (L2−) with an N2O2 donor atom and an apical chloride ligand as illustrated in Fig. 3. The coordination environment around the FeIII center can be best described as a distorted square pyramidal (sqp) geometry, confirmed by the geometric parameter τ, calculated using the relationship τ = |βα|/60, where β and α are the largest and second largest coordination angles, respectively. For this complex, the angles β = N1–Fe1–O2 and α = N2–Fe1–O1 yield a τ value of 0.34, indicative of a moderate distortion from ideal square pyramidal geometry (where τ = 0 corresponds to ideal square pyramidal, and τ = 1 represents ideal trigonal bipyramidal geometry).


image file: d5cy01107h-f3.tif
Fig. 3 Molecular structure (at 100 K) of (A) FeIII-complex (1). The thermal ellipsoids of the ORTEP representation are set at 50% probability. The solvated water molecule has been omitted for clarity.

In the distorted square pyramidal arrangement, the FeIII ion is displaced out of the mean basal plane formed by C14N2O2 ligand core (Δ18Fe) by 0.64 Å towards the axial chloride ligand, reinforcing the square pyramidal assignment. The average bond lengths within the coordination sphere: Fe–N = 2.0925(3) Å, Fe–O = 1.8995(2) Å, and Fe–Cl = 2.2273(8) Å, align well with expected values for FeIII complexes in analogous coordination environments.38 Additionally, the asymmetric unit contains a discrete water molecule, which resides within an open compartment formed by the coordinated ligand framework. This water molecule participates in stabilizing hydrogen-bonding interactions with the methoxy (–OCH3) oxygen atoms of the ligand, contributing to the overall crystal packing stability.

The structural representation of FeIII-complex (1) is presented in Fig. 3, showing displacement ellipsoids drawn at the 50% probability level. Molecular arrangements within the crystal lattice and detailed atom-numbering schemes are provided in Fig. S2 and S3, respectively. Comprehensive crystal data, including full data collection parameters and bond lengths and angles, are tabulated in Tables S1 and S3–S5, respectively.

Chemical ammonia oxidation

At the onset of this research, we evaluated the ammonia oxidation activity of an FeIII-complex under oxidative conditions following the standard procedure reported by Nakajima et al.19 Encouragingly, we discovered that the FeIII-complex (1) exhibited ammonia oxidation potential at room temperature (Fig. 4, Table 1). The reactions were conducted using ammonium trifluoromethanesulfonate (NH4OTf, 3.0 mmol) as the ammonia source, tris(4-bromophenyl)ammoniumyl hexachloroantimonate (0.90 mmol) as the oxidant, and 2,4,6-collidine (3.0 mmol) as a base. Varying catalytic amounts of the FeIII-complex (0.010, 0.05, 0.25, and 0.50 mmol) were tested in 15 mL of acetonitrile at room temperature for 2 hours (Fig. 4, Table 1). The experiments were conducted in a 50 mL reactor equipped with an inlet helium gas port connected to a mass flow controller (Bronkhorst), a silicone rubber septum port for 2,4,6-collidine injection, and an outlet gas port connected to a gas chromatograph (Shimadzu Nexis GC-2030) with a BID detector for on-line nitrogen gas analysis.
image file: d5cy01107h-f4.tif
Fig. 4 Ammonia oxidation (AO) activity under varying conditions. (A) Time-dependent N2 production rates at different catalyst loadings (0.5, 0.25, 0.05, and 0.01 mmol). (B) Total amount of N2 evolved at each catalyst loading. (C) N2 production rates using different ammonia sources/solvents (NH4NO3, NH3, and MeOH). (D) Total amount of N2 evolved for each nitrogen source/solvent. In part A and C, the N2 production rate is normalized to the catalyst mass.
Table 1 Catalytic ammonia oxidation with FeIII-complex (1)

image file: d5cy01107h-u1.tif

Entry Catalyst amount (mmol) NH3 source Solvent N2a (equiv.) N2b (%)
Reactions of ammonium triflate (3.0 mmol), oxidant (0.9 mmol), and base (3.0 mmol) were carried out in the presence of iron catalyst under an He atmosphere at room temperature for 2 h.a Equivalents are produced N2 per Fe atom.b Yields are based on oxidant. Selectivity for N2 >99%, determined by on-line GC (BID detector).
1 0.5 NH4OTf MeCN 0.05 17.83
2 0.25 NH4OTf MeCN 0.18 31.20
3 0.05 NH4OTf MeCN 2.20 73.3
4 0.01 NH4OTf MeCN 3.01 20.05
5 0.05 NH4OTf MeOH 0.48 16.05
6 0.05 NH4NO3 MeCN 0.61 20.53
7 0.05 NH3(MeCN) MeCN 0.73 24.36
8 0.00 NH4OTf MeCN 0.00 0.00


Prior to initiating the catalytic investigation, the reactor's headspace was purged with helium to eliminate air. Subsequently, the 2,4,6-collidine was injected, and the resultant N2 was quantified by directly analyzing the gas stream at a flow rate of 20 mL min−1. Under optimum conditions (0.05 mmol catalyst), the reaction yielded 0.109 mmol N2, equating to 2.20 equivalents per Fe atom and attaining a 73% yield relative to the oxidant. At higher catalyst loadings, decreased N2 production is attributed to possible catalyst–catalyst interactions such as aggregation or dimerization under oxidative conditions, which reduce the number of active sites.39–41

Subsequent experiments demonstrated considerable effects resulting from changes in ammonia supplies and solvents. Substituting NH4OTf with NH4NO3 led to significantly diminished N2 production (0.61 equivalents and a 20.53% yield, Table 1), underscoring the pivotal impact of ammonia source selection on catalytic efficacy. NH4OTf presumably provided superior solubility and enhanced contact with the Fe(III) catalyst relative to other investigated ammonium salts, hence augmenting overall catalytic efficacy. The selection of solvent significantly affected the results; acetonitrile was much more effective than methanol, as seen by a considerable reduction in catalytic activity with methanol (0.48 equivalents and a 16.05% yield). The higher performance with acetonitrile may stem from its advantageous dielectric constant, better catalyst solubility, and optimum stability of reactive intermediates.

Furthermore, trials conducted in the absence of a base resulted in negligible N2 generation, so unequivocally demonstrating the essential function of the base in promoting the ammonia oxidation process, presumably by efficiently deprotonating intermediates. These thorough investigations emphasize the ideal catalytic conditions attained with NH4OTf as the ammonia source, acetonitrile as the solvent, and a catalyst concentration of 0.05 mmol, effectively balancing high catalytic efficiency and significant yield, thereby underscoring the potential applicability of the Fe(III) complex in practical ammonia oxidation processes. Gas-phase selectivity, showed N2 as the sole nitrogen-containing product with no detectable N2O, NO, or NO2, giving >99% selectivity toward N2 under all conditions.

For comparison, Nakajima and co-workers reported Ru-based catalysts capable of oxidizing ammonia to N2 with up to 11.9 equiv. per Ru atom (79% yield), and mechanistic studies supported a nitride–nitride coupling pathway.19 More recently, Toda and co-workers demonstrated Mn-salen complexes as the first earth-abundant homogeneous systems for ammonia oxidation, producing up to 17.1 equiv. of N2 per Mn atom under chemical conditions.42 Against this backdrop, the observed activity of the Fe(III) complex underscores its potential as a sustainable alternative for ammonia oxidation catalysis.

DFT-based electronic structure and mechanistic analysis of ammonia oxidation

Density functional theory (DFT) calculations were performed to elucidate both the electronic structure of the FeIII-complex (1) and the mechanistic pathway of ammonia oxidation to dinitrogen. Geometry optimizations for the doublet (S = 1/2), quartet (S = 3/2), and sextet (S = 5/2) spin states, followed by frequency analyses, confirmed that all optimized structures correspond to true minima, with computed geometries closely matching experimental data. Spin-density analysis revealed predominant localization of unpaired electrons on the Fe center, confirming its +III oxidation state and high-spin (sextet) configuration (Fig. S4, Table S3). To further investigate the catalytic mechanism of ammonia oxidation to dinitrogen, we performed density functional theory (DFT) calculations at the B3LYP level, incorporating Grimme's D3 dispersion correction and the SMD solvation model with MeCN as the solvent. The reaction was modelled using an iron-based catalyst and evaluated through two alternative mechanistic routes, designated pathway A and pathway B. Each pathway comprises a sequence of oxidation (electron transfer) and deprotonation (proton transfer) steps, and the thermodynamic feasibility of each transformation was assessed by calculating the corresponding free energy changes (ΔG). The computed catalytic cycles are depicted in Fig. 5A and B, corresponding to pathways A and B, respectively, with the free energy changes (ΔGrxn) annotated for each step. Among the two, pathway B (the red pathway, which proceeds without ligand exchange) is thermodynamically more favorable, as shown in Fig. 6.
image file: d5cy01107h-f5.tif
Fig. 5 Proposed catalytic cycle for ammonia oxidation by (A) the complex via a ligand exchange pathway and (B) without ligand exchange. Free energy changes (ΔG) for individual steps are given in kcal mol−1. Oxidation steps are highlighted in red and deprotonation steps in blue.

image file: d5cy01107h-f6.tif
Fig. 6 Calculated free energy profile (kcal mol−1) for the conversion of NH3 to N2 (A to O) via the catalytic cycle in Fig. 6, mediated by the FeIII-complex (1). The ligand exchange pathway is shown in green, and the pathway without ligand exchange is shown in red. Free energies for individual steps were computed in SMD-MeCN continuum solvent.

This preference arises from its avoidance of high-energy intermediates and a more accessible overall energy landscape. Accordingly, the following mechanistic discussion focuses primarily on pathway B. For each intermediate, geometry optimizations were carried out by considering three possible spin states of the Fe(III) center in low-spin, intermediate-spin, and high-spin, and the reported free energy values correspond to the most stable (lowest-energy) spin configuration. Furthermore, detailed electronic structure analyses were performed for all key intermediates (6I to 6XV) involved in the multistep proton and electron transfer events of mechanistic pathway B (Fig. 7).


image file: d5cy01107h-f7.tif
Fig. 7 DFT-computed mechanism for FeIII-complex (1) catalyzed ammonia oxidation, showing key intermediates (A to O), spin states, spin density distributions, and free energy changes (in kcal mol−1). The calculations highlight electron and proton transfer steps and the evolution of metal- and ligand-centered radical character throughout the catalytic cycle. All intermediates shown were obtained from DFT calculations.

Additionally, a possible bimetallic N–N coupling pathway was evaluated. Four spin states: singlet, triplet, septet, and 11-et were examined, with the S = 11 state identified as the most stable configuration for the bimetallic intermediate. In this high-spin configuration, both Fe centres display parallel spins, resulting in an overall high-spin bimetallic species. As shown in the Fig. S5, the nucleophilic attack of NH3 on the Fe(IV)–NH· species is highly exergonic (ΔG = −32.39 kcal mol−1), whereas the bimetallic amido-coupling pathway is considerably less favorable (ΔG = 4.73 kcal mol−1). These findings indicate that the mononuclear nucleophilic attack mechanism is energetically preferred, while the bimetallic pathway is not favoured under the studied conditions.

Coordination of NH3 to the catalyst (steps A to B). Ammonia coordinates to the Fe(III) center of the complex [FeIII(L)Cl], forming the ammine complex [FeIII(L)(NH3)Cl]. This is a straightforward ligand coordination event without any change in oxidation state or spin multiplicity. Spin-density analysis confirms that the Fe center remains in a high-spin Fe(III) state, with the spin density slightly increasing from +4.07 in A to +4.12 in B, consistent with a sextet (S = 5/2) ground state. The process is thermodynamically favorable, with a computed free-energy change of −3.71 kcal mol−1, suggesting spontaneous NH3 binding under standard conditions. This coordination marks the first step in substrate activation and sets the stage for subsequent redox and bond-forming events in the catalytic pathway.
First oxidation/deprotonation (steps B to D). The first electron/proton transfer step begins with the oxidation of [FeIII(L)(NH3)Cl], a paramagnetic d5 complex. In the step from B → C, the complex [FeIII(L)(NH3)Cl] undergoes a one-electron oxidation, forming [FeIII(L+˙)(NH3)Cl]+. Spin-density analysis reveals that the oxidation is ligand-centered, as evidenced by the emergence of significant negative spin density on the ligand backbone (−0.25), indicating formation of a ligand π–cation radical (L+˙) with β-spin character. The iron center retains a high positive spin density, decreasing only slightly from +4.12 (6II) to +3.81 (5III), confirming that it remains in the Fe(III) oxidation state. This shift in spin distribution alters the magnetic properties of the intermediate (5III). Unlike 6II, which is a sextet (S = 5/2) due to high-spin Fe(III), intermediate 5III exhibits a quintet (S = 2) ground state. This arises from antiferromagnetic coupling between the high-spin Fe(III) center (S = 5/2, α-spin) and the ligand radical (S = 1/2, β-spin), resulting in a net spin of S = 2. The NH3 ligand remains essentially spin-inactive, with minimal spin density (0.12), confirming it does not participate in the redox event. The overall oxidation is exergonic, with a free-energy change of −3.80 kcal mol−1, suggesting a spontaneous nature under standard conditions. This step represents a key early redox activation of the ligand framework, setting up subsequent proton and electron transfer events critical for N–N bond formation in the catalytic cycle.

In the C → D step, the first deprotonation of the formally [FeIII(L+˙)(NH3)Cl]+ complex by 2,4,6-collidine is calculated to be endergonic, with a free-energy change of +19.80 kcal mol−1, indicating that this step requires external driving force under reaction conditions. The proton abstraction from the coordinated NH3 generates an Fe(III)-aminyl radical species [FeIII(L)(NH2˙)Cl], depending on the resulting electronic distribution. Spin-density analysis of the lowest-energy intermediate (3IV) reveals significant unpaired electron density on both the Fe center (+1.83) and the NH2 ligand (+0.17), with nearly zero spin density on the ligand backbone, confirming that the ligand radical character (L+˙) has been neutralized. This occurs through internal electron redistribution, where electron density from the more electron-rich NH2 group compensates for the ligand radical, restoring the ligand to a closed-shell (L) state. As a result, the radical character shifts from the ligand backbone to the NH2 group, forming an aminyl radical. These features clearly support an electronic structure best described as an Fe(III)–aminyl radical, where the iron remains in the Fe(III) oxidation state. The ground state of this intermediate is a triplet (S = 1), arising from ferromagnetic coupling between the low-spin Fe(III) center (S = 1/2) and the aminyl radical (S = 1/2) on the NH2 ligand. This configuration reflects a substantial reorganization of spin density relative to intermediate 5III and highlights the role of the deprotonation step in converting a ligand-centered radical into a metal–ligand radical pair, setting the stage for subsequent redox and bond-forming processes in the catalytic pathway.

Second oxidation/deprotonation (steps D to F). The second electron/proton transfer step begins with the oxidation of Fe(III)-aminyl radical, [FeIII(L)(NH2˙)Cl], a low spin d5 complex. The calculated free energy change for this oxidation (step D → E in Fig. 4) is exergonic by −4.30 kcal mol−1. Spin density analysis (Fig. 5a) indicates that the oxidation is primarily ligand-centered, leading to the formation of a [FeIII(L+˙)(NH2˙)Cl]+ species. The lowest-energy electronic configuration corresponds to an overall doublet spin state (2S + 1 = 2), with the FeIII center adopting a low-spin configuration (S = 1/2). The resulting spin distribution suggests ferromagnetic coupling between the aminyl radical (NH2˙) and the low-spin FeIII fragment. The overall electron transfer process is exergonic, with a computed free-energy change of −4.30 kcal mol−1, underscoring its thermodynamic feasibility and electronic driving force.

Upon deprotonation of [FeIII(L+˙)(NH2˙)Cl]+ (step D → E in Fig. 4), the electronic structure undergoes an intramolecular electron redistribution process. When a proton is removed from the NH2 ligand, the resulting electron redistribution triggers internal electron transfer from the Fe–NH2˙ moiety to the ligand radical (L+˙). This internal electron transfer neutralizes the ligand radical cation, converting L+˙ into a neutral ligand (L). Concurrently, the iron center is oxidized from Fe(III) to Fe(IV), and the deprotonated nitrogen ligand is transformed into an iminyl radical (NH˙) species. Consequently, the final resulting complex, [FeIV(L)(NH˙)Cl], is best described as a low-spin Fe(IV) species coupled antiferromagnetically with an iminyl radical (NH˙) ligand, exhibiting an overall doublet (2S + 1 = 2) ground state. This interpretation aligns closely with DFT-computed spin densities, which show a moderate spin density localized at iron (+1.83), spin density at nitrogen (0.17), and essentially no residual radical character on the ligand backbone (L). The overall deprotonation and electron transfer process is endergonic, with a computed free-energy change of 11.56 kcal mol−1, underscoring its thermodynamic feasibility and electronic driving force.

Third oxidation/deprotonation, and nucleophilic attack of NH3 on iminyl nitrogen (steps F–I). Upon examining the nucleophilic attack step (F → G), intermediate 2VI ([FeIV(L)(NH˙)Cl]), initially in a low-spin doublet ground state, undergoes radical coupling via direct nucleophilic attack by ammonia (NH3). The iminyl nitrogen ligand (spin density 0.17 at NH˙) in intermediate 2VI acts as an electrophilic site, facilitating the nucleophilic addition and formation of a stable N–N bond, yielding intermediate 6VII ([FeIII(L)(NH–NH3)Cl]). This coupling event transfers electron density from ammonia to the iminyl nitrogen ligand, neutralizing its radical character. The resultant electron redistribution reduces the Fe(IV) center to Fe(III), the iron transitions from a low-spin (S = 1) to a more favorable high-spin (S = 5/2, sextet) ground state, clearly indicated by the substantial increase in spin density at iron (from +1.13 to +4.06). The overall radical coupling and electron transfer process is highly exergonic, with a computed free-energy change of −32.39 kcal mol−1, underscoring its strong thermodynamic and electronic driving force.

In the third oxidation step (G → H), the complex undergoes a one-electron oxidation from ([FeIII(L)(NH–NH3)Cl]) to ([FeIV(L)(NH–NH3)Cl]+) which is indicated by the spin-density redistribution and electronic structural changes. Initially, intermediate 6VII is in a high-spin Fe(III) sextet state (spin density at Fe: +4.06), reflecting an electron-rich and stable Fe(III) center coordinated to the hydrazine-type ligand (NH–NH3). Upon oxidation, intermediate 3VIII is formed, and the spin density at iron significantly decreases (from +4.06 to +2.42), indicating oxidation of Fe(III) to Fe(IV). The ligand remains non-radical, as evidenced by negligible spin density on nitrogen atoms (0.05–0.06). This oxidation step therefore clearly occurs at the iron center, converting it from a high-spin Fe(III) (d5) into a low-spin Fe(IV) (d4) species. Consequently, intermediate 3VIII adopts a triplet (S = 1) ground state with the Fe(IV) oxidation state. This one-electron oxidation is exergonic, with a computed free-energy change of −2.70 kcal mol−1, highlighting that the step is thermodynamically feasible.

In the third deprotonation step from intermediate 3VIII ([FeIV(L)(NH–NH3)Cl]+) to 3IX ([FeIV(L)(NH–NH2)Cl]), the proton is abstracted from the hydrazine-type ligand (NH–NH3), forming an Fe(IV)-hydrazido species. The spin density distribution shows minimal changes at the iron center (spin density slightly decreases from +2.42 in 3VIII to +1.17 in 3IX), confirming that the oxidation state (Fe(IV)) and spin delocalized over the Fe(IV) center and NH–NH2 (spin density 0.14) moiety during this step. The nitrogen atoms also show negligible spin density (0.02), confirming no radical character arises on the ligand upon proton removal. Thus, the step H → I primarily represent a ligand-centered deprotonation without significant electron-transfer processes or spin-state transitions at the metal center. The resultant species (3IX) remains in a low-spin Fe(IV) triplet (S = 1) ground state. This step is highly thermodynamically favorable, as indicated by the large negative free-energy change (−23.90 kcal mol−1), clearly demonstrating that deprotonation of the hydrazine ligand is energetically favorable and readily occurs under reaction conditions.

Fourth oxidation/deprotonation (step I to K). In the fourth oxidation step from intermediate 3IX ([FeIV(L)(NH–NH2)Cl]) to intermediate 2X ([FeIV(L+˙)(NH–NH2˙)Cl]+), a one-electron oxidation occurs that introduces significant radical character simultaneously on the ligand backbone and the hydrazido ligand. Spin-density analysis reveals three distinct spin centers: positive spin density on the iron center (+1.04, α-spin), substantial positive spin density on the hydrazido ligand (+0.85, α-spin), and negative spin density on the ligand backbone (−0.62, β-spin). The positive spin densities at iron and hydrazido ligand indicate a strong ferromagnetic coupling between Fe(IV) and the hydrazido radical (NH–NH2˙). However, this combined Fe–hydrazido fragment is antiferromagnetically coupled with the ligand π–cation radical (L+˙), carrying opposite β-spin, resulting in partial spin cancellation and stabilization of the intermediate in a doublet (S = 1/2) ground state. The one-electron oxidation step is exergonic (−33.12 kcal mol−1), remains thermodynamic feasible.

In the fourth deprotonation step (J → K), intermediate 2X ([FeIV(L+˙)(NH–NH2˙)Cl]+) undergoes proton removal from the hydrazido radical ligand (NH–NH2˙), forming intermediate 2XI ([FeIII(L)(N–NH2)Cl]). Spin-density analysis clearly indicates a significant redistribution of spin density, with the iron center having a positive spin density of +1.09 (α-spin), and essentially negligible spin density on both the ligand backbone (L) and the diazenido ligand (N[double bond, length as m-dash]NH2), indicating that the radical character previously localized on these ligands is effectively quenched upon proton loss. Thus, this deprotonation step is accompanied by internal electron transfer, which reduces the ligand radical (L+˙) and hydrazido radical (NH–NH2˙) back to their closed-shell states, stabilizing a purely Fe(III)-centered species. The iron center, therefore, remains in a low-spin Fe(III) oxidation state, and the intermediate 2XI maintains a doublet (S = 1/2) ground state, now originating solely from a single unpaired electron localized primarily on Fe. The free-energy change for this deprotonation and electron redistribution step is moderately endergonic (11.29 kcal mol−1), yet remains feasible under reaction conditions.

Fifth oxidation/deprotonation (step K to M). In the fifth oxidation step (K → L), intermediate 2XI ([FeIII(L)(N–NH2)Cl]), initially a doublet species (Fe spin density = +1.09), undergoes one-electron oxidation to form intermediate 3XII ([FeIII(L+˙)(N–NH2)Cl]+). Spin-density analysis of intermediate 3XII shows substantial spin density localized on the iron center (+1.24) and on the ligand backbone (+0.53), with negligible spin density on the diazenido ligand (N–NH2). These data clearly indicate that oxidation occurs primarily on the ligand backbone, forming a ligand π–cation radical (L+˙), while the Fe center maintains a low-spin Fe(IV) state. Crucially, intermediate 3XII now adopts a triplet (S = 1) ground state, resulting from ferromagnetic coupling between the Fe(III) center (α-spin) and ligand radical (also α-spin). Thus, the spins on Fe and ligand align in parallel, increasing the overall spin multiplicity to a triplet. The computed free-energy change for this ligand-centered oxidation step is significantly exergonic (−8.08 kcal mol−1), identifying it as thermodynamically feasible under reaction conditions.

In the deprotonation step (L → M), intermediate 3XII ([FeIII(L+˙)(N–NH2)Cl]+), initially in a triplet ground state (S = 1), undergoes proton removal from the diazenido ligand (N–NH2), forming intermediate 5XIII ([FeIII(L)(N[double bond, length as m-dash]NH˙)Cl]). Spin-density analysis clearly shows a major electronic rearrangement: spin density at the iron center significantly increases from +1.24 in 3XII to +2.70 in 5XIII, indicating a spin switching of low-spin Fe(III) to intermediate spin (S = 3/2) Fe(III). Concurrently, the nitrogen ligand develops notable radical character (spin density = +0.98), corresponding to the formation of a diazenyl radical (N[double bond, length as m-dash]NH˙). Negligible spin density remains on the ligand backbone (L), demonstrating the ligand π–cation radical (L+˙) is fully neutralized to its closed-shell form during this process. As a consequence, intermediate 5XIII adopts an intermediate-spin Fe(III) quintet ground state (S = 2), resulting from the ferromagnetic coupling between the Fe(III) center and the diazenyl radical ligand. The computed free-energy change for this step is significantly exergonic (−28.56 kcal mol−1), highlighting the strong thermodynamic driving force favoring the formation of intermediate 5XIII under experimental conditions.

Sixth oxidation/deprotonation (steps M to O) and N2 dissociation (catalytic cycle restart). In the sixth oxidation step (M → N), intermediate 5XIII ([FeIII(L)(N[double bond, length as m-dash]NH˙)Cl]), initially in a quintet ground state (S = 2), undergoes a one-electron oxidation to form intermediate 4XIV ([FeIII(L+˙)(N[double bond, length as m-dash]NH˙)Cl]+). Spin-density analysis clearly indicates oxidation occurring predominantly on the ligand backbone, reflected by the emergence of significant negative (β-spin) density on ligand L (−0.56). The iron spin density decreases slightly from +2.70 (5XIII) to +2.47 (5XIV), while spin density on the diazenyl radical ligand slightly increases from +0.98 to +1.10, consistent with a persistent radical character. Thus, intermediate 4XIV features three distinct spin centers: the high-spin Fe(III) (α-spin), the diazenyl radical ligand (N[double bond, length as m-dash]NH˙, α-spin), and the ligand π–cation radical (L+˙, β-spin). The quartet ground state (S = 3/2) in intermediate 4XIV arises from strong ferromagnetic coupling between the Fe(III) center and diazenyl radical ligand (both α-spin), forming a combined high-spin fragment, which is further coupled antiferromagnetically to the ligand radical cation (β-spin), resulting the net quartet (S = 3/2) spin state. This oxidation step is nearly thermoneutral, having only a minor free-energy increase of +0.59 kcal mol−1, making it energetically feasible under typical catalytic conditions.

In the sixth deprotonation step (N → O), intermediate 4XIV ([FeIII(L+˙)(N[double bond, length as m-dash]NH˙)Cl]+), initially in a quartet ground state (S = 3/2), undergoes proton removal from the diazenyl radical ligand (N[double bond, length as m-dash]NH˙), forming intermediate 6XV ([FeIII(L)(N[triple bond, length as m-dash]N)Cl]). This proton abstraction triggers internal electron transfer, where electron density previously localized as radical character on the ligand backbone (L+˙) and diazenyl ligand (N[double bond, length as m-dash]NH˙) is redistributed back onto the iron center. Consequently, both ligand radicals become closed-shell species, losing their radical nature, as evidenced by negligible spin densities on these fragments. Simultaneously, the iron center is formally switch from intermediate spin (3/2) Fe(III) to high-spin Fe(III), confirmed by a significantly increased spin density (+3.97). Intermediate 6XV thus adopts a high-spin sextet (S = 5/2) ground state arising from the d5 electronic configuration of Fe(III). The computed free-energy change for this internal electron transfer and deprotonation step is highly exergonic (−92.42 kcal mol−1), underscoring a strong thermodynamic driving force stabilizing the formation of intermediate 6XV. This high exergonicity can be attributed to the formation of the N2 triple bond. In the final step, the loss of dinitrogen (N2) is exergonic, by −2.55 kcal mol−1 for 6XV intermediate. The catalytic loop can be restarted by the detachment of the N2 to form intermediate 6I.

Hence, in the overall ammonia oxidation catalytic cycle, the spin-state transition from low-spin to intermediate-spin at the Fe centre plays a crucial role in facilitating N–N bond formation by enhancing both metal–ligand orbital interactions and electronic flexibility. In the low-spin state, Fe d-orbitals are mostly paired, which restricts σ/π overlap with the nitrogen donors. Upon transition to an intermediate-spin configuration, partial population of the antibonding eg* orbitals increase Fe–N/O covalency and allows spin density delocalization onto the –NHx ligands. This delocalization promotes radical-type coupling between Fe–NH· intermediates or nucleophilic attack of NH3 on electrophilic Fe–NH species, thereby lowering the barrier for N–N bond formation.43,44 Furthermore, Zott and Peters have demonstrated that spin-state flexibility is crucial for efficient ammonia oxidation catalysis, as it allows the Fe center to access multiple oxidation and spin states during turnover. This flexibility enables the Fe center to mediate multielectron transformations with reduced reorganization energy, thereby promoting faster and more efficient N–N bond formation.32 Accordingly, the low-spin to intermediate-spin transition in our system provides an optimal balance between redox flexibility, Fe–N covalency, and spin delocalization, all of which are essential for effective Fe-mediated ammonia oxidation.

Conclusions

In conclusion, we report a high-spin FeIII-complex with a trans-1,2-bis(2-hydroxy-3-methoxyphenylmethaniminyl)cyclohexane salen-type ligand that efficiently catalyzes ammonia oxidation to dinitrogen under oxidative conditions. In the presence of a triarylaminium oxidant and 2,4,6-collidine base, the catalyst delivers up to 2.20 equivalents of N2 per iron center. Integrated spectroscopic, crystallographic, and computational analyses establish the structural and electronic features of the complex and reveal a nucleophilic attack pathway for N–N bond formation. Together, these findings advance molecular ammonia oxidation catalysis with earth-abundant iron, offering key mechanistic insight for the rational design of future catalysts and reinforcing the potential of ammonia as a practical carbon-free energy carrier.

Conflicts of interest

The Authors declare no competing financial interest.

Data availability

The related data have been provided in the form of supplementary information (SI). Supplementary information: the additional, experimental details, characterizations and computational details are available in the SI. See DOI: https://doi.org/10.1039/d5cy01107h.

CCDC 2384217 contains the supplementary crystallographic data for this paper.45

Acknowledgements

This research was supported by the King Fahd University of Petroleum and Minerals (KFUPM). The authors gratefully acknowledge the support provided by the Deanship of Research (DR) at King Fahd University of Petroleum & Minerals (KFUPM) for funding this work through project No.: INHT2410.

References

  1. A. Grinberg Dana, O. Elishav, A. Bardow, G. E. Shter and G. S. Grader, Nitrogen-based fuels: a power-to-fuel-to-power analysis, Angew. Chem., Int. Ed., 2016, 55, 8798–8805 CrossRef CAS PubMed.
  2. O. Elishav, B. Mosevitzky Lis, E. M. Miller, D. J. Arent, A. Valera-Medina, A. Grinberg Dana, G. E. Shter and G. S. Grader, Progress and prospective of nitrogen-based alternative fuels, Chem. Rev., 2020, 120, 5352–5436 CrossRef CAS PubMed.
  3. K. Morita, Ammonia − a fuel of the future?, available at: https://www.euractiv.com/section/energy-environment/opinion/ammonia-a-fuelof-the-future/, (accessed 10 February 2021).
  4. A. H. Tullo, Is ammonia the fuel of the future? Industry sees the agricultural chemical as a convenient means to transport hydrogen, available at: https://cen.acs.org/business/petrochemicals/ammonia-fuel-future/99/i8, (accessed 2 September 2025).
  5. P. L. Dunn, B. J. Cook, S. I. Johnson, A. M. Appel and R. M. Bullock, Oxidation of ammonia with molecular complexes, J. Am. Chem. Soc., 2020, 142, 17845–17858 CrossRef CAS PubMed.
  6. J. G. Chen, R. M. Crooks, L. C. Seefeldt, K. L. Bren, R. M. Bullock, M. Y. Darensbourg, P. L. Holland, B. Hoffman, M. J. Janik, A. K. Jones, M. G. Kanatzidis, P. King, K. M. Lancaster, S. V. Lymar, P. Pfromm, W. F. Schneider and R. R. Schrock, Beyond fossil fuel–driven nitrogen transformations, Science, 2018, 360, 6611 CrossRef.
  7. X. Zhu, J. Du, Z. Yu, Y.-B. Cheng and Y. Wang, NOx emission and control in ammonia combustion: state-of-the-art review and future perspectives, Energy Fuels, 2024, 38, 43–60 CrossRef CAS.
  8. N. Gruber and J. N. Galloway, An Earth-system perspective of the global nitrogen cycle, Nature, 2008, 451, 293–296 CrossRef CAS PubMed.
  9. C. Ferousi, S. H. Majer, I. M. DiMucci and K. M. Lancaster, Biological and bioinspired inorganic N–N bond-forming reactions, Chem. Rev., 2020, 120, 5252–5307 CrossRef CAS PubMed.
  10. X. Zhang, B. B. Ward and D. M. Sigman, Global nitrogen cycle: critical enzymes, organisms, and processes for nitrogen budgets and dynamics, Chem. Rev., 2020, 120, 5308–5351 CrossRef CAS PubMed.
  11. J. J. Warren, T. A. Tronic and J. M. Mayer, Thermochemistry of proton-coupled electron transfer reagents and its implications, Chem. Rev., 2010, 110, 6961–7001 CrossRef CAS.
  12. J. Zhao, A. S. Goldman and J. F. Hartwig, Oxidative addition of ammonia to form a stable monomeric amido hydride complex, Science, 2005, 307, 1080–1082 CrossRef CAS.
  13. N. V. Rees and R. G. Compton, Carbon-free energy: a review of ammonia- and hydrazine-based electrochemical fuel cells, Energy Environ. Sci., 2011, 4, 1255–1260 RSC.
  14. E. K. Dann, E. K. Gibson, R. H. Blackmore, C. R. A. Catlow, P. Collier, A. Chutia, T. E. Erden, C. Hardacre, A. Kroner, M. Nachtegaal, A. Raj, S. M. Rogers, S. F. R. Taylor, P. Thompson, G. F. Tierney, C. D. Zeinalipour-Yazdi, A. Goguet and P. P. Wells, Mechanistic insights into ammonia oxidation over transition-metal catalysts, Nat. Catal., 2019, 2, 157–163 CrossRef CAS.
  15. (a) M. Usman, A. Ali, Z. H. Yamani and M. N. Shaikh, Catalytic pathways for efficient ammonia-to-hydrogen conversion towards a sustainable energy future, Sustainable Energy Fuels, 2024, 8, 5329–5351 RSC; (b) M. Usman, A. Ali, A. Jedidi, A. Ajeebi, M. M. Hossain, K. M. Yau, H. Alghamdi, M. A. Aziz and M. N. Shaikh, Rare-earth-metal promoters (La, Ce, Nd, Sm) on nickel-supported Al2O3 catalysts for ammonia decomposition, Fuel, 2025, 396, 135272 CrossRef CAS.
  16. (a) P. L. Dunn, B. J. Cook, S. I. Johnson, A. M. Appel and R. M. Bullock, Oxidation of ammonia with molecular complexes, J. Am. Chem. Soc., 2020, 142, 17845–17858 CrossRef CAS PubMed; (b) J. Li, F. Zhang, H. Xiong, Y. Cai and B. Zhang, Molecular catalysts for electrocatalytic ammonia oxidation, Sci. China: Chem., 2024, 67, 3976–3993 CrossRef CAS.
  17. (a) M. S. Thompson and T. J. Meyer, Oxidation of coordinated ammonia, J. Am. Chem. Soc., 1981, 103, 5577–5579 CrossRef CAS; (b) J. P. Collman, J. E. Hutchison, M. S. Ennis, M. A. Lopez and R. Guilard, Reduced nitrogen hydride complexes of a cofacial metallodiporphyrin and their oxidative interconversion, J. Am. Chem. Soc., 1992, 114, 8074–8080 CrossRef CAS; (c) J. P. Collman, J. E. Hutchison, M. S. Ennis, M. A. Lopez and R. Guilard, Reduced nitrogen hydride complexes of a cofacial metallodiporphyrin and their oxidative interconversion: an analysis of ammonia oxidation and prospects for a dinitrogen electroreduction catalyst based on cofacial metallodiporphyrins, J. Am. Chem. Soc., 1992, 114, 8074–8080 CrossRef CAS.
  18. F. Habibzadeh, S. L. Miller, T. W. Hamann and M. R. Smith III, Homogeneous electrocatalytic oxidation of ammonia to N2 under mild conditions, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 2849–2853 CrossRef CAS.
  19. K. Nakajima, H. Toda, K. Sakata and Y. Nishibayashi, Ruthenium-catalysed oxidative conversion of ammonia into dinitrogen, Nat. Chem., 2019, 11, 702–709 CrossRef CAS.
  20. G. W. Margulieux, Z. R. Turner and P. J. Chirik, Synthesis and ligand modification chemistry of a molybdenum dinitrogen complex: redox and chemical activity of a bis(imino)pyridine ligand, Angew. Chem., Int. Ed., 2014, 53, 14211–14215 CrossRef CAS PubMed.
  21. M. J. Bezdek, S. Guo and P. J. Chirik, Coordination-induced weakening of ammonia, water and hydrazine X–H bonds in a molybdenum complex, Science, 2016, 354, 730–733 CrossRef CAS PubMed.
  22. (a) J. A. Gutiérrez-Orgaz, M. Nicaso, A. M. Beiler, M. Martínez-Belmonte, J. Benet-Buchholz, M. Segado-Centellas, X. Sala, C. Bo and A. Llobet, Exploring copper β-diketiminate complexes for heterogeneous ammonia oxidation anchored on graphitic surfaces via CH–π and π–π interactions, Inorg. Chem. Front., 2025, 12, 7234–7249 RSC; (b) H. Xiong, J. Yang, J. Li, Y. Cai, F. Zhang, J.-Y. Chen, R.-Z. Liao, L. Sun and B. Zhang, Electrochemical Ammonia Oxidation Catalyzed by a Ruthenium Complex with a Dangling Sulfonate Group, ACS Catal., 2025, 15, 8633–8642 CrossRef CAS; (c) T. Gupta, Sanyam, S. Saraswat, A. Mondal and B. Mondal, Electrochemical ammonia oxidation with a homogeneous molecular redox mediator, Chem. Sci., 2025, 16, 14377–14383 RSC; (d) J. Li, X. Shi, F. Zhang, X. Lu, Y. Zhang, R. Liao and B. Zhang, Electrocatalytic Ammonia Oxidation by a Ruthenium Complex Bearing a 2,6-Pyridinedicarboxylate Ligand, JACS Au, 2025, 5, 1812–1821 CrossRef CAS.
  23. G. W. Margulieux, M. J. Bezdek, Z. R. Turner and P. J. Chirik, Ammonia activation, H2 evolution and nitride formation from a molybdenum complex with a chemically and redox non-innocent ligand, J. Am. Chem. Soc., 2017, 139, 6110–6113 CrossRef CAS.
  24. P. Bhattacharya, Z. M. Heiden, E. S. Wiedner, S. Raugei, N. A. Piro, W. S. Kassel, R. M. Bullock and M. T. Mock, Ammonia oxidation by abstraction of three hydrogen atoms from a Mo–NH3 complex, J. Am. Chem. Soc., 2017, 139, 2916–2919 CrossRef CAS PubMed.
  25. M. J. Bezdek and P. J. Chirik, Interconversion of molybdenum imido and amido complexes by proton-coupled electron transfer, Angew. Chem., Int. Ed., 2018, 57, 2224–2228 CrossRef CAS PubMed.
  26. B. J. Cook, S. I. Johnson, G. M. Chambers, W. Kaminsky and R. M. Bullock, Triple hydrogen atom abstraction from Mn–NH3 complexes results in cyclophosphazenium cations, Chem. Commun., 2019, 55, 14058–14061 RSC.
  27. P. Bhattacharya, Z. M. Heiden, G. M. Chambers, S. I. Johnson, R. M. Bullock and M. T. Mock, Catalytic ammonia oxidation to dinitrogen by hydrogen atom abstraction, Angew. Chem., Int. Ed., 2019, 58, 11618–11624 CrossRef CAS PubMed.
  28. P. L. Dunn, S. I. Johnson, W. Kaminsky and R. M. Bullock, Diversion of catalytic C–N bond formation to catalytic oxidation of NH3 through modification of the hydrogen atom abstractor, J. Am. Chem. Soc., 2020, 142, 3361–3365 CrossRef CAS.
  29. (a) B. Kartal and J. T. Keltjens, Anammox biochemistry: a tale of heme c proteins, Trends Biochem. Sci., 2016, 41, 998–1011 CrossRef CAS; (b) K. M. Lancaster, J. D. Caranto, S. H. Majer and M. A. Smith, Alternative bioenergy: updates to and challenges in nitrification metalloenzymology, Joule, 2018, 2, 421–441 CrossRef CAS.
  30. B. M. Hoffman, D. Lukoyanov, Z.-Y. Yang, D. R. Dean and L. C. Seefeldt, Mechanism of nitrogen fixation by nitrogenase: the next stage, Chem. Rev., 2014, 114, 4041–4062 CrossRef CAS PubMed.
  31. D. Lukoyanov, N. Khadka, Z.-Y. Yang, D. R. Dean, L. C. Seefeldt and B. M. Hoffman, Reductive elimination of H2 activates nitrogenase to reduce the N≡N triple bond: characterization of the E4(4H) Janus intermediate in wild-type enzyme, J. Am. Chem. Soc., 2016, 138, 10674–10683 CrossRef CAS.
  32. (a) M. D. Zott, P. Garrido-Barros and J. C. Peters, Electrocatalytic ammonia oxidation mediated by a polypyridyl iron catalyst, ACS Catal., 2019, 9, 10101–10108 CrossRef CAS; (b) M. D. Zott and J. C. Peters, Enhanced ammonia oxidation catalysis by a low-spin iron complex featuring cis coordination sites, J. Am. Chem. Soc., 2021, 143, 7612–7616 CrossRef CAS PubMed; (c) M. D. Zott and J. C. Peters, Improving Molecular Iron Ammonia Oxidation Electrocatalysts via Substituent Effects that Modulate Standard Potential and Stability, ACS Catal., 2023, 13, 14052–14057 CrossRef CAS PubMed.
  33. (a) C. Baleizão and H. Garcia, Chiral salen complexes: an overview to recoverable and reusable homogeneous and heterogeneous catalysts, Chem. Rev., 2006, 106, 3987–4043 CrossRef PubMed; (b) G. Olivo, O. Lanzalunga and S. Di Stefano, Non-heme imine-based iron complexes as catalysts for oxidative processes, Adv. Synth. Catal., 2016, 358, 843–863 CrossRef CAS; (c) W. Sun and Q. Sun, Bioinspired manganese and iron complexes for enantioselective oxidation reactions: ligand design, catalytic activity, and beyond, Acc. Chem. Res., 2019, 52, 2370–2381 CrossRef CAS PubMed; (d) B. Stanje, P. Traar, J. A. Schachner, F. Belaj and N. C. Mösch-Zanetti, Iron-catalyzed oxidation of benzylic alcohols to benzoic acids, Dalton Trans., 2018, 47, 6412–6420 RSC; (e) C. B. Provis-Evans, S. Lau, V. Krewald and R. L. Webster, Regioselective alkyne cyclotrimerization with an in situ-generated [Fe(II)H(salen)]·Bpin catalyst, ACS Catal., 2020, 10, 10157–10168 CrossRef CAS.
  34. M. Usman, S. Tabassum, F. Arjmand, R. A. Khan, M. S. Ali, H. A. Al-Lohedan, A. Alsalme, M. A. Farah, K. M. Al-Anazi and M. Ahmad, CuII–NaI heteronuclear complex as anticancer entity against human breast cancer cell lines: DNA binding, cleavage, and computational studies, Inorg. Chim. Acta, 2018, 479, 229–239 CrossRef CAS.
  35. V. K. Sivasubramanian, M. Ganesan, S. Rajagopal and R. Ramaraj, Photoinduced electron transfer reactions of an oxovanadium(IV) Schiff base complex with aromatic amines, J. Org. Chem., 2002, 67, 1506–1510 CrossRef CAS.
  36. (a) J. Du, G. Liu, F. Li, Y. Zhu and L. Sun, Iron–salen complex and Co2+ ion-derived cobalt–iron hydroxide/carbon nanohybrid as an efficient oxygen evolution electrocatalyst, Adv. Sci., 2019, 6, 1900117 CrossRef PubMed; (b) M. C. Biesinger, B. P. Payne, A. P. Grosvenor, L. W. M. Lau, A. R. Gerson and R. S. C. Smart, Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni, Appl. Surf. Sci., 2011, 257, 2717–2730 CrossRef CAS; (c) S. J. Gerber and E. Erasmus, Electronic effects of metal hexacyanoferrates: An XPS and FTIR study, Mater. Chem. Phys., 2018, 203, 73–81 CrossRef CAS; (d) A. P. Grosvenor, B. A. Kobe, M. C. Biesinger and N. S. McIntyre, Investigation of multiplet splitting of Fe 2p XPS spectra and bonding in iron compounds, Surf. Interface Anal., 2004, 36, 1564–1574 CrossRef CAS.
  37. T. Kurahashi, Y. Kobayashi, S. Nagatomo, T. Tosha, T. Kitagawa and H. Fujii, A functional model for intradiol-cleaving dioxygenases: iron–salen complexes, Inorg. Chem., 2005, 44, 8156–8166 CrossRef CAS.
  38. (a) Y.-X. Zhou, X.-F. Zheng, D. Han, H.-Y. Zhang, X.-Q. Shen, C.-Y. Niu, P.-K. Chen, H.-W. Hou and Y. Zhu, Syntheses and crystal structures of Fe(III) and Mn(III) complexes with salen ligands: [Fe(salen)Cl]·CH3CN and [Mn(salen)(SCN)(CH3OH)]·CH3CN, Synth. React. Inorg., Met.-Org., Nano-Met. Chem., 2006, 36, 693–699 CrossRef CAS; (b) E. D. Portakal, Y. Kaya, E. Demirayak, E. K. Yeldir, A. Erçağ and İ. Kaya, Synthesis, characterization, and investigation of some properties of the new symmetrical bisimine Ni(II), Zn(II), and Fe(III) complexes derived from the monoimine ligand, Appl. Organomet. Chem., 2021, 35, e6265 CrossRef; (c) M. Cozzolino, V. Leo, C. Tedesco, M. Mazzeo and M. Lamberti, Salen, salan and salalen iron(III) complexes as catalysts for CO2/epoxide reactions and ROP of cyclic esters, Dalton Trans., 2018, 47, 13229–13238 RSC; (d) H.-L. Shyu, H.-H. Wei, G.-H. Lee and Y. Wang, Structure, magnetic properties, and epoxidation activity of iron(III) salicylaldimine complexes, J. Chem. Soc., Dalton Trans., 2000, 915–918 RSC.
  39. H. Toda, K. Kuroki, R. Kanega, S. Kuriyama, K. Nakajima, Y. Himeda, K. Sakata and Y. Nishibayashi, Manganese-catalyzed ammonia oxidation into dinitrogen under chemical or electrochemical conditions, ChemPlusChem, 2021, 86, 1511–1516 CrossRef CAS.
  40. W. Yang, G. A. Filonenko and E. A. Pidko, Performance of homogeneous catalysts viewed in dynamics, Chem. Commun., 2023, 59, 1757–1768 RSC.
  41. R. H. Crabtree, G. Ertl, H. Knözinger, F. Schüth and J. Weitkamp, Deactivation in Homogeneous Transition Metal Catalysis: Causes, Avoidance, and Cure, Chem. Rev., 2015, 115, 127–150 CrossRef CAS PubMed.
  42. Kinetic study and reaction mechanisms in homogeneous catalysis, in Comprehensive Coordination Chemistry III, ed. A. S. Borovik, Elsevier, Oxford, 2021, vol. 10, pp. 249–278 Search PubMed.
  43. Y. Li, J.-Y. Chen, Q. Miao, X. Yu, L. Feng, R.-Z. Liao, S. Ye, C.-H. Tung and W. Wang, A Parent Iron Amido Complex in Catalysis of Ammonia Oxidation, J. Am. Chem. Soc., 2022, 144, 4365–4375 CrossRef CAS.
  44. L. Nurdin, Y. Yang, P. G. N. Neate, W. E. Piers, L. Maron, M. L. Neidig, J.-B. Lin and B. S. Gelfand, Activation of ammonia and hydrazine by electron rich Fe (ii) complexes supported by a dianionic pentadentate ligand platform through a common terminal Fe (iii) amido intermediate, Chem. Sci., 2021, 12, 2231–2241 RSC.
  45. CCDC 2384217: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2l0z7v.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.