Investigating the formation of metal nitride complexes employing a tetradentate bis-carbene bis-phenolate ligand

Romain Kunert ab, Diego Martelino a, Samyadeb Mahato a, Nicholas M. Hein a, Jason Pulfer a, Christian Philouze b, Olivier Jarjayes *b, Fabrice Thomas *b and Tim Storr *a
aDepartment of Chemistry, Simon Fraser University, Burnaby, British Columbia V5A 1S6, Canada. E-mail: tim_storr@sfu.ca
bUniv. Grenoble Alpes, CNRS, DCM, F-38000, Grenoble, France. E-mail: Olivier.Jarjayes@univ-grenoble-alpes.fr; Fabrice.Thomas@univ-grenoble-alpes.fr

Received 17th June 2024 , Accepted 1st November 2024

First published on 8th November 2024


Abstract

The synthesis of MnV and CrV nitride complexes of a pro-radical tetradentate bis-phenol bis-N-heterocyclic carbene ligand H2LC2O2 was investigated. Employing either azide photolysis of the MnIII precursor complex MnLC2O2(N3) or a nitride exchange reaction between MnLC2O2(Br) and the nitride exchange reagent Mnsalen(N) failed to provide a useful route to the target nitride MnLC2O2(N). Experimental results support initial formation of the target nitride MnLC2O2(N), however, the nitride rapidly inserts into a Mn–CNHC bond. A second insertion reaction results in the isolation of the doubly inserted ligand product [H2LC2O2(N)]+ in good yield. In contrast, the Cr analogue CrLC2O2(N) was readily prepared and characterized by a number of experimental methods, including X-ray crystallography. Theoretical calculations predict a lower transition state energy for nitride insertion into the M–CNHC bond for Mn in comparison to Cr, and in addition the N-inserted product is stabilized for Mn while destabilized for Cr. Natural bond order (NBO) analysis predicts that the major bonding interaction (π M[triple bond, length as m-dash]N → σ* M–CNHC) promotes nucleophilic attack of the nitride on the carbene as the major reaction pathway. Finally, one-electron oxidation of CrLC2O2(N) affords a relatively stable cation that is characterized by experimental and theoretical analysis to be a metal-oxidized d0 CrVI species.


1. Introduction

High valent metal nitrido (N3−) species are of significant interest as proposed intermediates in biological and industrial processes,1–3 as well as N-atom transfer chemistry.4–13 A number of bioinspired Fe nitride complexes have been studied in the context of nitrogen fixation,14–20 and recent advances understanding synthetic FeS clusters have provided important information on the identity of activated Fe–N species.21–25 The stability and reactivity of discrete transition metal nitrides depends on factors such as metal identity and oxidation state, ligand identity, and overall geometry.26–28 Early transition metal nitrides are generally stabilized,29,30 and exhibit nucleophilic reactivity from either filled M[triple bond, length as m-dash]N π orbitals or the nitride lone pair.27,31–33 In contrast, late transition metal nitrides can be either isolated or described as transient intermediates, exhibiting electrophilic reactivity via M[triple bond, length as m-dash]N π* orbitals.28,34–42 Interestingly, a number of metal nitride complexes display ambiphilic reactivity,43–45 and subtle changes to the coordination environment can result in tunable reactivity at the nitride.33,35,46–49

We have reported that oxidation of nitrido metal salen complexes (see Scheme 1 for example salen H2SalR) results in an electronic structure that can be modulated via alteration of the electron-donating ability of para-R phenolate substituents, without changing the geometry at the metal center.33,50–52 Salen ligands are redox-active and have been documented to undergo oxidation or reduction at the ligand in place of the metal.53–56 In certain exceptional cases both metal and ligand oxidation can occur.57–59 We, and others, have also reported that the formation of a MnVI salen nitride results in rapid homocoupling of the nitride to form N2,52,60,61 however formation of a ligand radical (employing strongly electron-donating para-R substituents such as R = NMe2) precludes this homocoupling.52 Recent work with Ru62 and Mo63,64 nitrides has explored the homocoupling pathway and documented the formation of μ-nitrido products. In addition to homocoupling, Mn salen nitrides have been investigated in the context of the hydrogen atom bond dissociation free energy of the associated imido complexes,65 generation of ammonia by proton-coupled electron transfer,66–68 and as a catalyst for ammonia oxidation.69


image file: d4dt01765j-s1.tif
Scheme 1 Representative Salen (H2SalR) and N-heterocyclic carbene arylphenoxide ligands (HLCO, H2LOCO, H2LC2O2) and target metal nitride complexes of a tetradentate bis-carbene bis-phenolate ligand MLC2O2(N) M = Mn, Cr.

In the current work we investigate the formation and stability of Mn and Cr nitride complexes using a tetradentate bis-phenol bis-N-heterocyclic carbene ligand H2LC2O2,70 and compare to our recently published work using salen ligands (Scheme 1).33,50–52 The strong σ-donating capability (and potential for π-backbonding) of NHC ligands results in facile binding to electron-rich metals for catalytic applications.71,72 A number of high oxidation state metal complexes employing NHC ligands have also been characterized, including those of VV,73 CoIV,38,74 CuIII,75 NiIII,70,76,77 MnIV/V,78 FeIV–VII,14,16,19,20 MoVI,79 and NbV.80 Of particular relevance to the work herein is the use of tripodal carbene ligand frameworks by Smith and Meyer to stabilize high oxidation state Mn, Fe, and Co nitrides.14,16,19,20,78 In certain cases the nitride ligand has been shown to insert into the M–CNHC bond to form an imine.16,20,38

Redox-active NHC ligands have been described,81–88 and the presence of arylphenoxide units results in a significant lowering of the oxidation potential of the ligand. Indeed, a series of NiII ligand radical complexes of the HLCO, H2LOCO, and H2LC2O2 (Scheme 1) ligands have been characterized in the solid state.70,76,77 The different bonding properties of the imine and NHC donors in the salen and H2LC2O2 ligands (Scheme 1) juxtaposed with the similar metal binding geometry and capacity for ligand-based oxidation presents an interesting opportunity to compare the formation and stability of the respective Cr and Mn nitride complexes.

2. Results and discussion

2.1. Synthesis of Mn precursors

Reaction of the bromide salt of the tetradentate bis-carbene bis-phenolate ligand H2LC2O2·2HBr with Mn(OAc)2 and NEt3 in CH3CN solvent afforded the MnIII complex MnLC2O2(Br) (Scheme 2). The synthesis of MnLC2O2(Br) using CH3OH as the solvent was recently reported,89 and a MnIII complex of the H2LOCO ligand (Scheme 1) was previously studied by Bellemin-Laponnaz and co-workers.73 The complex was characterized by ESI-MS and EA. Dark brown crystals suitable for X-ray crystallography were obtained by slow evaporation of a concentrated CH2Cl2 solution. The solid-state structure of MnLC2O2(Br) was recently reported,89 however, certain metrical parameters differ due to the crystallization solvent. Thus, we will briefly describe the metrical parameters (Fig. S1 and Table S1) in order to compare with the other solid-state structures herein. The Mn ion lies in a distorted square pyramidal geometry with the Mn center coordinated by two phenolate oxygen atoms (O1/O2) and two carbons of the NHC units (C7/C17) in equatorial positions. The anionic bromide ligand occupies the apical position, with the Mn atom 0.417 Å above the C2O2 ligand plane. The angle between the NHC and phenolate rings are 25° and 28° in the structure, and in addition the ligand backbone adopts an umbrella shape with an angle between the two NHC units of 32°. This is in contrast to the neutral NiII complex of the same ligand reported by us, in which the angle between the two NHC units is only 3°, the angle between the NHC and phenolate rings are ∼9° and the complex is essentially flat.70 In addition, the Ni–CNHC bonds are significantly shorter (1.843 Å) in comparison to the Mn–CNHC bonds in MnLC2O2(Br) (2.03 Å). The ligand structural changes demonstrate significant binding flexibility, thereby accommodating different metal cations and different geometries.
image file: d4dt01765j-s2.tif
Scheme 2 Synthesis of MnIII precursor complexes MnLC2O2(Br) and MnLC2O2(N3). Conditions: (X = Br) Mn(OAc)2, NEt3, 95% yield; (X = N3) MnLC2O2(Br), NaN3, 93% yield.

The azido complex MnLC2O2(N3) was synthesized from a concentrated CH3CN solution of MnLC2O2(Br) in the presence of 1.5 equiv. of NaN3 (Scheme 2). Solvent removal, addition of CH2Cl2 and filtration removed the excess NaN3 affording MnLC2O2(N3) as a brown solid. IR (ν(N3) = 2040 cm−1), ESI-MS, and EA analysis confirmed product formation. Red/brown crystals suitable for X-ray diffraction were obtained via slow evaporation of a concentrated CH3CN solution of MnLC2O2(N3). The solid-state structure of MnLC2O2(N3) is presented in Fig. 1 (see ESI, Table S1 for selected crystallographic data). The distorted square pyramidal structure is similar to MnLC2O2(Br), with the azide ligand in the apical position, and an angle between the two NHC units of 33°. When dissolved in CH2Cl2 both MnLC2O2(Br) and MnLC2O2(N3) display a brownish color and broad absorptions in their UV-vis-NIR spectra (Table 1 and Fig. S2).


image file: d4dt01765j-f1.tif
Fig. 1 POV-ray representation of MnLC2O2(N3). Thermal ellipsoids shown at 50% probability level. Hydrogen atoms were omitted for clarity. Mn, pink; C, gray; O, red; N, blue. Select interatomic distances [Å]: Mn(1)–O(1): 1.889(2), Mn(1)–O(2): 1.885(2), Mn(1)–C(7): 2.026(2), Mn(1)–C(17): 2.022(2), Mn(1)–N(1): 2.110(4).
Table 1 UV-vis-NIR data for the indicated complexesa
Complex λ max [cm−1] (ε [M−1 cm−1])
a In CH2Cl2 solution. sh: shoulder. b 250 K.
MnLC2O2(Br) 34[thin space (1/6-em)]200 (11[thin space (1/6-em)]660, sh), 29[thin space (1/6-em)]200 (7550), 22[thin space (1/6-em)]600 (2510, sh)
MnLC2O2(N3) 28[thin space (1/6-em)]400 (3830, sh), 21[thin space (1/6-em)]900 (1270, sh)
CrLC2O2(N) 30[thin space (1/6-em)]770 (6200), 28[thin space (1/6-em)]000 (270)
[CrLC2O2(N)]+ 26[thin space (1/6-em)]500 (7900, sh), 11[thin space (1/6-em)]200 (2800)


2.2. Attempted synthesis of Mn nitride complex MnLC2O2(N)

With the two MnIII complexes in hand, we next investigated two different synthetic methods to prepare the same Mn nitride product MnLC2O2(N) (Scheme 1). Due to the susceptibility of nucleophilic attack at the methylene position between the two NHC rings of the ligand, we did not attempt to use NH4OH/bleach as described by Carreira and co-workers.90 Initially, we employed azide photolysis of MnLC2O2(N3) as this is a common method to produce nitrides (Scheme 3). Using the same protocol as one we had successfully used for Mn salen variants (λ = 312 nm, benzene solvent)52 resulted in solution color changes, however, the Mn nitride complex MnLC2O2(N) could not be isolated. Evans method analysis of the crude reaction mixture afforded a magnetic susceptibility of μeff ∼ 4 (or ∼3 unpaired electrons) consistent with the presence of a paramagnetic Mn species, and an absence of the expected diamagnetic d2 MnV nitride complex. While ESI-MS analysis of the crude reaction mixture showed the expected MnLC2O2(N) peak at 624.3 m/z, attempted purification via chromatography and/or recrystallization was unsuccessful (vide infra). Irradiation at lower energy (λ > 350 nm), or low temperature (λ = 312 nm, CH3CN, 235 K), resulted in no reaction.
image file: d4dt01765j-s3.tif
Scheme 3 Attempted synthesis of MnLC2O2(N) from either MnLC2O2(N3) (λ = 312 nm, benzene solvent) or MnLC2O2(Br) with a nitride exchange reagent Mnsalen(N) resulted in a mixture of products including the nitride inserted ligand [H2LC2O2(N)]+ (ESI-MS = 570.38 m/z).

We thus turned to a nitride exchange reaction, using a MnV salen nitride (Mnsalen(N)) complex as the nitride exchange reagent, which has been used to synthesize a variety of metal nitrides,91–93 including nitride exchange to other MnIII complexes (Scheme 3).94 Mixing of 1 equiv. of the nitride exchange reagent with MnLC2O2(Br) in CH3CN resulted in the formation of a brown precipitate identified as MnIIIsalen(Br) (ESI-MS, 321.04 m/z) which was removed via filtration. ESI-MS analysis of the filtrate exhibited the expected MnLC2O2(N) peak at 624.3 m/z (Fig. S3), however, we were again unable to isolate the desired nitride complex via chromatography and/or recrystallization. Attempting the nitride insertion reaction at low temperature (198 K, CH2Cl2) afforded the same result. Intriguingly, in these reactions a peak in the ESI-MS at 570.38 m/z indicated N-insertion into the ligand framework (Fig. S3), pointing to a possible decomposition pathway for the initially formed MnLC2O2(N) complex. Nitrido ligands have been reported to undergo intramolecular N-insertion into the ligand backbone of Fe,16,20 Co,38,95 U,96 and Ni41 complexes. In the majority of cases, population of antibonding π* M[triple bond, length as m-dash]N orbitals destabilizes the nitride, and thus intramolecular nitride insertion was unexpected for the putative d2 MnV complex in this work. Meyer in a series of elegant studies has demonstrated intramolecular nitride insertion at a carbenic center in a series of Co38 and Fe16,20 complexes employing tripodal carbene-based ligands. Interestingly, the analogous MnV nitride complexes did not undergo the same insertion chemistry.78

2.3 Isolation of nitride insertion product

To further investigate the potential nitride insertion product, we solubilized the crude reaction mixture from a representative nitride exchange reaction in diethyl ether and washed with aqueous 10% HCl and then EDTA. Analysis by ESI-MS showed the loss of the Mn complex peak at 624.3 m/z (Fig. S4). Further washing of a CH3CN solution with hexane, and recrystallization by slow evaporation of CH3CN/CH2Cl2 solution afforded pure ligand insertion product in 47% isolated yield. The same product can be isolated from the azide photolysis reaction, however in lower (∼15%) yield. 1H NMR of the product (Fig. S5) showed the compound to be symmetric in solution (Fig. 2A), with the expected mass of 570.38 m/z (Fig. S6). Further NMR analysis (13C, 13C DEPT and 1H–13C gHSQC NMR) provides confirmation of the structure in solution (Fig. S7–S9). Interestingly, 1H NMR in CDCl3 results in broadened resonances indicative of hindered rotation of the phenol due to dynamics in the fast-exchange regime (Fig. S10). Analysis of the product by X-ray crystallography showed that the nitride insertion product is a cation [H2LC2O2(N)]+ with chloride as the counterion (Fig. 2B). In addition, nitride insertion into both carbenic carbons has occurred, resulting in three fused rings including two imidazoles and a central triazone core. The nitride nitrogen (N1) exhibits similar short bond lengths to both carbenic carbons (N1–C7 and N1–C17: 1.335 and 1.334 Å respectively) and the three fused rings are essentially flat, with an 11° angle between the planes of the two imidazole rings. These structural features indicate the presence of a conjugated π system located across the N2–C7–N1–C17–N5 atoms (Fig. 2C). The hydrogen bond network involving the two phenols, a lattice water, and the chloride counterion ensures the secondary structure of the crystal (Fig. S11). Interestingly, Bullock and co-workers have reported that H-atom abstraction from a bound NH3 in [Mn(dmpe)2(CO)(NH3)]+ (dmpe = 1,2-bis(dimethylphosphino)ethane) results in double insertion of the N-atom into the diphosphine ligand to form a cyclophosphazenium cation,97 with theoretical calculations supporting insertion of an NHx species rather than the nitride herein.
image file: d4dt01765j-f2.tif
Fig. 2 (A) Structure of the nitride insertion product and charge delocalization. (B) POV-ray representation of [H2LC2O2(N)]Cl. Thermal ellipsoids shown at 50% probability level. Hydrogen atoms were omitted for clarity. (C) Selected bond lengths demonstrating the symmetric nature (and delocalization) of the product.

15N isotopic labeling was employed to further investigate the N-atom insertion into the carbene ligand backbone. We used the 15N-labeled nitride exchange reagent Mnsalen(15N) in a reaction with MnLC2O2(Br) and subsequent ESI-MS analysis showed a one Dalton m/z shift for the two ions at 571.4 (100%) and 625.3 (20%) (Fig. S12) in comparison to the unlabeled reaction (Fig. S3). The 15N-inserted product [H2LC2O2(15N)]Cl was isolated in a similar manner to the unlabeled analogue in a comparative yield (56%). To further confirm that the N-insertion is a result of intramolecular N-insertion from a transient MnLC2O2(N) nitride complex, and not for example from an intermolecular process, we synthesized the 15N-labeled azide complex MnLC2O2(14/15N3) using 50% terminally 15N-labeled NaN3. Photolysis of MnLC2O2(14/15N3) (λ = 312 nm) afforded a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture of N-inserted ligand products [H2LC2O2(15N)]Cl and [H2LC2O2(14N)]Cl by ESI-MS as well as an additional peak at 573.4 m/z likely corresponding to an O-inserted ligand product (Fig. S13). The O-inserted product was not investigated further in this work. Overall, the two different isotopic labeling experiments provide further evidence that intramolecular N-atom insertion from an unstable MnLC2O2(N) nitride complex results in formation of the [H2LC2O2(N)]Cl product.

2.4. Synthesis of the Cr nitride complex

Based on our previous work exploring the electronic structure and reactivity of the analogous Mn and Cr nitride complexes with salen ancillary ligands,33,51,52 and the increased stability of the Cr derivatives more generally,33,52 we investigated the preparation of the Cr nitride complex CrLC2O2(N). Following our previous work with salen ligands,33 and the general preparation outlined by Bendix,92 we added a solution of the H2LC2O2·2HBr ligand to an easily complexed [Cr[triple bond, length as m-dash]N]2+ synthon with labile auxillary ligands (CH3CN and Cl) (Scheme 4).
image file: d4dt01765j-s4.tif
Scheme 4 Synthesis of the Cr nitride complex CrLC2O2(N) using the in situ generated [Cr[triple bond, length as m-dash]N]2+ synthon. See experimental section for details.

A bright orange powder was isolated and confirmed to be the CrLC2O2(N) complex based on IR (ν(N[triple bond, length as m-dash]Cr) = 1025 cm−1), ESI-MS (621.32 m/z, [CrL + H]+, 100%; Fig. S14), and X-ray crystallography. The UV-vis absorption spectrum of CrLC2O2(N) displayed typical features of a d1 CrV species in a square pyramidal geometry, including an intense LMCT transition at 30[thin space (1/6-em)]770 cm−1, and a weak transition (ε < 300 M−1 cm−1) at 20[thin space (1/6-em)]800 cm−1 corresponding to a transition from the non-bonding d1xy into empty dxz or dyz π* orbitals (Fig. S15). In addition, frozen solution X-band EPR measurements displayed an axial pattern, with simulation parameters (g = 1.990, A53Cr = 49; g|| = 1.949, A53Cr = 125) indicative of a d1 CrV complex (Fig. 3).98


image file: d4dt01765j-f3.tif
Fig. 3 Frozen solution EPR spectra of concentration-matched CrLC2O2(N) (black) and oxidized [CrLC2O2(N)]+ (red). Grey line indicates simulation of the experimental data for the neutral complex. Inset is a magnification of the oxidized complex signal. Conditions: 0.5 mM complex in CH2Cl2; 0.1 M TBAP; freq. = 9.64 GHz; power = 2.0 mW; T = 11 K.

Single crystals suitable for X-ray diffraction were obtained by slow evaporation of a concentrated CH2Cl2/CH3CN solution of CrLC2O2(N) (Fig. 4 and Table S2). The Cr ion is in a distorted square pyramidal environment with the Cr center coordinated to two phenolate oxygens (O1/O2), two NHC carbons (C7/C17), and the nitride ligand (N1). The Cr–nitride bond length of 1.558 Å is similar to that reported for other CrV nitrides.33,99 The Cr center is shifted out of the ligand C2O2 plane by ca. 0.56 Å, which is slightly more than that reported for the salen analogues.33,99 Similarly to the X-ray structures for both MnLC2O2(Br) and MnLC2O2(N3) the ligand backbone for CrLC2O2(N) adopts an umbrella shape with an angle between the two NHC units of 25°. Based on the marked difference in stability of the MnLC2O2(N) and CrLC2O2(N) complexes investigated herein we turned to theoretical calculations to further investigate their predicted electronic structure and reactivity differences.


image file: d4dt01765j-f4.tif
Fig. 4 POV-ray representation of CrLC2O2(N). Thermal ellipsoids shown at the 50% probability level. Hydrogen atoms were omitted for clarity. Cr, pink; C, gray; O, red; N, blue. Select interatomic distances [Å]: Cr(1)–O(1): 1.935(1), Cr(1)–O(2): 1.919(1), Cr(1)–C(7): 2.058(1), Cr(1)–C(17): 2.062(1), Cr(1)–N(1): 1.558(1).

2.5. Theoretical calculations on CrLC2O2(N) and MnLC2O2(N)

The predicted metrical parameters for CrLC2O2(N) were determined to be within ±0.03 Å of the experimental X-ray data (Table 2). Similarly to the X-ray structure, a square pyramidal shape is predicted with the Cr atom out of plane towards the nitride (Fig. S16). The Cr[triple bond, length as m-dash]N bond length is predicted to be ca. 0.02 Å shorter in comparison to the X-ray data. Overall, the predicted structure for MnLC2O2(N) (Fig. S17) is similar to the Cr analogue, however, certain key differences are noted. For example, the computed Mn[triple bond, length as m-dash]N bond length is ca. 0.03 Å shorter in comparison to Cr[triple bond, length as m-dash]N, however the Mayer bond order for the Mn derivative is slightly lower (2.72 vs. 2.79). In addition, the Mn center is closer to the C2O2 plane, by ca. 0.05 Å (Table 2). By moving closer to the C2O2 plane the Mn center interacts more strongly with the carbene ligands, and the predicted M–C(7/17) carbene bond lengths are ca. 0.08 Å shorter for the Mn complex in comparison to Cr. The shorter metal–carbene bond lengths in MnLC2O2(N) likely facilitates the observed nitride insertion reactivity.
Table 2 X-ray metrical parameters (and predicted values in brackets) for MLC2O2(N) in Åa
Complex CrLC2O2(N) MnLC2O2(N)
a Opt: B3LYP-D3, 6-31g*, PCM(CH2Cl2). b Distance from M to O1–O2–C7–C17 plane.
M–O1 1.937 (1.921) (1.924)
M–O2 1.919 (1.918) (1.925)
M–C7 2.058 (2.063) (1.979)
M–C17 2.065 (2.048) (1.974)
M–N1 1.558 (1.536) (1.505)
M out of planeb 0.556 (0.543) (0.510)


We compared the relative stability of MnLC2O2(N) and CrLC2O2(N) by investigating the initial insertion of the nitride into one of the M–CNHC bonds. We considered this as the first step in the reaction pathway to eventually form the doubly inserted product [H2LC2O2(N)]+ (Scheme 3). The reaction pathway is shown in Fig. 5 with the triplet transition state (3TS) calculated to be 24.17 kcal mol−1 higher in energy than the singlet MnLC2O2(N) starting material. Note that by analyzing the potential energy surface (PES) for the reaction (Fig. S18), it is apparent that the triplet energy surface (unpaired electrons in d1xy and d1xz/yz rather than d2xy) is slightly lower in energy at the TS, with the 1TS slightly higher in energy at 26.12 kcal mol−1. Further analysis of the relative energies of the 1TS versus the 3TS using different DFT functionals/basis sets also predicts the 3TS to be of lower energy (Table S3), however such comparisons should be viewed with caution.100 The nitride N is bridging the Mn–CNHC bond at the TS (Fig. 5) supporting a N-migratory insertion pathway. The quintet nitride inserted MnIII imido reaction product MnLCNO2 is stabilized by 9.72 kcal mol−1 relative to the MnLC2O2(N) starting material (Fig. 5). In contrast, for the doublet CrLC2O2(N) complex the 2TS is calculated to be significantly higher at 35.22 kcal mol−1, with the resulting quartet nitride inserted CrIII imido reaction product CrLCNO2 de-stabilized by 3.03 kcal mol−1 relative to the CrLC2O2(N) starting material (Fig. 5). Spin crossover to the quartet manifold is calculated to occur after the 2TS (Fig. S18). These calculations highlight the greater susceptibility of MnLC2O2(N) to nitride insertion relative to CrLC2O2(N), as observed experimentally. We note that Meyer et al. isolated a protonated imine reaction product in their studies,16,38,101 however, we did not further evaluate this possibility as we have no experimental verification for either the imido or imine species herein. Further, we did not investigate the second N-carbene insertion to form the [H2LC2O2(N)]+ product due to the uncertainty in the identity of the final Mn product(s).


image file: d4dt01765j-f5.tif
Fig. 5 Predicted reaction profile for nitride insertion from MLC2O2(N) into the M–CNHC bond to form the inserted product MLCNO2 (M = Mn (blue), Cr (red)). In the 3TS for MnLC2O2(N) (top left) the nitride N is bridging the Mn–CNHC bond. Note the 1TS is slightly higher in energy and spin crossover is predicted to occur before the TS (see Fig. S18 and Table S3). The π M[triple bond, length as m-dash]N → σ* M–CNHC orbital interaction approaching the 3TS is shown (top right). Predicted mono-inserted product MnLCNO2 bottom right. See Experimental section for calculation details.

We next investigated the orbitals involved in the nitride insertion reaction to gain further insight. We analyzed the stabilization energy (E(2)) in the equilibrium structures of MLC2O2(N) (M = Mn, Cr) between different donating (filled) and accepting (empty) orbitals associated with the M[triple bond, length as m-dash]N and M–CNHC bonds using second-order perturbation theory (SOPT) from natural bond order (NBO) calculations.102,103 This analysis predicts that while a donor–acceptor interaction occurs from filled NHC orbitals to the π* M[triple bond, length as m-dash]N orbitals, the most significant interaction is in the opposite direction, and in particular π M[triple bond, length as m-dash]N → σ* M–CNHC (Table 3 and Fig. 5). It is interesting to note that the predicted donor–acceptor interactions are significantly increased for the Mn derivative in comparison to Cr (Table 3), which is in line with the shorter predicted M–CNHC bond distance for the Mn derivative, and the observed reactivity. Further, we investigated the change in predicted orbital occupancy via NBO analysis for MnLC2O2(N) as the N–Mn–CNHC bond angle is decreased from the equilibrium structure (ca. 90°) to the TS (ca. 60°) along the reaction coordinate, and in line with the SOPT analysis, the most significant decrease occurs for the π M[triple bond, length as m-dash]N orbitals, with a concomitant increase in occupancy for both σ* Mn–CNHC and π* NHC (Fig. S19). Overall, these calculations support the observed nitride insertion reactivity for the MnLC2O2(N) complex, and lack thereof for CrLC2O2(N), and that the reactivity is primarily driven by nitride donation into σ*/π* orbitals associated with the NHC ligand.

Table 3 Second-order perturbation theory (SOPT) analysis of orbital interactions relevant to nitride insertion for MLC2O2(N) (M = Mn, Cr)a
Reactivity Donor → acceptor interaction E (2) (kcal mol−1)
MnLC2O2(N) CrLC2O2(N)
a From NBO calculations (BP86-GD3/TZVP/PCM(CH2Cl2)) b Total of two π M[triple bond, length as m-dash]N orbitals.
Nitride → NHC π M[triple bond, length as m-dash]Nb → σ* M–CNHC 21.4 9.6
π M[triple bond, length as m-dash]Nb → π* NHC 6.0
NHC → Nitride σ M–CNHC → π* M[triple bond, length as m-dash]N 9.3 3.7


2.6. Oxidation of Cr nitride complex

Upon isolation and characterization of the neutral nitride complex CrLC2O2(N) we next investigated its oxidation chemistry to better understand the stability and electronic structure of the one-electron oxidized form. Cyclic voltammetry experiments exhibited a quasi-reversible one-electron redox process at 0.14 V vs. Fc+/Fc (Fig. 6), with no other redox processes observed in the electrochemical window (Fig. S20). Based on previous work, additional redox processes would be expected in the case of ligand-based oxidation.33,70 Interestingly, this redox potential is between that reported for CrNSalOiPr (E1/2 = 0.38 V vs. Fc+/Fc) and CrNSalNMe2 (E1/2 = −0.04 V vs. Fc+/Fc) with the former characterized as a CrVI species and the latter as a CrV ligand radical, upon oxidation.33
image file: d4dt01765j-f6.tif
Fig. 6 Cyclic voltammogram of CrLC2O2(N) showing a quasi-reversible redox process at 0.14 V vs. Fc+/Fc. Peak to peak difference (EpaEpc = 0.093 V). EpaEpc = 0.070 V for Fc+/Fc. Conditions: 0.5 mM complex, CH2Cl2, 0.1 M tetrabutylammonium perchlorate (TBAP), scan rate = 100 mV s−1, 298 K.

Based on the quasi-reversible oxidation process, and the relatively low redox potential, we further investigated the bulk oxidation of CrLC2O2(N) to [CrLC2O2(N)]+ with tris(2,4-dibromophenyl)aminium hexafluoroantiminoate ([N(C6H3Br2)3+[SbF6]; E1/2 = 1.1 V vs. Fc+/Fc).104 Upon sequential addition of a solution of the oxidant to CrLC2O2(N) at 253 K in CH2Cl2 a new band is observed at 26[thin space (1/6-em)]500 cm−1 along with a lower energy transition at 11[thin space (1/6-em)]200 cm−1 (2800 M−1 cm−1) (Fig. 7). The oxidized complex is stable for at least one hr at 253 K (Fig. S21). The UV-vis-NIR spectrum of [CrLC2O2(N)]+ is similar to that reported for the CrVI nitride complexes with salen ancillary ligands, in which a comparable low energy feature was assigned as a ligand-to-metal charge transfer (LMCT) band.33 It is unlikely that the 11[thin space (1/6-em)]200 cm−1 transition is a ligand radical intervalence charge transfer (IVCT) band as the single one-electron redox feature discounts a Class II regime, and analysis of the band properties (Δν1/2 = 6500 cm−1, ε = 2800 M−1 cm−1) are not in line with a Class III ligand radical (Δν1/2 ≤ 2000 cm−1, ε > 5000 M−1 cm−1).105


image file: d4dt01765j-f7.tif
Fig. 7 Chemical oxidation of CrLC2O2(N) using [N(C6H3Br2)3+[SbF6] monitored by UV-vis-NIR spectroscopy. Black: neutral; red: oxidized. Intermediate grey lines represent increasing aliquots of oxidant added until one equiv. was reached. Conditions: 0.45 mM complex, CH2Cl2, 253 K. Blue bar represents most intense TD-DFT predicted low energy transition at 12[thin space (1/6-em)]098 cm−1, and donor/acceptor orbitals indicating LMCT character for the CrVI electronic structure. See Experimental section for calculation details.

EPR analysis of [CrLC2O2(N)]+ shows ca. 5% signal intensity in comparison to a concentration-matched sample of CrLC2O2(N), with the remaining signal mostly due to unoxidized CrLC2O2(N) (Fig. 3). Loss of the EPR signal can be attributed to (1) formation of a d0 CrVI complex, (2) an antiferromagnetically coupled CrV ligand radical species (S = 0, open-shell singlet), or (3) a ferromagnetically coupled CrV ligand radical species (S = 1, triplet) exhibiting large zero-field splitting.106,107 We further investigated the predicted change in metrical parameters upon oxidation and relative energy of the different plausible electronic structures using theoretical calculations. The metal oxidized CrVI species was predicted to be lowest in energy, with the antiferromagnetically coupled CrV ligand radical species (broken-symmetry, BS) higher in energy by +6.03 kcal mol−1, and the ferromagnetically coupled CrV ligand radical species (triplet, T) highest in energy by +13.95 kcal mol−1 (Table 4). The ligand radical is predicted to be localized to form a phenoxyl/phenolate species for the BS and T solutions (Fig. S22), with the expected difference in Cr–O1/Cr–O2 bond lengths (Table 5). We further investigated the predicted low energy transitions for the three possible electronic structures based on the experimentally observed UV-vis-NIR feature at 11[thin space (1/6-em)]200 cm−1 (Fig. 7). Both ligand radical electronic structures (BS, T) predict low energy transitions in the NIR (Table 4), and no transitions of appreciable intensity are observed experimentally in this energy region (Fig. 7). However, the CrVI singlet solution correctly predicts a ligand to metal charge transfer (LMCT) transition at 12[thin space (1/6-em)]098 cm−1 (Fig. 7 and Table 4), providing further support that this is the correct electronic structure of [CrLC2O2(N)]+.

Table 4 Predicted relative energies for possible electronic structures of [CrLC2O2(N)]+ and associated time-dependent density functional theory (TD-DFT) predicted low energy transitionsa
Electronic structure Relative energy (kcal mol−1) Predicted low energy transition
a Single point: BP86-GD3/TZVP/PCM(CH2Cl2).
[CrLC2O2(N)]+ (S) 0.00 HOMO/HOMO−1 → LUMO (12[thin space (1/6-em)]098 cm−1, f = 0.0870)
[CrLC2O2(N)]+ (BS) +6.03 HOMO/HOMO−1 → LUMO (4727 cm−1, f = 0.0521)
[CrLC2O2(N)]+ (T) +13.95 HOMO/HOMO−1 → LUMO (3957 cm−1, f = 0.0949)


Table 5 Predicted metrical parameters for CrLC2O2(N) and [CrLC2O2(N)]+ in Åa
Complex CrLC2O2(N) [CrLC2O2(N)]+ (S) [CrLC2O2(N)]+ (T) [CrLC2O2(N)]+ (BS)
a Opt: B3LYP-D3, 6-31g*, PCM(CH2Cl2). b Distance from Cr to O1–O2–C7–C17 plane.
Cr–O1 (1.921) (1.819) (1.988) (1.960)
Cr–O2 (1.918) (1.802) (1.886) (1.876)
Cr–C7 (2.063) (2.044) (2.059) (2.062)
Cr–C17 (2.048) (2.072) (2.024) (2.028)
Cr–N1 (1.536) (1.516) (1.531) (1.532)
Cr out of planeb (0.543) (0.472) (0.505) (0.491)


3. Summary

In this work we studied the synthesis and properties of the Mn and Cr nitrides of a tetradentate bis-phenol bis-N-heterocyclic carbene ligand H2LC2O2. The tetradentate platform of H2LC2O2 exhibits similarities to the common salen analogue H2SalR (Scheme 1), however the different donating/accepting properties of the NHC units in comparison to imine functionalities of the salen framework provide for an interesting comparison. Indeed, while the MnV nitride salen complexes have been reported extensively,108,109 the MnV nitride MnLC2O2(N) could not be isolated herein. Our results suggest that MnLC2O2(N) forms, yet is unstable, and the nitride rapidly inserts into a Mn–CNHC bond. A second insertion reaction results in the isolation of the doubly inserted ligand product [H2LC2O2(N)]+. Interestingly, the Cr analogue CrLC2O2(N) can be readily prepared and does not exhibit nitride N-insertion reactivity. Theoretical investigations predict that the transition state for the first N-insertion is ca. 10 kcal mol−1 lower in energy for the Mn derivative, and the reactivity is primarily driven by nucleophilic attack of the terminal nitride on the carbene (π M[triple bond, length as m-dash]N → σ* M–CNHC). Theoretical calculations predict the Mn–CNHC bond lengths to be ∼0.08 Å shorter in comparison to the Cr–CNHC bonds for MLC2O2(N) (M = Mn, Cr) (Table 2), providing rationale for the facile N-insertion reactivity observed for MnLC2O2(N). In addition, the observed N-insertion reactivity is in accord with the general lower stability and more extensive reactivity of Mn nitrides in comparison to Cr nitrides.33,52,91–93 Finally, we investigated the one-electron oxidation of the CrLC2O2(N) complex, and determined that metal-based oxidation occurs to form the d0 CrVI product [CrVILC2O2(N)]+.

4. Experimental section

4.1. Materials and methods

All chemicals were of the highest quality grade and purified whenever necessary. The ligand H4LC2O2Br2 was synthesized following a literature procedure.70 The atom-transfer reagents Mnsalen(N) and CrCl3(THF)3 were also prepared according to published protocols.90,92 Dichloromethane and acetonitrile were dried by refluxing over calcium hydride and distilled prior to use. The tris(2,4-dibromophenyl)aminium hexafluoroantimonate radical oxidant [N(C5H3Br2)3][SbF6] was synthesized according to published protocols.110 Electronic spectra were obtained using a Cary 5000 spectrophotometer. Mass spectrometry (ESI positive mode) was performed on an Agilent 6210 TOF ESI-MS system. 1H nuclear magnetic resonance (NMR) spectroscopy as well as magnetic susceptibility via Evans Method were carried out on a Bruker AVANCE III 500 MHz instrument. Elemental analysis (C, H, N) were performed at Simon Fraser University on a Carlo Erba EA1110 CHN elemental analyser. All electron paramagnetic resonance (EPR) were recorded on a Bruker EMXplus spectrometer operating with a premium X-band microwave bridge and an HS resonator. EPR spectra were simulated using the EasySpin package in Matlab.111 Solid state infrared spectra (IR) were measured on a Thermo Nicolet Nexus 670 FT-IR spectrometer equipped with a Pike MIRacle attenuated total reflection (ATR) sampling accessory. Cyclic voltammetry (CV) was performed on a PAR-263A potentiometer equipped with a silver wire reference electrode, a platinum disk counter electrode and glassy carbon working electrode. Tetrabutylammonium perchlorate (0.1 M) was used as the supporting electrolyte, and decamethylferrocene was used as an internal standard in CH2Cl2.112 One-electron processes were confirmed via comparison of the electrochemical response of one equiv. of the Cr complex versus decamethylferrocene.

4.2. Synthesis

4.2.1 Synthesis of MnLC2O2(Br). This procedure differs slightly from that recently reported.89 The pro-ligand H4LC2O2Br2 (152 mg, 212 μmol) was suspended in a solution of manganese(II) acetate tetrahydrate (52.1 mg, 213 μmol) in dry acetonitrile (25 mL). Triethylamine (120 μL, 0.86 mmol) was added and the resulting mixture was stirred at 80 °C for 3 h. The solution gradually homogenized as it turned a brown color. It was then evaporated and the crude product was solubilized in toluene, then filtered over Celite. The filtrate was evaporated and dried under high vacuum overnight to afford MnLC2O2(Br) as a brown solid (147 mg, 95%). MS (HRMS): m/z = 609.29 [MnL − Br]+. Anal. calcd (%) C35H46MnBrN4O2·0.5H2O: C 60.17, H 6.78, N 8.02; found (%): C 60.19, H 7.25, N 7.68. Dark brown crystals suitable for X-Ray diffraction were obtained by slow evaporation of a concentrated dichloromethane solution of MnLC2O2(Br).
4.2.2 Synthesis of MnLC2O2(N3). MnLC2O2(Br) (135 mg, 196 μmol) was solubilized in a solution acetonitrile (8 mL) and H2O (1.5 mL). A solution of sodium azide 1 M in H2O (215 μL, 215 μmol) was added dropwise and the mixture was stirred at room temperature for 1 h. A large spatula of sodium sulfate was added and the volatiles were removed under vacuum. The crude product was solubilized in dichloromethane and filtered over Celite. The filtrate was evaporated to afford MnLC2O2(N3) as a brown solid (122 mg, 93%). IR: ν(N3) = 2040 cm−1. ESI-MS: m/z = 609.30 [MnL − N3]+. Anal. calcd (%) C35H46MnBrN7O2: C 64.50, H 7.11, N 15.04; found (%): C 64.11, H 7.23, N 14.75. Red/brown crystals suitable for X-Ray diffraction were obtained by slow evaporation of a concentrated acetonitrile solution of MnLC2O2(N3).
4.2.3 Synthesis of inserted ligand [H2LC2O2(N)]Cl. A solution of MnLC2O2(Br) (102.8 mg, 149 μmol) in dichloromethane (8 mL) was added dropwise to Mnsalen(N) (50.4 mg, 150 μmol) suspended in dichloromethane (5 mL). A brown precipitate formed instantly and the solution was stirred under air for 2 h. The precipitate was removed by filtration and the brown filtrate was evaporated. The crude mixture was solubilized in 20 mL of diethyl ether. The organic phase was washed three times with a 10% HCl solution, then three times by a saturated H2O solution of tetrasodium EDTA, followed by two times with H2O and one final time with brine. The organic phase was dried over Na2SO4 and evaporated. The resulting pale brown solid was solubilized in acetonitrile and washed 5 times with hexanes. Finally, the acetonitrile solution was evaporated and dried under high vacuum overnight, affording [H2LC2O2(N)]Cl as a beige powder (42.3 mg, 47%). 1H NMR (500 MHz, CD3OD) δ 1.28 (s, 18H, t-Bu), 1.34 (s, 18H, t-Bu), 6.30 (s, 2H, CH2), 7.11 (s, 2H, aryl–H), 7.29 (s, 2H, aryl–H), 7.32 (s, 2H, aryl–H), 7.41 (s, 2H, aryl–H). 13C (375 MHz) δ 31.04 (CH3), 32.64 (CH3), 36.14 (Cquat.), 37.27 (Cquat.), 61.65 (CH2), 115.70 (CH), 121.35 (CH), 123.37 (CH), 125.40 (Cquat.), 127.19 (CH), 141.91 (Cquat.), 145.18 (Cquat.), 148.26 (Cquat.), 149.96 (Cquat.). MS (HRMS): m/z = 570.38 [M − Cl]+.
4.2.4 Synthesis of inserted ligand [H2LC2O2(15N)]Cl. A solution of MnLC2O2(Br) (101.7 mg, 147 μmol) in dichloromethane (7 mL) was added dropwise to Mnsalen(15N) (50.0 mg, 149 μmol) suspended in dichloromethane (4 mL). A brown precipitate formed instantly and the solution was stirred under air for 3 h. The precipitate was removed by filtration and the filtrate was evaporated. The crude mixture was solubilized in diethyl ether. The organic phase was washed three times by a 10% HCl solution, then three times by a saturated H2O solution of tetrasodium EDTA, followed by two times with H2O and one final time with brine. The organic phase is dried over Na2SO4 and evaporated. The resulting pale brown solid was solubilized in acetonitrile and washed 5 times with hexanes. Finally, the acetonitrile solution was evaporated and dried under high vacuum overnight, affording [H2LC2O2(15N)]Cl as a beige powder (50.0 mg, 56%). MS (HRMS): m/z = 571.38 [M − Cl]+. Colorless crystals suitable for X-Ray diffraction were obtained by slow evaporation of a CH2Cl2/CH3CN (1[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v) solution of [H2LC2O2(15N)]Cl.
4.2.5 Synthesis of CrLC2O2(N). The procedure was adapted from a previous report by Bendix.92 A solution of CrCl3(THF)3 (219.1 mg, 585 μmol) was prepared in dry acetonitrile (5 mL) under N2 atmosphere in a glovebox. To this solution Mnsalen(N) was then added (193.4 mg, 577 μmol) while stirring. The purple solution immediately turned brown and a brown precipitate formed. The mixture was stirred under air at room temperature for 1 h then was filtered over Celite to afford a yellow-brown solution of [N[triple bond, length as m-dash]Cr]2+ complex. The solution containing the [N[triple bond, length as m-dash]Cr]2+ complex was then added dropwise to an ethanol solution (40 mL) of H4LC2O2Br2 (376.0 mg, 523 μmol) and Et3N (300 μL, 2.15 mmol) stirred at 60 °C. The resulting mixture was stirred under reflux under air overnight then evaporated. The orange brown crude solid was suspended in cold methanol, collected and washed with cold methanol to afford CrLC2O2(N) as a bright orange powder (127.6 mg, 39%). IR: ν(N[triple bond, length as m-dash]Cr) = 1025 cm−1. MS (HRMS): m/z = 621.32 [CrL + H]+. Orange crystals suitable for X-Ray diffraction were obtained by slow evaporation of a concentrated solution of CrLC2O2(N) in a dichloromethane/acetonitrile solvent system.
4.2.6 Synthesis of CrLC2O2(15N). Identical procedure to 2.5 except Mnsalen(15N) used. Yield 57.5 mg, 35%. IR: ν(15N[triple bond, length as m-dash]Cr) = 998 cm−1. MS (HRMS): m/z = 622.31 [CrL + H]+.

4.3. X-ray crystallography

Diffraction data for brown needle single crystals of MnLC2O2Br, red blade crystals of MnLC2O2(N3), and orange prism crystals of CrLC2O2(N) were collected at the University of British Columbia (Canada) by Dr Brian Patrick on a Bruker X8 Apex II diffractometer with graphite monochromated Mo-Kα radiation (λ = 0.71073 Å) at 296 K. Data were integrated using the Bruker SAINT software package (Bruker, V8.40B, 2016) and absorption corrections were performed using the multiscan technique SADABS-2016/2 (Bruker, 2016/2). The structures were solved by Dr Nicholas Hein and Jason Pulfer using the SHELX software113 implemented by Olex2.114 All non-hydrogen atoms were anisotropically refined and hydrogen atoms were placed at calculated positions and refined as riding atoms with isotropic displacement parameters. It is noted that for MnLC2O2(N3) a residual peak was modeled at 2.55 Å as a Br from the starting material. The refined occupancies are azide (0.967(3)) and Br (0.033(3)). Further data is shown in Tables S1 and S2. Diffraction data for the yellow prism single crystal of solvated [H2LC2O2(N)]Cl were collected at Université de Grenoble Alpes (France) by Dr Christian Philouze on a Bruker-AXS-Enraf-Nonius Kappa Apex II diffractometer with multilayer mirrors monochromated Mo-Kα radiation (λ = 0.71073 Å) from an Incoatec high brilliance micro-source at 210 K. Data were integrated using the Bruker EvalCCD software package and absorption corrections were performed using the multiscan technique SADABS-2004/1. The structure was solved by using the SHELX software113 implemented by Olex2.114 All non-hydrogen atoms were anisotropically refined and hydrogen atoms were placed at calculated positions and refined as riding atoms with isotropic displacement parameters. The structure displays a solvation site which may be modeled with two dichloromethane and one acetonitrile molecules spread over 3 different positions and with respective partial occupancy: 0.3968, 0.35167 and 0.25 Further data is shown in Table S1 and Fig. S11. The crystallographic data for the new structures have been deposited with Cambridge Crystallographic Data Centre as supplementary publication no. CCDC 2359452, 2359453, 2359455, and 2359456.

4.4. Theoretical calculations

Geometry optimizations were all performed using the Gaussian 16 program (Revision A.03)115 employing the B3LYP functional, the D3 dispersion correction116 in combination with the 6-31g* basis set. Frequency calculations performed on the same functional/basis set confirmed optimized structures were at a global minimum. Single-point calculations as well as time-dependant DFT calculations were performed using the BP86-GD3 functional and the TZVP basis set of Ahlrichs.117,118 Natural bond order (NBO) calculations were completed using the single-point calculations in the Gaussian 16 program.119 For reaction profile calculations single-point energies were converted to Gibbs free energies using corrections from the frequency calculations on the optimized coordinates. Calculated transition states were confirmed to have one negative frequency associated with N–C bond formation. All calculations employed a polarizable continuum model (PCM) for CH2Cl2 (ε = 8.93) for all atoms.120

Data availability

Compound characterization, 1H NMR and ESI-MS data, as well as X-ray crystallography details are available in the ESI. Calculation data is compiled in a zip folder.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by Natural Sciences and Engineering Research Council (NSERC) Discovery Grants (RGPIN-2019-06749 and RGPAS-2019-00054 to T. S.). Digital Alliance Canada is thanked for access to computational resources. D. M. acknowledges Mitacs Canada for a Globalink Fellowship (FR43327). Dr Brian Patrick is thanked for assistance with solving the X-ray structures. This work is also supported by the French National Research Agency in the framework of the “Investissements d'avenir” program (ANR-15-IDEX-02) and Labex ARCANE and CBH-EUR-GS (ANR-17-EURE-0003) for financial support. The NanoBio-ICMG platforms (FR 2607) are acknowledged for their support.

References

  1. D. V. Yandulov and R. R. Schrock, Catalytic reduction of dinitrogen to ammonia at a single molybdenum center, Science, 2003, 301, 76 CrossRef CAS .
  2. S. J. K. Forrest, B. Schluschaß, E. Y. Yuzik-Klimova and S. Schneider, Nitrogen Fixation via Splitting into Nitrido Complexes, Chem. Rev., 2021, 121, 6522 CrossRef CAS PubMed .
  3. G. Ertl, Catalytic Ammonia Synthesis, Plenum, New York, 1991 Search PubMed .
  4. M. N. Cosio and D. C. Powers, Prospects and challenges for nitrogen-atom transfer catalysis, Nat. Rev. Chem., 2023, 7, 424 CrossRef CAS PubMed .
  5. J. J. Curley, E. L. Sceats and C. C. Cummins, A Cycle for Organic Nitrile Synthesis via Dinitrogen Cleavage, J. Am. Chem. Soc., 2006, 128, 14036 CrossRef CAS PubMed .
  6. J. Du Bois, C. S. Tomooka, J. Hong and E. M. Carreira, Nitridomanganese(V) Complexes: Design, Preparation, and Use as Nitrogen Atom-Transfer Reagents, Acc. Chem. Res., 1997, 30, 364 CrossRef CAS .
  7. P. F. Kuijpers, J. I. van der Vlugt, S. Schneider and B. de Bruin, Nitrene Radical Intermediates in Catalytic Synthesis, Chem. – Eur. J., 2017, 23, 13819 CrossRef CAS PubMed .
  8. K. M. Carsch, I. M. DiMucci, D. A. Iovan, A. Li, S.-L. Zheng, C. J. Titus, S. J. Lee, K. D. Irwin, D. Nordlund, K. M. Lancaster and T. A. Betley, Synthesis of a copper-supported triplet nitrene complex pertinent to copper-catalyzed amination, Science, 2019, 365, 1138 CrossRef CAS PubMed .
  9. S. Kim, D. Kim, S. Y. Hong and S. Chang, Tuning Orbital Symmetry of Iridium Nitrenoid Enables Catalytic Diastereo- and Enantioselective Alkene Difunctionalizations, J. Am. Chem. Soc., 2021, 143, 3993 CrossRef CAS PubMed .
  10. Y. Gao, V. Carta, M. Pink and J. M. Smith, Catalytic Carbodiimide Guanylation by a Nucleophilic, High Spin Iron(II) Imido Complex, J. Am. Chem. Soc., 2021, 143, 5324 CrossRef CAS PubMed .
  11. B. Bagh, D. L. J. Broere, V. Sinha, P. F. Kuijpers, N. P. van Leest, B. de Bruin, S. Demeshko, M. A. Siegler and J. I. van der Vlugt, Catalytic Synthesis of N-Heterocycles via Direct C(sp3)–H Amination Using an Air-Stable Iron(III) Species with a Redox-Active Ligand, J. Am. Chem. Soc., 2017, 139, 5117 CrossRef CAS PubMed .
  12. T. Schmidt-Räntsch, H. Verplancke, J. N. Lienert, S. Demeschko, M. Otte, G. P. Van Trieste III, K. A. Reid, J. H. Reibenspies, D. C. Powers, M. C. Holthausen and S. Schneider, Nitrogen Atom Transfer Catalysis by Metallonitrene C–H Insertion: Photocatalytic Amidation of Aldehydes, Angew. Chem., Int. Ed., 2022, 61, e202115626 CrossRef PubMed .
  13. M. Goswami, V. Lyaskovskyy, S. R. Domingos, W. J. Buma, S. Woutersen, O. Troeppner, I. Ivanovic-Burmazovic, H. J. Lu, X. Cui, X. P. Zhang, E. J. Reijerse, S. DeBeer, M. M. van Schooneveld, F. F. Pfaff, K. Ray and B. de Bruin, Characterization of Porphyrin-Co(III)-'Nitrene Radical’ Species Relevant in Catalytic Nitrene Transfer Reactions, J. Am. Chem. Soc., 2015, 137, 5468 CrossRef CAS PubMed .
  14. J. J. Scepaniak, C. S. Vogel, M. M. Khusniyarov, F. W. Heinemann, K. Meyer and J. M. Smith, Synthesis, Structure, and Reactivity of an Iron(V) Nitride, Science, 2011, 331, 1049 CrossRef CAS PubMed .
  15. J. F. Berry, E. Bill, E. Bothe, S. D. George, B. Mienert, F. Neese and K. Wieghardt, An octahedral coordination complex of iron(VI), Science, 2006, 312, 1937 CrossRef CAS .
  16. M. Keilwerth, L. Grunwald, W. Mao, F. W. Heinemann, J. Sutter, E. Bill and K. Meyer, Ligand Tailoring Toward an Air-Stable Iron(V) Nitrido Complex, J. Am. Chem. Soc., 2021, 143, 1458 CrossRef CAS PubMed .
  17. N. B. Thompson, M. T. Green and J. C. Peters, Nitrogen Fixation via a Terminal Fe(IV) Nitride, J. Am. Chem. Soc., 2017, 139, 15312 CrossRef CAS .
  18. A. K. Maity, J. Murillo, A. J. Metta-Magaña, B. Pinter and S. Fortier, A Terminal Iron(IV) Nitride Supported by a Super Bulky Guanidinate Ligand and Examination of Its Electronic Structure and Reactivity, J. Am. Chem. Soc., 2017, 139, 15691 CrossRef CAS .
  19. M. Keilwerth, W. Mao, M. Malischewski, S. A. V. Jannuzzi, K. Breitwieser, F. W. Heinemann, A. Scheurer, S. DeBeer, D. Munz, E. Bill and K. Meyer, The synthesis and characterization of an iron(VII) nitrido complex, Nat. Chem., 2024, 16, 514 CrossRef CAS PubMed .
  20. M. Keilwerth, W. Mao, S. A. V. Jannuzzi, L. Grunwald, F. W. Heinemann, A. Scheurer, J. Sutter, S. DeBeer, D. Munz and K. Meyer, From Divalent to Pentavalent Iron Imido Complexes and an Fe(V) Nitride via N–C Bond Cleavage, J. Am. Chem. Soc., 2023, 145, 873 CrossRef CAS PubMed .
  21. E. D. Badding, S. Srisantitham, D. A. Lukoyanov, B. M. Hoffman and D. L. M. Suess, Connecting the geometric and electronic structures of the nitrogenase iron-molybdenum cofactor through site-selective 57Fe labelling, Nat. Chem., 2023, 15, 658 CrossRef CAS .
  22. Y. Kim, A. Sridharan and D. L. M. Suess, The Elusive Mononitrosylated Fe4S4 Cluster in Three Redox States, Angew. Chem., Int. Ed., 2022, 61, e202213032 CrossRef CAS PubMed .
  23. L. Grunwald, M. Clémancey, D. Klose, L. Dubois, S. Gambarelli, G. Jeschke, M. Wörle, G. Blondin and V. Mougel, A complete biomimetic iron-sulfur cubane redox series, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2122677119 CrossRef CAS .
  24. A. Sridharan, A. C. Brown and D. L. M. Suess, A Terminal Imido Complex of an Iron–Sulfur Cluster, Angew. Chem., Int. Ed., 2021, 60, 12802 CrossRef CAS PubMed .
  25. A. McSkimming and D. L. M. Suess, Dinitrogen binding and activation at a molybdenum–iron–sulfur cluster, Nat. Chem., 2021, 13, 666 CrossRef CAS .
  26. J. M. Smith, Reactive Transition Metal Nitride Complexes, Prog. Inorg. Chem., 2014, 58, 417 CAS .
  27. R. A. Eikey and M. M. Abu-Omar, Nitrido and imido transition metal complexes of Groups 6–8, Coord. Chem. Rev., 2003, 243, 83 CrossRef CAS .
  28. J. F. Berry, Terminal Nitrido and Imido Complexes of the Late Transition Metals, Comments Inorg. Chem., 2009, 30, 28 CrossRef CAS .
  29. C. E. Laplaza and C. C. Cummins, Dinitrogen Cleavage by a 3-Coordinate Molybdenum(III) Complex, Science, 1995, 268, 861 CrossRef CAS PubMed .
  30. C. E. Laplaza, M. J. A. Johnson, J. C. Peters, A. L. Odom, E. Kim, C. C. Cummins, G. N. George and I. J. Pickering, Dinitrogen cleavage by three-coordinate molybdenum(III) complexes: Mechanistic and structural data, J. Am. Chem. Soc., 1996, 118, 8623 CrossRef CAS .
  31. R. Thompson, B. L. Tran, S. Ghosh, C. H. Chen, M. Pink, X. Gao, P. J. Carroll, M. H. Baik and D. J. Mindiola, Addition of Si-H and B-H bonds and redox reactivity involving low-coordinate nitrido-vanadium complexes, Inorg. Chem., 2015, 54, 3068 CrossRef CAS PubMed .
  32. L. N. Grant, B. Pinter, T. Kurogi, M. E. Carroll, G. Wu, B. C. Manor, P. J. Carroll and D. J. Mindiola, Molecular titanium nitrides: nucleophiles unleashed, Chem. Sci., 2017, 8, 1209 RSC .
  33. D. Martelino, S. Mahato, W. VandeVen, N. M. Hein, R. M. Clarke, G. A. MacNeil, F. Thomas and T. Storr, Chromium Nitride Umpolung Tuned by the Locus of Oxidation, J. Am. Chem. Soc., 2022, 144, 11594 CrossRef CAS PubMed .
  34. S. N. Brown, Insertion of a metal nitride into carbon-carbon double bonds, J. Am. Chem. Soc., 1999, 121, 9752 CrossRef CAS .
  35. T. J. Crevier and J. M. Mayer, Direct Attack of Phenyl Anion at an Electrophilic Osmium−Nitrido Ligand, J. Am. Chem. Soc., 1998, 120, 5595 CrossRef CAS .
  36. W.-L. Man, W. W. Y. Lam, H.-K. Kwong, S.-M. Yiu and T.-C. Lau, Ligand-Accelerated Activation of Strong C-H Bonds of Alkanes by a (Salen)ruthenium(VI)–Nitrido Complex, Angew. Chem., Int. Ed., 2012, 51, 9101 CrossRef CAS .
  37. W.-L. Man, T.-M. Tang, T.-W. Wong, T.-C. Lau, S.-M. Peng and W.-T. Wong, Highly Electrophilic (Salen)ruthenium(VI) Nitrido Complexes, J. Am. Chem. Soc., 2004, 126, 478 CrossRef CAS PubMed .
  38. E. M. Zolnhofer, M. Käβ, M. M. Khusniyarov, F. W. Heinemann, L. Maron, M. van Gastel, E. Bill and K. Meyer, An Intermediate Cobalt(IV) Nitrido Complex and its N-Migratory Insertion Product, J. Am. Chem. Soc., 2014, 136, 15072 CrossRef CAS .
  39. M. G. Scheibel, Y. L. Wu, A. C. Stuckl, L. Krause, E. Carl, D. Stalke, B. de Bruin and S. Schneider, Synthesis and Reactivity of a Transient, Terminal Nitrido Complex of Rhodium, J. Am. Chem. Soc., 2013, 135, 17719 CrossRef CAS PubMed .
  40. J. Schoffel, A. Y. Rogachev, S. DeBeer George and P. Burger, Isolation and hydrogenation of a complex with a terminal iridium-nitrido bond, Angew. Chem., Int. Ed., 2009, 48, 4734 CrossRef .
  41. V. Vreeken, M. A. Siegler, B. de Bruin, J. N. Reek, M. Lutz and J. I. van der Vlugt, C-H Activation of Benzene by a Photoactivated Ni(II)(azide): Formation of a Transient Nickel Nitrido Complex, Angew. Chem., Int. Ed., 2015, 54, 7055 CrossRef CAS .
  42. P. Q. Kelly, A. S. Filatov and M. D. Levin, A Synthetic Cycle for Heteroarene Synthesis by Nitride Insertion**, Angew. Chem., Int. Ed., 2022, 61, e202213041 CrossRef CAS PubMed .
  43. H. Shi, H. K. Lee, Y. Pan, K.-C. Lau, S.-M. Yiu, W. W. Y. Lam, W.-L. Man and T.-C. Lau, Structure and Reactivity of a Manganese(VI) Nitrido Complex Bearing a Tetraamido Macrocyclic Ligand, J. Am. Chem. Soc., 2021, 143, 15863 CrossRef CAS PubMed .
  44. F. S. Schendzielorz, M. Finger, C. Volkmann, C. Würtele and S. Schneider, A Terminal Osmium(IV) Nitride: Ammonia Formation and Ambiphilic Reactivity, Angew. Chem., Int. Ed., 2016, 55, 11417 CrossRef CAS .
  45. T. J. Crevier, B. K. Bennett, J. D. Soper, J. A. Bowman, A. Dehestani, D. A. Hrovat, S. Lovell, W. Kaminsky and J. M. Mayer, C−N Bond Formation on Addition of Aryl Carbanions to the Electrophilic Nitrido Ligand in TpOs(N)Cl2, J. Am. Chem. Soc., 2001, 123, 1059 CrossRef CAS PubMed .
  46. C. Schiller, D. Sieh, N. Lindenmaier, M. Stephan, N. Junker, E. Reijerse, A. A. Granovsky and P. Burger, Cleavage of an Aromatic C–C Bond in Ferrocene by Insertion of an Iridium Nitrido Nitrogen Atom, J. Am. Chem. Soc., 2023, 145, 11392 CrossRef CAS PubMed .
  47. G. P. Connor, B. Q. Mercado, H. M. C. Lant, J. M. Mayer and P. L. Holland, Chemical Oxidation of a Coordinated PNP-Pincer Ligand Forms Unexpected Re–Nitroxide Complexes with Reversal of Nitride Reactivity, Inorg. Chem., 2019, 58, 10791 CrossRef CAS .
  48. T.-W. Wong, T.-C. Lau and W.-T. Wong, Osmium(VI) Nitrido and Osmium(IV) Phosphoraniminato Complexes Containing Schiff Base Ligands, Inorg. Chem., 1999, 38, 6181 CrossRef CAS .
  49. A. Dehestani, W. Kaminsky and J. M. Mayer, Tuning the Properties of the Osmium Nitrido Group in TpOs(N)X2 by Changing the Ancillary Ligand, Inorg. Chem., 2003, 42, 605 CrossRef CAS PubMed .
  50. S. Mahato, W. VandeVen, G. A. MacNeil, J. M. Pulfer and T. Storr, Untangling ancillary ligand donation versus locus of oxidation effects on metal nitride reactivity, Chem. Sci., 2024, 15, 2211 RSC .
  51. N. M. Hein, G. A. MacNeil and T. Storr, Elaboration on the Electronics of Salen Manganese Nitrides: Investigations into Alkoxy-Substituted Ligand Scaffolds, Inorg. Chem., 2021, 60, 16895 CrossRef CAS .
  52. R. M. Clarke and T. Storr, Tuning Electronic Structure To Control Manganese Nitride Activation, J. Am. Chem. Soc., 2016, 138, 15299 CrossRef CAS .
  53. F. Thomas, Ligand-centred oxidative chemistry in sterically hindered salen complexes: an interesting case with nickel, Dalton Trans., 2016, 45, 10866 RSC .
  54. R. M. Clarke, K. Herasymchuk and T. Storr, Electronic structure elucidation in oxidized metal–salen complexes, Coord. Chem. Rev., 2017, 352, 67 CrossRef CAS .
  55. T. Takeyama and Y. Shimazaki, Diversity of oxidation state in copper complexes with phenolate ligands, Dalton Trans., 2024, 53, 3911 RSC .
  56. J. Andrez, V. Guidal, R. Scopelliti, J. Pécaut, S. Gambarelli and M. Mazzanti, Ligand and Metal Based Multielectron Redox Chemistry of Cobalt Supported by Tetradentate Schiff Bases, J. Am. Chem. Soc., 2017, 139, 8628 CrossRef CAS .
  57. M. Kawai, T. Yamaguchi, S. Masaoka, F. Tani, T. Kohzuma, L. Chiang, T. Storr, K. Mieda, T. Ogura, R. K. Szilagyi and Y. Shimazaki, Influence of Ligand Flexibility on the Electronic Structure of Oxidized Ni(III)-Phenoxide Complexes, Inorg. Chem., 2014, 53, 10195 CrossRef CAS PubMed .
  58. A. Awasthi, S. C. Mallojjala, R. Kumar, R. Eerlapally, J. S. Hirschi and A. Draksharapu, Altering the Localization of an Unpaired Spin in a Formal Ni(V) Species, Chem. – Eur. J., 2024, 30, e202302824 CrossRef CAS .
  59. A. Awasthi, I. F. Leach, S. Engbers, R. Kumar, R. Eerlapally, S. Gupta, J. Klein and A. Draksharapu, Formation and Reactivity of a Fleeting Ni(III) Bisphenoxyl Diradical Species, Angew. Chem., Int. Ed., 2022, 61, e202211345 CrossRef CAS PubMed .
  60. M. Keener, M. Peterson, R. Hernández Sánchez, V. F. Oswald, G. Wu and G. Ménard, Towards Catalytic Ammonia Oxidation to Dinitrogen: A Synthetic Cycle by Using a Simple Manganese Complex, Chem. – Eur. J., 2017, 23, 11479 CrossRef CAS .
  61. T. Chantarojsiri, A. H. Reath and J. Y. Yang, Cationic Charges Leading to an Inverse Free-Energy Relationship for N−N Bond Formation by MnVI Nitrides, Angew. Chem., Int. Ed., 2018, 57, 14037 CrossRef CAS PubMed .
  62. S. V. Park, A. R. Corcos, A. N. Jambor, T. Yang and J. F. Berry, Formation of the N≡N Triple Bond from Reductive Coupling of a Paramagnetic Diruthenium Nitrido Compound, J. Am. Chem. Soc., 2022, 144, 3259 CrossRef CAS PubMed .
  63. C. C. Almquist, T. Rajeshkumar, H. Jayaweera, N. Removski, W. Zhou, B. S. Gelfand, L. Maron and W. E. Piers, Oxidation-induced ambiphilicity triggers N-N bond formation and dinitrogen release in octahedral terminal molybdenum(V) nitrido complexes, Chem. Sci., 2024, 15, 5152 RSC .
  64. C. C. Almquist, N. Removski, T. Rajeshkumar, B. S. Gelfand, L. Maron and W. E. Piers, Spontaneous Ammonia Activation Through Coordination-Induced Bond Weakening in Molybdenum Complexes of a Dianionic Pentadentate Ligand Platform, Angew. Chem., Int. Ed., 2022, 61, e202203576 CrossRef CAS PubMed .
  65. N. G. Léonard, T. Chantarojsiri, J. W. Ziller and J. Y. Yang, Cationic Effects on the Net Hydrogen Atom Bond Dissociation Free Energy of High-Valent Manganese Imido Complexes, J. Am. Chem. Soc., 2022, 144, 1503 CrossRef .
  66. S. Kim, H. Y. Zhong, Y. Park, F. Loose and P. J. Chirik, Catalytic Hydrogenation of a Manganese(V) Nitride to Ammonia, J. Am. Chem. Soc., 2020, 142, 9518 CrossRef CAS PubMed .
  67. F. Loose, D. A. Wang, L. Tian, G. D. Scholes, R. R. Knowles and P. J. Chirik, Evaluation of excited state bond weakening for ammonia synthesis from a manganese nitride: stepwise proton coupled electron transfer is preferred over hydrogen atom transfer, Chem. Commun., 2019, 55, 5595 RSC .
  68. D. Wang, F. Loose, P. J. Chirik and R. R. Knowles, N–H Bond Formation in a Manganese(V) Nitride Yields Ammonia by Light-Driven Proton-Coupled Electron Transfer, J. Am. Chem. Soc., 2019, 141, 4795 CrossRef CAS PubMed .
  69. 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 CrossRef CAS .
  70. R. Kunert, C. Philouze, O. Jarjayes and F. Thomas, Stable M(II)-Radicals and Nickel(III) Complexes of a Bis(phenol) N-Heterocyclic Carbene Chelated to Group 10 Metal Ions, Inorg. Chem., 2019, 58, 8030 CrossRef CAS PubMed .
  71. E. Peris, Smart N-Heterocyclic Carbene Ligands in Catalysis, Chem. Rev., 2018, 118, 9988 CrossRef CAS PubMed .
  72. P. de Frémont, N. Marion and S. P. Nolan, Carbenes: Synthesis, properties, and organometallic chemistry, Coord. Chem. Rev., 2009, 253, 862 CrossRef .
  73. S. Bellemin-Laponnaz, R. Welter, L. Brelot and S. Dagorne, Synthesis and structure of V(V) and Mn(III) NHC complexes supported by a tridentate bis-aryloxide-N-heterocyclic carbene ligand, J. Organomet. Chem., 2009, 694, 604 CrossRef CAS .
  74. J. A. Bellow, S. A. Stoian, J. van Tol, A. Ozarowski, R. L. Lord and S. Groysman, Synthesis and Characterization of a Stable High-Valent Cobalt Carbene Complex, J. Am. Chem. Soc., 2016, 138, 5531 CrossRef CAS .
  75. Z. S. Ghavami, M. R. Anneser, F. Kaiser, P. J. Altmann, B. J. Hofmann, J. F. Schlagintweit, G. Grivani and F. E. Kühn, A bench stable formal Cu(iii) N-heterocyclic carbene accessible from simple copper(II) acetate, Chem. Sci., 2018, 9, 8307 RSC .
  76. R. Kunert, C. Philouze, O. Jarjayes, T. Storr and F. Thomas, Single-Oxidation of Ni(II) Complexes of Bidentate Fused N-Heterocyclic Carbene–Phenol Conjugates, Organometallics, 2023, 42, 1550 CrossRef CAS .
  77. C. Gandara, C. Philouze, O. Jarjayes and F. Thomas, Coordination chemistry of a redox non-innocent NHC bis(phenolate) pincer ligand with nickel(II), Inorg. Chim. Acta, 2018, 482, 561 CrossRef CAS .
  78. H. Kropp, A. E. King, M. M. Khusniyarov, F. W. Heinemann, K. M. Lancaster, S. DeBeer, E. Bill and K. Meyer, Manganese nitride complexes in oxidation states III, IV, and V: synthesis and electronic structure, J. Am. Chem. Soc., 2012, 134, 15538 CrossRef CAS PubMed .
  79. S. Li, Z. Wang, T. S. A. Hor and J. Zhao, First crystallographic elucidation of a high-valent molybdenum oxo N-heterocyclic carbene complex [CpMoVIO2(IBz)]2[Mo6O19], Dalton Trans., 2012, 41, 1454 RSC .
  80. G. Ciancaleoni, L. Belpassi and F. Marchetti, Back-Donation in High-Valent d0 Metal Complexes: Does It Exist? The Case of NbV, Inorg. Chem., 2017, 56, 11266 CrossRef CAS PubMed .
  81. E. L. Rosen, C. D. Jr. Varnado, A. G. Tennyson, D. M. Khramov, J. W. Kamplain, D. H. Sung, P. T. Cresswell, V. M. Lynch and C. W. Bielawski, Redox-Active N-Heterocyclic Carbenes: Design, Synthesis, and Evaluation of Their Electronic Properties, Organometallics, 2009, 28, 6695 CrossRef CAS .
  82. W. I. Dzik, X. Xu, X. P. Zhang, J. N. H. Reek and B. de Bruin, ‘Carbene Radicals’ in CoII(por)-Catalyzed Olefin Cyclopropanation, J. Am. Chem. Soc., 2010, 132, 10891 CrossRef CAS PubMed .
  83. A. G. Tennyson, V. M. Lynch and C. W. Bielawski, Arrested Catalysis: Controlling Kumada Coupling Activity via a Redox-Active N-Heterocyclic Carbene, J. Am. Chem. Soc., 2010, 132, 9420 CrossRef CAS PubMed .
  84. W. I. Dzik, X. P. Zhang and B. de Bruin, Redox Noninnocence of Carbene Ligands: Carbene Radicals in (Catalytic) C−C Bond Formation, Inorg. Chem., 2011, 50, 9896 CrossRef CAS PubMed .
  85. E. V. Ilyakina, A. I. Poddel'sky, A. V. Piskunov, G. K. Fukin, A. S. Bogomyakov, V. K. Cherkasov and G. A. Abakumov, The interaction of N,N′-bis(2,6-dimethylphenyl)imidazol-2-ylidene with o-benzosemiquinonato zinc(II) and indium(III) complexes, New J. Chem., 2012, 36, 1944 RSC .
  86. C. Romain, S. Choua, J.-P. Collin, M. Heinrich, C. Bailly, L. Karmazin-Brelot, S. Bellemin-Laponnaz and S. Dagorne, Redox and Luminescent Properties of Robust and Air-Stable N-Heterocyclic Carbene Group 4 Metal Complexes, Inorg. Chem., 2014, 53, 7371 CrossRef CAS .
  87. C. F. Harris, M. B. Bayless, N. P. van Leest, Q. J. Bruch, B. N. Livesay, J. Bacsa, K. I. Hardcastle, M. P. Shores, B. de Bruin and J. D. Soper, Redox-Active Bis(phenolate) N-Heterocyclic Carbene [OCO] Pincer Ligands Support Cobalt Electron Transfer Series Spanning Four Oxidation States, Inorg. Chem., 2017, 56, 12421 CrossRef CAS PubMed .
  88. J. Beerhues, M. Neubrand, S. Sobottka, N. I. Neuman, H. Aberhan, S. Chandra and B. Sarkar, Directed Design of a AuI Complex with a Reduced Mesoionic Carbene Radical Ligand: Insights from 1,2,3-Triazolylidene Selenium Adducts and Extensive Electrochemical Investigations, Chem. – Eur. J., 2021, 27, 6557 CrossRef CAS PubMed .
  89. G. Meloni, L. Beghetto, M. Baron, A. Biffis, P. Sgarbossa, M. Mba, P. Centomo, L. Orian, C. Graiff and C. Tubaro, Manganese(III) complexes with tetradentate O^C^C^O ligands: Synthesis, characterization and catalytic studies on the CO2 cycloaddition with epoxides, Mol. Catal., 2023, 538, 113006 CrossRef CAS .
  90. J. DuBois, J. Hong, E. M. Carreira and M. W. Day, Nitrogen transfer from a nitridomanganese(V) complex: Amination of silyl enol ethers, J. Am. Chem. Soc., 1996, 118, 915 CrossRef CAS .
  91. J. Bendix, [Cr(N)Cl4]2-: A Simple Nitrido Complex Synthesized by Nitrogen-Atom Transfer, J. Am. Chem. Soc., 2003, 125, 13348 CrossRef CAS PubMed .
  92. T. Birk and J. Bendix, Atom transfer as a preparative tool in coordination chemistry. Synthesis and characterization of Cr(V) nitrido complexes of bidentate ligands, Inorg. Chem., 2003, 42, 7608 CrossRef CAS .
  93. G. Golubkov and Z. Gross, Chromium(V) and Chromium(VI) Nitrido Complexes of Tris(pentafluorophenyl)corrole, Angew. Chem., Int. Ed., 2003, 42, 4507 CrossRef CAS .
  94. G. Golubkov and Z. Gross, Nitrogen atom transfer between manganese complexes of salen, porphyrin, and corrole and characterization of a (nitrido)manganese(VI) corrole, J. Am. Chem. Soc., 2005, 127, 3258 CrossRef CAS PubMed .
  95. C. C. Hojilla Atienza, A. C. Bowman, E. Lobkovsky and P. J. Chirik, Photolysis and Thermolysis of Bis(imino)pyridine Cobalt Azides: C−H Activation from Putative Cobalt Nitrido Complexes, J. Am. Chem. Soc., 2010, 132, 16343 CrossRef CAS .
  96. D. M. King, F. Tuna, E. J. McInnes, J. McMaster, W. Lewis, A. J. Blake and S. T. Liddle, Isolation and characterization of a uranium(VI)-nitride triple bond, Nat. Chem., 2013, 5, 482 CrossRef CAS PubMed .
  97. 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 RSC .
  98. K. Meyer, J. Bendix, E. Bill, T. Weyhermüller and K. Wieghardt, Molecular and Electronic Structure of Nitridochromium(V) Complexes with Macrocyclic Amine Ligands, Inorg. Chem., 1998, 37, 5180 CrossRef CAS .
  99. J. Bendix, S. R. Wilson and T. Prussak-Wieckowska, The Improved Synthesis of [Cr(N)(salen)].CH3NO2, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1998, 54, 923 CrossRef .
  100. C. J. Cramer, J. R. Gour, A. Kinal, M. Wtoch, P. Piecuch, A. R. M. Shahi and L. Gagliardi, Stereoelectronic effects on molecular geometries and state-energy splittings of ligated monocopper dioxygen complexes, J. Phys. Chem. A, 2008, 112, 3754 CrossRef CAS PubMed .
  101. M. Keilwerth, J. Hohenberger, F. W. Heinemann, J. Sutter, A. Scheurer, H. Fang, E. Bill, F. Neese, S. Ye and K. Meyer, A Series of Iron Nitrosyl Complexes {Fe–NO}6–9 and a Fleeting {Fe–NO}10 Intermediate en Route to a Metalacyclic Iron Nitrosoalkane, J. Am. Chem. Soc., 2019, 141, 17217 CrossRef CAS PubMed .
  102. F. Weinhold, C. R. Landis and E. D. Glendening, What is NBO analysis and how is it useful? Int, Rev. Phys. Chem., 2016, 35, 399 CAS .
  103. F. Weinhold and C. R. Landis, Discovering Chemistry with Natural Bond Orbitals, Wiley, Hoboken, NJ, 2012 Search PubMed .
  104. N. G. Connelly and W. E. Geiger, Chemical Redox Agents for Organometallic Chemistry, Chem. Rev., 1996, 96, 877 CrossRef CAS PubMed .
  105. D. M. D'Alessandro and F. R. Keene, Current trends and future challenges in the experimental, theoretical and computational analysis of intervalence charge transfer (IVCT) transitions, Chem. Soc. Rev., 2006, 35, 424 Search PubMed .
  106. M. Orio, O. Jarjayes, H. Kanso, C. Philouze, F. Neese and F. Thomas, X-Ray, Structures of Copper(II) and Nickel(II) Radical Salen Complexes: The Preference of Galactose Oxidase for Copper(II), Angew. Chem., Int. Ed., 2010, 49, 4989 CrossRef CAS .
  107. A. Dei, D. Gatteschi, L. Pardi, A. L. Barra and L. C. Brunel, Millimeter band EPR spectra reveal large zero-field splittings in copper(II)—semiquinonato complexes, Chem. Phys. Lett., 1990, 175, 589 CrossRef CAS .
  108. J. DuBois, C. S. Tomooka, J. Hong and E. M. Carreira, Nitridomanganese(V) complexes: Design, preparation, and use as nitrogen atom-transfer reagents, Acc. Chem. Res., 1997, 30, 364 CrossRef CAS .
  109. S.-M. Yiu, W. W. Y. Lam, C.-M. Ho and T.-C. Lau, Facile N⋯N Coupling of Manganese(V) Imido Species, J. Am. Chem. Soc., 2007, 129, 803 CrossRef CAS PubMed .
  110. Y. Murata, F. Cheng, T. Kitagawa and K. Komatsu, Generation of Fullerenyl Cation (EtO)2P+(OH)CH2−C60+from RC60−H and from RC60−C60R (R = CH2P(O)(OEt)2), J. Am. Chem. Soc., 2004, 126, 8874 CrossRef CAS PubMed .
  111. S. Stoll and A. Schweiger, EasySpin, a comprehensive software package for spectral simulation and analysis in EPR, J. Magn, Reson., 2006, 178, 42 CrossRef CAS .
  112. I. Noviandri, K. N. Brown, D. S. Fleming, P. T. Gulyas, P. A. Lay, A. F. Masters and L. Phillips, The Decamethylferrocenium/Decamethylferrocene Redox Couple: A Superior Redox Standard to the Ferrocenium/Ferrocene Redox Couple for Studying Solvent Effects on the Thermodynamics of Electron Transfer, J. Phys. Chem. B, 1999, 103, 6713 CrossRef CAS .
  113. G. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C: Struct. Chem., 2015, 71, 3 Search PubMed .
  114. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, OLEX2: a complete structure solution, refinement and analysis program, J. Appl. Crystallogr., 2009, 42, 339 CrossRef CAS .
  115. M. J. Frisch, et al., Gaussian 16, 2016 Search PubMed .
  116. L. Goerigk and S. Grimme, Efficient and Accurate Double-Hybrid-Meta-GGA Density Functionals—Evaluation with the Extended GMTKN30 Database for General Main Group Thermochemistry, Kinetics, and Noncovalent Interactions, J. Chem. Theory Comput., 2011, 7, 291 CrossRef CAS .
  117. A. Schäfer, H. Horn and R. Ahlrichs, Fully optimized contracted Gaussian basis sets for atoms Li to Kr, J. Chem. Phys., 1992, 2571 CrossRef .
  118. A. Schäfer, C. Huber and R. Ahlrichs, Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr, J. Chem. Phys., 1994, 100, 5829 CrossRef .
  119. F. Weinhold and J. E. Carpenter, in The Structure of Small Molecules and Ions, ed. R. Naaman and Z.Yager, Springer, 1988, p. 227 Search PubMed .
  120. J. Tomasi, B. Mennucci and E. Cancès, The IEF version of the PCM solvation method: an overview of a new method addressed to study molecular solutes at the QM ab initio level, J. Mol. Struct.: THEOCHEM, 1999, 464, 211 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available. CCDC 2359452, 2359453, 2359455 and 2359456. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt01765j

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.