Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Sep 20.
Published in final edited form as: Inorg Chem. 2021 May 27;60(18):13811–13820. doi: 10.1021/acs.inorgchem.1c00683

[2Fe–2S] Cluster Supported by Redox-Active o-Phenylenediamide Ligands and Its Application toward Dinitrogen Reduction

Qiuming Liang 1, Joshua C DeMuth 2, Aleksa Radović 2, Nikki J Wolford 2, Michael L Neidig 2, Datong Song 3
PMCID: PMC8453056  NIHMSID: NIHMS1713270  PMID: 34043353

Abstract

As prevalent cofactors in living organisms, iron–sulfur clusters participate in not only the electron-transfer processes but also the biosynthesis of other cofactors. Many synthetic iron–sulfur clusters have been used in model studies, aiming to mimic their biological functions and to gain mechanistic insight into the related biological systems. The smallest [2Fe–2S] clusters are typically used for one-electron processes because of their limited capacity. Our group is interested in functionalizing small iron–sulfur clusters with redox-active ligands to enhance their electron storage capacity, because such functionalized clusters can potentially mediate multielectron chemical transformations. Herein we report the synthesis, structural characterization, and catalytic activity of a diferric [2Fe–2S] cluster functionalized with two o-phenylenediamide ligands. The electrochemical and chemical reductions of such a cluster revealed rich redox chemistry. The functionalized diferric cluster can store up to four electrons reversibly, where the first two reduction events are ligand-based and the remainder metal-based. The diferric [2Fe–2S] cluster displays catalytic activity toward silylation of dinitrogen, affording up to 88 equiv of the amine product per iron center.

Graphical Abstract

graphic file with name nihms-1713270-f0001.jpg

INTRODUCTION

Iron–sulfur clusters are ubiquitous and multipurpose cofactors in various living organisms.1,2 As the simplest ones in the iron–sulfur cluster family, the [2Fe–2S] clusters have important biological functions. For example, the [2Fe–2S] clusters can combine to form [4Fe–4S] cubanes, which can couple further to generate clusters of even higher nuclearities.35 Moreover, the [2Fe–2S] clusters, such as those in ferredoxin- and Rieske-type proteins, play a key role in one-electron-transfer processes,1,2 in which the [2Fe–2S] clusters cycle between the oxidized diferric [2Fe–2S]2+ and mixed-valence [2Fe–2S]+ states.69 The diferric [2Fe–2S]2+ cluster consists of two antiferromagnetically coupled iron centers, leading to an S = 0 ground state.69 When the diferric cluster is reduced by one electron, the additional electron can be localized at one of the iron centers to give an S = 1/2 state or valence-delocalized to give an S = 1/2 or S = 9/2 state.69 The influence of the amino acid residues in the first and second coordination spheres on the redox potentials of these clusters has been studied using mutations.1014

Many biomimetic [2Fe–2S] clusters have been investigated in order to gain information on their electronic structures and electron-transfer processes.1534 A few biomimetic [2Fe–2S] clusters supported by N,N-bidentate ligands have emerged recently (Chart 1). Meyer and co-workers isolated and crystallographically characterized a series of [2Fe–2S]n+ clusters (n = 2, 1, 0) supported by two 2,2′-(phenylmethylene)bis(benzimidazolide) ligands.2527 Analogous [2Fe–2S] complexes have also shown proton-coupled electron-transfer reactivity.2831 Driess and Holland reported a series of [2Fe–2S]n+ (n = 2, 1, 0) clusters supported by β-diketiminate ligands.32,33 Remarkably, the [2Fe–2S]+ clusters with diethylphenyl-substituted β-diketiminato ligands CH-[CMeN(2,6-Et2C6H3)]2 are extensively delocalized, as suggested by the 57Fe Mössbauer and X-ray absorption spectroscopy/X-ray emission spectroscopy spectroscopic data and density functional theory calculations.32 Jones reported a diferric [2Fe–2S]2+ cluster supported by a guanidinato ligand.34 No [2Fe–2S]n cluster supported by redox-active N,N-bidentate ligands has been reported to date.

Chart 1.

Chart 1.

Examples of [2Fe–2S] Supported by N,N-Bidentate Ligands

Redox-active ligands have multiple energetically accessible oxidation states.3547 Their coordination to metal centers induces radical reactivity and electron reservoir behavior, which is often utilized to develop exciting new examples of catalysis.3554 The N,N-chelating o-phenylenediamide (pda) is a classic example of a redox-active ligand, whose coordination chemistry has been explored toward a large number of transition metals and main-group elements.5567 It has been clearly established that the pda ligand exists in three different redox states in coordination compounds: the closed-shell dianion, open-shell π-radical monoanion (S = 1/2), and closed-shell neutral form.5567 We have been investigating the coordination chemistry of the pda ligand toward transition metals and main-group elements.6467 Our group is interested in coordinating redox-active pda ligands to the diferric [2Fe–2S]2+ cluster. We envision that the resulting complex can store multiple electrons (i.e., reduction equivalents) and mediate multielectron redox processes. Herein we report the synthesis and rich redox chemistry of a diferric [2Fe–2S]2+ cluster supported by two monoanionic pda ligands. The combination of spectroscopic methods and X-ray crystallographic analysis was used to characterize the complexes in various oxidation states. The catalytic activity toward dinitrogen (N2) reduction is also reported herein.

RESULTS AND DISCUSSION

The precursor [Fe(Xylpda•−)(η6-toluene)] [1; Xylpda = N,N′-bis(2,6-dimethylphenyl)-o-phenylenediamide] was prepared from N,N′-bis(2,6-dimethylphenyl)phenylenediamine68 and Fe(HMDS)269 (HMDS = hexamethyldisilazide) in toluene at room temperature. The spectroscopic data and metric parameters of 1 are similar to those of the previously reported Fe(Dipppda•−)(η6-toluene) [Dipppda = N,N′-bis(2,6-diisopropylphenyl)-o-phenylenediamide].66,67 The reaction of 1 with 1/8 equiv of S8 in tetrahydrofuran (THF) at room temperature affords dark-purple complex 2 in 78% yield (Scheme 1). The 1H NMR spectrum of complex 2 in THF-d8 at 25 °C shows five paramagnetically broadened and shifted resonances between +14 and +4 ppm. The chemical shift span becomes narrower at lower temperatures (Figures S8 and S9). The molecular structure of 2 was established by X-ray diffraction, featuring a rhombic [Fe2(μ-S)2] core coordinated by two pda ligands (Figure 1). The metric parameters of the [Fe2(μ-S)2] core are similar to those of the synthetic and biological [2Fe–2S]2+ clusters in the literature.1534 The average C–N bond length of the Xylpda ligand is 1.341(4) Å (Table 1), between the typical C–N single and double bond lengths. The Cα–Cα and Cγ–Cγ bond lengths are 1.452(1) and 1.416(1) Å, respectively, while the other four bonds of the phenylene backbone feature the long Cα–Cβ [avg. 1.418(2) Å] and short Cβ–Cγ [avg. 1.358(7) Å] bonds. The pattern in these metric parameters suggests that the pda ligands are open-shell π-radical anions Xylpda•− (S = 1/2). The intramolecular Fe–Fe distance [2.633(1) Å] is slightly longer than the sum of the covalent radii (2.50 Å).70 The solution magnetic moment of 2 is 1.2 μB in THF-d8 at 25 °C using the Evans method,71 close to an STotal = 0 ground state. The observed magnetic moment is similar to those of the analogous literature compounds.25,32 The zero-field Mössbauer spectrum of a solid sample of 2 recorded at 80 K shows one quadrupole doublet (Figure 3) with an isomer shift of δ = 0.23 mm s−1 and a quadrupole splitting of |ΔEQ| = 1.68 mm s−1 (Table 2), confirming the presence of equivalent high-spin Fe3+ centers. These values are similar to those observed for other synthetic and biological diferric [2Fe–2S]2+ clusters.1534 The cyclic voltammetry experiment of 2 in THF at 23 °C revealed three reversible reduction events at −0.98, −1.39, and −2.42 V (with respect to Fc+/Fc), respectively (Figure 2). In addition, there is a reduction event with limited chemical reversibility at −3.31 V (with respect to Fc+/Fc), which suggests that the four-electron reduction of 2 leads to a species that has limited stability. The cyclic voltammogram of 2 shows that the cluster can store up to four electrons and that three of the reduced clusters might be stable enough for isolation.

Scheme 1.

Scheme 1.

Synthesis of 2

Figure 1.

Figure 1.

X-ray structure of 2. The thermal ellipsoids are shown at 30% probability. Hydrogen atoms are omitted for clarity.

Table 1.

Selected Bond Distances and Angles of Complexes 2–6

graphic file with name nihms-1713270-t0002.jpg
2 3 4 5 6
Fe–N 1.973(4) 1.954(3) 1.955(5) 1.978(4)
1.959(3) 1.944(4) 1.965(4) 1.992(4) 2.092(2)
1.973(3) 1.994(3) 1.962(4) 2.049(5) 2.060(2)
1.952(5) 1.977(3) 1.954(4) 2.033(4)
N–Cα 1.338(6) 1.385(5) 1.403(7) 1.380(7)
1.344(7) 1.394(5) 1.385(8) 1.375(7) 1.387(2)
1.336(7) 1.343(5) 1.396(6) 1.387(7) 1.378(3)
1.344(5) 1.345(5) 1.392(6) 1.387(7)
Cα–Cα 1.453(6) 1.417(5) 1.429(7) 1.445(7) 1.432(3)
1.451(6) 1.432(5) 1.423(8) 1.428(8)
Cα–Cβ 1.420(8) 1.400(6) 1.386(9) 1.384(7)
1.420(6) 1.397(6) 1.390(8) 1.391(7) 1.397(3)
1.417(5) 1.415(5) 1.387(7) 1.386(8) 1.400(3)
1.414(7) 1.428(6) 1.396(7) 1.388(8)
Cβ–Cγ 1.353(6) 1.387(7) 1.408(8) 1.415(8)
1.356(8) 1.403(6) 1.402(9) 1.395(8) 1.399(3)
1.370(7) 1.372(5) 1.401(7) 1.388(8) 1.395(3)
1.353(5) 1.357(6) 1.401(7) 1.395(8)
Cγ–Cγ 1.416(6) 1.370(7) 1.379(8) 1.378(8) 1.377(3)
1.415(6) 1.406(5) 1.383(8) 1.374(9)
Fe–Fe 2.633(1) 2.6865(8) 2.764(1) 2.709(1) 2.7306(6)
2.7486(9)
Fe–S 2.186(1) 2.251(2) 2.264(2)
2.191(2) 2.243(1) 2.253(1) 2.207(2) 2.2804(7)
2.184(2) 2.238(1) 2.254(2) 2.266(2) 2.2819(6)
2.196(1) 2.245(2) 2.254(2)
∠S–Fe–S 105.98(5) 102.93(4) 104.30(6) 106.61(6) 106.47(2)
105.91(5) 104.69(6) 104.94(6)

Figure 3.

Figure 3.

Zero-field 57Fe Mössbauer spectrum of 26 in the solid state at 80 K.

Table 2.

Solution Magnetic Moment at 298 K and Solid-State Zero-Field Mössbauer Data at 80 K

complex μ eff a S Total b S Fe c δ,c mm s−1 EQ|, mm s−1
1 0 1/2 0.38 0.78
2 1.2d 0 5/2 0.23 1.68
3 1.9 1/2 5/2 0.24 1.60
4 1.6d 0 5/2 0.26 1.68
5 2.3 1/2 5/2 0.33 1.84
2 0.62 2.84
6 1.7d 0 2 0.70 2.61
a

Solution magnetic moment by the Evans method.

b

Spin state of the molecule.

c

Intrinsic spin state of the iron center.

d

The residual magnetic moment data are similar to those of the analogous literature compounds;25,32 the 1H NMR chemical shift span becomes narrower at lower temperatures (Figures S8S13).

Figure 2.

Figure 2.

Cyclic voltammogram of 2 (2 mM in THF/0.2 M TBAPF6) versus Fc+/Fc at a scan rate of 100 mV s−1 (top) and the corresponding redox couples (bottom).

Reducing 2 using 1 equiv of KC8 gives the new [2Fe–2S] cluster 3 as a dark-purple crystalline solid in 73% yield (Scheme 2). In the crystal lattice of 3, the complex anion [Fe2(μ-S)2(Xylpda)2] and [K(THF)6]+ cation do not display any close contacts (Figure 4). One of the Xylpda ligands resembles those in complex 2 [Cα–N bond lengths, 1.343(5) and 1.345(5) Å; Table 1], while the other displays longer Cα–N distances [1.385(5) and 1.394(5) Å]. These metric parameters suggest that the former is closer to Xylpda•− and the latter is closer to Xylpda2−; i.e., the reduction occurs at one of the Xylpda ligands in 2. The Fe–S distances [avg. 2.241(3) Å in 3 vs avg. 2.189(4) Å in 2] and Fe–Fe distances [2.6865(8) Å in 3 vs 2.633(1) Å in 2] are slightly longer than those of 2, respectively. The magnetic susceptibility measurements of 3 in THF-d8 using the Evans method71 revealed a magnetic moment of 1.9 μB at 298 K (Table 2). The zero-field Mössbauer spectrum of 3 displays only one quadrupole doublet with an isomer shift of 0.24 mm s−1 and a quadrupole splitting of 1.60 mm s−1 (Figure 3), suggesting that the two iron centers are indistinguishable under the experimental conditions. The similarity between the Mössbauer parameters of 2 and 3 suggests that the oxidation states of the iron centers remain unchanged upon one-electron reduction of 2, consistent with the ligand-based reduction indicated by the metric parameters of the crystal structures.

Scheme 2.

Scheme 2.

Syntheses of 3–6

Figure 4.

Figure 4.

Crystal structure of 3 with the thermal ellipsoids plotted at the 50% probability level. The [K(THF)6]+ counterion and all hydrogen atoms are omitted for clarity.

Treating 2 with 2 equiv of KC8 leads to the formation of cluster 4 as a dark-purple crystalline solid in 76% yield (Scheme 2). As shown in Figure 5, each molecule of 4 possesses a crystallographically imposed inversion center. The two potassium ions in each molecule are sandwiched between two xylyl groups of the pda ligands in a double η6 fashion and further coordinated by a bridging sulfide and a THF molecule. The Xylpda ligands display long Cα–N distances [1.385(8)–1.403(7) Å; Table 1] consistent with Xylpda2−, suggesting that the second reduction events for 2 are also ligand-based. The Fe–Fe distances in 4 [avg. of the two independent molecules: 2.756(8) Å] are slightly longer than those in 2 and 3 [2.6865(8) Å in 3 and 2.633(1) Å in 2], while the Fe–S distances are similar among all three compounds [avg. 2.251(3) Å in 4 vs avg. 2.241(3) Å in 3 and avg. 2.189(4) Å in 2]. The observed solution magnetic moment of 1.6 μB (Table 2) for 4 at 298 K in THF-d8 suggests a strong antiferromagnetic coupling. The Mössbauer spectra of 4 display one quadrupole doublet (δ = 0.26 mm s−1 and |ΔEQ| = 1.68 mm s−1; Figure 3), similar to those of 2 and 3. All results are consistent with ligand-based reduction, in which the metal oxidation states remain unchanged.

Figure 5.

Figure 5.

Crystal structure of 4 with the thermal ellipsoids plotted at the 50% probability level. All hydrogen atoms are omitted for clarity.

Complex 4 can be further reduced by 1 equiv of KC8, affording a dark-brown [2Fe–2S]+ cluster 5 in 81% yield (Scheme 2). In the crystal structure of 5 (Figure 6), the [Fe2(μ-S)2(Xylpda)2]3− fragment is associated with three potassium cations to balance the charge. One of the three potassium ions (K3) is coordinated by a bridging sulfide (S2), three THF molecules, and the diamide moiety (N3–C23–C28–N4) of one Xylpda ligand in an η4 fashion. The other two potassium ions (K1 and K2) are bound to the other bridging sulfide (S1) with a K1–S1–K2 angle of 175.34(6)°. In addition, K1 is sandwiched between two xylyl groups of the pda ligands in a double η6 fashion and coordinated by an additional THF molecule, whereas K2 is sandwiched between an η6-xylyl group and an η2-xylyl group and further coordinated by a THF molecule. Furthermore, K2 and C3 in each molecule of 5 are bonded to two adjacent molecules of 5 through C3–K2 interactions, leading to the formation of a 1D coordination polymer in the crystal lattice (Figure S24). The Cα–N bond distances of the Xylpda ligands are similar to those of 4 (Table 1), consistent with iron-based reduction. The Fe–S distances in 5 [avg. 2.248(2) Å] are similar to those in 4 [avg. 2.251(2) Å], whereas the Fe–Fe distance shortens slightly upon going from 4 [avg. 2.756(8) Å] to 5 [avg. 2.709(1) Å]. Two quadrupole doublets are observed in the Mössbauer spectrum of 5 (δ = 0.33 mm s−1 and |ΔEQ| = 1.84 mm s−1; δ = 0.62 mm s−1 and |ΔEQ| = 2.84 mm s−1; Figure 3), which suggests a mixed-valence Fe3+–Fe2+ cluster with two different iron environments, i.e., no delocalization on the Mössbauer time scale.

Figure 6.

Figure 6.

X-ray structure of one repeating unit of 5, with the thermal ellipsoids plotted at the 50% probability level. The minor component of the two-site disordering and all hydrogen atoms are omitted for clarity.

Although the cyclic voltammetry experiment suggests that the diferrous cluster from a further reduction of 5 may have limited chemical stability, we next attempted to prepare and isolate the diferrous cluster for structural characterization. Reducing 4 using 2 equiv of KC8 gives the orange diferrous [2Fe–2S]0 cluster 6 in 62% yield (Scheme 2). As shown in Figure 7, X-ray crystallographic analysis shows that, unlike 5, compound 6 has a nonpolymeric structure and is formulated as K4(THF)6Fe2(μ-S)2(Xylpda)2 in the solid state. There is a crystallographically imposed 2-fold rotation axis going through the two bridging sulfur atoms. Each bridging sulfide of the [Fe2(μ-S)2(Xylpda)2]4− unit is further bonded to a pair of symmetry related potassium cations. Each of the potassium ions bound to S1 is further coordinated by one THF molecule and the xylyl rings of the pda ligands in a double η6 fashion, whereas each of the potassium ions bound to S2 is further coordinated by two THF molecules and the N–C–C–N moiety of a pda ligand. The Cα–N bond lengths of the Xylpda ligands display negligible changes compared to those in 4 (Table 1), consistent with iron-based reduction. The Fe–S distance in 6 [avg. 2.2812(7) Å] is slightly longer than those in 4 and 5, while the Fe–Fe distance in 6 [2.7306(6) Å] is longer than that in 5 and shorter than that in 4. Compound 6 slowly decomposes in solution but is stable in the solid state for weeks at −35 °C. During the crystallization of 6, we observed a few crystals of [Fe(Xylpda)]2 (7; Figure S25), which might result from the decomposition of 6. The zero-field Mössbauer spectrum of 6 shows one quadrupole doublet with δ = 0.70 mm s−1 and |ΔEQ| = 2.61 mm s−1 (Figure 3), indicating the presence of two equivalent high-spin Fe2+ sites. The Mössbauer parameters of 5 and 6 are close to those observed for the reduced and superreduced biological and synthetic [2Fe–2S] clusters.1534

Figure 7.

Figure 7.

Crystal structure of 6 with the thermal ellipsoids plotted at the 50% probability level. All hydrogen atoms are omitted for clarity.

Prompted by compound 2’s capability of storing up to four electrons, we assessed the catalytic activity of compound 2 toward a multielectron process, namely, N2 reduction. The results are summarized in Table 3. The reaction with 500 equiv of KC8 and Me3SiCl per iron center in THF generated 17 ± 2 equiv of N(SiMe3)3 per iron center in 24 h (entry 1). The reactions in dimethyl ether (DME) and dioxane under otherwise identical conditions gave 31 ± 1 and 32 ± 2 equiv of N(SiMe3)3 per iron center, respectively (entries 2 and 3). With 1000 equiv of KC8 and Me3SiCl per iron center, the reaction in DME gave 36 ± 3 equiv of N(SiMe3)3 per iron center (entry 4), whereas using dioxane as the solvent increased the amount of N(SiMe3)3 to 51 ± 2 equiv per iron center (entry 5). When the amounts of KC8 and Me3SiCl are increased to 1500 equiv per ion center, the reaction in dioxane produced 60 ± 2 and 88 ± 2 equiv of N(SiMe3)3 per iron center in 24 and 72 h, respectively (entries 6 and 7). Compared to the iron-based N2 silylation catalysts in the literature,7283 the catalytic performance of 2 is moderate. Several iron–sulfur clusters have been used in N2 reduction chemistry.84 Our study reported herein, however, represents the first example using a [2Fe–2S] cluster in N2 reduction. Last, we examined the catalytic performance of complex 1. The reaction with 1000 equiv of KC8 and Me3SiCl per iron center in dioxane generated 28 ± 2 equiv of N(SiMe3)3 per iron center in 24 h (entry 8). The lower catalytic activity of compound 1 (entry 8 vs entry 5) suggests that the sulfide ligands play an important role in the catalytic reaction. Although the exact role of the sulfide ligands is not well understood, we speculate that they may enhance the stability of the iron-containing species under the reaction conditions. To probe the nature (i.e., homogeneous vs heterogeneous) of the catalysis, we conducted a filtration test, whereby the soluble and insoluble fractions are separated by filtration and then independently tested for catalytic activity (see the Supporting Information for the detailed procedure). The filtrate displayed much higher catalytic activity than the insoluble fraction, suggesting that the active species is soluble. However, the formation of catalytically active small iron nanoparticles cannot be ruled out.

Table 3.

Catalytic N 2 Silylation with Me 3 SiCl and KC 8 Using 2 a

graphic file with name nihms-1713270-t0003.jpg
entry solvent KC8 and Me3SiCl (equiv per iron center) N(SiMe3)3 (equiv per iron center)
1 THF 500 17 ± 2
2 DME 500 31 ± 1
3 dioxane 500 32 ± 2
4 DME 1000 36 ± 3
5 dioxane 1000 51 ± 2
6 dioxane 1500 60 ± 2
7b dioxane 1500 88 ± 2
8c dioxane 1000 28 ± 2
a

N(SiMe3)3 was first converted to NH4+ and then quantified via 1H NMR experiments with 1,3,5-trimethoxybenzene as the internal standard.

b

72 h.

c

1 was used as the catalyst.

CONCLUSION

We have synthesized a diferric [2Fe–2S] cluster (2) supported by two redox-active Xylpda ligands from the oxidation of [Fe(Xylpda)(toluene)] with S8. The cyclic voltammetry studies of the diferric complex [Fe2S2(Xylpda)2] (2) show its rich redox chemistry. Chemical reductions were used to synthesize [Fe2S2(Xylpda)2]n in four additional oxidation states (n = 4–, 3–, 2–, 1–). The combination of crystallographic and spectroscopic studies revealed that the first two reduction events are ligand-based, whereas the third and fourth reduction events are metal-based. The third reduction afforded a mixed-valence Fe3+Fe2+ cluster, where the electron is localized. Finally, the fourth reduction event gave a superreduced diferrous cluster, which showed limited stability. We have also further demonstrated the catalytic activity of 2 toward the catalytic silylation of N2 using Me3SiCl and KC8, producing up to 88 ± 2 equiv of N(SiMe3)3 per iron center. Unfortunately, we were unable to observe N2 binding on any of the [2Fe–2S] species prepared herein. Investigations into a further reduction of the diferrous cluster and the N2 reduction mechanism are underway in our laboratories.

EXPERIMENTAL SECTION

Materials and Methods.

All reactions were carried out in a N2-filled glovebox or using standard Schlenk techniques under N2. Glassware was dried in a 180 °C oven overnight. Diethyl ether, hexanes, pentane, and toluene solvents were dried by refluxing and distilling over sodium under N2. The THF solvent was dried by refluxing and distilling over sodium benzophenone ketyl under N2. C6D6 and THF-d8 were degassed through three consecutive freeze–pump–thaw cycles. All solvents were stored over 3 Å molecular sieves prior to use. Unless otherwise noted, all NMR spectra were recorded on an Agilent DD2 600 MHz spectrometer at 25 °C. Chemical shifts are referenced to the solvent signals. Solution magnetic moments were measured at 25 °C by the method originally described by Evans71 with stock and experimental solutions containing a known amount of a cyclohexane standard. Elemental analyses were carried out by ANALEST at the University of Toronto. Cyclic voltammograms were recorded at 295 K with a 600E Series electrochemical analyzer/workstation. Unless otherwise noted, all chemicals were purchased from commercial sources and used as received. The 1H NMR data for 36, where no coupling was resolved because of paramagnetic broadening, are presented in the following format: chemical shift (the peak width at half-height in hertz, integration value, and partial assignment based on integration). Although the crystal structures of 36 contain several coordinating and/or lattice THF molecules, the loss of THF molecules occurs under vacuum based on the 1H NMR and elemental analysis results. Therefore, the number of moles and percentage yields of these compounds were calculated using the formula provided in the elemental analysis section.

57Fe Mössbauer Spectroscopy.

All measurements for 57Fe Mössbauer spectroscopy were performed using nonenriched solids of the as-isolated complexes. All samples were prepared in an inert-atmosphere glovebox equipped with a liquid-nitrogen fill port to enable sample freezing to 77 K within the glovebox. Each sample was loaded into a Delrin Mössbauer sample cup for measurements and loaded under liquid nitrogen. Low-temperature 57Fe Mössbauer measurements were performed using a See Co. MS4 Mössbauer spectrometer integrated with a Janis SVT-400T He/N2 cryostat for measurements at 80 K. Isomer shifts were determined relative to α-iron at 298 K. All Mössbauer spectra were fit using the program WMoss (See Co). Errors of the fit analyses were the following: δ ± 0.02 mm s−1 and ΔEQ ± 3%. For multicomponent fits, the quantitation errors were ±3% (e.g., 50 ± 3%).

1. To a solid mixture of N,N′-bis(2,6-dimethylphenyl)-phenylenediamine (1.58 g, 3.00 mmol) and Fe(HMDS)2 (1.13 g, 3.00 mmol) was added 15 mL of toluene. The reaction mixture was stirred at room temperature overnight, yielding a red-purple solution, and then filtered through Celite. The filtrate was concentrated to dryness to afford a metallic green crystalline solid of 1 (1.28 g, 92%). Crystals suitable for X-ray crystallography were obtained by cooling a concentrated diethyl ether solution at −35 °C. 1H NMR (600 MHz, C6D6): δ 7.31 (m, 6H, overlapping, Xyl-H), 6.74 (dd, J = 6.5 and 3.2 Hz, 2H, Ar-H), 6.51 (dd, J = 6.3 and 3.3 Hz, 2H, Ar-H), 5.11 (td, J = 5.8 and 0.9 Hz, 1H, η6-tol-H), 4.86 (t, J = 6.0 Hz, 2H, η6-tol-H), 4.78 (d, J = 6.1 Hz, 2H, η6-tol-H), 2.17 (s, 12H, Xy-CH3), 1.74 (s, 3H, η6-tol-CH3). 13C NMR (151 MHz, C6D6): δ 156.18 (Xyl-C), 149.80 (Ar-C), 132.77 (Xyl-C), 128.50 (Xyl-C), 125.43 (Xyl-C), 119.73 (Ar-C), 111.95 (Ar-C), 95.31 (η6-tol-C), 83.17 (η6-tol-C), 83.10 (η6-tol-C), 82.04 (η6-tol-C), 19.33 (η6-tol-CH3), 17.99 (Xy-CH3). Anal. Calcd for C29H30N2Fe: C, 75.33; H, 6.54; N, 6.06. Found: C, 74.33; H, 6.64; N, 6.20.

2. To a solution of 1 (924.8 mg, 2.00 mmol) in 10 mL of THF was added S8 (64.1 mg, 0.25 mmol) as a solid. The reaction mixture was stirred at room temperature overnight and then filtered through Celite. Volatiles were removed under vacuum, leaving a black-purple solid. The solid was stirred with 5 mL of hexanes for 1 h and then collected on a fritted funnel, which was washed with hexanes (3 × 1 mL) and dried under a high vacuum (628.6 mg, 78%). Crystals suitable for X-ray crystallography were obtained by cooling a concentrated diethyl ether solution at −35 °C. 1H NMR (600 MHz, THF-d8): δ 13.99 (d, J = 6.4 Hz, 4H, Ar-H), 7.45 (d, J = 7.3 Hz, 8H, Xyl-H), 7.01 (m, 4H, Ar-H), 4.68 (s, 24H, CH3), 4.00 (t, J = 7.2 Hz, 4H, Xyl-H). The Evans method (298 K, THF-d8): μeff = 1.2 μB. Anal. Calcd for C44H44N4S2Fe2: C, 65.68; H, 5.51; N, 6.96. Found: C, 65.14; H, 5.50; N, 6.86.

3. To a solution of 2 (201.2 mg, 0.25 mmol) in 5 mL of THF was added a suspension of KC8 (33.8 mg, 0.25 mmol) in 3 mL of THF. The reaction mixture was stirred for 1 h at room temperature and then filtered through Celite. The filtrate was concentrated to ~2 mL, top-layered with 5 mL of pentane, and cooled to −35 °C overnight to yield dark-purple crystals that were suitable for X-ray crystallography. The supernatant was decanted off, and the solid was washed with diethyl ether (3 × 1 mL) and pentane (3 × 1 mL) and then dried under vacuum to afford 3 (193.1 mg, 73%). 1H NMR (600 MHz, THF-d8): δ 16.37 (Δ1/2 = 18.71 Hz, 8H, Xyl-H), 7.61 (Δ1/2 = 87.47 Hz, 24H, CH3), −1.44 (Δ1/2 = 19.56 Hz, 4H), −16.47 (Δ1/2 = 62.60 Hz, 4H), −22.99 (Δ1/2 = 90.67 Hz, 4H). The Evans method (298 K, THF-d8): μeff = 1.9 μB. Anal. Calcd for C44H44N4S2Fe2K·(C4H8O)3: C, 63.45; H, 6.47; N, 5.29. Found: C, 63.39; H, 6.24; N, 5.23.

4. To a solution of 2 (402.4 mg, 0.50 mmol) in 5 mL of THF was added a suspension of KC8 (135.2 mg, 1.00 mmol) in 3 mL of THF. The reaction mixture was stirred for 1 h at room temperature and then filtered through Celite. The filtrate was concentrated to ~2 mL, top-layered with 5 mL of pentane, and cooled to −35 °C overnight to yield dark-purple crystals that were suitable for X-ray crystallography. The supernatant was decanted off, and the solid was washed with diethyl ether (3 × 1 mL) and pentane (3 × 1 mL) and then dried under vacuum to afford 4 (363.2 mg, 76%). 1H NMR (600 MHz, THF-d8): δ 11.90 (Δ1/2 = 18.23 Hz, 8H, Xyl-H), 7.25 (Δ1/2 = 91.02 Hz, 24H, CH3), 0.97 (Δ1/2 = 59.72 Hz, 4H), −0.22 (Δ1/2 = 19.42 Hz, 4H), −2.50 (Δ1/2 = 59.59 Hz, 4H). The Evans method (298 K, THF-d8): μeff = 1.6 μB. Anal. Calcd for C44H44N4S2Fe2K2·C4H8O: C, 60.37; H, 5.49; N, 5.87. Found: C, 60.05; H, 5.52; N, 5.84.

5. To a solution of 4 (238.7 mg, 0.25 mmol) in 5 mL of THF was added a suspension of KC8 (33.8 mg, 0.25 mmol) in 3 mL of THF. The reaction mixture was stirred for 1 h at room temperature, yielding a brown solution, and then filtered through Celite. The filtrate was concentrated to ~2 mL, top-layered with 5 mL of pentane, and cooled to −35 °C overnight to yield brown crystals that were suitable for X-ray crystallography. The supernatant was decanted off, and the solid was washed with diethyl ether (3 × 1 mL) and pentane (3 × 1 mL) and then dried under vacuum to afford 5·THF (201.6 mg, 81%). 1H NMR (600 MHz, THF-d8): δ 12.24 (Δ1/2 = 111.5 Hz, 8H), 8.10 (Δ1/2 = 1229.87 Hz, 24H, CH3), −0.10 (Δ1/2 = 164.90 Hz, 4H), −1.31 (Δ1/2 = 92.42 Hz, 4H). The Evans method (298 K, THF-d8): μeff = 2.3 μB. Anal. Calcd for C44H44N4S2Fe2K3·C4H8O: C, 58.00; H, 5.27; N, 5.64. Found: C, 57.90; H, 5.02; N, 5.75.

6. To a solution of 4 (238.7 mg, 0.25 mmol) in 5 mL of THF was added a suspension of KC8 (67.6 mg, 0.50 mmol) in 3 mL of THF. The reaction mixture was stirred for 1 h at room temperature, yielding an orange–brown solution, and then filtered through Celite. The filtrate was concentrated to ~2 mL, top-layered with 5 mL of pentane, and cooled to −35 °C overnight to yield orange crystals that were suitable for X-ray crystallography. The supernatant was decanted off, and the solid was washed with diethyl ether (3 × 1 mL) and pentane (3 × 1 mL) and then dried under vacuum to afford 6 (205.8 mg, 62%). 1H NMR (600 MHz, THF-d8): δ 7.87 (Δ1/2 = 51.34 Hz, 4H), 7.25 (Δ1/2 = 131.14 Hz, 12H, CH3), 6.77 (Δ1/2 = 20.82 Hz, 4H), 6.43 (Δ1/2 = 7.47 Hz, 4H), 6.20 (Δ1/2 = 92.92 Hz, 4H), 1.98 (Δ1/2 = 183.01 Hz, 12H, CH3), 0.50 (Δ1/2 = 15.10 Hz, 4H). The Evans method (298 K, THF-d8): μeff = 1.7 μB. Because of the limited stability and high air sensitivity of complex 6, we were unable to obtain satisfactory elemental analysis results. The best elemental analysis results obtained so far are provided below. Anal. Calcd for C44H44N4S2Fe2K4·(C4H8O)5: C, 58.16; H, 6.41; N, 4.24. Found: C, 56.55; H, 6.37; N, 3.87.

Standard Procedure for the Catalytic Conversion of N2 to N(TMS)3 Using 2.

KC8 (90.0 mg, 666 μmol) was suspended in dioxane (1.8 mL) in a 2-dram vial equipped with a stir bar. Me3SiCl (85.0 μL, 666 μmol) was added, followed by a solution of 2 in dioxane (200 μL, 3.3 mM, 0.66 μmol). The reaction mixture was stirred at room temperature for 24 h and then filtered through a plug of Celite. To the filtrate was added a solution of HCl in Et2O (2 M, 5 mL). The resulting mixture was stirred for 1 h before removal of the solvent under reduced pressure. The resulting white solids were dissolved in DMSO-d6 with 1,3,5-trimethoxybenzene as an internal standard. The ammonium was quantified using 1H NMR spectroscopy. The experiments were performed in triplicate. The control experiment was performed in the absence of 2 under otherwise identical conditions. The control experiment gave only 3 equiv of N(TMS)3 with respect to the 1 equiv of 2 that is absent (i.e., 1.5 equiv per iron center that is absent).

Supplementary Material

SI

ACKNOWLEDGMENTS

We thank the Natural Sciences and Engineering Research Council of Canada (grants to D.S.) and the National Institutes of Health (Grant R01GM111480 to M.L.N.) for funding. Q.L. acknowledges support by the Department of Chemistry, University of Toronto, with a Special Opportunity Graduate Travel Fellowship.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.1c00683.

NMR and Mössbauer spectra, cyclic voltammetry data, X-ray crystallographic data and plots, and further experimental details (PDF)

Accession Codes

CCDC 20682962068302 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.inorgchem.1c00683

The authors declare no competing financial interest.

Contributor Information

Qiuming Liang, Davenport Chemical Research Laboratories, Department of Chemistry, University of Toronto, Toronto, Ontario M5S 3H6, Canada.

Datong Song, Davenport Chemical Research Laboratories, Department of Chemistry, University of Toronto, Toronto,Ontario M5S 3H6, Canada.

REFERENCES

  • (1).Beinert H; Holm RH; Münck E Iron-Sulfur Clusters: Nature’s Modular, Multipurpose Structures. Science 1997, 277, 653–659. [DOI] [PubMed] [Google Scholar]
  • (2).Solomon EI; Xie X; Dey A Mixed valent sites in biological electron transfer. Chem. Soc. Rev 2008, 37, 623–638. [DOI] [PubMed] [Google Scholar]
  • (3).Hu Y; Ribbe MW Biosynthesis of Nitrogenase FeMoco. Coord. Chem. Rev 2011, 255, 1218–1224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).Holm RH; Lo W Structural Conversions of Synthetic and Protein-Bound Iron-Sulfur Clusters. Chem. Rev 2016, 116, 13685–13713. [DOI] [PubMed] [Google Scholar]
  • (5).Burén S; Jiménez-Vicente E; Echavarri-Erasun C; Rubio LM Biosynthesis of Nitrogenase Cofactors. Chem. Rev 2020, 120, 4921–4968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).Crouse BR; Meyer J; Johnson MK Spectroscopic Evidence for a Reduced Fe2S2 Cluster with a S = 9/2 Ground State in Mutant Forms of Clostridium pasteurianum 2Fe Ferredoxin. J. Am. Chem. Soc 1995, 117, 9612–9613. [Google Scholar]
  • (7).Achim C; Golinelli M-P; Bominaar E; Meyer J; Münck E Mössbauer Study of Cys56Ser Mutant 2Fe Ferredoxin from Clostridium Pasteurianum: Evidence for Double Exchange in an [Fe2S2]+ Cluster. J. Am. Chem. Soc 1996, 118, 8168–8169. [Google Scholar]
  • (8).Achim C; Bominaar E; Meyer J; Peterson J; Münck E Observation and Interpretation of Temperature-Dependent Valence Delocalization in the [2Fe-2S]+ Cluster of a Ferredoxin from Clostridium pasteurianum. J. Am. Chem. Soc 1999, 121, 3704–3714. [Google Scholar]
  • (9).Subramanian S; Duin EC; Fawcett SEJ; Armstrong FA; Meyer J; Johnson MK Spectroscopic and Redox Studies of Valence-Delocalized [Fe2S2]+ Centers in Thioredoxin-like Ferredoxins. J. Am. Chem. Soc 2015, 137, 4567–4580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Bak DW; Elliott SJ Alternative FeS cluster ligands: tuning redox potentials and chemistry. Curr. Opin. Chem. Biol 2014, 19, 50–58. [DOI] [PubMed] [Google Scholar]
  • (11).Zuris JA; Halim DA; Conlan AR; Abresch EC; Nechushtai R; Paddock ML; Jennings PA Engineering the Redox Potential over a Wide Range within a New Class of FeS Proteins. J. Am. Chem. Soc 2010, 132, 13120–13122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).Kimura S; Kikuchi A; Senda T; Shiro Y; Fukuda M Tolerance of the Rieske-type [2Fe-2S] cluster in recombinant ferredoxin BphA3 from Pseudomonas sp. KKS102 to histidine ligand mutations. Biochem. J 2005, 388, 869–878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Holden HM; Jacobson BL; Hurley JK; Tollin G; Oh BH; Skjeldal L; Chae YK; Cheng H; Xia B; Markley JL Structure-function studies of [2Fe-2S] ferredoxins. J. Bioenerg. Biomembr 1994, 26, 67–88. [DOI] [PubMed] [Google Scholar]
  • (14).Liu J; Chakraborty S; Hosseinzadeh P; Yu Y; Tian S; Petrik I; Bhagi A; Lu Y Metalloproteins Containing Cytochrome, Iron-Sulfur, or Copper Redox Centers. Chem. Rev 2014, 114, 4366–4469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Rao PV; Holm RH Synthetic Analogues of the Active Sites of Iron-Sulfur Proteins. Chem. Rev 2003, 104, 527–560. [DOI] [PubMed] [Google Scholar]
  • (16).Mascharak PK; Papaefthymiou GC; Frankel RB; Holm RH Evidence for the localized iron(III)/iron(II) oxidation state configuration as an intrinsic property of [Fe2S2(SR)4]3− clusters. J. Am. Chem. Soc 1981, 103, 6110–6116. [Google Scholar]
  • (17).Beardwood P; Gibson JF Iron-sulfur dimers with benzimidazolate-thiolate, -phenolate or bis(benzimidazolate) terminal chelating ligands. Models for Rieske-type proteins. J. Chem. Soc., Dalton Trans 1992, 2457–2466. [Google Scholar]
  • (18).Ding X-Q; Bill E; Trautwein AX; Winkler H; Kostikas A; Papaefthymiou V; Simopoulos A; Beardwood P; Gibson JF Exchange interactions, charge delocalization, and spin relaxation in a mixed-valence di-iron complex studied by Mössbauer spectroscopy. J. Chem. Phys 1993, 99, 6421–6428. [Google Scholar]
  • (19).Albers A; Bayer T; Demeshko S; Dechert S; Meyer F A Functional Model for the Rieske Center: Full Characterization of a Biomimetic N-Ligated [2Fe-2S] Cluster in Different Protonation States. Chem. - Eur. J 2013, 19, 10101–10106. [DOI] [PubMed] [Google Scholar]
  • (20).Ballmann J; Albers A; Demeshko S; Dechert S; Bill E; Bothe E; Ryde U; Meyer F A Synthetic Analogue of Rieske-Type [2Fe-2S] Clusters. Angew. Chem., Int. Ed 2008, 47, 9537–9541. [DOI] [PubMed] [Google Scholar]
  • (21).Fuchs MGG; Dechert S; Demeshko S; Meyer F N-Coordinated [2Fe-2S] and [4Fe-4S] Clusters: Synthesis, Structures and Spectroscopic Characterisation. Eur. J. Inorg. Chem 2010, 2010, 3247–3251. [Google Scholar]
  • (22).Ballmann J; Dechert S; Demeshko S; Meyer F Tuning Electronic Properties of Biomimetic [2Fe-2S] Clusters by Ligand Variations. Eur. J. Inorg. Chem 2009, 2009, 3219–3225. [Google Scholar]
  • (23).Bergner M; Roy L; Dechert S; Neese F; Ye S; Meyer F Ligand Rearrangements at Fe/S Cofactors: Slow Isomerization of a Biomimetic [2Fe-2S] Cluster. Angew. Chem., Int. Ed 2017, 56, 4882–4886. [DOI] [PubMed] [Google Scholar]
  • (24).Saouma CT; Kaminsky W; Mayer JM Decomposition of a mixed-valence [2Fe-2S] cluster gives linear tetra-ferric and ferrous clusters. Polyhedron 2013, 58, 60–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Albers A; Demeshko S; Dechert S; Bill E; Bothe E; Meyer F The Complete Characterization of a Reduced Biomimetic [2Fe-2S] Cluster. Angew. Chem., Int. Ed 2011, 50, 9191–9194. [DOI] [PubMed] [Google Scholar]
  • (26).Albers A; Demeshko S; Pröpper K; Dechert S; Bill E; Meyer F A Super-Reduced Diferrous [2Fe-2S] Cluster. J. Am. Chem. Soc 2013, 135, 1704–1707. [DOI] [PubMed] [Google Scholar]
  • (27).Kowalska JK; Hahn AW; Albers A; Schiewer CE; Bjornsson R; Lima FA; Meyer F; DeBeer S X-ray Absorption and Emission Spectroscopic Studies of [L2Fe2S2]n Model Complexes: Implications for the Experimental Evaluation of Redox States in Iron-Sulfur Clusters. Inorg. Chem 2016, 55, 4485–4497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Albers A; Demeshko S; Dechert S; Saouma CT; Mayer JM; Meyer F Fast Proton-Coupled Electron Transfer Observed for a High-Fidelity Structural and Functional [2Fe-2S] Rieske Model. J. Am. Chem. Soc 2014, 136, 3946–3954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (29).Saouma CT; Kaminsky W; Mayer JM Protonation and Concerted Proton-Electron Transfer Reactivity of a Bis-Benzimidazolate Ligated [2Fe-2S] Model for Rieske Clusters. J. Am. Chem. Soc 2012, 134, 7293–7296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Bergner M; Dechert S; Demeshko S; Kupper C; Mayer JM; Meyer F Model of the MitoNEET [2Fe-2S] Cluster Shows Proton Coupled Electron Transfer. J. Am. Chem. Soc 2017, 139, 701–707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (31).Saouma CT; Pinney MM; Mayer JM Electron Transfer and Proton-Coupled Electron Transfer Reactivity and Self-Exchange of Synthetic [2Fe-2S] Complexes: Models for Rieske and mitoNEET Clusters. Inorg. Chem 2014, 53, 3153–3161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Yao S; Meier F; Lindenmaier N; Rudolph R; Blom B; Adelhardt M; Sutter J; Mebs S; Haumann M; Meyer K; Kaupp M; Driess M Biomimetic [2Fe-2S] Clusters with Extensively Delocalized Mixed-Valence Iron Centers. Angew. Chem., Int. Ed 2015, 54, 12506–12510. [DOI] [PubMed] [Google Scholar]
  • (33).Reesbeck ME; Rodriguez MM; Brennessel WW; Mercado BQ; Vinyard D; Holland PL Oxidized and reduced [2Fe-2S] clusters from an iron(I) synthon. JBIC, J. Biol. Inorg. Chem 2015, 20, 875–883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Fohlmeister L; Vignesh KR; Winter F; Moubaraki B; Rajaraman G; Pöttgen R; Murray KS; Jones C Neutral diiron(III) complexes with Fe2(μ-E)2 (E = O, S, Se) core structures: reactivity of an iron(I) dimer towards chalcogens. Dalton Trans. 2015, 44, 1700–1708. [DOI] [PubMed] [Google Scholar]
  • (35).Chirila A; Das BG; Kuijpers PF; Sinha V; de Bruin B Non-Noble Metal Catalysis: Molecular Approaches and Reactions; John Wiley & Sons, Ltd., 2018; pp 1–31. [Google Scholar]
  • (36).de Bruin B; Gualco P; Paul ND Ligand Design in Metal Chemistry; John Wiley & Sons, Ltd., 2016; pp 176–204. [Google Scholar]
  • (37).van der Vlugt JI Radical-Type Reactivity and Catalysis by Single-Electron Transfer to or from Redox-Active Ligands. Chem. - Eur. J 2019, 25, 2651–2662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (38).Jacquet J; Desage-El Murr M; Fensterbank L Metal-Promoted Coupling Reactions Implying Ligand-Based Redox Changes. ChemCatChem 2016, 8, 3310–3316. [Google Scholar]
  • (39).Suarez AIO; Lyaskovskyy V; Reek JNH; van der Vlugt JI; de Bruin B Complexes with Nitrogen-Centered Radical Ligands: Classification, Spectroscopic Features, Reactivity, and Catalytic Applications. Angew. Chem., Int. Ed 2013, 52, 12510–12529. [DOI] [PubMed] [Google Scholar]
  • (40).Luca OR; Crabtree R Redox-active ligands in catalysis. Chem. Soc. Rev 2013, 42, 1440–1459. [DOI] [PubMed] [Google Scholar]
  • (41).Munhá RF; Zarkesh RA; Heyduk AF Group transfer reactions of d0 transition metal complexes: redox-active ligands provide a mechanism for expanded reactivity. Dalton Trans. 2013, 42, 3751–3766. [DOI] [PubMed] [Google Scholar]
  • (42).Lyaskovskyy V; de Bruin B Redox Non-Innocent Ligands: Versatile New Tools to Control Catalytic Reactions. ACS Catal. 2012, 2, 270–279. [Google Scholar]
  • (43).Caulton KG Systematics and Future Projections Concerning Redox-Noninnocent Amide/Imine Ligands. Eur. J. Inorg. Chem 2012, 2012, 435–443. [Google Scholar]
  • (44).Praneeth VKK; Ringenberg MR; Ward TR Redox-Active Ligands in Catalysis. Angew. Chem., Int. Ed 2012, 51, 10228–10234. [DOI] [PubMed] [Google Scholar]
  • (45).Kaim W The Shrinking World of Innocent Ligands: Conventionaland Non-Conventional Redox-Active Ligands. Eur. J. Inorg. Chem 2012, 2012, 343–348. [Google Scholar]
  • (46).Chirik PJ Preface: Forum on Redox-Active Ligands. Inorg. Chem 2011, 50, 9737–9740. [DOI] [PubMed] [Google Scholar]
  • (47).Chirik PJ; Wieghardt K Radical Ligands Confer Nobility on Base-Metal Catalysts. Science 2010, 327, 794–795. [DOI] [PubMed] [Google Scholar]
  • (48).Rummelt SM; Zhong H; Korobkov I; Chirik PJ Iron-Mediated Coupling of Carbon Dioxide and Ethylene: Macrocyclic Metallalactones Enable Access to Various Carboxylates. J. Am. Chem. Soc 2018, 140, 11589–11593. [DOI] [PubMed] [Google Scholar]
  • (49).Wilding MJT; Iovan DA; Wrobel AT; Lukens JT; MacMillan SN; Lancaster KM; Betley TA Direct Comparison of C-H Bond Amination Efficacy through Manipulation of Nitrogen-Valence Centered Redox: Imido versus Iminyl. J. Am. Chem. Soc 2017, 139, 14757–14766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (50).Wilding WJT; Iovan DA; Betley TA High-Spin Iron Imido Complexes Competent for C-H Bond Amination. J. Am. Chem. Soc 2017, 139, 12043–12049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51).Hoyt JM; Schmidt VA; Tondreau AM; Chirik P Science 2015, 349, 960. [DOI] [PubMed] [Google Scholar]
  • (52).Tondreau AM; Atienza CCH; Weller KJ; Nye SA; Lewis KM; Delis JGP; Chirik PJ Iron Catalysts for Selective Anti-Markovnikov Alkene Hydrosilylation Using Tertiary Silanes. Science 2012, 335, 567–570. [DOI] [PubMed] [Google Scholar]
  • (53).Iovan DA; Wilding MJT; Baek Y; Hennessy ET; Betley TA Diastereoselective C-H Bond Amination for Disubstituted Pyrrolidines. Angew. Chem., Int. Ed 2017, 56, 15599–15602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Hennessy ET; Betley TA Complex N-Heterocycle Synthesis via Iron-Catalyzed, Direct C-H Bond Amination. Science 2013, 340, 591–595. [DOI] [PubMed] [Google Scholar]
  • (55).Pierpont CG Studies on charge distribution and valence tautomerism in transition metal complexes of catecholate and semiquinonate ligands. Coord. Chem. Rev 2001, 216–217, 99–125. [Google Scholar]
  • (56).Broere DLJ; Plessius R; van der Vlugt JI New avenues for ligand-mediated processes - expanding metal reactivity by the use of redox-active catechol, o-aminophenol and o-phenylenediamine ligands. Chem. Soc. Rev 2015, 44, 6886–6915. [DOI] [PubMed] [Google Scholar]
  • (57).Monakhov KY; van Leusen J; Kögerler P; Zins E-L; Alikhani ME; Tromp M; Danopoulos AA; Braunstein P Linear, Trinuclear Cobalt Complexes with o-Phenylene-bis-Silylamido Ligands. Chem. - Eur. J 2017, 23, 6504–6508. [DOI] [PubMed] [Google Scholar]
  • (58).van der Meer M; Rechkemmer Y; Peremykin I; Hohloch S; van Slageren J; Sarkar B (Electro)catalytic C-C bond formation reaction with a redox-active cobalt complex. Chem. Commun 2014, 50, 11104–11106. [DOI] [PubMed] [Google Scholar]
  • (59).Muresan N; Lu CC; Ghosh M; Peters JC; Abe M; Henling LM; Weyhermöller T; Bill E; Wieghardt K Bis(α-diimine)iron Complexes: Electronic Structure Determination by Spectroscopy and Broken Symmetry Density Functional Theoretical Calculations. Inorg. Chem 2008, 47, 4579–4590. [DOI] [PubMed] [Google Scholar]
  • (60).Khusniyarov MM; Bill E; Weyhermüller T; Bothe E; Harms K; Sundermeyer J; Wieghardt K Characterization of Three Members of the Electron-Transfer Series [Fe(pda)2]n (n = 2-, 1-, 0) by Spectroscopy and Density Functional Theoretical Calculations [pda = Redox Non-innocent Derivatives of N,N′-Bis-(pentafluorophenyl)-o-phenylenediamide(2-, 1-, 0)]. Chem. - Eur. J 2008, 14, 7608–7622. [DOI] [PubMed] [Google Scholar]
  • (61).Khusniyarov MM; Weyhermüller T; Bill E; Wieghardt K Tuning the Oxidation Level, the Spin State, and the Degree of Electron Delocalization in Homo- and Heteroleptic Bis(α-diimine) iron Complexes. J. Am. Chem. Soc 2009, 131, 1208–1221. [DOI] [PubMed] [Google Scholar]
  • (62).Khusniyarov MM; Weyhermüller T; Bill E; Wieghardt K Reversible Electron Transfer Coupled to Spin Crossover in an Iron Coordination Salt in the Solid State. Angew. Chem., Int. Ed 2008, 47, 1228–1231. [DOI] [PubMed] [Google Scholar]
  • (63).Hernán-Gómez A; Rodríguez M; Parella T; Costas M Electrophilic Iron Catalyst Paired with a Lithium Cation Enables Selective Functionalization of Non-Activated Aliphatic C-H Bonds via Metallocarbene Intermediates. Angew. Chem., Int. Ed 2019, 58, 13904–13911. [DOI] [PubMed] [Google Scholar]
  • (64).Janes T; Xu M; Song D Synthesis and reactivity of Li and TaMe3 complexes supported by N,N′-bis(2,6-diisopropylphenyl)-o-phenylenediamido ligands. Dalton Trans. 2016, 45, 10672–10680. [DOI] [PubMed] [Google Scholar]
  • (65).Janes T; Zatsepin P; Song D Reactivity of heavy carbene analogues towards oxidants: a redox active ligand-enabled isolation of a paramagnetic stannylene. Chem. Commun 2017, 53, 3090–3093. [DOI] [PubMed] [Google Scholar]
  • (66).Janes T; Rawson JM; Song D Syntheses and structures of Li, Fe, and Mo derivatives of N,N′-bis(2,6-diisopropylphenyl)-o-phenylenediamine. Dalton Trans. 2013, 42, 10640–10648. [DOI] [PubMed] [Google Scholar]
  • (67).Liang Q; Lin JH; DeMuth JC; Neidig ML; Song D Syntheses and characterizations of iron complexes of bulky o-phenylenediamide ligand. Dalton Trans. 2020, 49, 12287–12297. [DOI] [PubMed] [Google Scholar]
  • (68).Wenderski T; Light KM; Ogrin D; Bott SG; Harlan CJ Pd catalyzed coupling of 1,2-dibromoarenes and anilines: formation of N,N-diaryl-o-phenylenediamines. Tetrahedron Lett. 2004, 45, 6851–6853. [Google Scholar]
  • (69).Andersen RA; Faegri K; Green JC; Haaland A; Lappert MF; Leung W-P; Rypdal K Synthesis of bis[bis(trimethylsilyl)-amido]iron(II). Structure and bonding in M[N(SiMe3)2]2 (M = manganese, iron, cobalt): two-coordinate transition-metal amides. Inorg. Chem 1988, 27, 1782–1786. [Google Scholar]
  • (70).Fehlhammer WP; Stolzenberg H In Comprehensive Organometallic Chemistry; Wilkinson G, Stone FGA, Abel FW, Eds.; Pergamon Press: Oxford, U.K., 1982; pp 515–524. [Google Scholar]
  • (71).Evans DF The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance. J. Chem. Soc 1959, 2003–2005. [Google Scholar]
  • (72).Kim S; Loose F; Chirik PJ Beyond Ammonia: Nitrogen-Element Bond Forming Reactions with Coordinated Dinitrogen. Chem. Rev 2020, 120, 5637–5681. [DOI] [PubMed] [Google Scholar]
  • (73).Tanabe Y; Nishibayashi Y Recent advances in catalytic silylation of dinitrogen using transition metal complexes. Coord. Chem. Rev 2019, 389, 73–93. [Google Scholar]
  • (74).Masero F; Perrin MA; Dey S; Mougel V Dinitrogen Fixation: Rationalizing Strategies Utilizing Molecular Complexes. Chem. - Eur. J 2021, 27, 3892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (75).Araake R; Sakadani K; Tada M; Sakai Y; Ohki Y [Fe4] and [Fe6] hydride clusters supported by phosphines: synthesis, characterization, and application in N2 reduction. J. Am. Chem. Soc 2017, 139, 5596–5606. [DOI] [PubMed] [Google Scholar]
  • (76).Ferreira RB; Cook BJ; Knight BJ; Catalano VJ; García-Serres R; Murray LJ Catalytic Silylation of Dinitrogen by a Family of Triiron Complexes. ACS Catal. 2018, 8, 7208–7212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (77).Piascik AD; Li R; Wilkinson HJ; Green JC; Ashley AE Fe-Catalyzed Conversion of N2 to N(SiMe3)3 via an Fe-Hydrazido Resting State. J. Am. Chem. Soc 2018, 140, 10691–10694. [DOI] [PubMed] [Google Scholar]
  • (78).Yuki M; Tanaka H; Sasaki K; Miyake Y; Yoshizawa K; Nishibayashi Y Iron-Catalysed Transformation of Molecular Dinitrogen into Silylamine under Ambient Conditions. Nat. Commun 2012, 3, 1254–1259. [DOI] [PubMed] [Google Scholar]
  • (79).McWilliams SF; Bill E; Lukat-Rodgers G; Rodgers KR; Mercado BQ; Holland PL Effects of N2 Binding Mode on Iron-Based Functionalization of Dinitrogen to Form an Iron(III) Hydrazido Complex. J. Am. Chem. Soc 2018, 140, 8586–8598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (80).Buratto WR; Murray LJ Coordination Chemistry of Iron-Dinitrogen Complexes with Relevance to Biological N2 FixationReference Module in Chemistry, Molecular Sciences and Chemical Engineering Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier, DOI: 10.1016/B978-0-12-409547-2.14822-X. [DOI] [Google Scholar]
  • (81).Cavaillé A; Joyeux B; Saffon-Merceron N; Nebra N; Fustier-Boutignon M; Mézailles N Triphos-Fe dinitrogen and ′ dinitrogen-hydride complexes: relevance to catalytic N2 reductions. Chem. Commun 2018, 54, 11953–11956. [DOI] [PubMed] [Google Scholar]
  • (82).Li S; Wang Y; Yang W; Li K; Sun H; Li X; Fuhr O; Fenske D Organometallics 2020, 39, 757–766. [Google Scholar]
  • (83).Yin J; Li J; Wang G-X; Yin Z-B; Zhang W-X; Xi Z Dinitrogen Functionalization Affording Chromium Hydrazido Complex. J. Am. Chem. Soc 2019, 141, 4241–4247. [DOI] [PubMed] [Google Scholar]
  • (84).Tanifuji K; Ohki Y Metal-Sulfur Compounds in N2 Reduction and Nitrogenase-Related Chemistry. Chem. Rev 2020, 120, 5194–5251. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES