Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Jan 28.
Published in final edited form as: Organometallics. 2021 Jun 4;40(12):1974–1996. doi: 10.1021/acs.organomet.1c00267

Is the Electrophilicity of the Metal Nitrene the Sole Predictor of Metal-Mediated Nitrene Transfer to Olefins? Secondary Contributing Factors as Revealed by a Library of High-Spin Co(II) Reagents

Anshika Kalra 1, Vivek Bagchi 2, Patrina Paraskevopoulou 3, Purak Das 4, Lin Ai 5, Yiannis Sanakis 6, Grigorios Raptopoulos 7, Sudip Mohapatra 8, Amitava Choudhury 8, Zhicheng Sun 9, Thomas R Cundari 9, Pericles Stavropoulos 10
PMCID: PMC8797515  NIHMSID: NIHMS1736859  PMID: 35095166

Abstract

Recent research has highlighted the key role played by the electron affinity of the active metal-nitrene/imido oxidant as the driving force in nitrene additions to olefins to afford valuable aziridines. The present work showcases a library of Co(II) reagents that, unlike the previously examined Mn(II) and Fe(II) analogues, demonstrate reactivity trends in olefin aziridinations that cannot be solely explained by the electron affinity criterion. A family of Co(II) catalysts (17 members) has been synthesized with the assistance of a trisphenylamido-amine scaffold decorated by various alkyl, aryl, and acyl groups attached to the equatorial amidos. Single-crystal X-ray diffraction analysis, cyclic voltammetry and EPR data reveal that the high-spin Co(II) sites (S = 3/2) feature a minimal [N3N] coordination and span a range of 1.4 V in redox potentials. Surprisingly, the Co(II)-mediated aziridination of styrene demonstrates reactivity patterns that deviate from those anticipated by the relevant electrophilicities of the putative metal nitrenes. The representative L4Co catalyst (−COCMe3 arm) is operating faster than the L8Co analogue (−COCF3 arm), in spite of diminished metal-nitrene electrophilicity. Mechanistic data (Hammett plots, KIE, stereocontrol studies) reveal that although both reagents follow a two-step reactivity path (turnover-limiting metal-nitrene addition to the Cb atom of styrene, followed by product-determining ring-closure), the L4Co catalyst is associated with lower energy barriers in both steps. DFT calculations indicate that the putative [L4Co]NTs and [L8Co]NTs species are electronically distinct, inasmuch as the former exhibits a single-electron oxidized ligand arm. In addition, DFT calculations suggest that including London dispersion corrections for L4Co (due to the polarizability of the tert-Bu substituent) can provide significant stabilization of the turnover-limiting transition state. This study highlights how small ligand modifications can generate stereoelectronic variants that in certain cases are even capable of overriding the preponderance of the metal-nitrene electrophilicity as a driving force.

Graphical Abstract

graphic file with name nihms-1736859-f0001.jpg

INTRODUCTION

The role of aziridines1 as intermediates and end products of synthetic and biological chemistry is hard to overstate. Not only do aziridines afford avenues for further structural development by taking advantage of the energetic content and stereochemical disposition of their strained three-atom ring (ring opening, expansion, or rearrangement),2 but also they constitute valuable functionalities in the framework of several natural products possessing antibiotic or antineoplastic activities.3 In addition to the central role exercised by aziridines as fine chemicals and pharmaceutical agents,4 their contribution to the chemistry of materials has been increasingly recognized,5 especially as key entities for the development and postmodification of polymeric scaffolds.

Synthetic protocols for the generation of aziridines abound, but largely rely on three major methodologies. The cyclization of 1,2-amino precursors constitutes a traditional approach that has been more recently complemented by addition of either C1 sources to imines or electrophilic N1 donors to alkenes.6 The latter “C2 + N1” addition is extensively implemented due to its operational simplicity and availability of a wide range of suitable substrates and catalysts. The N1 donors encompass a variety of nitrene/nitrenoid precursor oxidants such as iminoiodanes (ArI=NR),7 haloamines (RNNaX, X = Cl, Br),8 O/N-substituted hydroxylamines and N-tosyloxycarbamates (RN-(X)–OR′, X = H, leaving group)9 or atom-economical organic azides (RN3).10 As opposed to oxo-transfer chemistry, the corresponding nitrene/nitrenoid transfer relies significantly on the choice of the attendant R group to control the electrophilicity of the active moiety and provide activated (R = SO2R, CO2R, COR, carbamoyl, sulfamoyl) or nonactivated aziridines (R = H, alkyl, aryl, silyl) with differential reactivity.11

A wide range of catalysts has been explored to influence reactivity and selectivity outcomes in nitrene transfer to alkenes, including several organocatalytic12 and metal-mediated processes.13 In the latter case, the presumptive and rather elusive metal-nitrene (M=NR) active species are entities with rich and variable stereoelectronic attributes, inherent and/or ligand induced, whose operation vis-à-vis olefinic substrates is a matter of intense investigation. The variety of transition metals employed, both from the first-row (Mn, Fe, Co, Ni, Cu)1422 and from the heavier platinum-group2325 and coinage elements,26,27 coupled with a range of ancillary ligand frameworks (e.g., porphyrinoids, salens, bis-oxazolines, tetracarboxylate paddlewheels, trispyrazolyl-borates/methanes, polypyridines) is a testament to the vigorous activity in this field and that of the closely related C–H bond amination reactions.28

Among the late 3d transition elements, the case of cobalt is most intriguing, inasmuch as isolable or even putative Co=NR units have been invoked with a variety of oxidation states (from II to V), electronic ground-state spins (S = 0, 1/2, 1, 3/2, 2), and coordination numbers (from 2 to 5).29 The most common configuration is that of diamagnetic Co(III) imidos (S = 0),30 mostly supported by C3 or C2 symmetric ligands. In a handful of cases, open-shell spin-states were observed for Co(III) imidos, as for instance with (trispyrazolylborato)CoIII(NAd) (S = 1, at T > 280 K),31 (dipyrrin)CoIII(NR) (S = 1 for R = Mes; S = 0 or 0 → 2 transition, for R = tBu, 1-Ad, other alkyls),32 [(hmds)2CoIII(NtBu)] (S = 1; hmds = N(SiMe3)2),33 and possibly bimetallic Zr(μ-NMes)CoIII(NMes) (S = 0 → 2 transition, near room temperature).34 None of these compounds have been reported to mediate nitrene-transfer to alkenes. Observable reactivity includes (i) nitrene-transfer to carbon monoxide;30g,31a,35 (ii) insertion of nitrene into ligand-derived carbene residues;30e (iii) formal hydrogen-atom abstraction from a tBu or Mes ligand moiety by open-shell Co=NR, presumably generating an amido Co–NHR unit and a carbon centered radical; the latter can then recombine with the amido,31a dimerize,31b or generate a Co–C bond;31b,32a (iv) intramolecular C–H bond insertion into alkyl azides (source of imido), mediated by (dipyrrin)CoIII(NR), to generate substituted N-heterocycles;32b,c and (v) a rare instance of intermolecular hydrogen-atom abstraction from C–H bonds of various substrates with BDEC–H ≤ 92 kcal mol−1 by [(hmds)2CoIII(NtBu)],33 leading to the corresponding Co(II) amido; the amido can then react with another equivalent of substrate (C–H) to perform either proton transfer (frequently with the concomitant formation of CoII–C organometallics) or formal hydrogen-atom abstraction via stepwise proton/electron transfer or direct HAT, giving rise to Co(I) and substrate dehydrogenation product. In several instances noted above, the carbophilic character of cobalt is notable as a product-determining factor.

Cobalt(II) imidos are more recent additions to the repertoire of cobalt reagents, and encompass both high-spin (S = 3/2)36 and low-spin (S = 1/2)37 cases as two- and four-coordinate compounds, respectively. The high-spin examples have been reported to perform nitrene-transfer to ethylene to afford RN=CH—CH3, presumably due to a [2π + 2π] activation mode. Similarly, certain C(sp)–H and Si–H bonds are activated not via H atom abstraction, but by means of [2π + 2σ] interactions.36b On the other hand, the low-spin Co(II) imidos are unreactive versus alkenes, although they engage in nitrene-transfer and/or nitrene-exchange with O/S with respect to substrates such as CO, PMe3, PhCHO, and CS2.37 Finally, two examples of high-valent Co(IV) and Co(V) bis-imido complexes ([IMes]Co(NDipp)2]0/+), possessing low-spin ground states of S = 1/2 and 0, respectively, proved to be rather unreactive.38 The open-shell Co(IV) congener is the only one that exhibits intramolecular nitrene C–H insertion into the o-Me group of the Mes residue, possibly via an ortho-cobaltation intermediate (Co–C).29

Whereas the catalytic formation of new C–N bonds by means of the isolable cobalt imidos noted above is only rarely observed, the advent of a library of CoII(Por) complexes that give rise to CoIII–nitrenoid radicals [(Por)CoIIINR] or [(Por)CoIII–(NR)2] has provided numerous instances of highly effective catalytic systems for the stereo-, chemo-, and site-selective aziridination of alkenes and amination of C–H bonds.19 Starting with Co(TPP), and electron-deficient analogues, several generations of CoII(Por) reagents with richly decorated porphyrins have been introduced in the past two decades to facilitate the activation of various organic azides, leading to the generation of well characterized low-spin (S = 1/2) CoIIINR moieties, with spin density largely localized on the N atom.19f,39 These relatively long-lived CoIII–nitrene-radical intermediates owe their stability to hydrogen-bonding interactions of the nitrene moieties with porphyrin-appended amido residues (−NHCOR*), which can further introduce and metal-orient chiral auxiliaries via their R* functionality in D2-symmetric overall geometries. Detailed theoretical and experimental studies19f,39 have established that the mode of operation of [(Por)CoIIINR] metalloradicals vis-à-vis C=C or C–H bonds consists of a two-step process: initial formation of a new N–C bond with alkenes and relocation of the spin density on the distal carbon atom (CoIII–N(R)–C–C–) (or formation of a CoIII–NHR amido and a substrate-bound radical via hydrogen-atom abstraction from a C–H bond), followed by an essentially barrierless collapse of the carbon-centered radical with the N atom to generate the product of aziridination (or amination) along with CoII(Por).

More recently, the structurally related [CoIII(TAMLred)] and [CoIII(TAMLsq)] compounds, featuring the tetraamido macrocyclic ligand TAML in its intact reduced form TAMLred and one-electron oxidized variant TAMLsq (sometimes denoted as TAML+•), have been shown to give rise to [CoIII(TAMLq)(NR)2] (S = 1) and [CoIII(TAMLq)(NR)] (S = 1/2), respectively (TAMLq = doubly oxidized, diamagnetic ligand; CoIII site is low-spin, S = 0).40 These cobalt nitrenes have emerged as capable catalysts for the aziridination of largely styrene substrates by imidoiodinanes (PhINNs, PhINTs, PhINTces).41 Their mode of operation is considered to be unique, inasmuch as the turnover-limiting, initial N–C bond formation with styrenes features an asynchronous transition state, encompassing a partial electron-transfer to form a styrenyl radical cation, in turn undergoing a nucleophilic attack by the nitrene lone pair (see below for more details). This initial charge-transfer has also emerged as a central component of the operation of iron(IV) imido species developed by Latour and co-workers for alkene aziridinations, further underscoring the importance of the electron affinity of the metal nitrene as a commonly encountered driving force.42

Finally, of great interest are recently reported Co(II) organoazides,43 which either thermally or photolytically can extrude N2 to give rise to nitrenoids best described as iminyl [CoIIINR] units (R = aryl, alkyl). Although the electronics of these fleeting intermediates are not yet known, crystal structures of such species (R = alkyl) have been determined following N2 expulsion from single crystals of metal azides in the solid state. For R = aryl, the cobalt(II) organoazide can promote nitrene-transfer by means of unisolable [CoIIINR], both intramolecularly ([3 + 2] annulation) and intermolecularly (C–H allylic amination or styrene aziridination in modest yields). The reactivity of the Co(II) aliphatic azides is more complex and includes (i) α-H atom abstraction via the incipient [CoIIINCH2R] to generate the imine (RCH=NH), if strong δ-C–H bonds (sec, prim) are present; (ii) δ-H atom abstraction and amination of relatively weaker δ-C–H bonds (benzylic, tertiary) by the cobalt alkyl azide itself (initial N2 extrusion is not needed), leading to substituted pyrrolidines; and (iii) intramolecular 1,3-dipolar cycloaddition of cobalt-bound CH2=CH(CH2)4N3 to afford 1,2,3-dihydrotriazole.

In the present work, we examine a library of high-spin Co(II) reagents (S = 3/2), supported by a modular trisphenylamido-amine ligand framework, giving rise to a weak equatorial field. Previous DFT calculations44 on one member of this library of reagents indicated that exposure to a nitrene source (PhI = NTs) generates a Co(III)–nitrene radical (CoIIINTs) with a high-spin ground state (S = 5/2). The corresponding doublet and quartet [Co]NTs states lie slightly higher than the sextet by ΔG values of 0.4 and 1.5 kcal mol−1, respectively. The computed spin density for the S = 5/2 state places ~1.1 unpaired electron on the nitrene N atom, and ~3.3 unpaired electrons on Co, with the remaining spin density being distributed to other atoms. Similarly to the low-spin porphyrin-supported CoIII–nitrene-radical (S = 1/2) noted above, the high-spin congener is capable of performing alkene aziridinations in a two-step process (successive formation of two N–C bonds). Remarkably, the computed transition-state barrier (ΔG = 23.4 kcal mol−1 vs CoIIINTs/styrene) for the rate-determining, initial Cβ…NTs bond-forming step of the high-spin system is very similar to that reported for the corresponding low-spin Co(Por) (ΔG = 24.1 kcal mol−1) or Co(AmidoPor) (ΔG = 22.8 kcal mol−1) with respect to CoIIINSO2Ph/styrene.19f The present work significantly enlarges the scope of high-spin Co(II) compounds as nitrene-transfer reagents, and provides insights in their operational characteristics, not only vis-à-vis the reported low-spin (Por)Co(II) paradigms, but also in comparison with the previously examined libraries of Mn(II) and Fe(II) reagents, supported by the same trisphenylamido-amine ligand framework.44 Whereas the nitrene-transfer reactivity of the Mn(II) reagents in alkene aziridinations largely depends on the electrophilicity of the presumptive MnIIINR (S = 3/2) moiety, underscoring the role of the electron affinity of the metal nitrene as a dominant factor, the reactivity of the corresponding, more reactive, Co(II) reagents is affected by additional subtle electronic and steric factors. These most likely arise from the tighter disposition of the reaction cavity, resulting in ligand-coordination flexibility, electronic rearrangement, and secondary stabilizing interactions. In this publication we show that even an otherwise small change in ligand substitution can have a significant effect on nitrene-transfer reactivity in aziridination reactions, occasionally overriding the preponderance of the metal-nitrene electrophilicity as a driving force.

RESULTS AND DISCUSSION

Synthesis and Characterization of New Ligands and Co(II) Complexes.

The family of trisphenylamido-amine ligands (L1H3–L17H3) employed in this study is shown in Figure 1. The majority of these ligands (L1H3–L15H3) have been used and reported in previous studies.4448 They are all derivatives of the common 2,2′,2″-triaminotriphenylamine framework,45 featuring carbonaceous arm substituents (alkyl, aryl acyl). Ligand L16H3 is prepared by methylation of deprotonated (KH) 2,2′,2″-triaminotriphenylamine by MeI in THF, and ligand L17H3 is derived via condensation of the same triamine with the corresponding chiral acyl chloride in the presence of Et3N in dichloromethane. The solid-state structures of these two new ligands (Figure S1) are indicative of their favorable preorganization for metalation, in a cavity that is buttressed by alkyl and acyl arms, respectively.

Figure 1.

Figure 1.

Ligands employed in the present study.

CoII complexes were synthesized with all ligands, by reacting the deprotonated (KH) ligand with anhydrous beads of CoCl2 in THF (alkyl and aryl armed ligands) or N,N-dimethylacetamide (DMA) (acyl armed ligands). A subset of CoII compounds, namely L3Co (3), L5Co (5), L8Co (8a), L9Co (9), L10Co (10), and L13Co (13), has been previously reported in a study that examined the use of these catalysts in controlled radical polymerization of olefins.48 In addition, L8Co (8b) has been explored in conjunction with the L8 Mn and L8Fe congeners, toward establishing metal-dependent trends in catalytic nitrene transfer to olefins.44 Figure 2 depicts representations of the minimal coordination site of each CoII site, derived from single-crystal X-ray diffraction data. In all cases the ligand coordinates in a trigonal pyramidal [N3(amido)Namine] mode, exhibiting various degrees of distortion, although in two instances (7, 17) the coordination to the axial Namine residue can be best described as a long contact. Moreover, compound 11 features a noncoordinating amido residue that has been protonated. The dominant four-coordinate [N3N] pattern is retained as the sole ligand field of seven CoII compounds (4, 5, 8a, 8c, 9, 10, 12). Additional elements of metal coordination, essentially located trans to the axial Namine residue, are observed with all other compounds, and include solvent moieties, especially for compounds crystallized from MeCN (3, 6, 13) and THF (16), as well as carbonyl units (−C(R)=O–CoII) deriving from acyl residues belonging to the ligand (7, 14, 17). In a single case (15), a six-coordinate CoII site arises from the presence of two ligand-derived ether residues in the metal coordination sphere, in addition to the usual [N3N] framework. Importantly, most structures are polymeric, largely due to an intricate network of intermolecular interactions generated by K+ ions. Mononuclear (1, 2, 5, 12, 13) or oligonuclear (8c, 11, 17) compounds (molecular or ionic) are only encountered in a handful of cases. A more detailed description of the structural features of the new CoII compounds is provided below.

Figure 2.

Figure 2.

Minimal coordination of CoII compounds with ligands L1–L17 explored in this study.

CoII Compounds with Acyl-Armed Ligands.

The seven new compounds (4, 6, 7, 8c, 14, 15, 17; Figure 3) that belong in this category are polymeric, with the exception of [K(DMA)3(L8)CoII]2 (8c) and [K2(THF)2K2(L17)4CoII4] (17) that feature a dimeric and tetrameric molecular unit, respectively. The catalytically important {[K(L4)CoII]·Diethyl Ether}n (4) exhibits higher symmetry than all other compounds, consisting of a rigorous 3-fold axis along the Co–Namine direction as well as through K+ ions relating three different molecules in the crystal lattice. This compound is characterized by an exclusive four-coordinate [N3N] ligand field and an open metalated cavity fortified by the three −COtBu arms. The carbonyl residues are positioned exo with respect to the cavity and are further engaged in contacts with K+ ions, inasmuch as each potassium is coordinated by three oxygen (carbonyl) atoms belonging to different molecules, and is also involved in K+–arene π contacts. These structural features are largely retained in the structure of [K(NCMe)(L6)CoII–NCMe]·2MeCN (6), but important deviations also apply, mostly because of the presence of a coordinated MeCN molecule in a trans position versus the Namine (Namine–Co–NMeCN = 178.75(14)°) and the lack of a strict 3-fold crystallographic symmetry. Otherwise, each K+ ion is still coordinated by three carbonyl residues belonging to different molecules in addition to a single MeCN, in lieu of any K+–arene contacts.

Figure 3.

Figure 3.

ORTEP diagrams (from left to right) of {[K(L4)CoII]·Diethyl Ether}n (4), [K(MeCN)(L6)CoII–NCMe]·2MeCN (6), [K2(DMA)4]-[[K(L7)2CoII2]2·2DMA (7), [K2(DMA)3(L8)CoII]2 (8c), [K(DMA)(L14)CoII]·DMA (14), [K(THF)3(L15)CoII]·THF (15), and [K(THF)K(L17)2CoII2]2·3Pentane (17), drawn with 40% thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selective interatomic distances (Å) and angles (deg): 4, Co(1)–N(1) = 2.151(11), Co(1)–N(2) = 1.986(6), Co(1)–[N(2), N(2), N(2)] = 0.32(2) (distance of Co from mean plane) Å, N(2)–Co(1)–N(2) = 117.48(11), N(2)–Co(1)–N(1) = 80.8(2); 6, Co(1)–N(1) = 2.235(3), Co(1)–N(2) = 2.046(4), Co(1)–N(3) = 2.049(4), Co(1)–N(4) = 2.053(4), Co(1)–N(5) = 2.063(4), Co(1)–[N(2), N(3), N(4)] = 0.48(2) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 115.18(15), N(2)–Co(1)–N(3) = 117.17(15), N(3)–Co(1)–N(4) = 111.69(15), N(2)–Co(1)–N(1) = 76.70(14), N(4)–Co(1)–N(1) = 76.49(14), N(3)–Co(1)–N(1) = 76.15(14), N(3)–Co(1)–N(5) = 103.24(15), N(2)–Co(1)–N(5) = 104.55(14), N(4)–Co(1)–N(5) = 102.81(15), N(1)–Co(1)–N(5) = 178.75(14); 7, Co(1)–N(1) = 2.463(4), Co(1)–N(2) = 2.053(4), Co(1)–N(3) = 2.054(5), Co(1)–N(4) = 2.054(4), Co(1)–O(6) = 1.975(4), Co(1)–[N(2), N(3), N(4)] = 0.61(2) (distance of Co from mean plane), Co(2)–N(5) = 2.517(5), Co(2)–N(6) = 2.081(5), Co(2)–N(7) = 2.033(5), Co(2)–N(8) = 2.069(5), Co(2)–O(3) = 1.984(4), Co(2)–[N(6), N(7), N(8)] = 0.67(2) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 109.51(18), N(2)–Co(1)–N(3) = 113.66(17), N(3)–Co(1)–N(4) = 111.52(18), N(2)–Co(1)– N(1) = 73.14(16), N(4)–Co(1)–N(1) = 73.24(16), N(3)–Co(1)–N(1) = 71.83(16), N(3)–Co(1)–O(6) = 103.19(17), N(2)–Co(1)–O(6) = 106.85(17), N(4)–Co(1)–O(6) = 111.92(16), N(1)–Co(1)–O(6) = 174.16(15), N(6)–Co(2)–N(7) = 111.14(19), N(6)–Co(2)–N(8) = 112.01(19), N(7)–Co(2)–N(8) = 106.34(19), N(6)–Co(2)–N(5) = 70.48(19), N(7)–Co(2)–N(5) = 70.69(16), N(8)–Co(2)–N(5) = 71.42(17), N(6)–Co(2)–O(3) = 105.77(19), N(7)–Co(2)–O(3) = 105.15(18), N(8)–Co(2)–O(3) = 116.27(17), N(5)–Co(2)–O(3) = 172.25(16); 8c, Co(1)–N(1) = 2.1422(13), Co(1)–N(2) = 1.9941(14), Co(1)–N(3) = 1.9899(14), Co(1)–N(4) = 1.9872(13), Co(1)–[N(2), N(3), N(4)] = 0.306(4) (distance of Co from mean plane), K(1)–O(1) = 2.7298(14), N(2)–Co(1)–N(4) = 114.79(6), N(2)–Co(1)–N(3) = 118.04(6), N(3)–Co(1)–N(4) = 120.18(6), N(2)–Co(1)–N(1) = 80.83(5), N(4)–Co(1)–N(1) = 81.54(5), N(3)–Co(1)–N(1) = 81.06(5); 14, Co(1)–N(1) = 2.224(3), Co(1)–N(2) = 2.018(4), Co(1)–N(3) = 2.031(4), Co(1)–N(4) = 2.021(4), Co(1)–O(3) = 2.162(3), Co(1)–[N(2), N(3), N(4)] = 0.42(1) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 110.16(14), N(2)–Co(1)–N(3) = 121.11(15), N(3)–Co(1)–N(4) = 115.82(15), N(2)–Co(1)–N(1) = 76.31(14), N(4)–Co(1)–N(1) = 78.82(13), N(3)–Co(1)–N(1) = 78.58(14), N(3)–Co(1)–O(3) = 106.07(14), N(2)–Co(1)–O(3) = 91.35(13), N(4)–Co(1)–O(3) = 108.68(13), N(1)–Co(1)–O(3) = 167.33(12); 15, Co(1)–N(1) = 2.337(17), Co(1)–N(2) = 2.041(17), Co(1)–N(3) = 2.048(18), Co(1)–N(4) = 2.033(17), Co(1)–O(4) = 2.405(13), Co(1)–O(6) = 2.302(14), Co(1)–[N(2), N(3), N(4)] = 0.53(2) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 103.6(7), N(2)–Co(1)–N(3) = 118.1(7), N(3)–Co(1)–N(4) = 117.9(7), N(2)–Co(1)–N(1) = 74.3(6), N(4)–Co(1)–N(1) = 76.4(7), N(3)–Co(1)–N(1) = 73.5(6), N(3)–Co(1)–O(6) = 107.6(6), N(2)–Co(1)–O(6) = 127.7(6), N(4)–Co(1)–O(6) = 73.7(6), N(1)–Co(1)–O(6) = 146.4(6), N(3)–Co(1)–O(4) = 72.9(6), N(2)–Co(1)–O(4) = 89.5(6), N(4)–Co(1)–O(4) = 154.0(6), N(1)–Co(1)–O(4) = 129.3(6), O(4)–Co(1)–O(6) = 80.5(5); 17, Co(1)–N(1) = 2.428(5), Co(1)–N(2) = 2.055(5), Co(1)–N(3) = 2.038(5), Co(1)–N(4) = 2.047(5), Co(1)–O(6) = 1.990(4), Co(1)–[N(2), N(3), N(4)] = 0.58(1) (distance of Co from mean plane), Co(2)–N(5) = 2.375(5), Co(2)–N(6) = 2.049(5), Co(2)–N(7) = 2.026(5), Co(2)–N(8) = 2.100(5), Co(2)–O(3) = 2.027(4), Co(2)–[N(6), N(7), N(8)] = 0.56(1) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 110.2(2), N(2)–Co(1)– N(3) = 118.9(2), N(3)–Co(1)–N(4) = 107.2(2), N(2)–Co(1)–N(1) = 73.94(18), N(4)–Co(1)–N(1) = 74.25(18), N(3)–Co(1)–N(1) = 72.09(18), N(3)–Co(1)–O(6) = 109.47(18), N(2)–Co(1)–O(6) = 99.64(18), N(4)–Co(1)–O(6) = 110.20(18), N(1)–Co(1)–O(6) = 172.96(17), N(6)–Co(2)–N(7) = 112.3(2), N(6)–Co(2)–N(8) = 113.4(2), N(7)–Co(2)–N(8) = 112.71(19), N(6)–Co(2)–N(5) = 74.71(18), N(7)–Co(2)–N(5) = 74.52(19), N(8)–Co(2)–N(5) = 73.18(17), N(6)–Co(2)–O(3) = 97.07(18), N(7)–Co(2)–O(3) = 112.18(19), N(8)–Co(2)–O(3) = 108.06(17), N(5)–Co(2)–O(3) = 171.18(17).

The structure of [K2(DMA)4][K(L7)2CoII2]2·2DMA (7) is organized in a much more complex manner, featuring a “one-dimensional” array of a repeating unit, [Co(1)/Co(2)–K(1)–Co(3)/Co(4)–K(2)]n, connected to identical arrays via lateral links provided by a DMA solvated K(3)/K(4) dimer (K2(DMA)4). Within the repeating unit, the Co(II) sites are arranged in two similar dimers linked via K+ contacts. The dimeric unit is composed of two slightly different Co(II) centers, each featuring the usual [N3N] ligand coordination, but with a long axial Co–Namine interaction (av. 2.49 Å). In addition, each Co(II) is coordinated by an oxygen atom (carbonyl) positioned trans with respect to the axial Co–Namine direction (av. Namine–Co–OC=O = 173.2°, Co–OC=O = 1.98 Å). Importantly, the oxygen atom (carbonyl) coordinated to each CoII center belongs to the ligand surrounding the partner CoII site, hence giving rise to a dimer. The K+ ions interconnecting the dimers in a pseudo 1-D array are coordinated by two carbonyl and, more weakly, two Namido residues all belonging to one dimer, and by only a single carbonyl moiety (O(10)) belonging to the adjacent dimer. A much more simplified version of this structure is adopted by [K(DMA)3(L8)CoII]2 (8c), exhibiting a dimeric structure comprised of two inversion symmetry related [N3N]CoII units connected via carbonyl O atoms of acyl residues to a central K2(μ-DMA)2(DMA)4 core.

Compound [K(DMA)(L14)CoII]·DMA (14) retains the usual [N3N] trigonal-pyramidal coordination but exhibits an additional unique feature, inasmuch as one of the chiral arms generates a seven-member loop by positioning the ester carbonyl in the coordination sphere of CoII, trans to the axial Namine residue (Namine–Co–OC=O = 167.33(12)°, Co–OC=O = 2.162(3) Å). The polymeric nature of the compound arises again due to identical K+ ions, coordinated by one DMA, forming contacts with oxygen residues (amidato carbonyls, MeC(=O)O–) belonging to three different CoII sites in the crystal lattice. Compound [K(THF)3(L15)CoII]·THF (15) is the only six-coordinate species observed, inasmuch as two acyl arms generate five-membered loops that place two ether residues (ROMe) in the coordination sphere of the [N3N]CoII site (Co–O = 2.302(14), 2.405(13) Å; Namine–Co–O = 146.4(6), 129.3(6)°). Identical K+ ions are coordinated by three THF molecules and two acyl residues, each located in neighboring molecules, thus giving rise to pseudo 1-D polymeric structures.

Finally, the structure of [K2(THF)2][K(L17)2CoII2]2 (17) is very similar to that observed for 7 with respect to the formation of two interconnecting dimers, but the K+ ions are organized differently, to afford a molecular (tetranuclear) rather than a polymeric complex. First, two K+ ions link the two dimers in 17, by employing the same contact pattern noted for the single K+ ion connecting the two dimers in 7. Second, the remaining two K+ ions in 17 are terminated by THF molecules, and thus do not provide connections that could generate an 1-D array of repeating tetranuclear units as in 7. Otherwise, the coordination and arrangement of the CoII sites in 17 and 7 is very similar, with somewhat more pronounced contact for the Namine residue (av. Co–Namine = 2.40 Å), and concomitant weaker attachment of the oxygen (carbonyl) moiety (av. Co–OC=O = 2.00 Å), along the axial coordination of CoII sites in 17 versus that of 7.

CoII Compounds with Alkyl-Armed Ligands.

The new methyl-substituted compound 16 and the previously reported isopropyl congener (9)48 are the only members of the alkyl-armed category of Co(II) compounds explored in this study. Compound 16 (Figure 4) demonstrates the familiar [N3N] coordination, but unlike the more bulky isopropyl analogue, it exhibits a five-coordinate CoII site due to the presence of a coordinated THF molecule trans to the Namine residue (Namine–Co–OTHF = 175.75(13)°). The electron-rich alkyl substitution dictates a stronger equatorial ligand field (av. Co–Namido = 2.000 Å) by comparison to all other five-coordinate CoII sites investigated in this study. The polymeric nature of the compound arises by means of a repeating −[Co(1)–K(1)]–sequence, which features K(1) ions engaging in K–Namido and K+–arene contacts with both ligands of adjacent Co(1) sites.

Figure 4.

Figure 4.

ORTEP diagram of [K(L16)CoII–THF]·0.5Pentane (16) drawn with 40% thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selective interatomic distances (Å) and angles (deg): Co(1)–N(1) = 2.226(3), Co(1)–N(2) = 2.007(4), Co(1)–N(3) = 1.990(4), Co(1)–N(4) = 2.004(4), Co(1)–O(1) = 2.207(3), Co(1)–[N(2), N(3), N(4)] = 0.37(2) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 112.70(15), N(2)–Co(1)–N(3) = 114.04(15), N(3)–Co(1)–N(4) = 122.94(15), N(2)–Co(1)–N(1) = 79.40(13), N(4)–Co(1)–N(1) = 78.61(14), N(3)–Co(1)–N(1) = 79.64(14), N(3)–Co(1)–O(1) = 100.45(14), N(2)–Co(1)–O(1) = 104.32(13), N(4)–Co(1)–O(1) = 97.91(14), N(1)–Co(1)–O(1) = 175.75(13).

CoII Compounds with Aryl-Armed Ligands.

Among the seven aryl-supported Co(II) reagents shown in Figure 2 (1, 2, 11, 12 are new; 3, 5, 13 have been previously reported48), only those crystallized from MeCN solutions (3, 13) feature a solvent molecule coordinated to the Co(II) center. All others, crystallized from THF solutions, exhibit four-coordinate [N3N]Co(II) sites devoid of any axial THF residues, in sharp contrast to analogous five-coordinate Mn and Fe reagents previously reported.

Among the four new Co(II) complexes (Figure 5), (L1)Co (1) provides nice green crystals from concentrated THF solutions, but single-crystal specimens (albeit of low quality) were only obtained in the presence of the exceptional K+ binder 2.2.2-cryptand. The resulting ionic complex [K(2.2.2-cryptand)][(L1)CoII]·3THF (1) exhibits a distorted [N3N] coordination with an equatorial ligand field (av. Co–N = 1.927 Å) that is equal or stronger than that demonstrated by similar four-coordinate Co(II) sites supported by aryl substituents (2, 5, 12), presumably due to the electron-rich character of the 4-tBu-substituted phenyl arm. Similarly, the 3,5-tBu2 disubstituted compound (L2)Co (2) proved to be isolable only in the presence of 2.2.2-cryptand, in the form of green crystals of [K(2.2.2-cryptand)][(L2)CoII]·1.5Pentane (2) of marginal quality. Its structure is almost identical with that of 1, with a similarly strong equatorial field (av. Co–N = 1.927 Å). The corresponding 3,5-Me2 disubstituted compound [K(THF)3(L12)CoII]·THF (12) is monomeric and geometrically analogous to 2, with a weaker equatorial field (av. Co–N = 1.956 Å), but, as opposed to 1 and 2, can be isolated without the assistance of 2.2.2-cryptand. The K+ ion in 12 is supported by three THF molecules and a host of contacts with aromatic moieties and N atom residues.

Figure 5.

Figure 5.

ORTEP diagrams (from left to right) of [K(2.2.2-cryptand)][(L1)CoII]·3THF (1), [K(2.2.2-cryptand)][(L2)CoII]·1.5Pentane (2), [K(THF)(L11H)CoII–OH]2 (11), and [K(THF)3(L12)CoII]·THF (12) drawn with 40% thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selective interatomic distances (Å) and angles (deg): 1, Co(1)–N(1) = 2.102(18), Co(1)–N(2) = 1.946(18), Co(1)–N(3) = 1.92(2), Co(1)–N(4) = 1.915(19), Co(1)–[N(2), N(3), N(4)] = 0.17(2) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 122.7(8), N(2)–Co(1)–N(3) = 118.0(8), N(3)–Co(1)–N(4) = 117.0(8), N(2)–Co(1)–N(1) = 84.0(8), N(4)–Co(1)–N(1) = 85.1(8), N(3)–Co(1)–N(1) = 85.6(8); 2, Co(1)–N(1) = 2.085(8), Co(1)–N(2) = 1.921(7), Co(1)–N(3) = 1.933(8), Co(1)–N(4) = 1.926(7), Co(1)–[N(2), N(3), N(4)] = 0.14(1) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 120.4(3), N(2)–Co(1)–N(3) = 118.2(3), N(3)–Co(1)–N(4) = 119.8(3), N(2)–Co(1)–N(1) = 85.9(3), N(4)–Co(1)–N(1) = 86.0(3), N(3)–Co(1)–N(1) = 85.5(3); 11, Co(1)–N(1) = 2.260(5), Co(1)–N(2) = 1.956(5), Co(1)–N(3) = 1.939(5), Co(1)–O(1) = 1.947(4), K(1)–O(1) = 2.675(5), K(1)–O(2) = 2.677(5), Co(1)–[N(2), N(3), N(4)] = 0.62(2) (distance of Co from mean plane), N(2)–Co(1)–N(3) = 118.7(2), N(2)–Co(1)–N(1) = 81.6(2), N(3)–Co(1)–N(1) = 80.1(2), N(1)–Co(1)–O(1) = 139.41(18), N(2)–Co(1)–O(1) = 108.5(2), N(3)–Co(1)–O(1) = 122.9(2), Co(1)–O(1)–K(1) = 113.83(18), O(1)–K(1)–O(2) = 116.95(15); 12, Co(1)–N(1) = 2.121(7), Co(1)–N(2) = 1.950(8), Co(1)–N(3) = 1.964(8), Co(1)–N(4) = 1.955(7), Co(1)–[N(2), N(3), N(4)] = 0.20(1) (distance of Co from mean plane), N(2)–Co(1)–N(4) = 119.2(3), N(2)–Co(1)–N(3) = 121.5(3), N(3)–Co(1)–N(4) = 116.2(3), N(2)–Co(1)–N(1) = 84.3(3), N(4)–Co(1)–N(1) = 84.0(2), N(3)–Co(1)–N(1) = 84.0(3).

Although both the 3,5-Me2 and 2,6-Me2 disubstituted compounds 12 and 13, respectively, can be isolated and characterized, the corresponding 2,4,6-Me3 trisubstituted species (L11)CoII has proven to be difficult to synthesize, apparently due to extreme sensitivity to even traces of water. In contrast, the analogous [K(THF)3(L11)MnII–THF] can be readily prepared.44 Indeed, after initial formation of a deep green species in the reaction of deprotonated L11H3 and CoCl2 in THF, the color soon fades and ligand can be recovered intact along with separation of blue Co(OH)2. In one instance, under scrupulous water exclusion, a few green crystals of [K(THF)(L11H)CoII–OH)]2 (11) have been isolated, amounting to a species that can be viewed as the formal product of water addition to (L11)CoII. Indeed, 11 features protonation and dissociation of one nitrogen residue from the equatorial field, with concomitant formation of a CoII–OH moiety. The hydroxide is further coordinated by two K(THF)+ ions in an overall dimeric structure that connects two inversion-related (L11H)CoII–OH monomers by means of a K2(OH)2 rhomb.

Structures Featuring Ligand Rearrangement.

In two instances, we have isolated a few compounds that exhibit a characteristic oxidative ligand rearrangement in the presence of traces of dioxygen or one-electron oxidants. Similarly reorganized compounds, featuring electron-donor substituents, have been previously studied in our lab and attributed to the formation of an incipient aminyl radical.45b,49 Indeed, the electron-rich (L2)CoII (2) is highly sensitive to oxidative rearrangement, and provided two crystallographically characterized species, [K(THF)3(L2re)CoII–THF] (2b) and [(L2re,ox)-CoII–THF]·0.5 Pentane (2c) (Scheme 1 and Figure S2; bonds broken in 2 and formed in 2b and 2c are shown in red), formed in comparable amounts. Compound 2c is not only ligand-rearranged, but also one-electron oxidized, as noted by relevant metrical parameters associated with the phenylene ring between atoms N1 and N2 (Figure S2). A possible overall stoichiometry for this reaction can be written as 2[(L2)CoII] → [(L2re)CoII]+ [(L2re,ox) CoII] + e.

Scheme 1.

Scheme 1

Compound (L16)CoII (16) is also highly sensitive to the same type of ligand rearrangement in the presence of traces of dioxygen, affording the isolable dimer [K(THF)2(L16re)CoII]2 (16b, Figure S3), which is equivalent to 2b in terms of ligand reorganization. In this case, we were not able to isolate any other compound that might provide evidence for the location of the oxidizing equivalent(s).

Electrochemistry.

Ten Co(II) compounds possessing aryl (L3Co, L5Co, L13Co), alkyl (L9Co), and acyl arms (L4Co, L6Co, L7Co, L8Co, L10Co, L17Co) were selected as representative examples for examination by cyclic voltammetry. Electro-chemical data for a handful of these examples have been previously reported.48 Figure 6 provides a collective presentation of the corresponding waves (first oxidation event), and Table S2 summarizes relevant electrochemical data (potentials are reported versus the ferrocenium/ferrocene (Fc+/Fc) couple). All aryl- and alkyl-armed Co(II) compounds examined by cyclic voltammetry feature semireversible waves at negative potentials, ranging from −0.665 (L13Co) to −0.090 V (L3Co), in accordance with the electron-rich nature of the corresponding substituents. Specifically for the alkyl-armed L9Co, the two closely spaced, semireversible waves observed (−0.654, −0.500 V), may represent the two slightly different Co(II) sites in the crystal structure. Given the almost identical wave currents for all these aryl- and alkyl-armed Co(II) compounds (3.0 M), and their anodic shifts with respect to the analogous Mn(II)/Mn(III) and Fe(II)/Fe(III) couples44,46a by approximately 0.65 and 0.25 V, respectively, we assign the corresponding semireversible waves to essentially metal-centered Co(II)/Co(III) cycles.

Figure 6.

Figure 6.

Cyclic voltammograms of compounds [K(L3)CoII–NCMe] (3) and [K(NCMe)3(L13)CoII–NCMe] (13) in MeCN/(nBu4N)PF6, [K(L4)CoII]·Et2O (4), [K(THF)6][(L5)CoII]·1.5THF (5), [K(MeCN)(L6)CoII–NCMe]·2MeCN (6), [K2(DMA)4][K(L7)2CoII2]2·2DMA (7), [K(THF)2(L8)CoII] (8a), [K2(L9)2CoII2] (9), and [K(THF)K(L17)2CoII2]2·3Pentane (17) in DMF/(nBu4N)PF6, and [K2(DMA)3(L10)2CoII2]·0.5Et2O (10) in DMA/(nBu4N)PF6, with a Au disk electrode (1.6 mm in diameter); scan rate 0.1 V/s.

In sharp contrast, the acyl-armed Co(II) compounds examined demonstrate irreversible anodic ways with variable ip,a values (mostly large versus the aryl/alkyl-substituted congeners), suggesting significant ligand-centered contributions. In addition, the first anodic wave shown in Figure 6 overlaps with a subsequent oxidation wave (not shown), rendering any attempts to garner further information from exhaustive electrolysis futile. Nevertheless, all initial anodic waves for the acyl-armed Co(II) sites are shifted to positive potentials (0.032 (L17Co) to 0.719 V (L8Co)), reflecting the effect of the electron-withdrawing acyl arm on the Co(II) center, albeit in an order not always consistent with the electronic nature of each individual acyl group. The most significant deviation is observed for L10Co (Ep,a = 0.559 V), whose anodic shift versus that of L10 Mn (Ep,a = −0.108 V; tentatively assigned to Mn(II)/Mn(III))44 betrays very little, if any, metal-centered contributions to the anodic wave. The intervention of ligand-centered oxidation events does not permit any secure identification of the solution structure of species such as L7Co and L10Co, which exhibit single anodic waves even in the presence of dinuclear units in their polymeric solid-state structures. On the other hand, the dual anodic-wave feature for L17Co (solid-state tetramer) may signify a dinuclear structure in solution. Finally, the catalytically important L8Co (−COCF3 arm) possesses the indisputably most electrochemically stable Co(II) site encountered in this series of Co(II) compounds, in agreement with previous findings for the L8 Mn and L8Fe analogues.44,46a

EPR Spectroscopy.

EPR spectra of selected Co(II) compounds were recorded from frozen DMF solutions of 3, 4, 5, 6, 7, 8a, 9, 10, 13, 14, 15, and 17. In all cases, the samples give rise to signals that are consistent with isolated CoII (S = 3/2) species.50 Spectra recorded at 10 K are shown in Figure 7. The spectra can be interpreted within the framework of the spin Hamiltonian:51

H^Zfs=DS^z2SS+13+EDS^x2S^y2+S^AI^+βBg0S^ (1)

Figure 7.

Figure 7.

Experimental (black line) and theoretical (red-dashed line) EPR spectra from frozen DMF solutions of complexes 3, 4, 5, 6, 7, 8a, 9, 10, 13, 14, 15, and 17. For 3, 4, 6, 7, 8a, 10, and 17, the subspectra from sites 1 (green), and 2 (blue) are also shown, as described in the text. A spike signal at ca. 3400 G in the spectra of 5, 9, and 10 is due to an impurity in the cavity. EPR conditions: Temperature, 10 K; modulated amplitude 10 Gpp; microwave power, 2.0 mW; microwave frequency, 9.4 GHz.

In eq 1, D and E are the zero field splitting (zfs) parameters. A is the hyperfine tensor relevant to the hyperfine interactions of the electronic (S = 3/2) and nuclear (I = 7/2 for 59Co) spins, and g0 is the intrinsic g-tensor of the Co(II) ion. For simplicity we assume that the principle axes of the tensors are parallel to each other. Often, due to large spin orbit coupling effects for CoII (S = 3/2), the zfs parameter |D| is quite large, whereas A and g0 are characterized by significant anisotropy. Under the influence of zfs, the 4-fold degeneracy is partially lifted in zero magnetic field, yielding two Kramers’ doublets, |±1/2〉 and |±3/2〉 separated by 2|D|. The |±1/2〉 (or |±3/2〉) doublet is the ground state for positive (or negative) D. For |D| ≫ hν (≈ 0.31 cm−1 at X-band), each Kramers doublet can be described by an effective Seff = 1/2, giving rise to an EPR spectrum characterized by an anisotropic geff tensor.

The EPR signals shown in Figure 7 are consistent with the | ±1/2〉 doublet. No detailed temperature dependence of the spectra was pursued in the present work. However, spectra recorded at 4.2 and/or 20 K (not shown) indicate that the intensity of the signals (scaled as Intensity × Temperature) decreases as temperature increases. This suggests that the |±1/2〉 doublet is the ground state, implying a positive value for D. For |D| ≫ hν, the spectra observed do not depend on D but rather on the rhombic zfs parameter E/D, the values of the intrinsic g0-tensor,52 and the hyperfine term. With the exception of the spectrum for complex 15, the g0-tensor was assumed axial (g0x = g0y = g0⊥g0z = g). In several cases (3, 4, 6, 7, 8a, 10, 17), the spectra indicated the presence of two Co(II) species, characterized by a different value for the parameter E/D. Assuming a common value for |D|, the simulations determine the relative abundance of each Co(II) site. The specific line shape of the spectra results from a combination of factors, including distributions on the parameter E/D (E/D strain), unresolved hyperfine interactions, and residual line-broadening mechanisms.53 The EPR parameters for all samples, as well as the relative ratios of the two species when applicable, are presented in Table S3. Because an unequivocal deconvolution of the line broadening mechanisms is not feasible, the quoted values for the relevant parameters are indicative.

With the exception of the six-coordinate 15, all other complexes (four- or five-coordinate) exhibit rhombic parameters (E/D) that lie in the interval [0, 0.19], indicating different degrees of rhombicity. All complexes (except 15) demonstrate a valley-shaped signal at g ~ 2.0, which corresponds to the g component of the geff-tensor. This signal is broadened due to hyperfine interactions, and in some cases (7, 9, 17) the hyperfine lines are well resolved. The simulations indicate that A values in all cases are in the range of 200–300 MHz and that the hyperfine term has to be taken into account in order to reproduce the low field region of the spectra, corresponding to the perpendicular components of the tensors in eq 1. Due to restrictions in the determination of the line-broadening mechanisms, it is not possible to unambiguously evaluate the magnitudes of the Ax and Ay. Therefore, an average value, A = (Ax + Ay)/2, is quoted in Table S3. This value ranges in the 30–90 MHz range. Complex 15 exhibits a relatively sharp peak at g = 6.24 and an extremely broad derivative feature with an apparent zero crossing at ca. g = 2.07. This behavior indicates a rhombic system with E/D = 0.33. From this point of view, complex 15 exhibits unique EPR properties in this series, most likely due to its special, six-coordinate ligand field, featuring two O residues in addition to the familiar [N3N] coordination. The origin of dual EPR-active species in DMF solutions of some compounds cannot be ascertained at the present time, especially since structural data from crystals derived from DMF solutions are not available. Possible sources are small deviations in the coordination of Co(II) sites in polymeric species (as noted for 7, crystallized from DMA solutions), and also potential differentiation arising from partial DMF coordination to Co(II) and even K+ sites.

In summary, the EPR studies from frozen DMF solutions of the complexes indicate that all complexes feature a Co(II) (S = 3/2) ion with a large and positive zfs term, D, and variable degree of rhombicity. The existence of more than one species in some cases suggest that the coordination environment of the Co ion can be quite flexible in solution.

Catalytic Aziridinations of Olefins.

Styrene.

The 17 CoII compounds shown in Figure 2 were first explored as catalysts (5 mol %) for the aziridination of styrene (8.0 equiv) by PhINTs (1.0 equiv) in chlorobenzene at room temperature (Table 1). The high-yielding L4CoII was first investigated as catalyst in several solvents (MeCN, 50%; 2,2,2-Trifluoroethanol, 62%; PhCF3, 61%; CH2Cl2, 61%; Benzene, 53%; PhCl, 70%) and found to afford superior levels of aziridine in chlorobenzene (similar solvent-screening results were previously obtained with MnII reagents44). The yield drops significantly (25%) if equimolar amounts of styrene and PhINTs are employed in PhCl, whereas a 2-fold excess of styrene over PhINTs improves the yield to 46%. Increasing the amount of the L4Co catalyst to 10 mol % suppresses the yield to 53%. The styrene aziridination yields obtained with various CoII catalysts (Table 1) vary significantly (18–70%) as a function of the ligand employed. As opposed to the corresponding LxMnII reagents (x = 1–15)44 that reveal a dominant relationship between increasing styrene aziridination yields with anodically shifting MnII/MnIII redox potentials (and, by extension, with increasing electrophilicity of the putative MnIIINR), the CoII catalysts provide yields that indicate a more complex pattern of electronic and steric contributions. For example, although the comparatively electron-deficient, acyl-armed reagents remain, on average, more productive than aryl- or alkyl-armed congeners, a comparison between the −COCMe3 (L4) and −COCF3 (L8) supported CoII reagents furnishes essentially the same aziridination yields, in spite of the fact that the Co(II)/Co(III) couple for L8Co is anodically shifted by ~500 mV versus L4Co. In addition, L8Co is potentially less congested than L4Co. In contrast, the MnII and FeII congeners are associated with significantly diverging yields in favor of the more electron-deficient L8M reagents (L4Mn, 12%; L8Mn, 75%; L4Fe, 45%; L8Fe, 73%), correlating with an anodic potential shift of approximately 600 mV for both the L8Mn and L8Fe catalysts with respect to their L4M analogues. More importantly, as indicated below, L4Co (4) is faster than L8Co (8b, 8c) in mediating styrene aziridination. The closely related L7Co (−COiPr arm) is in principle slightly less electron rich and sterically congested than the L4Co congener, but affords lower yields than L4Co in a slower reaction (the opposite is true for the corresponding Mn(II) catalysts).44 Two other acyl-substituted CoII catalysts (L14Co, L17Co), exhibiting significantly lower yields, are indicative of how the metal-nitrene electron-affinity bias may be overridden by other electronic or steric factors in the fairly restricted reaction cavity of Co reagents. The corresponding L14 Mn catalyst, by contrast, is among the most productive MnII aziridination reagents examined (yield: 67%).44 As expected, the aryl-substituted ligands generate CoII reagents that are poor mediators of styrene aziridination. These sites are oxidatively and even hydrolytically sensitive and tend to generate thermodynamic sinks. Nevertheless, the muted role of the electron-affinity criterion can still be discerned in the series of the electron-rich, aryl-armed reagents (1, 2, 3, 5, 12, 13), inasmuch as the highest-yielding L13Co (13) is the most electron rich member of the group. Finally, the alkyl-substituted L9Co and L16Co reagents are only modestly productive, as anticipated for electron-rich CoII sites, but again the more electron rich, isopropyl-substituted L9Co provides higher yields than the methyl-substituted congener L16Co. However, the latter undergoes facile oxidative ligand rearrangement that may compromise its structural integrity.

Table 1.

Yields of Styrene Aziridination Mediated by CoII Reagents 1–17a

graphic file with name nihms-1736859-t0002.jpg
compound yield (%)
[K(2.2.2-cryptand)][(L1)CoII]·3THF (1) 18
[K(2.2.2-cryptand)][(L2)CoII]·1.5Pentane (2) 18
[K(L3)CoII−NCMe] (3) 36
[K(L4)CoII]·Diethyl Ether (4) 70
[K(THF)6](L5)CoII]·1.5THF (5) 32
[K(NCMe)(L6)CoII−NCMe]·MeCN·0.5H2O (6) 50
[K2(DMA)4][K(L7)2CoII2]2·2DMA (7) 59
[K(NCMe)(L8)CoII−NCMe] (8b) 69
[K(DMA)3(L8)CoII]2 (8c) 68
[K2(L9)2CoII2]n (9) 38
[K2(DMA)3(L10)2CoII2]·0.5Et2O (10) 49
[K(THF)(L11H)CoII−OH)]2 (11)
[K(THF)3(L12)CoII]·THF (12) 25
[K(NCMe)3(L13)CoII−NCMe] (13) 38
[K(DMA)(L14)CoII]·DMA (14) 36
[K(THF)3(L15)CoII]·THF (15) 45
[K(L16)CoII−THF] (16) 25
[K(THF)K(L17)2CoII2]2 (17) 35
a

Conditions: Catalyst, 0.0125 mmol (5 mol %); PhINTs, 0.25 mmol; styrene, 2.0 mmol; MS 5 Å, 20 mg; PhCl, 0.200 g; 30 °C; 12 h.

Alkene Aziridinations by [K(L4)CoII]·Diethyl Ether (4).

The highest yielding L4Co (4) catalyst was subsequently investigated as a nitrene-transfer (NTs) mediator with a panel of aromatic, cyclo (Table 2), and aliphatic alkenes (Table S4). Styrenes substituted at the para position with both electron-donor and electron-acceptor groups were examined first (entries 1–9) under the conditions noted in the previous section. Both methylene chloride and chlorobenzene were employed (as well as acetonitrile in a few instances), invariably resulting in better aziridination yields in chlorobenzene. In the majority of cases, yields above 70% were recorded irrespective of the electron-rich or poor character of the substituent, with the exception of two moderately yielding cases involving strong electron-withdrawing groups (p-CF3, p-NO2). Moreover, the product of 4-MeO-styrene aziridination is known to be unstable,54 and is thus associated with modest yields. Overall, the aziridination yields for these para-substituted styrenes, save for the parent styrene, are comparable to those previously reported for L8Co. Ortho-substitution (entry 10) affects aziridination yields, presumably interfering with nitrene-transfer both sterically and electronically (due to the orthogonal orientation of the aromatic versus the olefinic plane of the substrate55). Similar steric inhibition is also observed for α-substituted styrenes (methyl, phenyl; entries 11 and 12), especially for the bulkier α-phenyl-styrene. An allylic amination product is also obtained in both cases, ascribed to aziridine-ring opening,56 which is more pronounced for the α-phenyl-substituted product.54 Steric hindrance is also evident in the aziridination of β-substituted styrenes (entries 13–16), especially for the bulky cis- and trans-stilbene (entries 15, 16). In agreement with previous observations in the application of L8Co,44 the cis congeners are more productive (entries 14, 15), with significant loss of stereochemical integrity. For all these encumbered substrates, L8Co is on average more productive than L4Co, presumably reflecting the somewhat more voluminous reaction cavity of L8Co. Allylic or benzylic aminations compete successfully with aziridinations (entries 17–22), unless cis (entry 18) and/or electron-rich (entry 21) olefins are involved. Beyond cycloalkenes, other terminal or internal aliphatic olefins exhibit low aziridination yields (Table S4), in accordance with previous results pertaining to the application of L8Co in aziridinations of aromatic and aliphatic olefins.44 A competitive styrene versus 1-hexene (1.0 mmol each) aziridination by PhINTs (0.25 mmol) catalyzed by L4Co (5 mol %) in chlorobenzene provided a ratio of 25:1 in favor of the styrene aziridination product (combined aziridine yield: 72%), not unlike the L8Co catalyst (28:1, yield 73%).44

Table 2.

Yields of Aziridination/Amination of Olefins by [K(L4)CoII]·Et2O (4)a

graphic file with name nihms-1736859-t0003.jpg
a

Conditions: 4, 0.0125 mmol (5 mol %); PhINTs, 0.25 mmol; olefin, 2.0 mmol; MS 5 Å, 20 mg; solvent (MeCN/CH2Cl2/PhCl) 0.200 g; 30 °C; 12 h.

b

In MeCN.

Mechanistic Studies.

Comparative Reaction Profile.

The formation of the product of styrene aziridination was monitored as a function of time (Figure 8) for catalysts L4Co (4), L7Co (7), and L8Co (8c) (all crystallized from DMA/ether), under the conditions noted above (Table 1), with the exception of the amount of PhCl used (500 mg). Yields were determined after quenching the reaction at various time intervals. Surprisingly, the more electron rich and sterically congested L4Co (4) exhibits faster product generation than L8Co (8c) during the first hour, whereas L7Co (7) is kinetically comparable to 8c. At the 1.0 h mark, the reaction is complete by 88% for L4Co (4), 65% for L8Co (8c), and 76% for L7Co (7). Interestingly, the MeCN-crystallized version of L8Co, [K(NCMe)(L8)CoII–NCMe] (8b), has been previously shown44 to be significantly slower; only 14% of the reaction is complete after 1.0 h in d5-PhCl, indicating potential interference by the strongly coordinating acetonitrile. However, 8b is still much faster than the corresponding [K(NCMe)(L8)MII–NCMe] (M = Mn, Fe) reagents, presumably reflecting the superior electrophilicity of CoIIINR as discerned from Ep,a values associated with the MII/MIII couple of the divalent catalysts (Fe: 0.228 V, Mn: 0.518 V, Co: 0.837 V).44

Figure 8.

Figure 8.

Yield of aziridine (%) as a function of time (min) in the reaction of styrene (2.0 mmol) by PhI = NTs (0.25 mmol) mediated by L4Co (4), L8Co (8c), and L7Co (7) (0.0125 mmol with respect to Co) in chlorobenzene (0.500 g) at 30 °C.

Hammett Analysis.

Several para-substituted styrenes (para substituent: Me, tBu, F, Cl, CF3, NO2; 1.0 mmol each) were subjected to competitive aziridination (PhINTs, 0.25 mmol) versus styrene (1.0 mmol), mediated by L4Co (5 mol %) in chlorobenzene (0.200 g), in the presence of 5 Å molecular sieves (25 mg). Hammett plots of log(kX/kH) (determined by 1H NMR from the ratio of the corresponding aziridines) as a function of the substituent polar parameter σP or even the resonance-responsive parameter σ+ did not provide any reliable linear free-energy correlations (Figure S4, Table S5). In contrast, Jiang’s dual-parameter correlation that incorporates both polar (σmb) and spin-responsive (σJJ*) parameters (log(kX/kH) = ρmbσmb + ρJJ*σJJ* + C)57 provides a reasonable linear correlation (R2 = 0.98; Figure 9). The negative ρmb value is consistent with a small positive charge developing at the benzylic carbon, whereas the always positive ρJJ* value denotes an incipient radical character for the same site. The ratio |ρmb/ρJJ*| = 1.12 is similar to the one previously observed for L8Co catalyzed aziridinations (|ρmb/ρJJ*| = 1.0),44 and indicates competitive contributions of polar and spin-delocalization effects. Polar effects are dominant in many Rh,24d Cu,24d,58 and Fe18g,59,60 catalyzed aziridinations, for which Hammett plots can be fit with the assistance of polar parameters alone (σp, σ+), but the need for incorporating spin-delocalization parameters (σ*, σJJ*)57,61 with a wide range of |ρmb/ρJJ*| values (0.04–2.02) is also evident in many other metal-catalyzed aziridinations.18e,22d,g,23f,g More recently, larger |ρ+/ρJJ*| values have been reported for the aziridination of styrene by PhINNs mediated by [CoIII(TAMLred)] (5.71) and [CoIII(TAMLsq)] (8.64), in accordance with a novel mechanism that involves a partial single-electron transfer from the styrene to the metal-nitrene as a component of the turnover-determining step.41 A similar mechanistic scenario has also been advanced for Fe-mediated aziridinations, but in this case Hammett correlations can be successfully accommodated with polar parameters alone (σ+).42

Figure 9.

Figure 9.

Linear free energy correlation of log(kX/kH) vs σmb, σJJ* for the aziridination of para-substituted styrenes (X = Me, tBu, F, Cl, CF3, NO2) mediated by L4Co (4).

Kinetic Isotope Effect and Stereochemical Integrity.

Evaluation of the secondary kinetic isotope effect was accomplished by 1H and 2H NMR with the assistance of deuterated styrenes (α-d-styrene, cis- and trans-β-d-styrene; 1.0 mmol each) in competitive aziridinations (PhINTs, 0.25 mmol) with styrene (1.0 mmol) catalyzed by L4Co (4) or L8Co (8c) (5 mol %) in chlorobenzene (Table 3). KIE values close to 1.0 were obtained with α-d-styrene for both catalysts, indicating that the α-styrenyl is unlikely to be involved in the initial nitrene attack to styrene. In contrast, the β-styrenes are associated with inverse KIE values for L4Co (cis: 0.90 (±0.02), trans: 0.92 (±0.02)) that can be attributed to a limited sp2sp3 rehybridization of styrenes Cβ site upon development of the initial N–Cβ bond (Scheme 2). More modest inverse KIE values are also noted in cis- and trans-β-d-styrene aziridinations mediated by L8Co (8b44 or 8c) (cis: 0.96 (±0.02), trans: 0.98 (±0.02)), suggesting only minimal N–Cβ bond formation in the transition state.

Table 3.

Secondary KIE Values in Aziridination of Deuterated Styrenes vs Styrene

catalyst graphic file with name nihms-1736859-t0004.jpg graphic file with name nihms-1736859-t0005.jpg graphic file with name nihms-1736859-t0006.jpg
[K(L4)CoII]n (4) 1.01(±0.01) 0.90(±0.02) 0.92(±0.02)
[K(DMA)3(L8)CoII]2 (8c) 1.00(±0.01) 0.96(±0.02) 0.98(±0.02)
Scheme 2.

Scheme 2

The kinetics of the aziridine ring closure (formation of the second N–Cα bond) was further evaluated by 2H NMR in the aziridinations of cis- and trans-β-d-styrene (Table 4), by examining the degree of stereochemical scrambling in the resulting aziridines (cis/trans partitioning due to Cα–Cβ bond rotation; Scheme 2) in competition with N–Cα bond formation. The ratio of cis/trans aziridine (94:6) and trans/cis aziridine (92:8) resulting from the L4Co-mediated aziridination of cis-β-d-styrene and trans-β-d-styrene, respectively, signifies the interference of very small energy barriers in aziridine-ring closure. A slightly larger barrier is indicated for the aziridination of the more sensitive cis-β-d-styrene by L8Co (cis/trans aziridine: 89/11). On the other hand, the stereochemical scrambling observed in the aziridination of cis-β-methyl-styrene is more pronounced with L4Co than L8Co.

Table 4.

Exploration of Stereochemical Integrity in the Aziridination of cis- and trans-β-d-Styrene

catalyst graphic file with name nihms-1736859-t0007.jpg graphic file with name nihms-1736859-t0008.jpg
[K(L4)CoII]n (4) 94:6 (cis:trans aziridine) 92:8 (trans:cis aziridine)
[K(DMA)3(L8)CoII]2(8c) 89:11 (cis:trans aziridine) 92:8 (trans:cis aziridine)

Computational Studies.

The structure and electronic description of the presumptive [L4Co]NTs intermediate were explored by DFT calculations at the B3LYP/6–31+G(d) level of theory. Free energy calculations suggest that the intermediate-spin quartet ground state (S = 3/2) lies only 0.1 kcal mol−1 lower than the high-spin sextet state (S = 5/2), and 2.8 kcal mol−1 below the doublet state (S = 1/2). As mentioned above, the calculated free energies for [L8Co]NTs indicate that the high-spin sextet is the ground state, in agreement with the weaker ligand field provided by the L8 versus L4 ligand.

Calculated structures for the three spin-states of [L4Co]NTs along with key metrical parameters are presented in Figure 10. The most conspicuous feature of these structures is the dissociation of one arm from the equatorial coordination sphere of the metal (Co–N = 3.93 (quartet), 3.67 (sextet), 4.08 (doublet) Å). The axial Co–Namine bond is also elongated (Co–N = 2.43 (quartet), 2.63 (sextet), 2.99 (doublet) Å), if not dissociated, by comparison to that of L4Co (2.151(11) Å). These features have been previously noted in the DFT structure of [L8Co]NTs, although the latter exhibits an additional Co–F equatorial contact (Co–F = 2.37 Å).44

Figure 10.

Figure 10.

DFT structures (minimal metal coordination) for [L4Co]NTs active species in different spin states (from left to right: quartet, sextet, doublet) optimized at the B3LYP/6–31+G(d) level of theory. Hydrogen atoms were omitted from the figure for clarity.

Most importantly, the calculated spin densities for all three spin states of [L4Co]NTs place a full oxidizing equivalent over the dissociated arm, hence generating a widely delocalized N-aryl amidyl radical (Figure 11). For the ground-state quartet, the computed spin density consists of ~3.2 unpaired e on Co, 0.79 e on the nitrene N atom and −1.0 unpaired e on the noncoordinating arm. Similar spin densities are calculated for the sextet (Co: 2.9 e, N: 0.9 e, ligand arm: 1.0 e) and the doublet state (Co: 2.8 e, N: −0.66 e, ligand arm: −1.0 e). This spin density distribution is accommodated by an electronic structure such as [(L4)Co(II)–NTs], featuring a high-spin Co(II) center (S = 3/2) and two oxidizing equivalents on the noncoordinating ligand arm and the N atom of the nitrene residue, respectively. In sharp contrast, the spin density of the ground-state sextet of [L8Co]NTs (Figure 11) is largely distributed on Co (3.3 e) and the N atom of the nitrene (1.1 e). In this case, the noncoordinating arm is redox innocent, and the residual spin density is centered over ligating N atoms in a typical spin polarization fashion. Hence, the sextet state of [L8Co]NTs is better accommodated with an [(L8)Co(III)–NTs] electronic description (SCo = 2).

Figure 11.

Figure 11.

Spin density on the calculated lowest-energy spin state of the putative cobalt nitrenoid intermediates: quartet [L4Co]NTs (left) and sextet [L8Co]NTs (right).

A global electrophilicy index (GEI) was also computed for [L4Co]NTs and [L8Co]NTs by employing Stephan’s improved methodology.62 For the quartet spin state, GEI is calculated to be 5.0 eV for [L4Co]NTs and 5.7 eV for [L8Co]NTs. The corresponding values for the sextet spin state are 4.7 and 6.1 eV for [L4Co]NTs and [L8Co]NTs, respectively. Thus, [L8Co]NTs is more electrophilic than [L4Co]NTs on the basis of the GEI criterion.

Unfortunately, all efforts to map the aziridination reaction coordinate starting from [L4Co]NTs and styrene have not been successful in locating an acceptable transition state for the initial N–Cb bond formation. All three spin states (quartet, sextet, doublet) of [L4Co]NTs generated large activation barriers for this initial step (~50 kcal/mol). However, when dispersion-corrected DFT was applied,63 as appropriate for polarizable bulky groups such as the tert-Bu, the corresponding barriers were reduced by approximately 20 kcal/mol. These barriers are still significant by comparison to those we have previously identified for the reaction of [L8Co]NTs (sextet) and styrene (23.4 kcal/mol for the turnover-limiting N–Cb bond formation).44 Further experimentation and attendant DFT calculations will be required to unravel reliable trends and contributing factors with the assistance of catalysts that feature substituents spanning the CF3 to CMe3 range.

FURTHER DISCUSSION AND CONCLUSIONS

In a rigorous recent account, Latour and co-workers42 highlight the importance of the electron affinity (EA) of iron-nitrene/imido species as a guiding principle for predicting their reactivity in a wide range of iron-mediated aziridinations. In these catalytic reactions, the iconic substrate styrene undergoes aziridinations by various iron-nitrene compounds (Fe = NR), under a general mechanistic scheme that designates the formation of the initial N–Cb bond as the rate-determining step, usually encountered in aziridinations with a two-step mechanism ([M]NR radical addition to styrene, ring-closure radical rebound). More importantly, this first step is front-loaded by significant charge transfer from styrene to the iron-nitrene and, thus, is crucially influenced by the EA of the active oxidant. The applicability of the EA as a general predictor of reactivity seems to be wide, but at the present time is largely confined within the realm of catalysts that provide Hammett correlations for the aziridination of para-substituted styrenes that can be accommodated with polar parameters alone (σP, σ+), or by a combination of polar and spin-delocalization parameters (σ*, σJJ*) with dominant polar contribution.

In an almost concurrent publication, de Bruin and co-workers41 advance a similar argument with the assistance of electrophilic Co(III)-nitrene radical aziridination reagents, generated from the reaction of anionic [CoIII(TAMLred)] or the one-electron oxidized and neutral [CoIII(TAMLsq)] with PhINNs. Hammett plots for para-substituted styrene aziridinations are fitted with both polar (σ+) and spin-delocalization parameters (σJJ*) with large |ρ+/ρJJ*| values (5.71 for [CoIII(TAMLred)] and 8.64 for [CoIII(TAMLsq)]), hence these systems can also be considered as good candidates for testing the EA criterion. Indeed, the larger |ρ+/ρJJ*| value for [CoIII(TAMLsq)] and DFT calculations indicate that the energy barrier for the initial, rate-limiting reaction of the incipient [CoIII]NNs and styrene to generate the N–Cb bond is lower for [CoIII(TAMLsq)] versus [CoIII(TAMLred)], in agreement with an anticipated higher EA value for the nitrene species generated from [CoIII(TAMLsq)]. Surprisingly, the experimental rate for the aziridination of styrene by these two catalysts favors [CoIII(TAMLred)] versus [CoIII(TAMLsq)], but this has been attributed largely to the instability of the latter reagent. A distinctive feature of the Co(III) systems is associated with the redox noninnocent character of the tetraanionic ligand (TAMLred)4−, which can be successively oxidized in one-electron steps to (TAMLsq)3− and (TAMLq)2−. The authors argue convincingly that the emerging radical nitrene species [CoIII(TAMLsq)(NNs)]/CoIII(TAMLq)(NNs)2] and [CoIII(TAMLq)(NNs)], resulting from [CoIII(TAMLred)] and [CoIII(TAMLsq)], respectively, interact with styrene in the rate-limiting step by means of an asynchronous transition state, encompassing significant single-electron transfer from styrene to the oxidized TAML ligand and a nucleophilic attack by the nitrene lone pair (in lieu of the NNs p radical) at the incipient styrenyl radical cation. The attendant single-electron relocation (TAML → Co(III) → NNs) reestablishes the N lone pair and retains the Co(III) oxidation state. Similar participation of charge transfer in the rate-limiting transition state between high-valent metal nitrenes/imidos and sulfides has been recently showcased for many other nitrene-transfer catalysts64 and is now established as a common mechanistic feature.

In a previous comprehensive study from our lab,44 we have shown that a library of Mn(II) catalysts, supported by the vast majority of the ligands used in the present work (L1–15H3), mediates alkene aziridinations with reactivity that increases in parallel with increasing electrophilicity of the putative [MnIII]–NTs active oxidant. Moreover, the electrophilicity criterion holds across the base metals, inasmuch as the reactivity of the best performing Mn(II) reagent L8 Mn is inferior to that of the more acidic L8Co. Although the electron affinity of the metal-nitrene reigns supreme for all these reagents, the molecular interaction between [MIII]–NR and styrene in the rate-limiting formation of the initial N–Cb bond, is quite distinct with respect to the reagents explored by Latour and de Bruin. Indeed, Hammett correlations for the aziridination of styrenes mediated by our L8M reagents (M = Mn, Fe, Co) reveal rather modest positive charge buildup on the α-styrenyl carbon (increasing with metal acidity in the expected order: Fe < Mn < Co), and require the inclusion of competitive spin-delocalization contributions (|ρmb/ρJJ*| = 0.75 (Mn), 1.17 (Fe), 1.00 (Co); the correlation for Fe was rather weak). The fact that these reagents demonstrate more modest charge-transfer characteristics is consistent with the operation of presumptive metal nitrenes ([MIII]–NR) resting at a lower oxidizing level than the high-valent iron and cobalt nitrenes of Latour and de Bruin, respectively. Overall, these Mn(II) reagents and congeners can also be accommodated under the general EA criterion advanced by Latour (after all, they are catalysts engaged in typical electrophilic radical reactions), although they are not characterized by a dominant charge-transfer component. Incidentally, a strongly enhanced radical contribution, as in the case of Betley’s iron dipyrrinato complexes (|ρmb/ρJJ*| = 0.04 for NAd),18e has been interpreted42 as the result of a competitive energy barrier for the second, ring-closing step (radical rebound).

The library of the Co(II) reagents (S = 3/2) reported in this work showcases some surprising deviations from the EA criterion. Although the importance of the electrophilicity of the metal-nitrene can still be detected in the relative enhanced yields provided by the Co(II) compounds possessing acyl-versus aryl- or alkyl-substituted ligands, the trend is certainly not as smooth and predictable as that previously encountered with the Mn(II) reagents.44 Indeed, a closer inspection of the acyl-substituted subset of the Co(II) library of reagents reveals a wide range of yields in the aziridination of styrene that cannot be correlated with the anticipated electrophilicities of the corresponding cobalt-nitrene moieties. To further pinpoint the provenance of these disparities, we selected the high-yielding L8Co (−COCF3 arm) and L4Co (−COCMe3 arm) for further investigation. The L8Co was previously studied44 in tandem with the L8Mn and L8Fe congeners and found to be more reactive and selective than the other two base metal analogues. Mechanistic and computational studies showed that all three L8 M reagents follow a two-step styrene aziridination path (turnover-limiting addition of [L8 MIII]–NTs to the β-styrenyl carbon followed by product-determining ring-closure via radical rebound), with activation barriers in the order Fe > Mn > Co for both steps. The trend is consistent with the anticipated metal-nitrene electrophilicities (first step) and ease of reduction from M(III) to M(II) (second step), hence highlighting the dominant role of EA in both steps of styrene aziridination (aliphatic olefins do not follow the same trend for the second step).

The representative case of L4Co, however, presents a conundrum, inasmuch as its reactivity in terms of styrene aziridination yields is comparable to that provided by L8Co. More importantly, the rate of product buildup in the first hour of the reaction mediated by L4Co is superior to that of L8Co (Figure 8). These results are difficult to reconcile for a reagent such as L4Co, whose Co(II/III) couple is cathodically shifted by 500 mV versus that of L8Co, and its nitrene derivative [L4Co]NTs is computed to have a lower global electrophilicity index (GEI) than that of [L8Co]NTs, in agreement with the electronic nature of the CMe3 and CF3 substituents. In addition, the enhanced reactivity of L4Co deviates from that of the corresponding L4Mn and L4Fe reagents, which exhibit significantly lower yields (and rates) in styrene aziridinations by comparison to the L8Mn and L8Fe analogues, in line with their electrophilic characteristics.

Mechanistic analysis of the operation of L4Co in styrene aziridination indicate that both polar (σmb) and spin-delocalization (σJJ*) parameters are needed to fit Hammett plots, suggesting that both modest positive charge buildup and radical stabilization participate in the turnover-limiting step. The unexpected preponderance of the polar effect for L4Co by comparison to L8Co can be traced both in the slightly higher values of absolute ρmb (−0.58) and relative |ρmb/ρJJ*| (1.17) than those observed for L8Co (ρmb = −0.56, |ρmb/ρJJ*| = 1.0). The secondary KIE values obtained from the competition between styrene and selectively deuterated styrene in aziridinations confirm that both catalysts operate via an initial, turnover-limiting N–Cb bond formation step, but also indicate that the L4Co mediated pathway incorporates more significant rehybridization of the β-styrenyl carbon in the transition state, hence placing this TS energetically closer to the resulting radical intermediate [L4Co]N(Ts)–CH2CH2Ph. Moreover, the ring-closing step (radical rebound) seems to operate via a miniscule energy barrier for both L4Co and L8Co styrene aziridinations, but the one for L4Co is even more suppressed than that for L8Co, as judged by the superior retention of stereochemistry in the aziridination of the sensitive substrate cis-β-styrene. This runs counter to what is usually the main driving force for the aziridine-ring closure of styrene, namely the ease of reduction from Co(III) to Co(II).22g,44,65 All these mechanistic observations would have been perfectly in line, had the supporting L4 ligand been more electron-withdrawing than L8.

DFT calculations on the electronic and geometric disposition of the presumptive [L4Co]NTs vis-à-vis the previously explored [L8Co]NTs highlight how a small ligand modification can result in a major electronic rearrangement. First, the ground state of [L4Co]NTs is computed to be the quartet (S = 3/2), positioned slightly underneath the sextet (S = 5/2). The sextet is the clear ground state of [L8Co]NTs, presumably due to the weaker ligand field provided by the L8 ligand. Geometrically, both cobalt nitrenes are quite similar, their most outstanding feature being the elongation of one of the equatorial N residues to a noncoordinating position. However, spin-density calculations reveal that the noncoordinating arm of [L4Co]NTs is one-electron oxidized, whereas the corresponding arm of [L8Co]NTs is redox innocent. The single-electron distribution of the resulting N-aryl amidyl radical in [L4Co]NTs is spread throughout the noncoordinating arm, with almost half of the spin density being localized on the N atom. Apparently, the electron withdrawing CF3 residue protects the noncoordinating arm of [L8Co]NTs from a similar one-electron oxidation. The overall electronic picture for the ground state of the two cobalt nitrenes is schematically summarized in Figure 12. As noted above, the [L4Co]NTs sextet (α-spin on the N-aryl amidyl radical) is calculated to be only 0.1 kcal/mol higher in free energy relative to the quartet.

Figure 12.

Figure 12.

Schematic distribution of spins in the ground state of [L4Co]NTs and [L8Co]NTs.

Whereas a definitive justification for the higher reactivity of L4Co, in spite of lower electrophilicity, versus L8Co cannot be provided at the present time, the following observations should be taken into account:

  1. Although it cannot be excluded, it is deemed rather unlikely that the ease of formation of the cobalt-nitrene itself (presumably favoring [L4Co]NTs) will be a contributing factor, since our previous calculations for the reaction of [L8MII] (M = Mn, Fe, Co) and PhINTs indicate almost instantaneous generation of [L8M]NTs. Rate-limiting metal-nitrene formation is more common with organic azides (RN3).66

  2. The fact that L4M (M = Mn, Fe), as well as a wide range of other Mn(II) reagents, exhibit reactivities consistent with the EA criterion, whereas L4Co and other Co(II) reagents demonstrate deviations, suggest that ligand-centered contributions to the overall oxidizing ability of the reagent may enable more favorable reactivity channels. Indeed, [LCoIII]NTs is more likely to store oxidizing equivalents on ligand residues, as inferred by the cyclic voltammograms of the LCoII reagents, and anticipated due to the superior oxidizing power of Co(III) versus Mn(III) or Fe(III). Among other possibilities, N-aryl amidyl radicals are known to add to alkenes, at least intramolecularly,67 and more electrophilic N-aryl sulfonamidyl radicals can even add intermolecularly.68 Although they are not expected to outcompete the metal-bound nitrene radical, they might offer stabilizing interactions not yet realized. On the other hand, the similarity of the Hammett parameters for L4Co and L8Co suggests that the electronic differences in the ground states of [L4Co]NTs and [L8Co]NTs may have only a small effect on their reactivities, but this point requires further elaboration once more information is available for the corresponding transition states.

  3. Multiple spin-state reactivity channels,65c,69 such as those offered by the almost isoenergetic quartet and sextet states of [L4Co]NTs, may afford enhanced reactivity profiles in aziridinations70 by comparison to a potentially single spin-state operation by the [L8Co]NTs sextet.

  4. London dispersion (LD) interactions applying intramolecularly between highly polarizable alkyl substituents (also known as σσ interactions) are now well established stabilizing forces of sterically congested molecules in solution, by means of favorable enthalpic contributions.71 The tert-Bu group and other conformationally rigid alkyl groups (flexible alkyl groups have an unfavorable entropic impact)72 have been credited as “dispersion energy donors”,73 and deemed responsible for stabilizing many highly congested organic and inorganic compounds.7174 More importantly, LD forces have started receiving recognition as contributors to observed chemical reactivity and catalytic outcomes.75 In the more tight reaction cavity of cobalt reagents, the stabilization offered by tert-Bu groups via LD interactions can play a significant role, as already noted in our initial dispersion-corrected DFT calculations. Interestingly, the L7Co reagent, which carries the less polarizable i-Pr substituent, demonstrates lower aziridination rates, not unlike those of L8Co.

Future experimental and computational research will seek to disentangle and quantify the factors contributing to the enhancement of catalytic reactivity above and beyond the underlying electrophilic character of the active oxidant, and further explore whether reagents with other rigid alkyl substituents can be superior mediators of nitrene-transfer chemistry.

Supplementary Material

Supplementary Information

ACKNOWLEDGMENTS

This work was generously supported by the National Institute of General Medical Sciences of the National Institutes of Health under Award Numbers R15GM117508 and R15GM139071 (to P.S.), and in part by the National Science Foundation through grant CHE-0412959 (to P.S.). Authors Z.S. and T.R.C. further acknowledge the U.S. National Science Foundation for partial support of this research through grants CHE-1464943 and CHE-1953547. Authors G.R. and P.P. also thank the Special Account of Research Grants of the National and Kapodistrian University of Athens for partial support. We thank the reviewers for many insightful comments.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.organomet.1c00267.

Experimental procedures; physicochemical characterization of new compounds; crystallographic information; catalytic and mechanistic data (PDF)

Accession Codes

CCDC 2069526–2069535 and 2069537–2069542 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.organomet.1c00267

The authors declare no competing financial interest.

Contributor Information

Anshika Kalra, Department of Chemistry, Missouri University of Science and Technology, Rolla, Missouri 65409, United States.

Vivek Bagchi, Department of Chemistry, Missouri University of Science and Technology, Rolla, Missouri 65409, United States; Institute of Nano Science and Technology, Mohali, Punjab 160062, India.

Patrina Paraskevopoulou, Inorganic Chemistry Laboratory, Department of Chemistry, National and Kapodistrian University of Athens, Athens 15771, Greece;.

Purak Das, Department of Chemistry, Missouri University of Science and Technology, Rolla, Missouri 65409, United States.

Lin Ai, Department of Chemistry, Missouri University of Science and Technology, Rolla, Missouri 65409, United States; College of Chemistry, Beijing Normal University, Beijing 100875, People’s Republic of China;.

Yiannis Sanakis, Institute of Advanced Materials, Physicochemical Processes, Nanotechnology and Microsystems, NCSR “Demokritos”, Athens 15310, Greece.

Grigorios Raptopoulos, Inorganic Chemistry Laboratory, Department of Chemistry, National and Kapodistrian University of Athens, Athens 15771, Greece.

Pericles Stavropoulos, Department of Chemistry, Missouri University of Science and Technology, Rolla, Missouri 65409, United States;.

REFERENCES

  • (1).(a) Aziridines and Epoxides in Organic Synthesis; Yudin AK, Ed.; Wiley-VCH: Weinheim, 2006. [DOI] [PubMed] [Google Scholar]; (b) Padwa A Aziridines and Azirines: Monocyclic. In Comprehensive Heterocyclic Chemistry III; Katritzky AR, Ramsden CA, Scriven EFV, Taylor RJK, Eds.; Elsevier Science: Amsterdam, 2008; Vol. 1, pp 1–104. [DOI] [PubMed] [Google Scholar]; (c) Sweeney JB Aziridines: epoxides’ ugly cousins? Chem. Soc. Rev 2002, 31, 247–258. [DOI] [PubMed] [Google Scholar]; (d) Lapinsky DJ Three-Membered Ring Systems. Prog. Heterocycl. Chem 2015, 27, 61–85. [DOI] [PubMed] [Google Scholar]; (e) Botuha C; Chemla F; Ferreira F; Pérez-Luna A Aziridines in Natural Product Synthesis. In Heterocycles in Natural Product Synthesis; Majumdar KC, Chattopadhyay SK, Eds.; Wiley-VCH: Weinheim, 2011; pp 1–39. [DOI] [PubMed] [Google Scholar]; (f) Padwa A; Murphree SS Epoxides and aziridines – A mini review. ARKIVOC 2006, 2006, 6–33. [DOI] [PubMed] [Google Scholar]
  • (2).(a) Ouyang K; Hao W; Zhang W-X; Xi Z Transition-Metal-Catalyzed Cleavage of C–N Single Bonds. Chem. Rev 2015, 115, 12045–12090. [DOI] [PubMed] [Google Scholar]; (b) Rotstein BH; Zaretsky S; Rai V; Yudin AK Small Heterocycles in Multicomponent Reactions. Chem. Rev 2014, 114, 8323–8359. [DOI] [PubMed] [Google Scholar]; (c) Huang C-Y; Doyle AG The Chemistry of Transition Metals with Three-Membered Ring Heterocycles. Chem. Rev 2014, 114, 8153–8198. [DOI] [PubMed] [Google Scholar]; (d) Stankovic S; D’hooghe M; Catak S; Eum H; Waroquier M; Van Speybroeck V; De Kimpe N; Ha H-J Regioselectivity in the ring opening of non-activated aziridines. Chem. Soc. Rev 2012, 41, 643–665. [DOI] [PubMed] [Google Scholar]; (e) Cardoso AL; Pinho e Melo TMVD Aziridines in Formal [3 + 2] Cycloadditions: Synthesis of Five-Membered Heterocycles. Eur. J. Org. Chem 2012, 6479–6501. [DOI] [PubMed] [Google Scholar]; (f) Lu P Recent developments in regioselective ring opening of aziridines. Tetrahedron 2010, 66, 2549–2560. [DOI] [PubMed] [Google Scholar]; (g) Schneider C Catalytic, Enantioselective Ring Opening of Aziridines. Angew. Chem., Int. Ed 2009, 48, 2082–2084. [DOI] [PubMed] [Google Scholar]; (h) Hu XE Nucleophilic ring opening of aziridines. Tetrahedron 2004, 60, 2701–2743. [DOI] [PubMed] [Google Scholar]; (i) McCoull W; Davis FA Recent Synthetic Applications of Chiral Aziridines. Synthesis 2000, 2000, 1347–1365. [DOI] [PubMed] [Google Scholar]
  • (3).(a) Tomasz M; Mitomycin C Small, fast and deadly (but very selective). Chem. Biol 1995, 2, 575–579. [DOI] [PubMed] [Google Scholar]; (b) Coleman RS; Li J; Navarro A Total Synthesis of Azinomycin A. Angew. Chem., Int. Ed 2001, 40, 1736–1739. [DOI] [PubMed] [Google Scholar]; (c) Zhao Q; He Q; Ding W; Tang M; Kang Q; Yu Y; Deng W; Zhang Q; Fang J; Tang G; Liu W Characterization of the Azinomycin B Biosynthetic Gene Cluster Revealing a Different Iterative Type I Polyketide Synthase for Naphthoate Biosynthesis. Chem. Biol 2008, 15, 693–705. [DOI] [PubMed] [Google Scholar]; (d) Ogasawara Y; Liu H-W Biosynthetic Studies of Aziridine Formation in Azicemicins. J. Am. Chem. Soc 2009, 131, 18066–18068. [DOI] [PubMed] [Google Scholar]; (e) Nakao Y; Fujita M; Warabi K; Matsunaga S; Fusetani N Miraziridine A, a Novel Cysteine Protease Inhibitor from the Marine Sponge Theonella aff. Mirabilis. J. Am. Chem. Soc 2000, 122, 10462–10463. [DOI] [PubMed] [Google Scholar]; (f) Liu Y; Li M; Mu H; Song S; Zhang Y; Chen K; He X; Wang H; Dai Y; Lu F; Yan Z; Zhang H Identification and characterization of the ficellomycin biosynthesis gene cluster from Streptomyces ficellus. Appl. Microbiol. Biotechnol 2017, 101, 7589–7602. [DOI] [PubMed] [Google Scholar]; (g) Foulke-Abel J; Agbo H; Zhang H; Mori S; Watanabe CMH Mode of action and biosynthesis of the azabicycle-containing natural products azinomycin and ficellomycin. Nat. Prod. Rep 2011, 28, 693–704. [DOI] [PubMed] [Google Scholar]
  • (4).(a) Ismail FMD; Levitsky DO; Dembitsky VM Aziridine alkaloids as potential therapeutic agents. Eur. J. Med. Chem 2009, 44, 3373–3387. [DOI] [PubMed] [Google Scholar]; (b) Ballereau S; Andrieu-Abadie N; Saffon N; Génisson Y Synthesis and biological evaluation of aziridine-containing analogs of phytosphingosine. Tetrahedron 2011, 67, 2570–2578. [DOI] [PubMed] [Google Scholar]; (c) Dorr RT; Wisner L; Samulitis BK; Landowski TH; Remers WA Anti-tumor activity and mechanism of action for a cyanoaziridine-derivative, AMP423. Cancer Chemother. Pharmacol 2012, 69, 1039–1049. [DOI] [PubMed] [Google Scholar]; (d) Pillai B; Cherney MM; Diaper CM; Sutherland A; Blanchard JS; Vederas JC; James MNG Structural insights into stereochemical inversion by diaminopimelate epimerase: An antibacterial drug target. Proc. Natl. Acad. Sci. U. S. A 2006, 103, 8668–8673. [DOI] [PubMed] [Google Scholar]
  • (5).(a) Ujjain SK; Bhatia R; Ahuja P Aziridine-functionalized graphene: Effect of aromaticity for aryl functional groups on enhanced power conversion efficiency of organic photovoltaic cells. J. Saudi Chem. Soc 2019, 23, 655–665. [Google Scholar]; (b) Tuci G; Filippi J; Ba H; Rossin A; Luconi L; Pham-Huu C; Vizza F; Giambastiani G How to teach an old dog new (electrochemical) tricks: aziridine-functionalized CNTs as efficient electrocatalysts for the selective CO2 reduction to CO. J. Mater. Chem. A 2018, 6, 16382–16389. [Google Scholar]; (c) Moon HK; Kang S; Yoon HJ Aziridine-functionalized polydimethylsiloxanes for tailorable polymeric scaffolds: aziridine as a clickable moiety for structural modification of materials. Polym. Chem 2017, 8, 2287–2291. [Google Scholar]; (d) Tuci G; Luconi L; Rossin A; Berretti E; Ba H; Innocenti M; Yakhvarov D; Caporali S; Pham-Huu C; Giambastiani G Aziridine-Functionalized Multi-walled Carbon Nanotubes: Robust and Versatile Catalysts for the Oxygen Reduction Reaction and Knoevenagel Condensation. ACS Appl. Mater. Interfaces 2016, 8, 30099–30106.27768269 [Google Scholar]; (e) Chen R; Chen JS; Zhang C; Kessler MR Rapid room-temperature polymerization of bio-based multiaziridine-containing compounds. RSC Adv. 2015, 5, 1557–1563. [Google Scholar]; (f) Kim H-J; Chaikittisilp W; Jang K-S; Didas SA; Johnson JR; Koros WJ; Nair S; Jones CW Aziridine-Functionalized Mesoporous Silica Membranes on Polymeric Hollow Fibers: Synthesis and Single-Component CO2 and N2 Permeation Properties. Ind. Eng. Chem. Res 2015, 54, 4407–4413. [Google Scholar]
  • (6).(a) Degennaro L; Trinchera P; Luisi R Recent Advances in the Stereoselective Synthesis of Aziridines. Chem. Rev 2014, 114, 7881–7929. [DOI] [PubMed] [Google Scholar]; (b) Pellissier H Recent Developments in Asymmetric Aziridination. Adv. Synth. Catal 2014, 356, 1899–1935. [DOI] [PubMed] [Google Scholar]; (c) Pellissier H Recent developments in asymmetric aziridination. Tetrahedron 2010, 66, 1509–1555. [DOI] [PubMed] [Google Scholar]; (d) Singh GS; D’hooghe M; De Kimpe N Synthesis and Reactivity of C-Heteroatom-Substituted Aziridines. Chem. Rev 2007, 107, 2080–2135. [DOI] [PubMed] [Google Scholar]; (e) Watson IDG; Yu L; Yudin AK Advances in Nitrogen Transfer Reactions Involving Aziridines. Acc. Chem. Res 2006, 39, 194–206. [DOI] [PubMed] [Google Scholar]; (f) Osborn HMI; Sweeney J The asymmetric synthesis of aziridines. Tetrahedron: Asymmetry 1997, 8, 1693–1715. [DOI] [PubMed] [Google Scholar]; (g) Tanner D Chiral Aziridines–Their Synthesis and Use in Stereoselective Transformations. Angew. Chem., Int. Ed. Engl 1994, 33, 599–619. [DOI] [PubMed] [Google Scholar]
  • (7).(a) Darses B; Rodrigues R; Neuville L; Mazurais M; Dauban P Transition metal-catalyzed iodine(III)-mediated nitrene transfer reactions: efficient tools for challenging syntheses. Chem. Commun 2017, 53, 493–508. [DOI] [PubMed] [Google Scholar]; (b) Chang JWW; Ton TMU; Chan PWH Transition-metal-catalyzed aminations and aziridinations of C–H and C=C bonds with iminoiodinanes. Chem. Rec 2011, 11, 331–357. [DOI] [PubMed] [Google Scholar]; (c) Karila D; Dodd RH Recent Progress in Iminoiodane-Mediated Aziridination of Olefins. Curr. Org. Chem 2011, 15, 1507–1538. [DOI] [PubMed] [Google Scholar]
  • (8).(a) Chanda BM; Vyas R; Landge SS Synthesis of aziridines using new catalytic systems with bromamine-T as the nitrene source. J. Mol. Catal. A: Chem 2004, 223, 57–60. [Google Scholar]; (b) Chanda BM; Vyas R; Bedekar AV Investigations in the Transition Metal catalyzed Aziridination of Olefins, Amination, and Other Insertion Reactions with Bromamine-T as the Source of Nitrene. J. Org. Chem 2001, 66, 30–34.11429913 [Google Scholar]
  • (9).(a) Ma Z; Zhou Z; Kürti L Direct and Stereospecific Synthesis of N-H and N-Alkyl Aziridines from Unactivated Olefins Using Hydroxylamine-O-Sulfonic Acids. Angew. Chem., Int. Ed 2017, 56, 9886–9890. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Jat JL; Paudyal MP; Gao H; Xu Q-L; Yousufuddin M; Devarajan D; Ess DH; Kürti L; Falck JR Direct Stereospecific Synthesis of Unprotected N-H and N-Me Aziridines from Olefins. Science 2014, 343, 61–65. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Lebel H; Huard K; Lectard S N-Tosyloxycarbamates as a Source of Metal Nitrenes: Rhodium-Catalyzed C–H Insertion and Aziridination Reactions. J. Am. Chem. Soc 2005, 127, 14198–14199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).(a) Driver TG Recent advances in transition metal-catalyzed N-atom transfer reactions of azides. Org. Biomol. Chem 2010, 8, 3831–3846. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Intrieri D; Zardi P; Caselli A; Gallo E Organic azides: “energetic reagents” for the intermolecular amination of C–H bonds. Chem. Commun 2014, 50, 11440–11453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (11).(a) Hao W; Wu X; Sun JZ; Siu JC; MacMillan SN; Lin S Radical Redox-Relay Catalysis: Formal [3 + 2] Cycloaddition of N-Acylaziridines and Alkenes. J. Am. Chem. Soc 2017, 139, 12141–12144. [DOI] [PubMed] [Google Scholar]; (b) Goossens H; Vervisch K; Catak S; Stanković S; D’hooghe M; De Proft F; Geerlings P; De Kimpe N; Waroquier M; Van Speybroeck V Reactivity of Activated versus Nonactivated 2-(Bromomethyl)aziridines with respect to Sodium Methoxide: A Combined Computational and Experimental Study. J. Org. Chem 2011, 76, 8698–8709. [DOI] [PubMed] [Google Scholar]; (c) Li L; Zhang J Lewis Acid-catalyzed [3 + 2] Cycloaddition of Alkynes with N-Tosyl-aziridines via Carbon–Carbon Bond Cleavage: Synthesis of Highly Substituted 3-Pyrrolines. Org. Lett 2011, 13, 5940–5943. [DOI] [PubMed] [Google Scholar]; (d) Dauban P; Dodd RH PhI = NSes: A New Iminoiodinane Reagent for the Copper-Catalyzed Aziridination of Olefins. J. Org. Chem 1999, 64, 5304–5307. [DOI] [PubMed] [Google Scholar]; (e) Alonso DA; Andersson PG Deprotection of Sulfonyl Aziridines. J. Org. Chem 1998, 63, 9455–9461. [DOI] [PubMed] [Google Scholar]; (f) Bergmeier SC; Seth PP A general method for deprotection of N-toluenesulfonyl aziridines using sodium naphthalenide. Tetrahedron Lett. 1999, 40, 6181–6184. [DOI] [PubMed] [Google Scholar]; (g) Warner DL; Hibberd AM; Kalman M; Klapars A; Vedejs E N-Silyl Protecting Groups for Labile Aziridines: Application toward the Synthesis of N-H Aziridinomitosenes. J. Org. Chem 2007, 72, 8519–8522. [DOI] [PubMed] [Google Scholar]
  • (12).(a) Zhu Y; Wang Q; Cornwall RG; Shi Y Organocatalytic Asymmetric Epoxidation and Aziridination of Olefins and Their Synthetic Applications. Chem. Rev 2014, 114, 8199–8256. [DOI] [PubMed] [Google Scholar]; (b) Zhang Y; Lu Z; Wulff WD Catalytic Asymmetric Aziridination with Catalysts Derived from VAPOL and VANOL. Synlett 2009, 2009, 2715–2739. [DOI] [PubMed] [Google Scholar]; (c) Yoshimura A; Middleton KR; Zhu C; Nemykin VN; Zhdankin VV Hypoiodite-Mediated Metal-Free catalytic Aziridination of Alkenes. Angew. Chem., Int. Ed 2012, 51, 8059–8062. [DOI] [PubMed] [Google Scholar]; (d) Siu T; Yudin AK Practical Olefin Aziridination with a Broad Substrate Scope. J. Am. Chem. Soc 2002, 124, 530–531. [DOI] [PubMed] [Google Scholar]
  • (13).(a) Chandrachud PP; Jenkins DM Transition Metal Aziridination Catalysts. In Encyclopedia of Inorganic and Bioinorganic Chemistry; Wiley Online Library, 2017; pp 1–11. [DOI] [PubMed] [Google Scholar]; (b) Che C-M; Lo VK-Y; Zhou C-Y Oxidation by Metals (Nitrene). In Comprehensive Organic Synthesis II, 2nd ed.; Knochel P, Molander GA, Eds.; Elsevier: Amsterdam, 2014; Vol. 7, pp 26–85. [DOI] [PubMed] [Google Scholar]; (c) Jung N; Bräse S New Catalysts for the Transition-Metal-Catalyzed Synthesis of Aziridines. Angew. Chem., Int. Ed 2012, 51, 5538–5540. [DOI] [PubMed] [Google Scholar]; (d) Jiang H; Zhang XP Oxidation: C–N Bond Formation by Oxidation (Aziridines). In Comprehensive Chirality; Carreira EM, Yamamoto H, Eds.; Elsevier: Amsterdam, 2012; Vol. 5, pp 168–182. [DOI] [PubMed] [Google Scholar]; (e) Mößner C; Bolm C Catalyzed Asymmetric Aziridinations. In Transition Metals for Organic Synthesis, 2nd ed.; Beller M, Bolm C, Eds.; Wiley-VCH: Weinheim, Germany, 2004; pp 389–402. [DOI] [PubMed] [Google Scholar]; (f) Halfen JA Recent Advances in Metal-Mediated Carbon-Nitrogen Bond Formation Reactions: Aziridination and Amidation. Curr. Org. Chem 2005, 9, 657–669. [DOI] [PubMed] [Google Scholar]; (g) Müller P; Fruit C Enantioselective Catalytic Aziridinations and Asymmetric Nitrene Insertions into CH Bonds. Chem. Rev 2003, 103, 2905–2919. [DOI] [PubMed] [Google Scholar]
  • (14).Historic Mn/Fe porphyrinoid systems:; (a) Mansuy D; Mahy J-P; Dureault A; Bedi G; Battioni P Iron- and manganese-porphyrin catalysed aziridination of alkenes by tosyl- and acyl-iminoiodobenzene. J. Chem. Soc., Chem. Commun 1984, 1161–1163. [Google Scholar]; (b) Mahy J-P; Battioni P; Mansuy D Formation of an Iron(III) Porphyrin Complex with a Nitrene Moiety Inserted into a Fe–N Bond during Alkene Aziridination by [(Tosylimido)iodo]benzene Catalyzed by Iron(III) Porphyrins. J. Am. Chem. Soc 1986, 108, 1079–1080. [Google Scholar]; (c) Mahy J-P; Bedi G; Battioni P; Mansuy D Aziridination of alkenes catalysed by porphyrinirons: Selection of catalysts for optimal efficiency and stereospecificity. J. Chem. Soc., Perkin Trans. 2 1988, 2, 1517–1524. [Google Scholar]; (d) Groves JT; Takahashi T Activation and Transfer of Nitrogen from Nitridomanganese(V) Porphyrin Complex. The Aza Analogue of Epoxidation. J. Am. Chem. Soc 1983, 105, 2073–2074. [Google Scholar]
  • (15).For select reviews of metal porphyrinoid nitrene-transfer systems, see:; (a) Intrieri D; Carminati DM; Gallo E Recent Advances in Metal Porphyrinoid-Catalyzed Nitrene and Carbene Transfer Reactions. In Handbook of Porphyrin Science; Kadish KM, Smith KM, Guilard R, Eds.; World Scientific, 2016; Vol. 38, pp 1–99. [DOI] [PubMed] [Google Scholar]; (b) Anding BJ; Woo LK An Overview of Metalloporphyrin-Catalyzed Carbon and Nitrogen Group Transfer Reactions. In Handbook of Porphyrin Science; Kadish KM, Smith KM, Guilard R, Eds.; World Scientific, 2012; Vol. 21, pp 145–319. [DOI] [PubMed] [Google Scholar]; (c) Ruppel JV; Fields KB; Snyder NL; Zhang XP Metalloporphyrin-Catalyzed Asymmetric Atom/Group Transfer Reactions. In Handbook of Porphyrin Science; Kadish KM, Smith KM, Guilard R, Eds.; World Scientific, 2010; Vol. 10, pp 1–84. [DOI] [PubMed] [Google Scholar]; (d) Fantauzzi S; Caselli A; Gallo E Nitrene transfer reactions mediated by metallo-porphyrin complexes. Dalton Trans. 2009, 5434–5443. [DOI] [PubMed] [Google Scholar]
  • (16).For select reviews of engineered hemoproteins for nitrene transfer to alkenes, see:; (a) Brandenberg OF; Fasan R; Arnold FH Exploiting and engineering hemoproteins for abiological carbene and nitrene transfer reactions. Curr. Opin. Biotechnol 2017, 47, 102–111. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Farwell CC; Zhang RK; McIntosh JA; Hyster TK; Arnold FH Enantioselective Enzyme-Catalyzed Aziridination Enabled by Active-Site Evolution of a Cytochrome P450. ACS Cent. Sci 2015, 1, 89–93. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Gober JG; Brustad EM Non-natural carbenoid and nitrenoid insertion reactions catalyzed by heme proteins. Curr. Opin. Chem. Biol 2016, 35, 124–132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).For select examples of Mn-based nitrene-transfer to alkenes, see:; (a) Zdilla MJ; Abu-Omar MM Mechanism of Catalytic Aziridination with Manganese Corrole: The Often Postulated High-Valent Mn(V) Imido Is Not the Group Transfer Reagent. J. Am. Chem. Soc 2006, 128, 16971–16979. [DOI] [PubMed] [Google Scholar]; (b) Nishikori H; Katsuki T Catalytic and highly enantioselective aziridination of styrene derivatives. Tetrahedron Lett. 1996, 37, 9245–9248. [DOI] [PubMed] [Google Scholar]; (c) Minakata S; Ando T; Nishimura M; Ryu I; Komatsu M Novel Asymmetric and Stereospecific Aziridination of Alkenes with a Chiral Nitridomanganese Complex. Angew. Chem., Int. Ed 1998, 37, 3392–3394. [DOI] [PubMed] [Google Scholar]; (d) Du Bois J; Tomooka CS; Hong J; Carreira EM Nitridomanganese(V) Complexes: Design, Preparation, and Use as Nitrogen Atom-Transfer Reagents. Acc. Chem. Res 1997, 30, 364–372. [DOI] [PubMed] [Google Scholar]; (e) Lai T-S; Kwong H-L; Che C-M; Peng S-M Catalytic and asymmetric aziridination of alkenes catalyzed by a chiral manganese porphyrin complex. Chem. Commun 1997, 2373–2374. [DOI] [PubMed] [Google Scholar]; (f) Liang J-L; Huang J-S; Yu X-Q; Zhu N; Che C-M Metalloporphyrin-Mediated Asymmetric Nitrogen-Atom Transfer to Hydrocarbons: Aziridination of Alkenes and Amidation of Saturated C–H Bonds Catalyzed by Chiral Ruthenium and Manganese Porphyrins. Chem. - Eur. J 2002, 8, 1563–1572. [DOI] [PubMed] [Google Scholar]; (g) O’Connor KJ; Wey S-J; Burrows CJ Alkene aziridination and epoxidation catalyzed by chiral metal salen complexes. Tetrahedron Lett. 1992, 33, 1001–1004. [DOI] [PubMed] [Google Scholar]
  • (18).For select examples of Fe-based nitrene-transfer to alkenes, see:; (a) Damiano C; Intrieri D; Gallo E Aziridination of alkenes promoted by iron or ruthenium complexes. Inorg. Chim. Acta 2018, 470, 51–67. [Google Scholar]; (b) Fingerhut A; Serdyuk OV; Tsogoeva SB Non-heme iron catalysts for epoxidation and aziridination reactions of challenging terminal alkenes: towards sustainability. Green Chem. 2015, 17, 2042–2058. [Google Scholar]; (c) Correa A; Mancheño OG; Bolm C Iron-catalysed carbon-heteroatom and heteroatom-heteroatom bond forming processes. Chem. Soc. Rev 2008, 37, 1108–1117.18497924 [Google Scholar]; (d) Iovan DA; Betley TA Characterization of Iron-Imido Species Relevant for N-Group Transfer Chemistry. J. Am. Chem. Soc 2016, 138, 1983–1993.26788747 [Google Scholar]; (e) Hennessy ET; Liu RY; Iovan DA; Duncan RA; Betley TA Iron-mediated intermolecular N-group transfer chemistry with olefinic substrates. Chem. Sci 2014, 5, 1526–1532. [Google Scholar]; (f) King ER; Hennessy ET; Betley TA Catalytic C–H Bond Amination from High-Spin Iron Imido Complexes. J. Am. Chem. Soc 2011, 133, 4917–4923.21405138 [Google Scholar]; (g) Patra R; Coin G; Castro L; Dubourdeaux P; Clémancey M; Pécaut J; Lebrun C; Maldivi P; Latour J-M Rational design of Fe catalysts for olefin aziridination through DFT-based mechanistic analysis. Catal. Sci. Technol 2017, 7, 4388–4400. [Google Scholar]; (h) Avenier F; Latour J-M Catalytic aziridination of olefins and amidation of thioanisole by a non-heme iron complex. Chem. Commun 2004, 1544–1545. [Google Scholar]; (i) Liu Y; Che CM [FeIII(F20-tpp)Cl] Is an Effective Catalyst for Nitrene Transfer Reactions and Amination of Saturated Hydrocarbons with Sulfonyl and Aryl Azides as Nitrogen Source under Thermal and Microwave-Assisted Conditions. Chem. - Eur. J 2010, 16, 10494–10501.20648488 [Google Scholar]; (j) Klotz KL; Slominski LM; Hull AV; Gottsacker VM; Mas-Ballesté R; Que L Jr.; Halfen JA Non-heme iron(II) complexes are efficient olefin aziridination catalysts. Chem. Commun 2007, 2063–2065. [Google Scholar]; (k) Klotz KL; Slominski LM; Riemer ME; Phillips JA; Halfen JA Mechanism of the Iron-Mediated Alkene Aziridination reaction: Experimental and Computational Investigations. Inorg. Chem 2009, 48, 801–803.19102690 [Google Scholar]; (l) Chandrachud PP; Bass HM; Jenkins DM Synthesis of Fully Aliphatic Aziridines with a Macrocycle Tetracarbene Iron Catalyst. Organometallics 2016, 35, 1652–1657. [Google Scholar]; (m) Cramer SA; Jenkins DM Synthesis of Aziridines from Alkenes and Aryl Azides with a Reusable Macrocyclic Tetracarbene Iron Catalyst. J. Am. Chem. Soc 2011, 133, 19342–19345.22081884 [Google Scholar]; (n) Muñoz SB Jr.; Lee W-T; Dickie DA; Scepaniak JJ; Subedi D; Pink M; Johnson MD; Smith JM Styrene Aziridination by Iron(IV) Nitrides. Angew. Chem., Int. Ed 2015, 54, 10600–10603. [Google Scholar]; (o) Nakanishi M; Salit A-F; Bolm C Iron-Catalyzed Aziridination Reactions. Adv. Synth. Catal 2008, 350, 1835–1840. [Google Scholar]; (p) Mayer AC; Salit A-F; Bolm C Iron-catalysed aziridination reactions promoted by an ionic liquid. Chem. Commun 2008, 5975–5977. [Google Scholar]; (q) Heuss BD; Mayer MF; Dennis S; Hossain MM Iron mediated nitrenoid transfer: [(η5-C5H5)Fe-(CO)2(THF)]+[BF4] catalyzed aziridination of olefins. Inorg. Chim. Acta 2003, 342, 301–304. [Google Scholar]
  • (19).For select examples of Co-based nitrene-transfer to alkenes, see:; (a) Jiang H; Lang K; Lu H; Wojtas L; Zhang XP Asymmetric Radical Bicyclization of Allyl Azidoformates via Cobalt(II)-Based Metalloradical Catalysis. J. Am. Chem. Soc 2017, 139, 9164–9167. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Jiang H; Lang K; Lu H; Wojtas L; Zhang XP Intramolecular Radical Aziridination of Allylic Sulfamoyl Azides by Cobalt(II)-Based Metalloradical Catalysis: Effective Construction of Strained Heterobicyclic Structures. Angew. Chem., Int. Ed 2016, 55, 11604–11608. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Subbarayan V; Jin L-M; Cui X; Zhang XP Room temperature activation of aryloxysulfonyl azides by [Co(II)(TPP)] for selective radical aziridination of alkenes via metalloradical catalysis. Tetrahedron Lett. 2015, 56, 3431–3434. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Tao J; Jin L-M; Zhang XP Synthesis of chiral N-phosphoryl aziridines through enantioselective aziridination of alkenes with phosphoryl azide via Co(II)-based metalloradical catalysis. Beilstein J. Org. Chem 2014, 10, 1282–1289. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Jin L-M; Xu X; Lu H; Cui X; Wojtas L; Zhang XP Effective Synthesis of Chiral N-Fluoroaryl Aziridines through Enantioselective Aziridination of Alkenes with Fluoroaryl Azides. Angew. Chem., Int. Ed 2013, 52, 5309–5313. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Suarez AIO; Jiang H; Zhang XP; de Bruin B The radical mechanism of cobalt(II) porphyrin-catalyzed olefin aziridination and the importance of cooperative H-bonding. Dalton Trans. 2011, 40, 5697–5705. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Lu H; Jiang H; Hu Y; Wojtas L; Zhang XP Chemoselective intramolecular allylic C–H amination versus C=C aziridination through Co(II)-based metalloradical catalysis. Chem. Sci 2011, 2, 2361–2366. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Subbarayan V; Ruppel JV; Zhu S; Perman JA; Zhang XP Highly asymmetric cobalt-catalyzed aziridination of alkenes with trichloroethoxysulfonyl azide (TcesN3). Chem. Commun 2009, 4266–4268. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Jones JE; Ruppel JV; Gao G-Y; Moore TM; Zhang XP Cobalt-Catalyzed Asymmetric Olefin Aziridination with Diphenylphosphoryl Azide. J. Org. Chem 2008, 73, 7260–7265. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Ruppel JV; Jones JE; Huff CA; Kamble RM; Chen Y; Zhang XP A Highly Effective Cobalt Catalyst for Olefin Aziridination with Azides: Hydrogen Bonding Guided Catalyst Design. Org. Lett 2008, 10, 1995–1998. [DOI] [PMC free article] [PubMed] [Google Scholar]; (k) Gao G-Y; Jones JE; Vyas R; Harden JD; Zhang XP Cobalt-Catalyzed Aziridination with Diphenylphosphoryl Azide (DPPA): Direct Synthesis of N-Phosphorus-Substituted Aziridines from Alkenes. J. Org. Chem 2006, 71, 6655–6658. [DOI] [PMC free article] [PubMed] [Google Scholar]; (l) Gao G-Y; Harden JD; Zhang XP Cobalt-Catalyzed Efficient Aziridination of Alkenes. Org. Lett 2005, 7, 3191–3193. [DOI] [PMC free article] [PubMed] [Google Scholar]; (m) Caselli A; Gallo E; Fantauzzi S; Morlacchi S; Ragaini F; Cenini S Allylic Amination and Aziridination of Olefins by Aryl Azides Catalyzed by CoII(tpp): A Synthetic and Mechanistic Study. Eur. J. Inorg. Chem 2008, 2008, 3009–3019. [DOI] [PMC free article] [PubMed] [Google Scholar]; (n) Caselli A; Gallo E; Ragaini F; Ricatto F; Abbiati G; Cenini S Chiral porphyrin complexes of cobalt(II) and ruthenium(II) in catalytic cyclopropanation and amination reactions. Inorg. Chim. Acta 2006, 359, 2924–2932. [DOI] [PMC free article] [PubMed] [Google Scholar]; (o) Cenini S; Tollari S; Penoni A; Cereda C Catalytic amination of unsaturated hydrocarbons: reactions of p-nitrophenylazide with alkenes catalyzed by metallo-porphyrins. J. Mol. Catal. A: Chem 1999, 137, 135–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).For select examples of Ni-based nitrene-transfer to alkenes, see:; (a) Waterman R; Hillhouse GL Group Transfer from Nickel Imido, Phosphinidine, and Carbene Complexes to Ethylene with Formation of Aziridine, Phosphirane, and Cyclopropane Products. J. Am. Chem. Soc 2003, 125, 13350–13351. [DOI] [PubMed] [Google Scholar]; (b) Lin BL; Clough CR; Hillhouse GL Interactions of Aziridines with Nickel Complexes: Oxidative-Addition and Reductive-Elimination Reactions that Break and Make C–N Bonds. J. Am. Chem. Soc 2002, 124, 2890–2891. [DOI] [PubMed] [Google Scholar]
  • (21).For historic examples of Cu-based enantioselective aziridinations, see:; (a) Li Z; Quan RW; Jacobsen EN Mechanism of the (Diimine)copper-Catalyzed Asymmetric Aziridination of Alkenes. Nitrene Transfer via Ligand-Accelerated Catalysis. J. Am. Chem. Soc 1995, 117, 5889–5890. [Google Scholar]; (b) Li Z; Conser KR; Jacobsen EN Asymmetric Alkene Aziridination with Readily Available Chiral Diimine-Based Catalysts. J. Am. Chem. Soc 1993, 115, 5326–5327. [Google Scholar]; (c) Evans DA; Faul MM; Bilodeau MT; Anderson BA; Barnes DM Bis(oxazoline)-copper complexes as chiral catalysts for the enantioselective aziridination of olefins. J. Am. Chem. Soc 1993, 115, 5328–5329. [Google Scholar]
  • (22).For select examples of Cu-based nitrene-transfer to alkenes, see:; (a) Gephart RT III; Warren TH Copper-Catalyzed sp3 C–H Amination. Organometallics 2012, 31, 7728–7752. [Google Scholar]; (b) Lebel H; Parmentier M Copper-catalyzed enantioselective aziridination of styrenes. Pure Appl. Chem 2010, 82, 1827–1833. [Google Scholar]; (c) Lebel H; Parmentier M; Leogane O; Ross K; Spitz C Copper bis(oxazolines) as catalysts for stereoselective aziridination of styrenes with N-tosyloxycarbamates. Tetrahedron 2012, 68, 3396–3409. [Google Scholar]; (d) Maestre L; Sameera WMC; Díaz-Requejo MM; Maseras F; Pérez PJ A General Mechanism for the Copper- and Silver-Catalyzed Olefin Aziridination Reactions: Concomitant Involvement of the Singlet and Triplet Pathways. J. Am. Chem. Soc 2013, 135, 1338–1348.23276287 [Google Scholar]; (e) Mairena MA; Díaz-Requejo MM; Belderraín TR; Nicasio MC; Trofimenko S; Pérez PJ Copper-Homoscorpionate Complexes as Very Active Catalysts for the Olefin Aziridination Reaction. Organometallics 2004, 23, 253–256. [Google Scholar]; (f) Díaz-Requejo MM; Pérez PJ; Brookhart M; Templeton JL Substituent Effects on the Reaction Rates of Copper-Catalyzed Cyclopropanation and Aziridination of para-Substituted Styrenes. Organometallics 1997, 16, 4399–4402. [Google Scholar]; (g) Bagchi V; Paraskevopoulou P; Das P; Chi L; Wang Q; Choudhury A; Mathieson JS; Cronin L; Pardue DB; Cundari TR; Mitrikas G; Sanakis Y; Stavropoulos P A Versatile Tripodal Cu(I) Reagent for C–N Bond Construction via Nitrene-Transfer Chemistry: Catalytic Perspectives and Mechanistic Insights on C–H Aminations/Amidinations and Olefin Aziridinations. J. Am. Chem. Soc 2014, 136, 11362–11381.25025754 [Google Scholar]; (h) Lam TL; Tso KC-H; Cao B; Yang C; Chen D; Chang X-Y; Huang J-S; Che C-M Tripodal S-Ligand Complexes of Copper(I) as Catalysts for Alkene Aziridination, Sulfide Sulfimidation, and C–H Amination. Inorg. Chem 2017, 56, 4253–4257.28358495 [Google Scholar]; (i) Li Y; He J; Khankhoje V; Herdtweck E; Köhler K; Storcheva O; Cokoja M; Kühn FE Copper(II) complexes incorporating poly/perfluorinated alkoxyaluminate-type weakly coordinating anions: Syntheses, characterization and catalytic application in stereoselective olefin aziridination. Dalton Trans 2011, 40, 5746–5754.21523276 [Google Scholar]; (j) Comba P; Haaf C; Lienke A; Muruganantham A; Wadepohl H Synthesis, Structure, and Highly Efficient Copper-Catalyzed Aziridination with a Tetraaza-Bispidine Ligand. Chem. - Eur.J 2009, 15, 10880–10887.19746459 [Google Scholar]; (k) Wang X; Ding K One-Pot Enantioselective Aziridination of Olefins Catalyzed by a Copper(I) Complex of a Novel Diimine Ligand by Using PhI(OAc)2 and Sulfonamide as Nitrene Precursors. Chem. - Eur. J 2006, 12, 4568–4575.16598800 [Google Scholar]; (l) Robert-Peillard F; Di Chenna PH; Liang C; Lescot C; Collet F; Dodd RH; Dauban P Catalytic stereoselective alkene aziridination with sulfonimidamides. Tetrahedron: Asymmetry 2010, 21, 1447–1457. [Google Scholar]; (m) Kundu S; Miceli E; Farquhar E; Pfaff FF; Kuhlmann U; Hildebrandt P; Braun B; Greco C; Ray K Lewis Acid Trapping of an Elusive Copper–Tosylnitrene Intermediate Using Scandium Triflate. J. Am. Chem. Soc 2012, 134, 14710–14713.22928636 [Google Scholar]; (n) Ma L; Jiao P; Zhang Q; Du D-M; Xu J Ligand and substrate π-stacking interaction controlled enantioselectivity in the asymmetric aziridination. Tetrahedron: Asymmetry 2007, 18, 878–884. [Google Scholar]; (o) Vedernikov AN; Caulton KG Angular Ligand Constraint Yields an Improved Olefin Aziridination Catalyst. Org. Lett 2003, 5, 2591–2594.12868866 [Google Scholar]; (p) Evans DA; Faul MM; Bilodeau MT Development of the Copper-Catalyzed Olefin Aziridination Reaction. J. Am. Chem. Soc 1994, 116, 2742–2753. [Google Scholar]; (q) Kwart H; Khan AA Copper-Catalyzed Decomposition of Benzenesulfonyl Azide in Cyclohexene Solution. J. Am. Chem. Soc 1967, 89, 1951–1953. [Google Scholar]
  • (23).For select examples of Ru-based nitrene-transfer to alkenes, see:; (a) Uchida T; Katsuki T Asymmetric Nitrene Transfer Reactions: Sulfimidation, Aziridination and C–H Amination Using Azide Compounds as Nitrene Precursors. Chem. Rec 2014, 14, 117–129. [DOI] [PubMed] [Google Scholar]; (b) Kim C; Uchida T; Katsuki T Asymmetric olefin aziridination using a newly designed Ru(CO)(salen) complex as the catalyst. Chem. Commun 2012, 48, 7188–7190. [DOI] [PubMed] [Google Scholar]; (c) Kawabata H; Omura K; Uchida T; Katsuki T Construction of Robust Ruthenium(salen)(CO) Complexes and Asymmetric Aziridination with Nitrene Precursors in the Form of Azide Compounds That Bear Easily Removable N-Sulfonyl Groups. Chem. - Asian J 2007, 2, 248–256. [DOI] [PubMed] [Google Scholar]; (d) Harvey ME; Musaev D; Du Bois J A Diruthenium Catalyst for Selective, Intramolecular Allylic C–H Amination: Reaction Development and Mechanistic Insight Gained through Experiment and Theory. J. Am. Chem. Soc 2011, 133, 17207–17216. [DOI] [PubMed] [Google Scholar]; (e) Chan K-H; Guan X; Lo VK-Y; Che C-M Elevated Catalytic Activity of Ruthenium(II)–Porphyrin-Catalyzed Carbene/Nitrene Transfer and Insertion Reactions with N-Heterocyclic Carbene Ligands. Angew. Chem., Int. Ed 2014, 53, 2982–2987. [DOI] [PubMed] [Google Scholar]; (f) Leung SK-Y; Tsui W-M; Huang J-S; Che C-M; Liang J-L; Zhu N Imido Transfer from Bis(imido)ruthenium(VI) Porphyrins to Hydrocarbons: Effect of Imido Substituents, C–H Bond Dissociation Energies, and RuVI/V Reduction Potentials. J. Am. Chem. Soc 2005, 127, 16629–16640. [DOI] [PubMed] [Google Scholar]; (g) Au S-M; Huang J-S; Yu W-Y; Fung W-H; Che C-M Aziridination of Alkenes and Amidation of Alkanes by Bis(tosylimido)ruthenium(VI) Porphyrins. A Mechanistic Study. J. Am. Chem. Soc 1999, 121, 9120–9132. [DOI] [PubMed] [Google Scholar]; (h) Fantauzzi S; Gallo E; Caselli A; Piangiolino C; Ragaini F; Re N; Cenini S Rearrangement of N-Aryl-2-Vinylaziridines to Benzoazepines and Dihydropyrroles: A Synthetic and Theoretical Study. Chem. - Eur. J 2009, 15, 1241–1251. [DOI] [PubMed] [Google Scholar]; (i) Fantauzzi S; Gallo E; Caselli A; Piangiolino C; Ragaini F; Cenini S The (Porphyrin)ruthenium-Catalyzed Aziridination of Olefins Using Aryl Azides as Nitrogen Sources. Eur. J. Org. Chem 2007, 2007, 6053–6059. [DOI] [PubMed] [Google Scholar]; (j) Fantauzzi S; Gallo E; Caselli A; Ragaini F; Macchi P; Casati N; Cenini S Origin of the Deactivation in Styrene Aziridination by Aryl Azides, Catalyzed by Ruthenium Porphyrin Complexes. Structural Characterization of a Δ2-1,2,3-triazoline RuII(TPP)CO Complex. Organometallics 2005, 24, 4710–4713. [DOI] [PubMed] [Google Scholar]; (k) Zardi P; Pozzoli A; Ferretti F; Manca G; Mealli C; Gallo E A mechanistic investigation of the ruthenium porphyrin catalyzed aziridination of olefins by aryl azides. Dalton Trans. 2015, 44, 10479–10489. [DOI] [PubMed] [Google Scholar]; (l) Manca G; Gallo E; Intrieri D; Mealli C DFT Mechanistic Proposal of the Ruthenium Porphyrin-Catalyzed Allylic Amination by Organic Azides. ACS Catal. 2014, 4, 823–832. [DOI] [PubMed] [Google Scholar]
  • (24). For select examples of Rh-based nitrene-transfer to alkenes, see ref 9 and the following:; (a) Roizen JL; Harvey ME; Du Bois J Metal-Catalyzed Nitrogen-Atom Transfer Methods for the Oxidation of Aliphatic C–H Bonds. Acc. Chem. Res 2012, 45, 911–922. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Fiori KW; Espino CG; Brodsky BH; Du Bois J A mechanistic analysis of the Rh-catalyzed intramolecular C–H amination reaction. Tetrahedron 2009, 65, 3042–3051. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Fruit C; Robert-Peillard F; Bernardinelli G; Müller P; Dodd RH; Dauban P Diastereoselective rhodium-catalyzed nitrene transfer starting from chiral sulfonimidamide-derived iminoiodanes. Tetrahedron: Asymmetry 2005, 16, 3484–3487. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Müller P; Baud C; Jacquier Y The rhodium(III)-catalyzed aziridination of olefins with {[(4-Nitrophenyl)-sulfonyl]imino}phenyl-λ3-iodane. Can. J. Chem 1998, 76, 738–750. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Nageli I; Baud C; Bernardinelli G; Jacquier Y; Moraon M; Mullet P Rhodium(II)-Catalyzed CH Insertions with {[(4-Nitrophenyl)-sulfonyl]imino}phenyl-λ3-iodane. Helv. Chim. Acta 1997, 80, 1087–1105. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Catino AJ; Nichols JM; Forslund RE; Doyle MP Efficient Aziridination of Olefins Catalyzed by Mixed-Valent Dirhodium(II, III) Caprolactamate. Org. Lett 2005, 7, 2787–2790. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Liang J-L; Yuan S-X; Chan PWH; Che C-M Rhodium-(II,II) Dimer as an Efficient Catalyst for Aziridination of Sulfonamides and Amidation of Steroids. Org. Lett 2002, 4, 4507–4510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).For select examples of Pd-based nitrene-transfer to alkenes, see:; (a) Okamoto K; Oda T; Kohigashi S; Ohe K Palladium-catalyzed Decarboxylative Intramolecular Aziridination from 4H-Isoxazol-5-ones Leading to 1-Azabicyclo[3.1.0]hex-2-enes. Angew. Chem., Int. Ed 2011, 50, 11470–11473. [DOI] [PubMed] [Google Scholar]; (b) Han J; Li Y; Zhi S; Pan Y; Timmons C; Li G Palladium-catalyzed aziridination of alkenes using N,N-dichloro-ptoluenesulfonamide as nitrogen source. Tetrahedron Lett. 2006, 47, 7225–7228. [DOI] [PubMed] [Google Scholar]
  • (26).For select examples of Ag-based nitrene-transfer to alkenes, see:; (a) Mat Lani AS; Schomaker JM Site-Selective, Catalyst-Controlled Alkene Aziridination. Synthesis 2018, 50, 4462–4470. [Google Scholar]; (b) Ju M; Weatherly CD; Guzei IA; Schomaker JM Chemo- and Enantioselective Intramolecular Silver-Catalyzed Aziridinations. Angew. Chem., Int. Ed 2017, 56, 9944–9948. [Google Scholar]; (c) Dolan NS; Scamp RJ; Yang T; Berry JF; Schomaker JM Catalyst-Controlled and Tunable, Chemoselective Silver-catalyzed Intermolecular Nitrene Transfer: Experimental and Computational Studies. J. Am. Chem. Soc 2016, 138, 14658–14667.27726353 [Google Scholar]; (d) Rigoli JW; Weatherly CD; Alderson JM; Vo BT; Schomaker JM Tunable, Chemoselective Amination via Silver Catalysis. J. Am. Chem. Soc 2013, 135, 17238–17241.24187997 [Google Scholar]; (e) Huang M; Corbin JR; Dolan NS; Fry CG; Vinokur AI; Guzei IA; Schomaker JM Synthesis, Characterization, and Variable-Temperature NMR Studies of Silver(I) Complexes for Selective Nitrene Transfer. Inorg. Chem 2017, 56, 6725–6733.28509541 [Google Scholar]; (f) Weatherly C; Alderson JM; Berry JF; Hein JE; Schomaker JM Catalyst-Controlled Nitrene Transfer by Tuning Metal:Ligand Ratios: Inshight into the Mechanisms of Chemo-selectivity. Organometallics 2017, 36, 1649–1661. [Google Scholar]; (g) Scamp RJ; Rigoli JW; Schomaker JM Chemoselective silver-catalyzed nitrene insertion reactions. Pure Appl. Chem 2014, 86, 381–393. [Google Scholar]; (h) Rigoli JW; Weatherly CD; Vo BT; Neale S; Meis AR; Schomaker JM Chemoselective Allene Aziridination via Ag(I) Catalysis. Org. Lett 2013, 15, 290–293.23265391 [Google Scholar]; (i) Llaveria J; Beltrán A; Díaz-Requejo MM; Matheu MI; Castillón S; Pérez PJ Efficient Silver-Catalyzed Regioand Stereospecific Aziridination of Dienes. Angew. Chem., Int. Ed 2010, 49, 7092–7095. [Google Scholar]; (j) Li Z; He C Recent Advances in Silver-Catalyzed Nitrene, Carbene, and Silylene-Transfer Reactions. Eur. J. Org. Chem 2006, 2006, 4313–4322. [Google Scholar]; (k) Cui Y; He C Efficient Aziridination of Olefins Catalyzed by a Unique Disilver(I) Compound. J. Am. Chem. Soc 2003, 125, 16202–16203.14692757 [Google Scholar]
  • (27).For select examples of Au-based nitrene-transfer to alkenes, see:; (a) Li Z; Ding X; He C Nitrene Transfer Reactions Catalyzed by Gold Complexes. J. Org. Chem 2006, 71, 5876–5880. [DOI] [PubMed] [Google Scholar]; (b) Deng X; Baker TA; Friend CM A Pathway for NH Addition to Styrene Promoted by Gold. Angew. Chem., Int. Ed 2006, 45, 7075–7078. [DOI] [PubMed] [Google Scholar]
  • (28).(a) Singh R; Mukherjee A Metalloporphyrin Catalyzed C–H Amination. ACS Catal. 2019, 9, 3604–3617. [Google Scholar]; (b) Park Y; Kim Y; Chang S Transition Metal-Catalyzed C–H Amination: Scope, Mechanism, and Applications. Chem. Rev 2017, 117, 9247–9301.28051855 [Google Scholar]; (c) Stavropoulos P Metal-Catalyzed and Metal-Free Intermolecular Amination of Light Alkanes and Benzenes. Comments Inorg. Chem 2017, 37, 1–57. [Google Scholar]; (d) Hazelard D; Nocquet P-A; Compain P Catalytic C–H amination at its limits: challenges and solutions. Org. Chem. Front 2017, 4, 2500–2521. [Google Scholar]; (e) Collet F; Lescot C; Dauban P Catalytic C–H amination: the stereoselectivity issue. Chem. Soc. Rev 2011, 40, 1926–1936.21234469 [Google Scholar]; (f) Zalatan DN; Du Bois J Metal-Catalyzed Oxidations of C–H to C–N Bonds. Top. Curr. Chem 2009, 292, 347–378. [Google Scholar]; (g) Díaz-Requejo MM; Pérez PJ Coinage Metal Catalyzed C–H Bond Functionalization of Hydrocarbons. Chem. Rev 2008, 108, 3379–3394.18698739 [Google Scholar]
  • (29).Hu L; Chen H What Factors Control the Reactivity of Cobalt–Imido Complexes in C–H Bond Activation via Hydrogen Abstraction? ACS Catal. 2017, 7, 285–292. [Google Scholar]
  • (30).(a) Jones C; Schulten C; Rose RP; Stasch A; Aldridge S; Woodul WD; Murray KS; Moubaraki B; Brynda M; La Macchia G; Gagliardi L Amidinato– and Guanidinato–Cobalt(I) Complexes: Characterization of Exceptionally Short Co–Co Interactions. Angew. Chem., Int. Ed 2009, 48, 7406–7410. [DOI] [PubMed] [Google Scholar]; (b) Cowley RE; Bontchev RP; Sorrell J; Sarracino O; Feng Y; Wang H; Smith JM Formation of a Cobalt(III) Imido from a Cobalt(II) Amido Complex. Evidence for Proton-Coupled Electron Transfer. J. Am. Chem. Soc 2007, 129, 2424–2425. [DOI] [PubMed] [Google Scholar]; (c) Mehn MP; Brown SD; Jenkins DM; Peters JC; Que L, Jr Vibrational Spectroscopy and Analysis of Pseudo-tetrahedral Complexes with Metal Imido Bonds. Inorg. Chem 2006, 45, 7417–7427. [DOI] [PubMed] [Google Scholar]; (d) Dai X; Kapoor P; Warren TH [Me2NN]Co(η6-toluene): O = O, N = N, and O = N Bond Cleavage Provides β-Diketiminato Cobalt μ-Oxo and Imido Complexes. J. Am. Chem. Soc 2004, 126, 4798–4799. [DOI] [PubMed] [Google Scholar]; (e) Hu X; Meyer K Terminal Cobalt(III) Imido Complexes Supported by Tris(Carbene) Ligands: Imido Insertion into the Cobalt–Carbene Bond. J. Am. Chem. Soc 2004, 126, 16322–16323. [DOI] [PubMed] [Google Scholar]; (f) Betley TA; Peters JC Dinitrogen Chemistry from Trigonally Coordinated Iron and Cobalt Platforms. J. Am. Chem. Soc 2003, 125, 10782–10783. [DOI] [PubMed] [Google Scholar]; (g) Jenkins DM; Betley TA; Peters JC Oxidative Group Transfer to Co(I) Affords a Terminal Co(III) Imido Complex. J. Am. Chem. Soc 2002, 124, 11238–11239. [DOI] [PubMed] [Google Scholar]
  • (31).(a) Shay DT; Yap GPA; Zakharov LN; Rheingold AL; Theopold KH Intramolecular C–H Activation by an Open-Shell Cobalt(III) Imido Complex. Angew. Chem., Int. Ed 2005, 44, 1508–1510. [DOI] [PubMed] [Google Scholar]; (b) Thyagarajan S; Shay DT; Incarvito CD; Rheingold AL; Theopold KH Intramolecular C–H Activation by Inferred Terminal Cobalt Imido Intermediates. J. Am. Chem. Soc 2003, 125, 4440–4441. [DOI] [PubMed] [Google Scholar]
  • (32).(a) King ER; Sazama GT; Betley TA Co(III) Imidos Exhibiting Spin Crossover and C–H Bond Activation. J. Am. Chem. Soc 2012, 134, 17858–17861. [DOI] [PubMed] [Google Scholar]; (b) Baek Y; Betley TA Catalytic C–H Amination Mediated by Dipyrrin Cobalt Imidos. J. Am. Chem. Soc 2019, 141, 7797–7806. [DOI] [PubMed] [Google Scholar]; (c) Baek Y; Hennessy ET; Betley TA Direct Manipulation of Metal Imido Geometry: Key Principles to Enhance C–H Amination Efficacy. J. Am. Chem. Soc 2019, 141, 16944–16953. [DOI] [PubMed] [Google Scholar]
  • (33).Reckziegel A; Pietzonka C; Kraus F; Werncke CG C–H Bond Activation by an Imido Cobalt(III) and the Resulting Amido Cobalt(II) Complex. Angew. Chem., Int. Ed 2020, 59, 8527–8531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Wu B; Sánchez RH; Bezpalko MW; Foxman BM; Thomas CM Formation of Heterobimetallic Zirconium/Cobalt Diimido Complexes via a Four-Electron Transformation. Inorg. Chem 2014, 53, 10021–10023. [DOI] [PubMed] [Google Scholar]
  • (35).Chomitz WA; Arnold J Reactivity of a Co(I) [N2P2] complex with azides: evidence for a transient Co(III) imido species. Chem. Commun 2008, 3648–3650. [DOI] [PubMed] [Google Scholar]
  • (36).(a) Yao X-N; Du J-Z; Zhang Y-Q; Leng X-B; Yang M-W; Jiang S-D; Wang Z-X; Ouyang Z-W; Deng L; Wang B-W; Gao S Two-Coordinate Co(II) Imido Complexes as Outstanding Single-Molecule Magnets. J. Am. Chem. Soc 2017, 139, 373–380. [DOI] [PubMed] [Google Scholar]; (b) Du J; Wang L; Xie M; Deng L A Two-Coordinate Cobalt(II) Imido Complex with NHC Ligation: Synthesis, Structure, and Reactivity. Angew. Chem., Int. Ed 2015, 54, 12640–12644. [DOI] [PubMed] [Google Scholar]
  • (37).Liu Y; Du J; Deng L Synthesis, Structure, and Reactivity of Low-Spin Cobalt(II) Imido Complexes [(Me3P)3Co(NAr)]. Inorg. Chem 2017, 56, 8278–8286. [DOI] [PubMed] [Google Scholar]
  • (38).Zhang L; Liu Y; Deng L Three-Coordinate Cobalt(IV) and Cobalt(V) Imido Complexes with N-Heterocyclic Carbene Ligation: Synthesis, Structure, and Their Distinct Reactivity in C–H Bond Amination. J. Am. Chem. Soc 2014, 136, 15525–15528. [DOI] [PubMed] [Google Scholar]
  • (39).(a) Goswami M; Rebreyend C; de Bruin B Porphyrin Cobalt(III) “Nitrene Radical” Reactivity; Hydrogen Atom Transfer from Ortho-YH Substituents to the Nitrene Moiety of Cobalt-Bound Aryl Nitrene Intermediates (Y = O, NH). Molecules 2016, 21, 242. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Goswami M; Lyaskovskyy V; Domingos SR; Buma WJ; Woutersen S; Troeppner O; Ivanović-Burmazović I; Lu H; Cui X; Zhang XP; Reijerse EJ; DeBeer S; van Schooneveld MM; Pfaff FF; Ray K; de Bruin B Characterization of Porphyrin-Co(III)-′Nitrene Radical′ Species Relevant in Catalytic Nitrene Transfer Reactions. J. Am. Chem. Soc 2015, 137, 5468–5479. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Hopmann KH; Ghosh A Mechanism of Cobalt-Porphyrin–Catalyzed Aziridination. ACS Catal. 2011, 1, 597–600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (40).van Leest NP; Tepaske MA; Oudsen J-PH; Venderbosch B; Rietdijk NR; Siegler MA; Tromp M; van der Vlugt JI; de Bruin B Ligand Redox Noninnocence in [CoIII(TAML)]0/− Complexes Affects Nitrene Formation. J. Am. Chem. Soc 2020, 142, 552–563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (41).van Leest NP; Tepaske MA; Venderbosch B; Oudsen J-PH; Tromp M; van der Vlugt JI; de Bruin B Electronically Asynchronous Transition States for C–N Bond Formation by Electrophilic [CoIII(TAML)]-Nitrene Radical Complexes Involving Substrate-to-Ligand Single-Electron Transfer and a Cobalt-Centered Spin Shuttle. ACS Catal. 2020, 10, 7449–7463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (42).Coin G; Patra R; Rana S; Biswas JP; Dubourdeaux P; Clémancey M; de Visser SP; Maiti D; Maldivi P; Latour J-M Fe-Catalyzed Aziridination Is Governed by the Electron Affinity of the Active Imido-Iron Species. ACS Catal. 2020, 10, 10010–10020. [Google Scholar]
  • (43).Baek Y; Das A; Zheng S-L; Reibenspies JH; Powers DC; Betley TA C–H Amination Mediated by Cobalt Organoazide Adducts and the Corresponding Cobalt Nitrenoid Intermediates. J. Am. Chem. Soc 2020, 142, 11232–11243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (44).Bagchi V; Kalra A; Das P; Paraskevopoulou P; Gorla S; Ai L; Wang Q; Mohapatra S; Choudhury A; Sun Z; Cundari TR; Stavropoulos P Comparative Nitrene-Transfer Chemistry to Olefinic Substrates Mediated by a Library of Anionic Mn(II) Triphenylamido-Amine Reagents and M(II) Congeners (M = Fe, Co, Ni) Favoring Aromatic over Aliphatic Alkenes. ACS Catal. 2018, 8, 9183–9206. [Google Scholar]
  • (45).(a) Jones MB; MacBeth CE Tripodal Phenylamine-Based Ligands and Their CoII Complexes. Inorg. Chem 2007, 46, 8117–8119. [DOI] [PubMed] [Google Scholar]; (b) Çelenligil-Çetin R; Paraskevopoulou P; Dinda R; Staples RJ; Sinn E; Rath NP; Stavropoulos P Synthesis, Characterization, and Reactivity of Iron Trisamidoamine Complexes that Undergo both Metal- and Ligand-Centered Oxidative Transformations. Inorg. Chem 2008, 47, 1165–1172. [DOI] [PubMed] [Google Scholar]; (c) Çelenligil-Çetin R Synthesis, Characterization, and Reactivity of Iron Complexes with N-Donor Ligands in Relation to Oxygenation of Hydrocarbons, Ph.D. Thesis, Boston University, Boston, MA, 2004. [DOI] [PubMed] [Google Scholar]
  • (46).(a) Paraskevopoulou P; Lin A; Wang Q; Pinnapareddy D; Acharrya R; Dinda R;Çelenligil-Çetin R; Floros G; Sanakis Y; Choudhury A; Rath NP; Stavropoulos P Synthesis and Characterization of a Series of Structurally and Electronically Diverse Fe(II) Complexes Featuring a Family of Triphenylamido-amine Ligands. Inorg. Chem 2010, 49, 108–122. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Pinnapareddy D Synthesis and Characterization of Metal Reagents Mediating C–X Activation (X = Cl,H) of Hydrocarbons, Ph.D. Thesis, University of Missouri–Rolla, Rolla, MO, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).(a) Jones MB; Hardcastle KI; MacBeth CE Synthetic, spectral and structural studies of mononuclear tris(κ2-amidate) aluminium complexes supported by tripodal ligands. Polyhedron 2010, 29, 116–119. [Google Scholar]; (b) Jones MB; Hardcastle KI; Hagen KS; MacBeth CE Oxygen Activation and Intramolecular C–H Bond Activation by an Amidate-Bridged Diiron(II) Complex. Inorg. Chem 2011, 50, 6402–6404.21667986 [Google Scholar]; (c) Villanueva O; Weldy NM; Blakey SB; MacBeth CE Cobalt catalyzed sp3 C–H amination utilizing aryl azides. Chem. Sci 2015, 6, 6672–6675.29435216 [Google Scholar]
  • (48).Bagchi V; Raptopoulos G; Das P; Christodoulou S; Wang Q; Ai L; Choudhury A; Pitsikalis M; Paraskevopoulou P; Stavropoulos P Synthesis and characterization of a family of Co(II) triphenylamido-amine complexes and catalytic activity in controlled radical polymerization of olefins. Polyhedron 2013, 52, 78–90. [Google Scholar]
  • (49).Çelenligil-Çetin R; Paraskevopoulou P; Lalioti N; Sanakis Y; Staples RJ; Rath NP; Stavropoulos P Metalloradical Complexes of Manganese and Chromium Featuring an Oxidatively Rearranged Ligand. Inorg. Chem 2008, 47, 10998–11009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (50).(a) Bencini A; Benelli C; Gatteschi D; Zanchini C ESR Spectra of Low-Symmetry High-Spin Cobalt(II) Complexes. 2. Pseudotetrahedral Dichlorobis(triphenylphospine oxide)cobalt(II). Inorg. Chem 1979, 18, 2137–2140. [Google Scholar]; (b) Benelli C; Gatteschi D ESR Spectra of Low-Symmetry High-Spin Cobalt(II) Complexes. 10. Five-Coordinated Trigonal-Bipyramidal Complexes. Inorg. Chem 1982, 21, 1788–1790. [Google Scholar]; (c) Drulis H; Dyrek K; Hoffmann KP; Hoffmann SK; Weselucha-Birczynska A ESR Spectra of Low-Symmetry Tetrahedral High-Spin Cobalt(II) in a Cinchoninium Tetrachlorocobaltate(II) Dihydrate Single Crystal. Inorg. Chem 1985, 24, 4009–4012. [Google Scholar]; (d) Makinen MW; Kuo LC; Yim MB; Wells GB; Fukuyama JM; Kim JE Ground Term Splitting of High-Spin Co2+ as a Probe of Coordination Structure. 1. Dependence of the Splitting on Coordination Geometry. J. Am. Chem. Soc 1985, 107, 5245–5255. [Google Scholar]; (e) Bennett B EPR of Cobalt-Substituted Zinc Enzymes. In Metals in Biology: Applications of High-Resolution EPR to Metalloenzymes, Biological Magnetic Resonance; Hanson G, Berliner L, Eds.; Springer-Verlag: New York, 2010; Vol. 29, pp 345–366. [Google Scholar]; (f) Zolnhofer EM; Käß M; Khusniyarov MM; Heinemann FW; Maron L; van Gastel M; Bill E; Meyer K An Intermediate Cobalt(IV) Nitrido Complex and its N-Migratory Insertion Product. J. Am. Chem. Soc 2014, 136, 15072–15078.25243488 [Google Scholar]; (g) Antholine WE Resolved Hyperfine at L-band for High-Spin CoEDTA, A Model for Co Sites in Proteins. Int. J. Mol. Sci 2019, 20, 2385. [Google Scholar]
  • (51).Abragam A; Bleaney B Electron Paramagnetic Resonance of Transition Ions; Clarendon Press: Oxford, 1970. [Google Scholar]
  • (52).Pilbrow JR Effective g values for S = 3/2 and S = 5/2. J. Magn. Reson 1978, 31, 479–490. [Google Scholar]
  • (53).Petasis DT; Hendrich MP Quantitative Interpretation of Multifrequency Multimode EPR Spectra of Metal Containing Proteins, Enzymes, and Biomimetic Complexes. In Electron Paramagnetic Resonance Investigations of Biological Systems by Using Spin Labels, Spin Probes, and Intrinsic Metal Ions, Part A; Qin PZ, Warncke K, Eds.; Methods in Enzymology; Elsevier, 2015; Vol. 563, Chapter 8, pp 171–208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Müller P; Baud C; Jacquier Y A Method for Rhodium(II)-Catalyzed Aziridination of Olefins. Tetrahedron 1996, 52, 1543–1548. [Google Scholar]
  • (55).Al-Ajlouni A; Espenson JH Epoxidation of Styrenes by Hydrogen Peroxide As Catalyzed by Methylrhenium Trioxide. J. Am. Chem. Soc 1995, 117, 9243–9250. [Google Scholar]
  • (56).Souto JA; Zian D; Muñiz K Iodine(III)-Mediated Intermolecular Allylic Amination under Metal-Free Conditions. J. Am. Chem. Soc 2012, 134, 7242–7245. [DOI] [PubMed] [Google Scholar]
  • (57).Jiang X-K Establishment and Successful Application of the σJJ• Scale of Spin-Delocalization Substituent Constants. Acc. Chem. Res 1997, 30, 283–289. [Google Scholar]
  • (58).Han H; Park SB; Kim SK; Chang S Copper–Nitrenoid Formation and Transfer in Catalytic Olefin Aziridination Utilizing Chelating 2-Pyridylsulfonyl Moieties. J. Org. Chem 2008, 73, 2862–2870. [DOI] [PubMed] [Google Scholar]
  • (59).Liang S; Jensen MP Half-Sandwich Scorpionates as Nitrene Transfer Catalysts. Organometallics 2012, 31, 8055–8058. [Google Scholar]
  • (60).Liu P; Wong EL-M; Yuen AW-H; Che C-M Highly Efficient Alkene Epoxidation and Aziridination catalyzed by Iron(II) Salt + 4,4′,4″-Trichloro-2,2′:6′,2″-terpyridine/4,4″-Dichloro-4′-O-PEG-OCH3-2,2′:6′,2″-terpyridine. Org. Lett 2008, 10, 3275–3278. [DOI] [PubMed] [Google Scholar]
  • (61).Dinçtürk S; Jackson RA Free radical reactions in solution. Part 7. Substituent effects on free radical reactions: comparison of the σ• scale with other measures of radical stabilization. J. Chem. Soc., Perkin Trans. 2 1981, 2, 1127–1131. [Google Scholar]
  • (62).Jupp AR; Johnstone TC; Stephan DW Improving the Global Electrophilicity Index (GEI) as a Measure of Lewis Acidity. Inorg. Chem 2018, 57, 14764–14771. [DOI] [PubMed] [Google Scholar]
  • (63).(a) Johnson ER; Becke AD A post-Hartree–Fock model of intermolecular interactions. J. Chem. Phys 2005, 123, No. 024101. [DOI] [PubMed] [Google Scholar]; (b) Grimme S; Ehrlich S; Goerigk L Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem 2011, 32, 1456–1465. [DOI] [PubMed] [Google Scholar]; (c) Grimme S; Antony J; Ehrlich S; Krieg H A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys 2010, 132, 154104. [DOI] [PubMed] [Google Scholar]
  • (64).(a) Lu X; Li X-X; Lee Y-M; Jang Y; Seo MS; Hong S; Cho K-B; Fukuzumi S; Nam W Electron-Transfer and Redox Reactivity of High-Valent Iron Imido and Oxo Complexes with the Formal Oxidation States of Five and Six. J. Am. Chem. Soc 2020, 142, 3891–3904. [DOI] [PubMed] [Google Scholar]; (b) Sabenya G; Gamba I; Gómez L; Clémancey M; Frisch JR; Klinker EJ; Blondin G; Torelli S; Que L Jr.; Martin-Diaconescu V; Latour J-M; Lloret-Fillol J; Costas M Octahedral iron(iv)–tosylimido complexes exhibiting single electron-oxidation reactivity. Chem. Sci 2019, 10, 9513–9529. [DOI] [PubMed] [Google Scholar]; (c) Hong S; Lu X; Lee Y-M; Seo MS; Ohta T; Ogura T; Clémancey M; Maldivi P; Latour J-M; Sarangi R; Nam W Achieving One-Electron Oxidation of a Mononuclear Nonheme Iron(V)-Imido Complex. J. Am. Chem. Soc 2017, 139, 14372–14375. [DOI] [PubMed] [Google Scholar]; (d) Kumar S; Faponle AS; Barman P; Vardhaman AK; Sastri CV; Kumar D; de Visser SP Long-range Electron Transfer Triggers Mechanistic Differences between Iron(IV)-Oxo and Iron(IV)-Imido Oxidants. J. Am. Chem. Soc 2014, 136, 17102–17115. [DOI] [PubMed] [Google Scholar]
  • (65).(a) Ogliaro F; Harris N; Cohen S; Filatov M; de Visser SP; Shaik S A Model “Rebound” Mechanism of Hydroxylation by Cytochrome P450: Stepwise and Effectively Concerted Pathways, and Their Reactivity Patterns. J. Am. Chem. Soc 2000, 122, 8977–8989. [Google Scholar]; (b) Shaik S; de Visser SP; Ogliaro F; Schwarz H; Schröder D Two-state reactivity mechanisms of hydroxylation and epoxidation by cytochrome P-450 revealed by theory. Curr. Opin. Chem. Biol 2002, 6, 556–567.12413538 [Google Scholar]; (c) Moreau Y; Chen H; Derat E; Hirao H; Bolm C; Shaik S NR Transfer Reactivity of Azo-Compound I of P450. How Does the Nitrogen Substituent Tune the Reactivity of the Species toward C–H and C=C Activation? J. Phys. Chem. B 2007, 111, 10288–10299.17676893 [Google Scholar]
  • (66).(a) Kuijpers PF; van der Vlugt JI; Schneider S; de Bruin B Nitrene Radical Intermediates in Catalytic Synthesis. Chem. - Eur. J 2017, 23, 13819–13829. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Xiong T; Zhang Q New amination strategies based on nitrogen-centered radical chemistry. Chem. Soc. Rev 2016, 45, 3069–3087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (67).(a) Xiong P; Xu H-C Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res 2019, 52, 3339–3350. [DOI] [PubMed] [Google Scholar]; (b) Wang P; Zhao Q; Xiao W; Chen J Recent advances in visible-light photoredox-catalyzed nitrogen radical cyclization. Green Synth.Catal 2020, 1, 42–51. [DOI] [PubMed] [Google Scholar]; (c) Zheng S; Gutiérrez-Bonet Á; Molander GA Merging Photoredox PCET with Ni-Catalyzed Cross-Coupling: Cascade Amidoarylation of Unactivated Olefins. Chem. 2019, 5, 339–352. [DOI] [PubMed] [Google Scholar]; (d) Davies J; Svejstrup TD; Reina DF; Sheikh NS; Leonori D Visible-Light-Mediated Synthesis of Amidyl Radicals: Transition-Metal-Free Hydroamination and N-Arylation Reactions. J. Am. Chem. Soc 2016, 138, 8092–8095. [DOI] [PubMed] [Google Scholar]
  • (68).Wang Q; Wang P; Gao X; Wang D; Wang S; Liang X; Wang L; Zhang H; Lei A Regioselective/electro-oxidative intermolecular [3 + 2] annulation for the preparation of indolines. Chem. Sci 2020, 11, 2181–2186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (69).(a) Usharani D; Janardanan D; Li C; Shaik S A Theory for Bioinorganic Chemical Reactivity of Oxometal Complexes and Analogous Oxidants: The Exchange and Orbital-Selection Rules. Acc. Chem. Res 2013, 46, 471–482. [DOI] [PubMed] [Google Scholar]; (b) Swart M; Gruden M Spinning around in Transition-Metal Chemistry. Acc. Chem. Res 2016, 49, 2690–2697. [DOI] [PubMed] [Google Scholar]
  • (70).Ren Y; Cheaib K; Jacquet J; Vezin H; Fensterbank L; Orio M; Blanchard S; Desage-El Murr M Copper-Catalyzed Aziridination with Redox-Active Ligands: Molecular Spin Catalysis. Chem. - Eur. J 2018, 24, 5086–5090. [DOI] [PubMed] [Google Scholar]
  • (71).Schümann JM; Wagner JP; Eckhardt AK; Quanz H; Schreiner PR Intramolecular London Dispersion Interactions Do Not Cancel in Solution. J. Am. Chem. Soc 2021, 143, 41–45. [DOI] [PubMed] [Google Scholar]
  • (72).(a) Rösel S; Becker J; Allen WD; Schreiner PR Probing the Delicate Balance between Pauli Repulsion and London Dispersion with Triphenylmethyl Derivatives. J. Am. Chem. Soc 2018, 140, 14421–14432. [DOI] [PubMed] [Google Scholar]; (b) Rösel S; Balestrieri C; Schreiner PR Sizing the role of London dispersion in the dissociation of all-meta tert-butyl hexaphenylethane. Chem. Sci 2017, 8, 405–410. [DOI] [PubMed] [Google Scholar]; (c) Grimme S; Schreiner PR Steric Crowding Can Stabilize a Labile Molecule: Solving the Hexaphenylethane Riddle. Angew. Chem., Int. Ed 2011, 50, 12639–12642. [DOI] [PubMed] [Google Scholar]
  • (73).(a) Bursch M; Caldeweyher E; Hansen A; Neugebauer H; Ehlert S; Grimme S Understanding and Quantifying London Dispersion Effects in Organometallic Complexes. Acc. Chem. Res 2019, 52, 258–266. [DOI] [PubMed] [Google Scholar]; (b) Grimme S; Huenerbein R; Ehrlich S On the Importance of the Dispersion Energy for the Thermodynamic Stability of Molecules. ChemPhysChem 2011, 12, 1258–1261. [DOI] [PubMed] [Google Scholar]
  • (74).(a) Power PP An Update on Multiple Bonding between Heavier Main Group Elements: The Importance of Pauli Repulsion, Charge-Shift Character, and London Dispersion Force Effects. Organometallics 2020, 39, 4127–4138. [Google Scholar]; (b) Liptrot D; Power PP London dispersion forces in sterically crowded inorganic and organometallic molecules. Nat. Rev. Chem 2017, 1, No. 0004. [Google Scholar]
  • (75).Wagner JP; Schreiner PR London Dispersion in Molecular Chemistry – Reconsidering Steric Effects. Angew. Chem., Int. Ed 2015, 54, 12274–12296. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Information

RESOURCES