Abstract
We report the catalytic C–H amination mediated by an isolable CoIII imido complex (TrL)Co(NR) supported by a sterically demanding dipyrromethene ligand (TrL = 5-mesityl-1,9-(trityl)dipyrrin). Metalation of (TrL)Li with CoCl2 in THF afforded a high-spin (S = 3/2) three-coordinate complex (TrL)CoCl. Chemical reduction of (TrL)CoCl with potassium graphite yielded the high-spin (S = 1) CoI synthon (TrL)Co which is stabilized through an intramolecular η6-arene interaction. Treatment of (TrL)Co with a stoichiometric amount of 1-azidoadamantane (AdN3) furnished a three-coordinate, diamagnetic CoIII imide (TrL)Co(NAd) as confirmed by single-crystal X-ray diffraction, revealing a rare trigonal pyramidal geometry with an acute Co–Nimido–C angle 145.0(3)°. Exposure of 1–10 mol % of (TrL)Co to linear alkyl azides (RN3) resulted in catalytic formation of substituted N-heterocycles via intramolecular C–H amination of a range of C–H bonds, including primary C–H bonds. The mechanism of the C–N bond formation was probed via initial rate kinetic analysis and kinetic isotope effect experiments [kH/kD = 38.4(1)], suggesting a stepwise H−atom abstraction followed by radical recombination. In contrast to the previously reported C–H amination mediated by (ArL)Co(NR) (ArL = 5-mesityl-1,9-(2,4,6-Ph3C6H2)dipyrrin), (TrL)Co(NR) displays enhanced yields and rates of C–H amination without the aid of a cocatalyst, and no catalyst degradation to a tetrazene species was observed, as further supported by the pyridine inhibition effect on the rate of C–H amination. Furthermore, (TrL)Co(NAd) exhibits an extremely low one-electron reduction potential ( = −1.98 V vs [Cp2Fe]+/0) indicating that the highly basic terminal imido unit contributes to the driving force for H−atom abstraction.
Graphical Abstarct
1. INTRODUCTION
The development of a cost-effective and efficient transition-metal catalyst for direct C–H functionalization has been of interest to synthesize value-added commodity chemicals from simple hydrocarbon feedstocks.1,2 To this end, the use of bioinspired transition metal complexes featuring metal−ligand multiple bonds (MLMBs) has emerged as an appealing C–H functionalization methodology.3–5 In this regard, the iron-oxo complex (compound I) of Cytochrome P450 demonstrates nature’s key strategy to render the iron-oxo multiple bond highly reactive, thereby enhancing its O−atom transfer efficacy.6 Specifically, the open-shell electronic configuration, high oxidation state of compound I, and highly basic oxo unit of its one-electron reduced form have been proposed as leading features enabling facile C–H hydroxylation at ambient conditions.7,8
As such, the isoelectronic synthetic analogues, metal nitrenoid complexes (e.g., imido, M(NR2−);4,9–11 iminyl, M(2NR−);12–19 nitrene adducts, M(3NR)20–23) have received great attention as viable N−group transfer reagents into unactivated C–H bonds. In particular, late, first-row transition metal-based nitrenoid complexes have been extensively studied for direct C−N bond formation (Mn,24 Fe,18,19,25–34 Co,35–43 Ni,10,11,44,45 and Cu23,46,47). Among the structurally characterized metal nitrenoid complexes, however, only a few manifest productive C–H amination due to the typically strong MLMBs between the metal ion and nitrene unit.33,34,38,39,44,46 Thus, in order to improve the efficacy of C–H amination, understanding the key factors governing the lability of the terminal nitrenoid moiety is necessary. Akin to the enzymatic C–H hydroxylation, assessing redox potentials and basicity of relevant species (e.g., imides or amides) involved in C–H amination processes operating via H−atom abstraction (HAA) is essential to comprehend the thermodynamic and kinetic properties of the desired transformations.48–54
From a synthetic perspective, these properties ( and pKa) can be tuned by various elements such as identity or oxidation state of a metal center(s),55 electronic configuration of metal nitrenoid complexes,19,33,56 and the local geometry of the terminal nitrenoid moiety.57 Previously, we have demonstrated that manipulation of the spin state of the metal imide33 and the spin density at the nitrogen atom of the nitrenoid unit19 can serve as a synthetic approach to weaken the MLMBs and enable facile C–H amination.
In contrast, the direct manipulation of the geometry of a metal nitrenoid and its effect on C–H amination efficacy is relatively less explored. Nonetheless, Holland and co-workers have proposed that the geometry of the terminally bound imido unit can alter the basicity of the metal imido complex. They found enhanced C–H activation reactivity from the β-diketiminate FeIII imido complex upon pyridine coordination.57–59 The pyridine-bound iron imide was proposed as more bent and basic compared to the pyridine-free iron imide. However, structural characterization of the bent imide is absent, and thus, a direct correlation between the bent imido geometry and its basicity is lacking.
Recently, our group showcased that the geometry of the imido unit can also affect the rate of C–H amination.38 We have demonstrated the intramolecular C–H amination mediated by (ArL)Co(NR) (ArL = 5-mesityl-1,9-(2,4,6-Ph3C6H2)dipyrrin) and found that addition of pyridine improves both yields and rates of C–H amination. Exogenous pyridine coordinates to the trigonal planar (ArL)Co(NR) to generate the pyridine-bound imide (ArL)Co(NR)(py) providing steric protection to circumvent formation of a catalytically inactive cobalt tetrazido complex. More importantly, the rate of C–H amination was accelerated in the presence of pyridine as cocatalyst. We have ascribed the enhanced rate of C–H amination from (ArL)Co(NR)(py) to the bent and attenuated MLMB (Co−NR) as suggested by density functional theory (DFT) calculations, yet isolation of the bent (ArL)Co(NR)-(py) was hampered due to the weak binding affinity of pyridine.
Motivated by these precedents, we postulated that generating a metal-bound nitrenoid unit in a bent geometry can be an alternative strategy to engender a more reactive MLMB and thus improve C–H amination efficacy both kinetically and thermodynamically.52 Toward this, we sought to design a ligand that would enable synthesis of a cobalt imido complex featuring a significantly bent nitrenoid unit in the absence of exogenous ligands (e.g., pyridine). Specifically, we sought to investigate the following questions: (1) can we synthesize a dipyrrinato CoIII imido complex featuring a bent nitrenoid unit by modifying the substituents on the dipyrrin scaffold? If possible, (2) is the imide competent for catalytic C–H amination to afford substituted N-heterocycles? And (3) how does the geometry of the metal-bound nitrenoid moiety alter the efficacy and reactivity profile for C–H amination with respect to yields, rates, and reaction scope?
Toward this goal, we report herein the synthesis of a trityl-substituted dipyrromethene platform (TrL) (TrL = 5-mesityl-1,9-(trityl)dipyrrin) that provides access to a robust CoI catalyst (TrL)Co and the corresponding bent CoIII imido complex (TrL)Co(NR), which enables a facile CoI/III catalytic C–H amination. Detailed mechanistic studies on the C–H amination processes establish the improved C–H amination efficacy of (TrL)Co(NR) as compared to the previously reported linear (ArL)Co(NR), demonstrating the importance of modulating the primary coordination sphere to effect more reactive and synthetically applicable C–H amination catalysis.
2. RESULTS AND DISCUSSION
2.1. Synthesis and Characterization of Cobalt Synthons.
As discussed in the Introduction, the ligand design principle to synthesize a bent CoIII imide was informed by the previously studied ligand-accelerated C–H amination processes mediated by (ArL)Co(NR) (ArL = 5-mesityl-1,9-(2,4,6-Ph3 C6 H2 )dipyrrin).38 Specifically, we sought to emulate the primary coordination environment of (ArL)Co(NR)(py) by modifying the flanking groups of the dipyrrin scaffold. To this end, trityl moieties were considered to be more voluminous ligand flanking units than the 2,4,6-triphenylphenyl substituents of (ArL). Thus, we envisioned that the steric profile of the trityl groups may enforce a bent geometry on the nitrenoid unit and eliminate formation of cobalt tetrazido complexes even in the absence of exogenous pyridine. Additionally, we proposed that one of the phenyl rings of the trityl substituents may stabilize the low-coordinate cobalt center via an intramolecular η6-arene interaction affording a roboust CoI synthon akin to the previously reported iron and cobalt synthons such as (ArL)Fe33 and (ArL)Co.38
With these design principles in mind, we first prepared 2-trityl-1H-pyrrole via Friedel−Crafts alkylation of excess pyrrole with trityl chloride to avoid bis-alkylation. Following literature procedures,18 subsequent condensation with mesitylaldehyde dimethyl acetal followed by oxidation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) afforded (TrL)H (Scheme 1a). Subsequent deprotonation of (TrL)H with PhLi yielded the lithiated scaffold (TrL)Li. Metalation of (TrL)Li with CoCl2 in THF for 20 h at room temperature afforded a three-coordinate complex (TrL)CoCl (1) (Scheme 1b). Single crystals of 1 suitable for X-ray diffraction were grown from a layered 1:2 benzene/hexanes solution at room temperature. In contrast to the previously reported trigonal planar CoII complex (ArL)CoBr, 1 adopts a trigonal pyramidal geometry, displaying an η1-arene interaction between the cobalt and one of the ortho carbons from the trityl substituents (Figure 1a). The discrete η1-arene interaction [Co–Cortho = 2.345(7) Å] is of note as it occupies the equatorial position of the dipyrrin plane positioning the chloride on the axial site. Solid-state magnetic susceptibility of 1 was measured with super-conducting quantum interference device magnetometry (SQUID) and is consistent with a high-spin ground state (S = 3/2) (Figure 2a).
Scheme 1.
Ligand and Metal Complex Syntheses
Figure 1.
Solid-state structures of (a) (TrL)CoCl (1), (b) (TrL)Co (2), and (c, d) (TrL)Co(NAd) (3) at 100 K with thermal ellipsoids set at 50% probability (hydrogen atoms and solvent molecules are omitted for clarity; Co aquamarine, C gray, N blue, Cl green). Selected bond lengths (Å) and angles (deg) for 1: Co–Cortho, 2.345(7). 3: Co–Cortho, 2.756; Co–N, 1.636(3); ∠Co–N–C = 145.0(3).
Figure 2.
Variable-temperature susceptibility data of (a) (TrL)CoCl (1) and (b) (TrL)Co (2) collected at 1.0 T from 5 to 300 K. (inset) Reduced magnetization data collected at three fields (1, 4, 7 T) over the temperature range 1.8–10 K. Magnetization fit parameters obtained with PHI:60 for 1, S = 3/2, g = 2.4, D = −48.1 cm−1, |E/D| = 0.11 and, for 2, S = 1, g = 2.03, D = 25.98 cm−1, |E/D| = 0.085. (c) Variable-temperature susceptibility data of (TrL)Co(NAd) (3) collected at 0.5 and 1.0 T upon heating and cooling (5–300 K and 300–5 K).
Chemical reduction of 1 with KC8 in a thawing benzene solution cleanly furnished the reduced species (TrL)Co (2), which exhibits an intramolecular η6-arene interaction between the cobalt and one of the phenyl groups of the trityl substituents (Scheme 1b, Figure 1b). The solid-state magnetic susceptibility of 2 is consistent with an S = 1 high-spin ground state (Figure 2b). The cyclic voltammogram of 2 measured in a non-coordinating solvent 1,2-didfluorobenzene reveals two sequential one-electron oxidations (E1/2 = −0.80, 0.75 V vs [Cp2Fe]+/0) and a single one-electron reduction (−2.1 V, Figure 3a). The observed redox potentials are almost identical with the ones observed for its congener (ArL)Co, suggesting that 2 is a suitable alternative CoI synthon to further explore two-electron oxidative group transfer reactivity (Figure 3b).
Figure 3.
Cyclic voltammograms of (a) 2 and (b) (ArL)Co obtained in 1,2-difluorobenzene at room temperature using 0.1 M [nBu4N]-[PF6] electrolyte. Potentials were referenced to the [Cp2Fe]+/0 couple. Scan rate = 100 mV/s.
2.2. Synthesis and Characterization of (TrL)Co(NAd).
With the (TrL)CoI synthon in hand, we sought to investigate its reactivity toward organoazides. Treating a benzene-d6 solution of 2 with a stoichiometric amount of 1-azidoada-mantane (AdN3) immediately furnished a diamagnetic species as verified by 1H and 13C NMR spectroscopies. Single crystals of the product obtained from a 1:1 benzene/hexanes mixture stored at room temperature were analyzed by single crystal X-ray diffraction to reveal formation of a three-coordinate CoIII imido complex (TrL)Co(NAd) (3) (Scheme 2 and Figure 1c–d).
Scheme 2.
Reactivity Assessment of 2
Notably, 3 exhibits a unique distorted trigonal pyramidal geometry with a significantly bent imido moiety [∠Co–N–C = 145.0(3)°]. Akin to 1, a relatively long η1-arene interaction between the cobalt and one of the ortho carbons from the trityl groups is present (Co–Cortho = 2.756 Å) orienting the nitrenoid unit orthogonal to the dipyrrin plane. Furthermore, the sterically encumbered trityl substituents shield the trans position of the nitrenoid moiety, enforcing an acute ∠L–Co–NAd (119.7°) as well as blocking the fourth coordination site of 3 as further evidenced by the space-filling model (Figure S–28). Consequently, no bimolecular reactivity affording the four-coordinate tetrazido complex was detected upon heating a benzene-d6 solution of 3 with excess amount of AdN3 (Scheme 2).
We note that only a handful of structurally characterized metal imides display a substantially bent nitrenoid unit. For late, first-row transition metals, a four-coordinate bis(pyrrolyl)-pyridine FeIV imide61 and five-coordinate FeIV imide62 supported by a tetradentate N-heterocyclic (NHC) ligand display an Fe–N–C angle of 140.5(3)° and 150.07(10)°, respectively. Hillhouse and co-workers have shown the synthesis of a three-coordinate nickel imide featuring a bent nitrenoid moiety [∠Ni–N–C = 127.3(3)°] using a sterically encumbered aryl azide (e.g., 2,5-dimesitylphenylazide, 2,6-bis(2,6-diisopropylphenyl)phenylazide).45 Given that those bent M–N–C linkages were induced by either multidentate chelating ligands or steric properties of the R substituents on the metal-bound imido (NR) group, furnishing the bent imido complex 3 via direct manipulation of the primary coordination environment with the low-coordinate bidentate dipyrrin ligand is significant.
Having isolated the bent imido 3, we sought to understand its electronic structure. As discussed, the diamagnetic spin ground state of 3 is further supported by the temperature-independent susceptibility data (Figure 2c), which is consistent with the vast majority of CoIII imides including the analogous dipyrrin alkyl CoIII imide (ArL)Co(NR) reported by our group.37–43 However, unlike (ArL)Co(NR), no thermal population of spin excited states (e.g., S = 0 → 2) was observed (Figure 2c inset).37
To further elucidate the frontier molecular orbital interactions in 3, DFT calculations were conducted (SI–59).63,64 Interestingly, owing to the bent geometry, HOMO, HOMO − 1, and HOMO − 2 of 3 display almost non-bonding or significantly disrupted π* interactions with respect to the imido unit. Instead, σ* or π* contributions from the dipyrrin become dominant. The non-bonding interaction with and the slight π-overlap between the 2p orbitals of the Nimido and Co , 3dyz, and 3dxy are illustrated in Figure 4.
Figure 4.
Frontier molecular orbital description of 3.
2.3. Catalytic C–H Amination.
2.3.1. Steric Protection from (TrL) Leads to Higher Yields of C–H Amination.
Encouraged by the distorted MLMB of 3 and its steric shielding to prevent the catalyst degradation, we speculated that analogous (TrL)Co(NR) species featuring a linear alkyl moiety would greatly benefit the intramolecular C–H amination to afford N-hetetrocycles. To test this hypothesis, (4-azido-4-methylpentyl) benzene (5) was exposed to 10 mol % of 2 in benzene-d6. Gratifyingly, the desired product 2,2-dimethyl-5-phenylpyrrolidine (6) was quantitatively obtained at room temperature along with regenerated catalyst 2 (Figure S–2). We note that under the same conditions ([5] = 85.0 mM at 25 °C), 10 mol % of (ArL)Co yielded only 13% of 6 due to concomitant formation of a catalytically inactive tetrazido complex (ArL)Co(κ2-N4R2, R = C(CH3)2(CH2)3Ph),38 high-lighting that absence of the off-cycle pathway is critical to improve the yields of C–H amination. As a result, even with 1 mol % of 2 in benzene-d6,6 was obtained in 91% yield at 80°C.
The steric protection of the trityl substituents is further supported by the effect of pyridine on the rates of catalytic C–H amisnation. Given that the fourth-coordination site of 3 is blocked by the trityl substituents, we hypothesized that pyridine would behave as an inhibitor as opposed to an accelerant in contrast with the C–H amination mediated by (ArL)Co(NR). Indeed, unlike (ArL)Co(NR), which binds 1 equiv of pyridine in solution,38 3 showed no binding affinity toward pyridine. Moreover, at 50 °C, slower rates of C–H amination were observed when exposing 5 to 10 mol % of 2 at higher concentrations of pyridine-d5 (Figure S-7). A double-reciprocal plot suggests that pyridine-d5 and 5 compete for the same coordination site of 2,65 and thus, the desired C–H amination is hampered in the presence of pyridine-d5 (Figures 5a and S-9). Job plot analysis further supports that 2 readily binds pyridine in benzene-d6 to furnish the (TrL)Co(py) (4) adduct (Scheme 2 and Figure 5b).66 Furthermore, single-crystal X-ray diffraction analysis of 4 verifies that pyridine occupies the same axial position as the imido moiety of 3 (Figure 5c). The solid-state structure of 4 displays a η2-arene interaction with one of the ortho and meta carbons from the trityl groups [Co−Cortho = 2.0520(8) Å and Co−Cmeta = 2.239(8) Å]. These results are in line with the aforementioned steric environment of 3 suggesting that the fourth coordination site of (TrL)Co(NR) is no longer available.
Figure 5.
(a) Double-reciprocal plot with 10 mol % of 2 at varying concentrations of pyridine-d5 at 50 °C. (b) Job plot analysis between 2 and pyridine in benzene-d6 at 25 °C. (c) Solid-state structures of (TrL)Co(py) (4) at 100 K with thermal ellipsoids set at 50% probability (hydrogen atoms and solvent molecules are omitted for clarity; Co aquamarine, C gray, N blue). Selected bond lengths (Å): Co−Cortho, 2.050(8); Co−Cmeta, 2.239(8).
Thus, by simply switching the flanking groups of the CoI catalyst from the 2,4,6-triphenylphenyl to trityl, pyridine becomes a competitive inhibitor as opposed to a cocatalyst for the desired C–H amination. These results demonstrate that the pyridine ligation effect precluding tetrazene formation, which is observed for the previously studied amination catalysis with the (ArL) scaffold is successfully incorporated within the (TrL) ligand. Furthermore, considering that many other attempted N−group transfer reactions utilizing organoazides as a nitrene source are hampered by the reductive coupling of organoazides furnishing tetrazene37,38,67 or hexazene68 complexes, we believe that the study described herein provides a ligand design to circumvent loss of reactive metal nitrenoid intermediates.
2.3.2. Mechanism of the C–H Amination.
Next, we sought to investigate the mechanism of C–H amination mediated by (TrL)Co(NR) and explore how the unusual geometry of the cobalt imide affects its C–H amination reactivity as compared to that of (ArL)Co(NR).
To this end, the reaction progression was monitored by 1H NMR spectroscopy (Figure 6a). In the presence of a 10-fold excess amount of azide substrate 5, (TrL)Co (2) immediately converted to the corresponding CoIII imide (7) at 25 °C as verified by the diagnostic 1H NMR chemical shifts (Figure S-2). At early stages of the reaction, imido 7 is the resting species as the concentration of azide 5 is sufficient to continuously oxidize 2 as soon as 7 generates pyrrolidine 6 (Figure 6b inset). Indeed, the rate of the azide consumption (−d[5]/dt) is almost identical to the rate of product formation (d[6]/dt) at initial stages of the catalytic reaction (SI-17). Thus, while the concentration of 7 is constant at early stages of the reaction (6 < 10%), the rate of formation of 6 is zeroth order in [5] and first order in [7], corroborating that 7 is competent for C–H amination in a unimolecular fashion (Figure 6b–c).
Figure 6.
(a) Full reaction progression with 10 mol % of 2 at 25 °C in benzene-d6. (b) Zeroth-order dependence on the concentration of 5 at 25 °C in benzene-d6. (inset) Initial stages of reaction showing the concentration of 7 is constant. A dashed line indicates the region less than 10% of 6. (c) First-order dependence on the concentration of 7 at 25 °C benzene-d6.
Furthermore, a kinetic isotope effect (KIE) value was evaluated by comparing the initial rates of 10 mol % of 2 with 5 and (4-azido-4-methylpentyl-1,1-d2) benzene (5-d2). The observed ratio of the rate constants (kH/kD) was determined to be 38.4(1) at 25 °C (Figure 7), consistent with a rate-determining homolytic C–H bond cleavage.69 Given all the kinetic data, we propose the following CoI/III catalytic cycle (Scheme 3): fast oxidation of 2 by an alkyl azide affords a CoIII imide which undergoes intramolecular HAA; subsequent radical recombination of the resulting CoII amide affords the cyclized products. However, we note that a direct C–H insertion pathway from the Co imido complex cannot be fully excluded.
Figure 7.
Kinetic isotope effect experiments with 10 mol % of 2 at 25 °C in benzene-d6.
Scheme 3.
Proposed Catalytic Cycle
2.3.3. Enhancement in Rates of C–H Amination.
Surprisingly, the rate of C–H amination mediated by (TrL)Co(NR) is significantly faster compared to (ArL)Co(NR). Specifically, the rate constant (k) for the benzylic C–H amination mediated by 7 to afford 6 is 0.063 min−1 at 25 °C, whereas the same magnitude of k = 0.062 min−1 was observed for the analogous (ArL)Co(NR) at 50 °C.38 We propose the faster conversion is a result of the bent imido moiety of 7 (as observed for 3) that likely resembles the transition state geometry for HAA more closely than the linear (ArL)Co(NR). Therefore, the enhanced C–H amination rate of 7 supports our original hypothesis, where a bent nitrenoid unit may enhance the rate of C–H amination as less structural rearrangement is required for HAA. Overall, these results highlight that a bent MLMB linkage, which is enforced by the (TrL) scaffold, is critical to enhance the rate of C–H amination, thus kinetically improving the overall C–H amination process.
2.3.4. Scope of C–H Amination.
In addition to the enhanced rates, (TrL)Co(NR) is capable of functionalizing a range of C–H bonds, even primary C–H bonds in a catalytic fashion (Table 1). Previously, we were only able to functionalize benzylic C–H bonds from the analogous three-coordinate metal imides M(NR2−): (ArL)Fe(NAd)33 and (ArL)Co(NR).38 Thus, the ability of (TrL)Co(NR) to functionalize C–H bonds stronger than a benzylic C–H bond catalytically is noteworthy. As such, we sought to address the elements rendering (TrL)Co(NR) a more potent oxidant relative to the analogous (ArL)Co(NR). Given that both imides (TrL)Co(NR) and (ArL)Co(NR) accomplish intramolecular C–H amination via an initial rate-limiting HAA step, as suggested by the measured KIE values,38 we attempted to evaluate thermodynamic parameters such as ( and pKa as described in eq 1, where the constant C represents the standard potential for proton reduction in aprotic solvent.49 We hypothesized that the difference in driving force for HAA from the respective imides is proportional to the difference in bond dissociation energies (BDEs) of the N−H bonds of the corresponding CoII amide complexes (Scheme 4).
Table 1.
Catalytic C–H Amination Using 2 and Alkyl Azidesa
![]() |
Isolated yields from reactions conducted with 85.0 mM of azide substrate.
1 mol % of 2.
5 mol % of 2.
10 mol % of 2.
Scheme 4.
Thermodynamic Square Scheme for HAA from CoIII Imides 3 and 8
(1) |
To this end, we used 3 and its isolable congener (ArL)Co(NAd) (8) as model complexes to probe their one-electron reduction potentials ( values for CoIII/II couples) via electrochemical analysis. Cyclic voltammetry experiments of 3 and 8 were conducted using 0.1 M [nBu4N][PF6] electrolyte in the non-coordinating solvent 1,2-difluorobenzene to obviate exogenous ligand coordination that may affect the . Interestingly, both 3 and 8 display reversible one electron reductions at E1/2 of −1.98 and −1.93 V (vs [Cp2Fe]+/0), respectively (Figure 8).
Figure 8.
Cyclic voltammograms of 3 and 8 obtained in 1,2-difluorobenzene at room temperature using 0.1 M [nBu4N][PF6] electrolyte. Potentials were referenced to the [Cp2Fe]+/0 couple. Scan rate = 100 mV/s.
The observed low reduction potentials suggest that the driving force for HAA can be ascribed to the high basicity of the one-electron reduced anionic imides [LCo(NR)]−, which are equivalent to the conjugate base of amides [LCo(NHR)] 3a and 8a, respectively (Scheme 4).55 Furthermore, given the almost identical values of 3 and 8, we believe that the greater oxidizing power of (TrL)Co(NR) compared to (ArL)Co(NR) arises from a more basic imido moiety of its one-electron reduced form [(TrL)Co(NR)]−.
Unfortunately, our attempts to directly probe the pKa of the respective amide complexes (TrL)Co(NHAd) (3a) and (ArL)Co(NHAd) (8a) were unsuccessful due to their high propensity for subsequent HAA resulting in formation of 1-adamantylamine.52 Instead, the difference in pKa between 3a and 8a was estimated by accounting for the difference in BDEs of the C–H bonds (6 kcal/mol) that can be activated by the respective imides intramolecularly at 25 °C. Specifically at 25°C, (TrL)Co(NR) accomplishes amination of secondary C–H bonds (96 kcal/mol, R = C(CH3)2(CH2)3CH3) affording 2,2,5-trimethylpyrrolidine, whereas (ArL)Co(NR) can only functionalize benzylic C–H bonds (90 kcal/mol, R = C(CH3)2(CH2)3Ph) to afford 6. Accordingly, the conjugate base of 3a [(TrL)Co(NR)]− is estimated to be about 105 times more basic than the conjugate base of 8a [(ArL)Co(NR)]−. The observed difference in pKa is further supported by DFT calculations,52 yet evaluation of the absolute pKa of each amide complex needs to be further examined (SI-41).
This observed trend in basicity is in agreement with literature reports proposing that a bent imido moiety is more basic compared to a linear one.57 Thus, these results demonstrate that altering the local geometry of the terminally bound metal imido unit subsequently affects its basicity as well and, thereby, enhances its oxidizing power and broadens its C–H amination reaction scope.
3. CONCLUSIONS
The foregoing results illustrate how rational ligand design allowed the synthesis of a robust CoI catalyst and isolation of a bent CoIII imide that promotes facile catatlytic C–H amination. The previously studied pyridine-accelerated C–H amination process mediated by (ArL)Co(NR) informed the design principles for the synthesis of a dipyrromethene ligand (TrL) featuring sterically encumbered trityl substituents. The subtle modification of the supporting ligand drastically altered the primary coordination environment of a CoIII imide to force the terminal imido unit perpendicular to the dipyrrin ligand plane, imposing a bent geometry. Kinetic and thermodynamic analyses of the C–H amination mediated by either (TrL)Co(NR) or (ArL)Co(NR) highlight that the unique structure of (TrL)Co(NR) is key for the significantly enhanced C–H amination efficacy as manifested in yields, rates, and oxidizing power. In particular, the sterically demanding trityl substituents prevent catalyst degradation (via tetrazene formation), contributing to higher yields of C–H amination. More importantly, the atypical trigonal pyramidal geometry of the CoIII imide with a significantly bent imide resembles the transition state for HAA more closely, and thereby, (TrL)Co(NR) displays enhanced rates of C–H amination compared to the trigonal planar (ArL)Co(NR) possessing a linear imido geometry. Lastly, the more bent, and thus more basic, one-electron reduced species of (TrL)Co(NR) provides a greater thermodynamic driving force for HAA enabling C–H amination with a broader substrate scope at milder conditions. We believe that this study provides an alternative synthetic strategy to generate highly reactive metal imido complexes via direct manipulation of the geometry of MLMBs to accomplish more applicable catalytic transformations.
4. EXPERIMENTAL SECTION
General experimental considerations, synthesis and spectroscopic characterizations, X-ray diffraction techniques, kinetic analysis, electrochemistry measurements, and DFT results are all provided in the Supporting Information.
4.1. Preparation of (TrL)CoCl (1).
In an oven-dried 20 mL vial, anhydrous CoCl2 (0.105 g, 0.81 mmol, 1.2 equiv) was dissolved in THF and frozen in a liquid nitrogen cooled cold-well. To a frozen solution, a solution of (TrL)Li (0.52 g, 0.68 mmol, 1 equiv) in THF was added and stirred for 20 h at room temperature. The resulting dark red reaction mixture was filtered through a coarse glass frit with Celite to remove LiCl and the excess CoCl2. The Celite was washed with an additional 20 mL benzene. The collected filtrate was frozen and benzene was removed by sublimation in vacuo to afford 1 as a dark red powder (0.52 g, 90%). Crystals suitable for X-ray diffraction were grown from a 2:1 n-hexane/benzene layered solution at room temperature. 1H NMR (600 MHz, 295 K, C6D6): δ/ppm 68.14, 23.15, 20.56, 4.70, 0.64, −1.16, −10.13, −37.86. Anal. calc. for C56H45ClCoN2: C 80.04, H 5.40, N 3.33; Found: C 80.00, H 5.29, N 3.45.
4.2. Preparation of (TrL)Co (2).
In an oven-dried 20 mL vial, 1 (0.4 g, 0.476 mmol, 1 equiv) was dissolved in 10 mL of benzene and added to a frozen solution of potassium graphite (KC8, 0.071 g, 0.524 mmol, 1.1 equiv). The reaction mixture was slowly warmed up to room temperature and stirred for 3 h at room temperature. The dark red-orange mixture was filtered through a coarse glass frit with Celite to remove graphite. The Celite was washed with an additional 20 mL benzene. The collected filtrate was frozen and benzene was removed by sublimation in vacuo to afford 2 as a dark orange powder. (0.3 g, 78%). Crystals suitable for X-ray diffraction were grown from a 1:1 n-hexane/benzene mixture at room temperature. 1H NMR (600 MHz, 295 K, C6D6): δ/ppm 49.82, 21.04, 14.90, 5.84, 5.16, 4.54, 0.96, −2.51. Anal. calc. for C56H45CoN2: C 83.56, H 5.64, N 3.48; Found: C 83.45, H 5.50, N 3.39.
4.3. Preparation of (TrL)Co(NAd) (3).
In an oven-dried 20 mL vial, a solution of 1-azidoadamantane (22 mg, 0.12 mmol, 1.0 equiv) in benzene was added to a stirring solution of 2 (100 mg, 0.12 mmol, 1.0 equiv) in benzene to observe an immediate color change from dark red to dark purple. After stirring for 30 min at room temperature, the reaction mixture was frozen, and benzene was removed by sublimation in vacuo to afford 3 as a purple powder (109 mg, 92%). Crystals suitable for X-ray diffraction were grown from a 1:1 n-hexane/benzene mixture at room temperature. 1H NMR (600 MHz, 295 K, C6D6): δ/ppm 7.94 (d, 12 H), 7.23 (d, 2 H), 6.97 (t, 6 H), 6.84 (s, 2 H), 6.67 (t, 11 H), 6.03 (d, 2 H), 2.26 (s, 5 H), 2.20 (s, 3 H), 1.80 (s, 3 H), 1.63 (t, 6 H), 1.08 (s, 6 H). 13C{1H} NMR (125MHz, C6D6): δ/ppm 192.23, 153.36, 151.26, 146.88, 137.28, 135.37, 134.93, 132.14, 129.57, 128.59, 125.51, 124.41, 69.27, 61.65, 61.12, 37.54, 33.30, 31.03, 21.25, 20.69. Anal. calc. for C66H60CoN3: C 83.08, H 6.34, N 4.40; Found: C 82.72, H 6.52, N 4.71.
Supplementary Material
ACKNOWLEDGMENTS
T.A.B. acknowledges this work was supported by a grant from the NIH (GM-115815), the Dreyfus Foundation (Teacher-Scholar Award), and Harvard University. Y.B. is thankful for a Kwanjeong Graduate Student Fellowship.
Footnotes
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.9b09015.
General experimental considerations and procedures, multinuclear NMR data, electrochemistry data, kinetic analysis, magnetometry data, solid state molecular structures, computational details, and crystallography data (PDF)
Crystallographic CIF files for complexes 1–4 (CIF)
The authors declare no competing financial interest.
REFERENCES
- (1).Labinger JA; Bercaw JE Understanding and exploiting C–H bond activation. Nature 2002, 417, 507. [DOI] [PubMed] [Google Scholar]
- (2).Bergman RG C–H activation. Nature 2007, 446, 391. [DOI] [PubMed] [Google Scholar]
- (3).Davies HML; Manning JR Catalytic C–H functionalization by metal carbenoid and nitrenoid insertion. Nature 2008, 451, 417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Saouma CT; Peters JC M≡E and M=E complexes of iron and cobalt that emphasize three-fold symmetry (E = O, N, NR). Coord. Chem. Rev 2011, 255 (7), 920–937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Nugent WA; Mayer JM Metal-ligand multiple bonds: the chemistry of transition metal complexes containing oxo, nitrido, imido, or alkylidyne ligands; Wiley-Interscience: New York, 1988. [Google Scholar]
- (6).Ortiz de Montellano PR Cytochrome P450: structure, mechanism, and biochemistry, 4th ed.; Kluwer Academic/Plenum Publishers: New York, 2005. [Google Scholar]
- (7).Hill EA; Weitz AC; Onderko E; Romero-Rivera A; Guo Y; Swart M; Bominaar EL; Green MT; Hendrich MP; Lacy DC; Borovik AS Reactivity of an FeIV-oxo complex with protons and oxidants. J. Am. Chem. Soc 2016, 138 (40), 13143–13146. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Green MT; Dawson JH; Gray HB Oxoiron(IV) in chloroperoxidase compound II is basic: implications for P450 chemistry. Science 2004, 304 (5677), 1653. [DOI] [PubMed] [Google Scholar]
- (9).Zhang L; Deng L. C–H bond amination by iron-imido/nitrene species. Chin. Sci. Bull 2012, 57 (19), 2352–2360. [Google Scholar]
- (10).Waterman R; Hillhouse GL η2-Organoazide complexes of nickel and their conversion to terminal imido complexes via dinitrogen extrusion. J. Am. Chem. Soc 2008, 130 (38), 12628–12629. [DOI] [PubMed] [Google Scholar]
- (11).Mindiola DJ; Hillhouse GL Terminal amido and imido complexes of three-coordinate nickel. J. Am. Chem. Soc 2001, 123 (19), 4623–4624. [DOI] [PubMed] [Google Scholar]
- (12).Olivos Suarez AI; 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 (21), 5697–5705. [DOI] [PubMed] [Google Scholar]
- (13).Kuijpers PF; van der Vlugt JI; Schneider S; de Bruin B. Nitrene radical intermediates in catalytic synthesis. Chem. - Eur. J 2017, 23 (56), 13819–13829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Lyaskovskyy V; Suarez AIO; Lu H; Jiang H; Zhang XP; de Bruin B. Mechanism of cobalt(II) porphyrin-catalyzed C–H amination with organic azides: radical nature and H−atom abstraction ability of the key cobalt(III)−nitrene intermediates. J. Am. Chem. Soc 2011, 133 (31), 12264–12273. [DOI] [PubMed] [Google Scholar]
- (15).Suarez AIO; Lyaskovskyy V; Reek JNH; van der Vlugt JI; de Bruin B. Complexes with nitrogen-centered radical ligands: classification, spectroscopic features, reactivity, and catalytic applications. Angew. Chem., Int. Ed 2013, 52 (48), 12510–12529. [DOI] [PubMed] [Google Scholar]
- (16).Kuijpers PF; Tiekink MJ; Breukelaar WB; Broere DLJ; van Leest NP; van der Vlugt JI; Reek JNH; de Bruin B. Cobalt-porphyrin-catalysed intramolecular ring-closing C–H amination of aliphatic azides: a nitrene-radical approach to saturated heterocycles. Chem. - Eur. J 2017, 23 (33), 7945–7952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).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 (16), 5468–5479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).King ER; Hennessy ET; Betley TA Catalytic C–H bond amination from high-spin iron imido complexes. J. Am. Chem. Soc 2011, 133 (13), 4917–4923. [DOI] [PubMed] [Google Scholar]
- (19).Wilding MJT; Iovan DA; Wrobel AT; Lukens JT; MacMillan SN; Lancaster KM; Betley TA Direct comparison of C–H bond amination efficacy through manipulation of nitrogen-valence centered redox: imido versus iminyl. J. Am. Chem. Soc 2017, 139 (41), 14757–14766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Conradie J; Ghosh A. Electronic structure of an iron-porphyrin−nitrene complex. Inorg. Chem 2010, 49 (1), 243–248. [DOI] [PubMed] [Google Scholar]
- (21).Corona T; Ribas L; Rovira M; Farquhar ER; Ribas X; Ray K; Company A. Characterization and reactivity studies of a terminal copper−nitrene species. Angew. Chem., Int. Ed 2016, 55 (45), 14005–14008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Berry JF Terminal nitrido and imido complexes for the late transition metals. Comments Inorg. Chem 2009, 30 (1−2), 28–66. [Google Scholar]
- (23).Carsch KM; DiMucci IM; Iovan DA; Li A; Zheng S-L; Titus CJ; Lee SJ; Irwin KD; Nordlund D; Lancaster KM; Betley TA Synthesis of a copper-supported triplet nitrene complex pertinent to copper-catalyzed amination. Science 2019, 365 (6458), 1138–1143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Zdilla MJ; Dexheimer JL; Abu-Omar MM Hydrogen atom transfer reactions of imido manganese(V) corrole: one reaction with two mechanistic pathways. J. Am. Chem. Soc 2007, 129 (37), 11505–11511. [DOI] [PubMed] [Google Scholar]
- (25).Brown SD; Betley TA; Peters JC A Low-spin d5 iron imide: nitrene capture by low-coordinate iron(I) provides the 4-coordinate Fe(III) complex [PhB(CH2PPh2)3]Fe≡N-p-tolyl. J. Am. Chem. Soc 2003, 125 (2), 322–323. [DOI] [PubMed] [Google Scholar]
- (26).Brown SD; Peters JC Ground-state singlet L3Fe-(μ-N)-FeL3 and L3Fe(NR) complexes featuring pseudotetrahedral Fe(II) centers. J. Am. Chem. Soc 2005, 127 (6), 1913–1923. [DOI] [PubMed] [Google Scholar]
- (27).Betley TA; Peters JC Dinitrogen chemistry from trigonally coordinated ion and cobalt platforms. J. Am. Chem. Soc 2003, 125 (36), 10782–10783. [DOI] [PubMed] [Google Scholar]
- (28).Thomas CM; Mankad NP; Peters JC Characterization of the terminal iron(IV) imides {[PhBPtBu2(pz’)]FeIV⋮NAd}+. J. Am. Chem. Soc 2006, 128 (15), 4956–4957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Bart SC; Lobkovsky E; Bill E; Chirik PJ Synthesis and hydrogenation of bis(imino)pyridine iron imides. J. Am. Chem. Soc 2006, 128 (16), 5302–5303. [DOI] [PubMed] [Google Scholar]
- (30).Bowman AC; Milsmann C; Bill E; Turner ZR; Lobkovsky E; DeBeer S; Wieghardt K; Chirik PJ Synthesis and electronic structure determination of N-alkyl-substituted bis(imino)pyridine iron imides exhibiting spin crossover behavior. J. Am. Chem. Soc 2011, 133 (43), 17353–17369. [DOI] [PubMed] [Google Scholar]
- (31).Cowley RE; DeYonker NJ; Eckert NA; Cundari TR; DeBeer S; Bill E; Ottenwaelder X; Flaschenriem C; Holland PL Three-coordinate terminal imidoiron(III) complexes: structure, spectroscopy and mechanism of formation. Inorg. Chem 2010, 49 (13), 6172–6187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Wang L; Hu L; Zhang H; Chen H; Deng L. Three-coordinate iron(IV) bisimido complexes with aminocarbene ligation: synthesis, structure, and reactivity. J. Am. Chem. Soc 2015, 137 (44), 14196–14207. [DOI] [PubMed] [Google Scholar]
- (33).Wilding MJT; Iovan DA; Betley TA High-spin iron imido complexes competent for C–H bond amination. J. Am. Chem. Soc 2017, 139 (34), 12043–12049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Iovan DA; Betley TA Characterization of iron-imido species relevant for N-group transfer chemistry. J. Am. Chem. Soc 2016, 138 (6), 1983–1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).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 (43), 12640–12644. [DOI] [PubMed] [Google Scholar]
- (36).Liu Y; Du J; Deng L. Synthesis, structure, and reactivity of low-spin cobalt(II) imido complexes [(Me3P)3Co(NAr)]. Inorg. Chem 2017, 56 (14), 8278–8286. [DOI] [PubMed] [Google Scholar]
- (37).King ER; Sazama GT; Betley TA Co(III) imidos exhibiting spin crossover and C–H bond activation. J. Am. Chem. Soc 2012, 134 (43), 17858–17861. [DOI] [PubMed] [Google Scholar]
- (38).Baek Y; Betley TA Catalytic C–H amination mediated by dipyrrin cobalt imidos. J. Am. Chem. Soc 2019, 141 (19), 7797–7806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).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 (10), 1508–1510. [DOI] [PubMed] [Google Scholar]
- (40).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 (50), 16322–16323. [DOI] [PubMed] [Google Scholar]
- (41).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 (40), 7406–7410. [DOI] [PubMed] [Google Scholar]
- (42).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 (15), 4798–4799. [DOI] [PubMed] [Google Scholar]
- (43).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 (38), 11238–11239. [DOI] [PubMed] [Google Scholar]
- (44).Laskowski CA; Miller AJM; Hillhouse GL; Cundari TR A Two-coordinate nickel imido complex that effects C–H amination. J. Am. Chem. Soc 2011, 133 (4), 771–773. [DOI] [PubMed] [Google Scholar]
- (45).Harrold ND; Hillhouse GL Strongly bent nickel imides supported by a chelating bis(N-heterocyclic carbene) ligand. Chem. Sci 2013, 4 (10), 4011–4015. [Google Scholar]
- (46).Aguila MJB; Badiei YM; Warren TH Mechanistic insights into C–H amination via dicopper nitrenes. J. Am. Chem. Soc 2013, 135 (25), 9399–9406. [DOI] [PubMed] [Google Scholar]
- (47).Bakhoda A; Jiang Q; Bertke JA; Cundari TR; Warren TH Elusive terminal copper arylnitrene intermediates. Angew. Chem., Int. Ed 2017, 56 (23), 6426–6430. [DOI] [PubMed] [Google Scholar]
- (48).Stone KL; Borovik AS Lessons from nature: unraveling biological C–H bond activation. Curr. Opin. Chem. Biol 2009, 13 (1), 114–118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Warren JJ; Tronic TA; Mayer JM Thermochemistry of proton-coupled electron transfer reagents and its implications. Chem. Rev 2010, 110 (12), 6961–7001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Bordwell FG; Cheng JP; Harrelson JA Homolytic bond dissociation energies in solution from equilibrium acidity and electrochemical data. J. Am. Chem. Soc 1988, 110 (4), 1229–1231. [Google Scholar]
- (51).Dhar D; Tolman WB Hydrogen atom abstraction from hydrocarbons by a copper(III)-hydroxide complex. J. Am. Chem. Soc 2015, 137 (3), 1322–1329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Nieto I; Ding F; Bontchev RP; Wang H; Smith JM Thermodynamics of hydrogen atom transfer to a high-valent iron imido complex. J. Am. Chem. Soc 2008, 130 (9), 2716–2717. [DOI] [PubMed] [Google Scholar]
- (53).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 (9), 2424–2425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Scepaniak JJ; Young JA; Bontchev RP; Smith JM Formation of ammonia from an iron nitrido complex. Angew. Chem., Int. Ed 2009, 48 (17), 3158–3160. [DOI] [PubMed] [Google Scholar]
- (55).Parsell TH; Yang M-Y; Borovik AS C–H bond cleavage with reductants: re-investigating the reactivity of monomeric MnIII/IV−oxo complexes and the role of oxo ligand basicity. J. Am. Chem. Soc 2009, 131 (8), 2762–2763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).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 (44), 15525–15528. [DOI] [PubMed] [Google Scholar]
- (57).Cowley RE; Eckert NA; Vaddadi S; Figg TM; Cundari TR; Holland PL Selectivity and mechanism of hydrogen atom transfer by an isolable imidoiron(III) complex. J. Am. Chem. Soc 2011, 133 (25), 9796–9811. [DOI] [PubMed] [Google Scholar]
- (58).Cowley RE; Holland PL Ligand effects on hydrogen atom transfer from hydrocarbons to three-coordinate iron imides. Inorg. Chem 2012, 51 (15), 8352–8361. [DOI] [PubMed] [Google Scholar]
- (59).Eckert NA; Vaddadi S; Stoian S; Lachicotte RJ; Cundari TR; Holland PL Coordination-number dependence of reactivity in an imidoiron(III) complex. Angew. Chem., Int. Ed 2006, 45 (41), 6868–6871. [DOI] [PubMed] [Google Scholar]
- (60).Chilton NF; Anderson RP; Turner LD; Soncini A; Murray KS PHI: A powerful new program for the analysis of anisotropic monomeric and exchange-coupled polynuclear d- and f- block complexes. J. Comput. Chem 2013, 34 (13), 1164–1175. [DOI] [PubMed] [Google Scholar]
- (61).Searles K; Fortier S; Khusniyarov MM; Carroll PJ; Sutter J; Meyer K; Mindiola DJ; Caulton KG A cis-divacant octahedral and mononuclear iron(IV) imide. Angew. Chem., Int. Ed 2014, 53 (51), 14139–14143. [DOI] [PubMed] [Google Scholar]
- (62).Anneser MR; Elpitiya GR; Townsend J; Johnson EJ; Powers XB; DeJesus JF; Vogiatzis KD; Jenkins DM Unprecedented five-coordinate iron(IV) imides generate divergent spin states based on the imide R-groups. Angew. Chem., Int. Ed 2019, 58 (24), 8115–8118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Perdew JP Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B: Condens. Matter Mater. Phys 1986, 33 (12), 8822–8824. [DOI] [PubMed] [Google Scholar]
- (64).Becke AD Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A: At., Mol., Opt. Phys 1988, 38 (6), 3098–3100. [DOI] [PubMed] [Google Scholar]
- (65).Lineweaver H; Burk D. The determination of enzyme dissociation constants. J. Am. Chem. Soc 1934, 56 (3), 658–666. [Google Scholar]
- (66).Renny JS; Tomasevich LL; Tallmadge EH; Collum DB Method of continuous variations: applications of Job plots to the study of molecular associations in organometallic chemistry. Angew. Chem., Int. Ed 2013, 52 (46), 11998–12013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (67).Bowman AC; Tondreau AM; Lobkovsky E; Margulieux GW; Chirik PJ Synthesis and electronic structure diversity of pyridine(diimine)iron tetrazene complexes. Inorg. Chem 2018, 57 (16), 9634–9643. [DOI] [PubMed] [Google Scholar]
- (68).Cowley RE; Elhaïk J; Eckert NA; Brennessel WW; Bill E; Holland PL A bridging hexazene (RNNNNNNR) ligand from reductive coupling of azides. J. Am. Chem. Soc 2008, 130 (19), 6074–6075. [DOI] [PubMed] [Google Scholar]
- (69).Hammes-Schiffer S; Stuchebrukhov AA Theory of coupled electron and proton transfer reactions. Chem. Rev 2010, 110 (12), 6939–6960. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.