Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Mar 2.
Published in final edited form as: Angew Chem Int Ed Engl. 2020 Jun 9;59(35):15215–15219. doi: 10.1002/anie.202006391

N–H Bond Formation at a Diiron Bridging Nitride

Shaoguang Zhang [a], Peng Cui [a], Tianchang Liu [a], Qiuran Wang [a], Thomas J Longo [b], Laura M Thierer [a], Brian C Manor [a], Michael R Gau [a], Patrick J Carroll [a], Georgia C Papaefthymiou [b], Neil C Tomson [a]
PMCID: PMC7680347  NIHMSID: NIHMS1611925  PMID: 32441448

Abstract

Despite their connection to ammonia synthesis, little is known about the ability of iron-bound, bridging nitrides to form N–H bonds. Herein we report a linear diiron bridging nitride complex supported by a redox-active macrocycle. The unique ability of the ligand scaffold to adapt to the geometric preference of the bridging species was found to facilitate the formation of N–H bonds via proton-coupled electron transfer to generate a μ-amide product. The structurally analogous μ-silyl- and μ-borylamide complexes were shown to form from the net insertion of the nitride into the E–H bonds (E = B, Si). Protonation of the parent bridging amide produced ammonia in high yield, and treatment of the nitride with PhSH was found to liberate NH3 in high yield through a reaction that engages the redox-activity of the ligand during PCET.

Keywords: bimetallic complexes, iron nitrides, N-H bond formation, redox-active ligands, PCET


Ammonia (NH3) is a promising energy carrier for fuel cell, hydrogen storage, and alternative fuel applications. It is produced industrially by the Haber-Bosch process, which uses a catalyst that contains metallic iron and various main group promoters. Chemisorption of N2 at the catalyst surface is understood to form bridging nitrogen adatoms on the metallic surface (typically referred to as surface nitrides),[1] which then combine with surface hydrogen adatoms (surface hydrides) through successive N–H coupling events to generate ammonia.[2] Given the high covalency that has been proposed in surface and bulk iron nitride systems,[3] iron-bound, electrophilic μ-N species are of considerable interest.

Most examples of diiron bridging nitrides exhibit both high formal oxidation states (Fe≥4+) and saturated coordination spheres in the vicinity of the nitride.[4,5] And even though nitrides typically exhibit electrophilic reactivity when bound to oxidizing metal centers,[6] no examples have been provided in the literature of these species undergoing μ-N-based chemistry. Instead, the diiron bridging nitrides that undergo N–H bond formation make use of lower-valent iron centers (Fe≤3+); two such systems are known. Brown and Peters reported the ability of Fe2+2 and Fe2+Fe3+ bridging nitrides supported by tris(phosphino)borate ligands (PhBP3) to cleave H2, thereby yielding products that contain both bridging imide (μ-NH) and bridging hydride (μ-H) moieties (Scheme 1, top).[7] More recently, Holland and co-workers reported that an N2-derived μ-N between two Fe3+ ions acts as a strong base (Scheme 1, middle);[8] related μ-imide products, formed via intermolecular protonation, generate NH3 following net hydrogen atom transfer and protonolysis.[9]

Scheme 1.

Scheme 1.

Structures and reactivity profiles for isolated diiron and dicobalt bridging nitride complexes that show reactivity at μ-N.

We recently described the putative formation of a pair of transient bridging nitrides between two cobalt ions.[10] The isolation of products that result from intramolecular hydrogen atom abstraction, P=N bond formation, and N-atom insertion into an aliphatic C–H bond provided compelling evidence for the formation of highly electrophilic Co2(μ-N) cores. An example of this moiety was recently isolated by Fortier and co-workers and shown to undergo intermolecular activation of H2 (Scheme 1, bottom).[11] The results presented below describe an isolable diiron bridging nitride that is isoelectronic to our previously reported Co2(μ-N) species. Both the formation and the reactivity of the diiron nitride appear to rely on the unique ability of the supporting ligand to accommodate geometric changes of the bridging nitrogen.[12] This species undergoes intermolecular reactivity toward formation of N–H bonds via proton-coupled electron transfer (PCET), E–H bond insertion chemistry (E = Si, B), and protonation.

Treatment of the diamagnetic, bridging chloride complex [(3PDI2)Fe2(μ-Cl)(PMe3)2][OTf] ([Fe2Cl]+) with NaN3 in THF resulted in a color change from brown to purple at room temperature, corresponding to the formation of the diiron bridging nitride [(3PDI2)Fe2(μ-N)(PMe3)2][OTf] ([Fe2N]+, Scheme 2). A crystal structure of the product revealed conversion from the folded ligand architecture[13] of [Fe2Cl]+ to an unfolded ligand that supports a linear coordination environment about the nitride (∠ Fe–N–Fe= 177.1(2)°, Figure 1). The Fe–Nμ distances of 1.677(3) and 1.673(3) Å are indicative of multiple bond character,[7b, 14] and the Fe–NPDI distances (avg. Fe–Nim = 1.93 Å; avg. Fe–Npy = 1.86 Å) are consistent with other low- or intermediate-spin iron centers bound by PDI scaffolds.[15] The constraints of the Fe2N core induce a modest twist in the 3PDI2 macrocycle (∠py-py = 23.1°), which causes the two iron centers to be crystallographically distinct from one another (see Supporting Information). In order to evaluate the amount of electron density on the ligand, we make use of the empirically derived Δ parameter. Δ is defined as the difference between the average Cim–Cpy distance and the average of the Cim–Nim and Cpy–Npy distances on a pyridinediimine ligand,[16] and this parameter was recently adapted for use with 3PDI2.[13] Values near 0.105 Å represent (3PDI2)2−, and Δ decreases as the ligand is reduced. The value of 0.073 Å for [Fe2N]+ indicates the presence of three electron’s worth of electron density on the ligand, as (3PDI2)4− is not reached until Δ ≅ 0.055 Å.

Scheme 2.

Scheme 2.

Synthesis of a diiron bridging nitride complex and its N–H bond formation chemistry.

Figure 1.

Figure 1.

Depictions of cationic portions of the molecular structures of [Fe2N]+ (top left), [Fe2NH2]+ (top right), [Fe2NHSi]+ (bottom left) and [Fe2NHB]+ (bottom right); all counterions and hydrogen atoms were removed for clarity, except for the hydrogens on N and Si.

The [Fe2N]+ product is diamagnetic in solution. Its 1H NMR spectrum reveals overall C2h symmetry, and one singlet was observed in its 31P{1H} NMR spectrum at 4.11 ppm. Together, these data indicate that the phosphine ligands remain bound on the NMR timescale but that the inequivalence of the two iron centers observed crystallographically is not observable in solution. The use of 15N-labelled NaN3 (each N3 contains one terminal 15N atom) allows for isotopic enrichment at μ-N ([Fe2N*]+; ca. 50% 15N). A signal at 815 ppm was observed in the 15N NMR spectrum of this material, comparable to that reported by Peters for {[(PhBP3)Fe]2(μ-N)} (15N NMR δ = 801 ppm).[7b] These values are considerably upfield of those reported for terminal Fe nitrides (950–1150 ppm), all of which formally contain Fe4+ ions,[17] but the downfield chemical shifts of the nitrides likely result from temperature-independent paramagnetism, thereby complicating the assignment of trends in these data to changes in the physical oxidation state of the nitride.[7b] The IR spectrum of [Fe2N]+ reveals an Fe–N–Fe stretching mode at 991 cm−1 (965 cm−1 for [Fe2N*]+). These values are comparable to the handful of values that have been reported in the literature for terminal and bridging nitrides on late transition metals (between 910 to 1000 cm−1).[5b,18]

Initial investigations into the reactivity of [Fe2N]+ have revealed diverse reaction chemistry toward formation of N–H bonds. Monitoring the treatment of [Fe2N]+ with 4.0 equiv of Ph2PH in PhF at 60 °C revealed clean conversion (>85 %) to a diamagnetic bridging amide species, [(3PDI2)Fe2(μ-NH2)(PMe3)2][OTf] ([Fe2NH2]+, Scheme 2), along with the formation of 1.0 equiv of Ph2P–PPh2 over 4 h. This is consistent with the PCET of two equivalents of hydrogen atoms to μ-N. The resulting μ-NH2 protons were observed at 6.25 ppm by 1H NMR spectroscopy in C6D6 (5.37 ppm in CD3CN), a signal that significantly decreased in intensity with the use of Ph2PD (95 % D) as a reagent. A 2H NMR spectrum of the deuterated product confirmed deuteration at the bridging amide and gave no indication of deuteration at other positions within the molecule. Only a single N–H stretching feature was observed in the IR spectrum of [Fe2NH2]+.[19] This signal at 3334 cm−1 shifted to 2448 cm−1 for [Fe2ND2]+, and the incomplete deuteration of the sample ([Fe2NHD]+) fortuitously allowed for the observation of signals that result from DN–H and HN–D stretches at 3311 and 2447 cm−1, respectively (see Supporting Information). The identity of this product was further supported by independent synthesis via treatment of [Fe2Cl]+ with 1.5 equiv of NaNH2 to give a product with identical analytical data in 69 % yield.

A crystal structure of [Fe2NH2]+ revealed that the ligand had refolded to form a geometry similar to that of [Fe2Cl]+ (Figure 1).[13] The ligand isomerization between [Fe2Cl]+, [Fe2N]+, and [Fe2NH2]+ supports the view that the 3PDI2 ligand is able to adapt to the stereoelectronic requirements of the cluster core, from an unfolded geometry suitable for supporting an sp-hybridized μ-nitride to a folded ligand geometry that accommodates sp3-hybridized fragments like a μ-amide. Beyond this geometric change, we were interested to find that the the 3PDI2 ligand retains three electron’s worth of electron density (Δ = 0.087 Å, see Supporting Information),[20] and the Fe–NPDI (avg. Fe–Nim = 1.93 Å; avg. Fe–Npy = 1.82 Å) and Fe–P (avg. 2.26 Å) bond lengths are virtually identical to those for [Fe2N]+ (see above; avg. Fe–P = 2.25 Å).

Other main group hydrides – PhSiH3 and HBpin (pin = pinacolate) – were also found to react with the μ-nitride of [Fe2N]+. Both formed amide products, but in these cases, the amides resulted from insertion of the bridging nitride into the E-H bond (E = Si, B; Scheme 2).[21] Treatment of [Fe2N]+ with 1.0 eq. of either reagent in fluorobenzene at 80 °C cleanly generated the diiron silylamide [(3PDI2)Fe2(μ-N(H)SiH2Ph)(PMe3)2][OTf] ([Fe2NHSi]+) and borylamide [(3PDI2)Fe2(μ-NHBpin)(PMe3)2][OTf] ([Fe2NHB]+) complexes, respectively. The latter reaction is of particular interest given the high B–H BDE of ca. 110 kcal/mol.[22] As with the parent amide [Fe2NH2]+, the crystal structures of [Fe2NHSi]+ and [Fe2NHB]+ revealed folded ligand geometries (Figure 1), with apparent Fe–Fe bonding interactions (dFe–Fe = ca. 2.58 Å, see Supporting Information) and sp3-hybridized μ-NHE residues.[13] The silyl- and borylamide products were found to be diamagnetic in solution, as observed for both [Fe2NH2]+ and the isoelectronic, folded-ligand complex [Fe2Cl]+. The μ-NHR complexes were also found to exhibit Cs symmetry in solution, indicating that the folded ligand geometries observed in the solid state persist in solution. For [Fe2NHSi]+, the SiH2 and μ-NH protons were observed at 6.46 ppm (2H) and 3.23 ppm (1H) by 1H NMR spectroscopy, and for [Fe2NHB]+, the μ-NH proton was located at 3.93 ppm (1H). A very weak N–H stretch was observed in the IR spectrum of [Fe2NHSi]+ at 3274 cm−1, while [Fe2NHB]+ provided a more intense VN–H at 3328 cm−1. The Si–H stretching frequencies in [Fe2NHSi]+ were found at 2091 and 2125 cm−1, for the symmetric and asymmetric SiH2 vibrational modes.

Zero-field Mössbauer spectroscopic data were collected on [Fe2NH2]+, [Fe2Cl]+, and [Fe2N]+ at room temperature (Figure 2). All three spectra are best modelled by fitting the data with two doublets of equal intensity. This is consistent with the crystallographic data that showed the two iron centers within a molecule to be in comparable, albeit distinct, chemical environments. The data for [Fe2NH2]+ revealed doublets at δ = +0.10±0.03 and +0.22±0.03 mm/s, and those for [Fe2Cl]+ were found at +0.17±0.03 and +0.32±0.03 mm/s (see Supporting Information). Mononuclear (PDI)Fe species with intermediate-spin Fe2+ ions typically show δ > +0.30 mm/s when bound by ligands that undergo negligible π-backbonding,[23] suggesting that [Fe2NH2]+ and [Fe2Cl]+ may contain less electron density at the metal centers compared to an Fe2+ ion. This would be consistent with the electronic structure of [Fe2Cl]+ that was described previously as containing an (Fe2)5+ core with intermediate-spin Fe centers.[12] The Mössbauer data for [Fe2N]+ yielded doublets at δ = −0.15±0.03, −0.04±0.03 mm/s. The more negative isomer shifts for [Fe2N]+ compared to [Fe2NH2]+ could result from several factors, including an increase in oxidation state or a decrease in spin configuration. A change from intermediate- to low-spin configurations is known to decrease the value of δ for a single oxidation state by ca. 0.2–0.4 mm/s,[23] which is worth noting given the similarity in both the Δ values and the Fe–NPDI and Fe–P bond distances between the two structures.

Figure 2.

Figure 2.

Mössbauer spectroscopic data for [Fe2NH2]+, [Fe2Cl]+, and [Fe2N]+; experimental (●) and theoretical (solid lines).

Finally, it was found that treatment of [Fe2N]+ with 3.0 equiv of PhSH results in complete removal of the bridging nitrogen, forming NH3 in 71% isolated yield (Scheme 3). Limited examples are known of molecular transition-metal nitrides forming free NH3 following PCET reactions.[17, 24] The transition metal product in the present case is the previously described tris(thiophenolate) [(3PDI2)Fe2(μ-SPh)(SPh)2][OTf] ([Fe2(SPh)3]+),[12] which forms in 56% yield following crystallization. The oxidation of the (3PDI2)3− ligand in [Fe2N]+ to (3PDI2)0 in [Fe2(SPh)3]+ during this reaction indicates that intramolecular electron transfer occurs in addition to the formation of N–H bonds. Like Ph2PH (BDEP–H = ~61 kcal/mol; see Supporting Information), PhSH (BDES–H = 79 kcal/mol) is known to act as a hydrogen atom donor,[25] but treatment of [Fe2N]+ with non-Lewis basic H-atom donors like 1,4-cyclohexadiene or 9,10-dihydroanthracene (~76–80 kcal/mol) did not form [Fe2NH2]+.[26] This suggests either that coordination by the phosphine/thiol may precede PCET or that the PCET process may exhibit significant proton transfer character at the outset of the reaction.[27] Once the bridging amide is formed, a third equivalent of PhSH (pKa = 8.9 in MeCN) is able to effect the formation of NH3 via protonolysis.[28] This interpretation is supported by protonolysis of the phosphine-bound [Fe2NH2]+ with one equivalent of PhSH, which generates [(3PDI2)Fe2(μ-SPh)(PMe3)2][OTf] ([Fe2(SPh)]+; Scheme 3) in 62% isolated yield and NH3 in 89% isolated yield.

Scheme 3.

Scheme 3.

Formation of ammonia via N-H bond formation at diiron bridging nitride or amide complexes.

In conclusion, we have shown that the geometric flexibility of the 3PDI2 ligand supports the formation of a linear diiron bridging nitride. N–H bond formation via proton-coupled electron transfer and E–H bond insertion chemistry were both consistent with considerable electrophilic character of the bridging nitride. The formation of free NH3 was accomplished via treatment with PhSH, suggesting that both PCET and protonolysis contribute to the removal of the bridging nitrogen from the cluster core. Work is underway to determine the mechanism by which the μ-N in [Fe2N]+ performs PCET and E–H bond insertion chemistry, as we investigate the manner in which the unique electronic properties of the 3PDI2 ligand impart electrophilic character to [Fe2N]+.

Supplementary Material

Supporting Information

Acknowledgements

We thank Dr. Arthur Viescas for providing technical support on Mössbauer spectroscopy. We thank the University of Pennsylvania, the donors of the American Chemical Society Petroleum Research Fund (57346-DNI3), and the National Institutes of Health (R35GM128794) for support of this research.

References

  • [1].a) Strongin DR, Carrazza J, Bare SR, Somorjai GA, J. Catal 1987, 103, 213–215; [Google Scholar]; b) Falicov LM, Somorjai GA, Proc. Natl. Acad. Sci 1985, 82, 2207–2211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Schlögl R, in Handbook of Heterogeneous Catalysis (Eds.: Ertl G, Knözinger H, Schüth F, Weitkamp J), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2008, pp. 2501–2575. [Google Scholar]
  • [3].a) Navio C, Alvarez J, Capitan MJ, Ecija D, Gallego JM, Yndurain F, Miranda R, Phys. Rev. B 2007, 75, 125422; [Google Scholar]; b) Chen J-S, Yu C, Lu H, J. Alloys Compd 2015, 625, 224. [Google Scholar]
  • [4].a) Berry JF, Comments Inorg. Chem 2009, 30, 28–66; [Google Scholar]; b) Smith JM, in Progress in Inorganic Chemistry , Vol. 58 (Ed.: Karlin KD), John Wiley & Sons, Inc., Hoboken, 2014, pp. 417–470. [Google Scholar]
  • [5].a) Hohenberger J, Ray K, Meyer K, Nat. Commun 2012, 3, 720; [DOI] [PubMed] [Google Scholar]; b) Smith JM, Subedi D, Dalton Trans. 2012, 41, 1423–1429; [DOI] [PubMed] [Google Scholar]; c) Afanasiev P, Sorokin AB, Acc. Chem. Res 2016, 49, 583–593; [DOI] [PubMed] [Google Scholar]; d) Reiners M, Maekawa M, Daniliuc CG, Freytag M, Jones PG, White PS, Hohenberger J, Sutter J, Meyer K, Maron L, Walter MD, Chem. Sci 2017, 8, 4108–4122; [DOI] [PMC free article] [PubMed] [Google Scholar]; e) Jüstel T, Weyhermüller T, Wieghardt K, Bill E, Lengen M, Trautwein AX, Hildebrandt P, Angew. Chem 1995, 107, 744–747; Angew. Chem. Int. Ed. Engl. 1995, 34, 667–669; [Google Scholar]; f) Ercolani C, Jubb J, Pennesi G, Russo U, Trigiante G, Inorg. Chem 1995, 34, 2535–2541; [Google Scholar]; g) Summerville DA, Cohen IA, J. Am. Chem. Soc 1976, 98, 1747–1752. [DOI] [PubMed] [Google Scholar]
  • [6]. For electrophilic nitridyl complexes, see:; a) Suarez AIO, Lyaskovskyy V, Reek JNH, van der Vlugt JI, de Bruin B, Angew. Chem 2013, 125, 12740–12760; Angew. Chem. Int. Ed. 2013, 52, 12510–12529; [DOI] [PubMed] [Google Scholar]; b) Scheibel MG, Askevold B, Heinemann FW, Reijerse EJ, de Bruin B, Schneider S, Nat. Chem 2012, 4, 552–558; [DOI] [PubMed] [Google Scholar]; c) Gloaguen Y, Rebreyend C, Lutz M, Kumar P, Huber M, van der Vlugt JI, Schneider S, de Bruin B, Angew. Chem 2014, 126, 6932–6936; Angew. Chem. Int. Ed. 2014, 53, 6814–6818; [DOI] [PubMed] [Google Scholar]; d) Tran BL, Washington MP, Henckel DA, Gao X, Park H, Pink M, Mindiola DJ, Chem. Commun 2012, 48, 1529–1531; [DOI] [PubMed] [Google Scholar]; e) Scheibel MG, Wu Y, Stückl AC, Krause L, Carl E, Stalke D, de Bruin B, Schneider S, J. Am. Chem. Soc 2013, 135, 17719–17722; [DOI] [PubMed] [Google Scholar]; f) Rebreyend C, Mouarrawis V, Siegler MA, van der Vlugt JI, de Bruin B, Eur. J. Inorg. Chem 2019, 4249–4255; [Google Scholar]; g) Ghannam J, Sun Z, Cundari TR, Zeller M, Lugosan A, Stanek CM, Lee W-T, Inorg. Chem 2019, 58, 7131–7135; [DOI] [PubMed] [Google Scholar]
  • [7].a) Brown SD, Mehn MP, Peters JC, J. Am. Chem. Soc 2005, 127, 13146–13147; [DOI] [PubMed] [Google Scholar]; b) Brown SD, Peters JC, J. Am. Chem. Soc 2005, 127, 1913–1923. [DOI] [PubMed] [Google Scholar]
  • [8].MacLeod KC, Menges FS, McWilliams SF, Craig SM, Mercado BQ, Johnson MA, Holland PL, J. Am. Chem. Soc 2016, 138, 11185–11191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].MacLeod KC, McWilliams SF, Mercado BQ, Holland PL, Chem. Sci 2016, 7, 5736–5746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Cui P, Wang Q, McCollom SP, Manor BC, Carroll PJ, Tomson NC, Angew. Chem 2017, 129, 16195; Angew. Chem. Int. Ed. 2017, 56, 15979–15983. [DOI] [PubMed] [Google Scholar]
  • [11].Sengupta D, Sandoval-Pauker C, Schueller EC, Encerrado-Manriquez AM, J Metta-Magana A, Lee W-Y, Seshadri R, Pintér B, Fortier S. J. Am. Chem. Soc 2020, doi: 10.1021/jacs.0c00291. [DOI] [PubMed] [Google Scholar]
  • [12].Zhang S, Wang Q, Thierer LM, Weberg AB, Gau MR, Carroll PJ, Tomson NC, Inorg. Chem 2019, 58, 12234–12244. [DOI] [PubMed] [Google Scholar]
  • [13].Wang Q, Zhang S, Cui P, Weberg AB, Thierer LM, Manor BC, Gau MR, Carroll PJ, Tomson NC, Inorg. Chem 2019, doi: 10.1021/acs.inorgchem.9b02339. [DOI] [PubMed] [Google Scholar]
  • [14].a) Andersen RA, Faegri K, Green JC, Haaland A, Lappert MF, Leung WP, Rypdal K, Inorg. Chem 1988, 27, 1782–1786; [Google Scholar]; b) King ER, Hennessy ET, Betley TA, J. Am. Chem. Soc 2011, 133, 4917–4923. [DOI] [PubMed] [Google Scholar]
  • [15].a) Bouwkamp MW, Bowman AC, Lobkovsky E, Chirik PJ, J. Am. Chem. Soc 2006, 128, 13340–13341; [DOI] [PubMed] [Google Scholar]; b) Bart SC, Chłopek K, Bill E, Bouwkamp MW, Lobkovsky E, Neese F, Wieghardt K, Chirik PJ, J. Am. Chem. Soc 2006, 128, 13901–13912. [DOI] [PubMed] [Google Scholar]
  • [16]. For detail on the derivation of Δ, see:; Römelt C, Weyhermüller T, Wieghardt K, Coord. Chem. Rev 2019, 380, 287–317. [Google Scholar]
  • [17].Betley TA, Peters JC, J. Am. Chem. Soc 2004, 126, 6252–6254. [DOI] [PubMed] [Google Scholar]
  • [18].a) Jüstel T, Müller M, Weyhermüller T, Kressl C, Bill E, Hildebrandt P, Lengen M, Grodzicki M, Trautwein AX, Nuber B, Wieghardt K, Chem.: Eur. J 1999, 5, 793; [Google Scholar]; b) Schöffel J, Rogachev AY, DeBeer George S, Burger P, Angew. Chem 2009, 121, 4828–4832; Angew. Chem. Int. Ed. 2009, 48, 4734–4738. [DOI] [PubMed] [Google Scholar]
  • [19].a) Yang D, Xu S, Zhang Y, Li Y, Li Y, Wang B, Qu J, Inorg. Chem 2018, 57, 15198–15204; [DOI] [PubMed] [Google Scholar]; b) Schollhammer P, Pétillon P,FY, Poder-Guillou S, Saillard JY, Talarmin J, Muir KW, Chem. Commun 1996, 2633–2634. [Google Scholar]
  • [20].Römelt C, Weyhermüller T, Wieghardt K, Coord. Chem. Rev 2019, 380, 287–317. [Google Scholar]
  • [21].a) Thompson R, Tran BL, Ghosh S, Chen C-H, Pink M, Gao X, Carroll PJ, Baik M-H, Mindiola DJ, Inorg. Chem 2015, 54, 3068–3077; [DOI] [PubMed] [Google Scholar]; b) Sieh D, Burger P, J. Am. Chem. Soc 2013, 135, 3971; [DOI] [PubMed] [Google Scholar]; c) Maity AK, Murillo J, Metta-Magaña AJ, Pinter B, Fortier S, J. Am. Chem. Soc 2017, 139, 15691–15700. [DOI] [PubMed] [Google Scholar]
  • [22].a) Sun Z, Hull OA, Cundari TR, Inorg. Chem 2018, 57, 6807–6815; [DOI] [PubMed] [Google Scholar]; b) Procacci B, Jiao Y, Evans ME, Jones WD, Perutz RN, Whitwood AC, J. Am. Chem. Soc 2015, 137, 1258–1272. [DOI] [PubMed] [Google Scholar]
  • [23].Bart SC, Lobkovsky E, Bill E, Wieghardt K, Chirik PJ, Inorg. Chem 2007, 46, 7055–7063. [DOI] [PubMed] [Google Scholar]
  • [24].a) Scepaniak JJ, Young JA, Bontchev RP, Smith JM, Angew. Chem. Int. Ed 2009, 48, 3158–3160; [DOI] [PubMed] [Google Scholar]; b) Scepaniak JJ, Vogel CS, Khusniyarov MM, Heinemann FW, Meyer K, Smith JM, Science 2011, 331, 1049–1052. [DOI] [PubMed] [Google Scholar]
  • [25].a) Dénès F, Pichowicz M, Povie G, Renaud P, Chem. Rev 2014, 114, 2587–2693; [DOI] [PubMed] [Google Scholar]; b) Miller DC, Choi GJ, Orbe HS, Knowles RR, J. Am. Chem. Soc 2015, 137, 13492–13495; [DOI] [PMC free article] [PubMed] [Google Scholar]; c) Darcy JW, Koronkiewicz B, Parada GA, Mayer JM, Acc. Chem. Res 2018, 51, 2391–2399; [DOI] [PMC free article] [PubMed] [Google Scholar]; d) Saouma CT, Morris WD, Darcy JW, Mayer JM, Chem.: Eur. J 2015, 21, 9256–9260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Klein JEMN, Dereli B, Que L, Cramer CJ, Chem. Commun 2016, 52, 10509–10512. [DOI] [PubMed] [Google Scholar]
  • [27].Bezdek MJ, Guo S, Chirik PJ, Science 2016, 354, 730–733. [DOI] [PubMed] [Google Scholar]
  • [28].Nag S, Datta D, Indian. J. Chem. A 2007, 46, 1263–1265. [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Information

RESOURCES